Open Access
14 December 2023 Status report on emerging photovoltaics
Annick Anctil, Meghan N. Beattie, Christopher Case, Aditya Chaudhary, Benjamin D. Chrysler, Michael G. Debije, Stephanie Essig, David K. Ferry, Vivian E. Ferry, Marina Freitag, Isaac Gould, Karin Hinzer, Harald Hoppe, Olle Inganäs, Lethy Krishnan Jagadamma, Min Hun Jee, Raymond K. Kostuk, Daniel Kirk, Stephan Kube, Minyoung Lim, Joseph M. Luther, Lorelle Mansfield, Michael D. McGehee, Duong Nguyen Minh, Preeti Nain, Matthew O. Reese, Angèle Reinders, Ifor D. W. Samuel, Wilfried van Sark, Hele Savin, Ian R. Sellers, Sean E. Shaheen, Zheng Tang, Fatima Toor, Ville Vähänissi, Ella Wassweiler, Emily L. Warren, Vincent R. Whiteside, Han Young Woo, Gang Xiong, Xitong Zhu
Author Affiliations +
Abstract

This report provides a snapshot of emerging photovoltaic (PV) technologies. It consists of concise contributions from experts in a wide range of fields including silicon, thin film, III-V, perovskite, organic, and dye-sensitized PVs. Strategies for exceeding the detailed balance limit and for light managing are presented, followed by a section detailing key applications and commercialization pathways. A section on sustainability then discusses the need for minimization of the environmental footprint in PV manufacturing and recycling. The report concludes with a perspective based on broad survey questions presented to the contributing authors regarding the needs and future evolution of PV.

1.

Introduction

The field of photovoltaic (PV) solar power generation has grown to become a significant component of the global energy landscape, with 4.5% of the world’s electricity being generated and 240  GW of new installations in 2022.1,2 However, further acceleration of PV and other renewable energy sources is urgently required to mitigate the impacts of global greenhouse gas emissions. A recent analysis has concluded that PV needs to grow at 25% annually with a target of 75 TW of global installations by 2050, a 75x increase from current installed capacity.3 While the existing PV landscape is largely dominated by silicon (crystalline and polycrystalline) and CdTe, achieving these long-term goals can be greatly aided by the development of new materials, device concepts, and light management strategies that enable higher efficiencies and more scalable and sustainable manufacturing.

This article is intended to provide a snapshot of the current status of emerging PV approaches that show potential in helping to achieve the above goals. It is intended to be a convenient resource for people within and outside the field, including new researchers, students, technology managers, and program managers, who can play a role in accelerating the global effort. The article is structured in sections covering silicon (Sec. 2), thin film (Sec. 3), III-V (Sec. 4), perovskite (Sec. 5), organic (Sec. 6), and dye-sensitized solar cells (Sec. 7). Each section provides background, a technology status update, and challenges towards commercialization/scalability. Subsequent sections provide an overview of the applications and commercialization of emerging PV (Sec. 8), strategies for exceeding the detailed balance limit (Sec. 9), and concepts in light management (Sec. 10). A final section describes sustainability and environmental impact issues that apply to all the above technologies (Sec. 11). The article concludes with a perspectives section (Sec. 12) that first discusses common themes that appear throughout the article, and then also presents and draws conclusions from a survey of emerging PV, which was completed anonymously by contributing authors.

2.

Silicon Photovoltaics

2.1.

Bifacial Silicon PV

Section Author: Fatima Toor (University of Iowa)

In recent years, bifacial silicon (Si) photovoltaics (PV) technology has been growing in commercial PV systems because it results in higher performance yield and lower levelized cost of energy (LCOE) compared with conventional monofacial PV technology.4 Figure 1 illustrates the difference between monofacial and bifacial solar cell architectures. While traditional monofacial Si cells have an aluminum (Al)-based back surface field (BSF) that covers the entire backside of the cell precluding any light absorption on the backside, bifacial Si cells have point contacts that allow for backside light collection and absorption in the cells.

Fig. 1

Schematic of (a) monofacial and (b) bifacial solar cells.

JPE_13_4_042301_f001.png

The latest International Technology Roadmap for Photovoltaic (ITRPV) report5 suggests that Al BSF cell concept is expected to be phased out within 2023. The report predicts that mature concepts of diffused and passivated pn junction Si solar cells will be further used in the mainstream commercial PV panels with different rear side passivation technologies, such as passivated emitter and rear cell (PERC),6 passivated emitter, rear locally-diffused (PERL),7 passivated emitter, rear totally diffused (PERT),8 and tunnel oxide passivated contact (TOPCon).9 The ITRPV report suggests that in 2022 the market share of PERC p-type mono-Si was 80%, however the share of p-type mono-Si PERC will decrease to about 10% within the next ten years since upcoming, promising new cell technologies are using n-type material. TOPCon on n-material, using tunnel oxide passivation stacks at the rear side, will gain market share from about 10% in 2022 up to 60% within the next ten years. Based on the report’s analysis, TOPCon on n-type is expected to become the dominant cell concept after 2025.

All these cell architectures (PERC/PERL/PERT/TOPCon) support bifacial PV cell and module architecture. Therefore, not surprisingly, given the trends in advanced Si solar cell architectures, the ITRPV report predicts that by 2033, bifacial cells will make up 90% of the Si PV installations. Another notable trend presented in the ITRPV report is that, while in 2023, about 65% of modules are monofacial modules, the share of bifacial modules will grow to about 70% within the next years as shown in Fig. 2. This is because bifacial Si solar cells can be used in bifacial modules as well as in conventional, monofacial modules.

Fig. 2

ITRPV 2022-2033 predicted trends for monofacial and bifacial modules with bifacial Si solar cells.

JPE_13_4_042301_f002.png

The bifacial PV project supported by the US Department of Energy (DOE)10 performed and published extensive optical modeling based on RADIANCE11 to determine the performance gains of bifacial modules installed at various tilt angles, with different row spacing, and ground albedo.1220 The team showed the effect of installation parameters on bifacial gain in energy (BGE) defined as: BGE=(Eb/Em)1, where Eb and Em are the energy yield of the bifacial and monofacial module, respectively. This value shows the energy gain in using bifacial PV system over the equivalent monofacial system (with the same installation parameters). The study’s results indicate that bifacial PV arrays generate lower energy than expected due to horizon blocking and large shadowing area cast by the modules on the ground. For albedo of 21%, the center module in a large array generates up to 7% less energy than a single bifacial module – this single bifacial module is typically utilized for lab certification of BGE of bifacial modules. The team also proposed draft recommendations for IEC 60904 to include bifacial PV module testing so that bifacial PV gain measurements can be standardized.21 The standard has since been formalized and adopted by the international community as IEC TS 60904-1-2:2019. As of 2023, the standard has gone through a further revision with added modifications for dual-simulator measurements to be re-published by the end of the year.

2.2.

Black Silicon

Section Authors: Ville Vähänissi (Aalto University) and Hele Savin (Aalto University)

Nanostructured silicon, often called black silicon due to its dark appearance, eliminates surface reflectance over a wide wavelength range [Fig. 3(a)] due to gradual change in refractive index2325 and increases the absorption of long wavelengths due to extended optical path.2629 Therefore, black silicon is a promising alternative to minimize optical losses in PV applications. However, due to the rough nanoscale morphology, electrical passivation of such surfaces has been considered challenging. In 2012 this problem was tackled with atomic layer deposited Al2O3 that provides a conformal coating [Fig. 3(b)] combined with a high-quality interface with silicon.30,31 Indeed, it was demonstrated that surface passivation was equally good in black silicon as in the planar counterparts.32,33 The successful surface passivation of black silicon was also proven later in actual solar cells (IBC) resulting in an external quantum efficiency (EQE) close to unity in a broad wavelength range (300 nm-1000 nm) and a conversion efficiency well above 20%.34

Fig. 3

(a) Reflectance of black silicon and acidic-textured silicon with SiNx ARC as a function of wavelength.22 (b) Scanning electron microscope image of black silicon surface coated with a 50 nm atomic layer deposited Al2O3 surface passivation layer using Beneq TFS-500 system.

JPE_13_4_042301_f003.png

In PERC and TOPCON cells (or any front contact cell type) emitter formation brings an additional challenge as it needs to be made on the nanostructured surface. The dopant profile is difficult to control via a conventional diffusion process. Consequently, dopant diffusion on enhanced surface areas often results in too heavy Auger and SRH recombination inside the nanostructures.3542 This problem has recently been overcome via a more controlled doping method, i.e. using ion-implantation, which has resulted in emitter saturation currents less than 20  fA/cm2 43 combined with EQEs even above 100% in UV.44

While the original motivation to use black silicon in PV has been the improved optics, other benefits have been discovered later on. Firstly, the nanotexturing does not seem to have any substrate crystallinity related limitations,33 e.g., it works well also for the diamond-wire cut multicrystalline wafers that are difficult to texturize with conventional acidic solutions. This benefit is valid for all commonly known black silicon fabrication methods, i.e., dry etching,4547 metal assisted chemical etching,4855 and laser processing.5661 Another benefit is related to metal impurity gettering, i.e., black silicon has been shown to have a higher gettering efficiency for iron as compared to planar surfaces.62 The same phenomenon is expected to take place for other mobile impurities as well. Quite surprisingly, black Si can also reduce the harmful light-induced degradation although the exact mechanism is still under debate.63,64

While the aforementioned fundamental limitations regarding integration of black silicon into solar cells have been overcome in the laboratory environment, industrial viability raises other challenges. These include possibly harsh handling of the nanostructured surfaces in the production line, potential contact resistance issues when using nonconformal screen printing, as well as preserving excellent optics during module encapsulation. There are preliminary results showing that all these issues can be tackled efficiently in industrial production (Fig. 4).22 However, more extensive cost and lifecycle analysis would be required for black silicon to become the mainstream technology in PV. Fortunately, there are pioneers in the field who are already mass-producing black silicon solar cells. At the moment, metal assisted chemical etching seems to be the most viable fabrication method for the commercial cells as it requires fewer changes to the existing manufacturing lines.

Fig. 4

Dry-etched black silicon IBC solar panel without any antireflection coating fabricated in Naps Solar Systems, Finland. The energy conversion efficiency of 18.1% and 19.9% of the whole IBC module and the best individual cell of the module, respectively, demonstrate that fragile nanostructures can withstand standard module fabrication steps.65

JPE_13_4_042301_f004.png

3.

Thin Film PV

3.1.

Cu(In,Ga)Se2 PV: Fundamentals and Efficiency Limits

Section Author: Lorelle Mansfield (National Renewable Energy Laboratory and RASEI)

Cu(In,Ga)Se2 (CIGS) solar cells have the highest record efficiency among commercialized thin-film PV technologies at 23.6%.66 CIGS has the benefits of thin films,67 such as low materials costs, low embodied carbon, and varied form factors. The substrate design of most cells (Fig. 5) allows fabrication on flexible plastics and metal foils, which leads to a multitude of potential applications. It is suitable for architectural solar, military deployment, space power, and consumer applications. Flexible substrates also allow for roll-to-roll fabrication enabling rapid manufacturing of high-efficiency devices.

Fig. 5

Cross-section of a typical CIGS-based device.

JPE_13_4_042301_f005.png

CIGS technology has experienced several technological advances in recent history. One of the most impactful has been post-deposition treatments (PDT) with alkali metal fluorides, such as potassium fluoride (KF), rubidium fluoride (RbF), and cesium fluoride (CsF) which have significantly improved open-circuit voltage (VOC) and efficiency of devices, even those fabricated at lower temperatures. Lower temperature processing can be further enhanced through alloying CIGS with small amounts of silver (Ag), which also results in smoother surfaces, improved crystallinity, and increased carrier concentrations. Traditionally, chemical-bath deposited cadmium sulfide (CdS) was used as a buffer layer, whereas now zinc oxysulfide (Zn(O,S)) and other more transparent and more manufacturable (e.g. sputtered) layers are being employed. Research directions being explored include CIGS for tandem solar cells. Bandgap variations are key to efficient tandems, and the CIGS-based material system can be tuned from 1 eV to around 2.5 eV by alloying in the form of (Ag,Cu)(In,Ga)(Se,S)2 (Fig. 6). Existing high-efficiency low-bandgap CIGS cells are excellent bottom cells for perovskite and CdTe top cells, and improved wide-bandgap CIGS top cells could even be paired with Si bottom cells.

Fig. 6

(Ag,Cu)(In,Ga)(Se,S)2 material system with bandgaps of ternary compounds.

JPE_13_4_042301_f006.png

Although CIGS-based PV are efficient and stable, several challenges have delayed their ubiquitous adoption. As could be expected, initial CIGS installations were not as robust as subsequent generations, which led to an incorrect assumption that CIGS products were inherently inferior solar cells. Unfortunately, some potential investors still hold those views, and CIGS research and development funding has decreased compared to other thin-film technologies, particularly in the US. To date, CIGS technology has suffered from a greater gap in efficiency between record cells (23.6%) and record modules (19.2%) than other commercialized PV technologies. Closing this gap could be accomplished through using wider bandgap absorbers, designing new ways to assemble cells into modules, and developing better transparent conductive contacts.68 Unlike silicon technology, CIGS has yet to experience the full benefit of economies of scale. With the complexity of processing required to fabricate the absorber material, most companies have proprietary manufacturing equipment. That relegates the technology to smaller markets where premium prices are accepted for products with specific requirements such as flexibility and high specific power. Lower-cost processing methods and equipment designs have the potential to revolutionize the industry and bring CIGS to multi-GW scale.

3.2.

CdTe PV: Fundamentals and Efficiency Limits

Section Author: Matthew O. Reese (National Renewable Energy Laboratory)

CdTe is the most scaled thin-film PV technology. This is enabled by low material costs, high-throughput deposition, robust process space, low degradation, and low embodied energy. While CdTe is directly competing at utility-scale with Si, it has done so with relatively low demonstrated cell record efficiency, having only achieved 2/3 of its detailed balance limit. The current collection has been well-engineered, such that it is near the theoretical limit. Further monofacial improvements will largely stem from advances enabling improved photovoltage.

CdTe devices have eliminated the parasitically-absorbing CdS emitter with a deleterious “cliff” in its band conduction band offset (CBO), replacing it with a wider bandgap oxide, such as the tunable MgxZn1xO enabling a small “spike” in CBO with reduced recombination.69 Furthermore, the absorber transitioned from the congruently-sublimating, binary CdTe absorber to a graded ternary Cd(Se,Te). This has changed the effective device bandgap from 1.5 eV in pure CdTe (which is present now at the back) to 1.4  eV at the front. This increases the theoretical and effective current density in devices [Fig. 7(a)], while raising the overall theoretical efficiency limit by 1% absolute. These changes have dramatically improved achievable minority carrier lifetimes from a few ns to >100  ns.71,72 Due to this increase in lifetime, device photovoltage has seen a modest increase even though the bandgap has decreased 100  mV [Fig. 7(b)].

Fig. 7

(a) CdTe PV certified record cell current densities and (b) open-circuit voltages, relative to the detailed balance limit for the bandgap of CdTe (1.5 eV) and a bandgap of 1.4 eV, which is typical of Se-alloyed material. Data from the semi-annual publication up to Version 61 of “Solar cell efficiency tables” (Ref. 70).

JPE_13_4_042301_f007.png

While CdTe has demonstrated scalability and low cost, it must demonstrate continued efficiency improvement. Over the past several years, work has revolved around changing the defect (doping) chemistry to achieve this. Initial demonstrations with single crystals demonstrated an improvement in stable carrier concentration of 100x using a group V (GrV) dopant to replace traditional Cu to explicitly improve VOC.73 While GrV dopants have improved stability over Cu due to much lower diffusion and a reduction in complexing with Cl,74,75 much improved VOC and efficiency over Cu-doped polycrystalline thin-film devices have yet to be realized. This is for a few reasons. First, the large increase in absorber carrier concentration collapses the space charge region, making devices more sensitive to front interface recombination.76 Second, long lifetime from Se incorporation makes devices more sensitive to back interface recombination. Third, GrV-doped polycrystalline absorbers, while capable of high carrier concentration, have low activation, leading to significant losses from voltage fluctuations.77 Increasing VOC requires simultaneous improvements on multiple fronts. However, in less than eight years, insight from single crystals is influencing GW-scale decisions. GrV-doped CdTe PV promises increased energy yield due to its increased stability over its Cu-doped predecessor, with demonstrated similar initial efficiencies that are poised to increase over the next few years as these challenges are overcome.

3.3.

CdTe PV: Commercial Perspective

Section Authors: Ella Wassweiler (First Solar) and Gang Xiong (First Solar)

CdTe is the most commercially successful thin film PV technology. In 2022, CdTe module production capacity has exceeded 9 GW annually and is expected to reach 21 GW by 2025. At the start of 2023, First Solar, the largest CdTe solar module manufacturer, reached a milestone of 50GW accumulated production. Thin film CdTe manufacturing has an inherently better cost structure compared to Si. In addition, CdTe has an energy yield advantage due to its lower temperature coefficient, better spectra response, and slower long term degradation rate. These advantages allow CdTe to compete with Si in utility-scale markets. Roughly 40% of the US utility-scale PV market was served by CdTe in 2019.

Currently the record CdTe solar cell efficiency is 22.3%.78 Module efficiency has exceeded 19%. The state-of-art CdTe solar cell structure is outlined in Fig. 8. The absorber typically consists of a graded CdSexTe1x alloy. A buffer layer such as Zn1xMgxO can be inserted between the transparent conducting oxide (TCO) and absorber. With a small positive conduction band offset between Zn1xMgxO and absorber, carrier recombination at the front interface can be suppressed. ZnTe is a good back contact to CdTe with relatively low barrier height (0.2  eV).79

Fig. 8

State-of-art CdTe solar cell structure.

JPE_13_4_042301_f008.png

The absorber is typically p-type, traditionally doped by Cu with carrier concentration between 1014 and 1015  cm3. Increasing carrier concentration is a major pathway to further boost CdTe solar cell efficiency.

Carrier concentrations greater than 1016  cm3 have been demonstrated with group-V dopants such as As or P. For example, 22% As-doped CdTe devices have been reported.80 Compared to Cu-doped devices, As doping has additional energy yield advantages including even lower long-term degradation rate and lower temperature coefficients.81 The paradigm shift from Cu to As-doping, provided a new platform and opportunities to further improve CdTe solar cell performance. In order to do so, it is suggested that the reduction of non-radiative recombination in the device stack, as well as reduction of potential fluctuation in the absorber are of critical importance.77,80

First Solar’s Series 7 module is 2.8  m2 in size and has total area efficiency of up to 19.3%. CdTe modules are fabricated using a fully integrated inline process, with glass-in and module-out within 4 hours. Compared to Si, CdTe solar module has lower energy payback time and lower environmental footprint.

Improving module efficiency is the most effective way to reduce the cost per watt for manufacturing and levelized cost of electricity of a PV system. Bifacial CdTe solar cells have also been demonstrated.82 There is great synergy between cell efficiency and bifaciality improvement, as both require further reduction of back contact recombination loss. With less back contact recombination, absorber thickness can also be effectively reduced.

4.

III-V Tandem PV

4.1.

III-V Tandem PV

Section Authors: Meghan N. Beattie (University of Ottawa) and Karin Hinzer (University of Ottawa)

Bandgap engineering of III-V materials and vertical integration of multiple subcells in tandem devices has resulted in the highest efficiency PV. III-V semiconductors have high absorption, and many compositions have direct bandgaps. Multiple pn junction subcells absorb a larger portion of the solar spectrum and reduce thermalization losses compared to single junction devices. The higher energy photons from the solar spectrum are absorbed by the largest bandgap subcells at the top of the tandem stack, while the lower energy photons are transmitted through the top subcells to be absorbed by the smaller bandgap subcells below. Figure 9 displays common approaches.

Fig. 9

Schematic showing four different design architectures for III-V tandem solar cells. ARC, antireflection coating; MM, metamorphic.

JPE_13_4_042301_f009.png

The standard approach uses lattice-matched materials for the entire device, minimizing dislocations in the crystal lattice. Lattice-matched subcells are interconnected using transparent tunnel diodes [Fig. 9(a)]. In this series-connected configuration, the device current is limited by the smallest subcell current. The current generated in each subcell is maximized by tuning layer thicknesses and bandgaps so that each subcell absorbs the same amount of light. The lattice-matching requirement leads to the use of a variety of materials from Ge for the bottom subcell to ternary and quaternary III-V alloys or nanostructures83,84 for upper layers.

With limited semiconductor substrates to choose from, the lattice-matched approach has a limited material design space, which caps the efficiency. Several advanced strategies have been developed to overcome the lattice-matching constraint, enabling the development of all-III-V and III-V hybrid tandems. One such approach is wafer bonding, wherein subcells are grown on different substrates and directly bonded without an intermediate layer8587 [Fig. 9(b)]. A four-junction GaInP/GaInAs//GaInAsP/GaInAs wafer bonded device demonstrated the highest recorded solar cell efficiency to date: 47.6% under an irradiance of 665 suns.70,86

Another approach uses a graded metamorphic buffer to increment the lattice constant from that of the substrate to that of the subcell material84,8890 [Fig. 9(c)]. Excellent crystal quality can be achieved for III-V semiconductors grown on metamorphic buffer layers. An example is this six-junction AlGaInP/AlGaAs/GaAs/GaInAs(3) inverted metamorphic cell with a peak efficiency of 47.1% under 143 suns.89,91

A third technique to overcome lattice-matching involves mechanical stacking of fabricated subcells grown on different substrates using a transparent adhesive to join the upper and lower subcells9294 [Fig. 9(d)]. This technique allows for electrical configurations with two, three, or four terminals, overcoming current-matching constraints. This approach was used to fabricate a four terminal, three-junction mechanically stacked GaInP/GaAs//Si device having 35.9% efficiency under 1-sun illumination.94,95

These architectures can be combined with subcell segmentation to further enhance efficiencies.96 This design technique enables near-perfect current-matching using non-ideal absorber bandgaps by dividing each subcell into semitransparent pn junction segments. It is used in vertically segmented III-V photonic power converters (PPCs) that convert monochromatic laser light to electricity. Multijunction III-V PPCs with up to 30 absorbing segments have been demonstrated with efficiencies up to 66%.97100

Due to their highest efficiencies, long-lifetimes, radiation hardness, and low resistances, III-V multijunction PV power satellites, unmanned vehicles and concentrated solar systems. Table 1 lists selected III-V tandem cells of note with corresponding efficiencies from Green et al.70

Table 1

Selected III-V tandem cells and efficiencies from solar cell efficiency tables (version 63).70

DesignEfficiency (%)Intensity (suns)CategoryGroup(s)
2J GaAsP/Si23.41MM, SiOSU/UNSW/SolAero88
3J GaInP/GaAs/Si25.91MM, SiFraunhofer ISE101
3J GaInP/AlGaAs/CIGS28.11MS, CIGSAIST/Fraunhofer ISE93
2J GaInP/GaAs32.81LMLG Electronics102
2J GaAs//Si (4-terminal)32.81MSNREL/CSEM/EPFL102
2J GaInP/mqw-GaAs32.91mqwNREL/UNSW103
3J GaInP/GaAs//Si35.91MS, SiNREL/CSEM/EPFL94
3J GaInP/GaInAsP//Si36.11WB, SiFraunhofer ISE/AMOLF85
3J GaInP/GaAs/GaInAs (ELO)37.81MM, ELOMicrolink104
6J AlGaInP/AlGaAs/GaAs/GaInAs(3)39.21MMNREL89
3J GaInP/mqw-GaAs/GaInAs39.51MM, mqwNREL84
2J GaInAsP/GaInAs35.538MMNREL102
3J GaInP/GaAs/GaInAs44.4302MMSharp105
4J GaInP/GaAs/GaInAs/GaInAs45.7234MMNREL106
6J AlGaInP/AlGaAs/GaAs/GaInAs(3)47.1143MMNREL89
4J GaInP/GaInAs//GaInAsP/GaInAs47.6665WBFraunhofer ISE86
LM, lattice-matched; WB, wafer-bonded; MM, metamorphic; MS, mechanical stack; Si, III-V/silicon tandem; CIGS, III-V/CIGS tandem; mqw, multiquantum well; ELO, epitaxial lift-off.

4.2.

III-V/Si Tandem Junction Solar Cells

Section Authors: Stephanie Essig (ipv, University of Stuttgart) and Emily L. Warren (National Renewable Energy Laboratory)

Research and development of III-V/Si tandem solar cells is motivated by the efficiency enhancement achievable with multijunction architectures and the low cost and current market dominance of Si. Theoretically, the detailed balance efficiency of Si-based two (2J) and three-junction (3J) solar cells reaches 45% and almost 50%, respectively.107 From a detailed balance perspective, the ideal III-V top cell material for a Si-based 2J device is GaInP with a bandgap of 1.8 eV and record 1-sun single-junction efficiency of 22.0%.70 Fabrication of III-V solar cells is based on epitaxy, yet the direct growth of III-Vs with suitable bandgap energies on Si is challenging due to a mismatch in lattice constant and thermal expansion.108 Therefore, high threading dislocation densities occur, causing significant non-radiative recombination losses in the top cell, which limit the tandem cell performance.90 To overcome this difficulty, alternative manufacturing methods, including mechanical stacking109112 and wafer-bonding,85,113,114 of the III-V and Si cells, have been investigated.

The efficiency of monolithic III-V/Si dual-junction solar cells fabricated by metal organic vapor phase epitaxy (MOCVD) or molecular beam epitaxy (MBE) on Si has so far been limited to 25.0%.115 This record device comprises a 1.7 eV wide-bandgap GaAsP top cell on a graded GaAsP buffer, that effectively reduces the dislocation density. In comparison, 3J cells realized by direct epitaxy on Si have reached higher efficiencies up to 25.9%,101 but have not surpassed the record values for Si single-junction devices. Other possible ways to reduce the dislocation density and cost of polished surfaces include use of nanopatterned substrates, and dislocation filtering approaches such as strained superlattices and thermal cycle annealing.116118

In contrast, mechanically stacked and wafer-bonded III-V/Si tandems push the efficiency of Si-based PV beyond 30%. Their few-μm thick III-V top cells are grown lattice-matched on GaAs parent wafers and transferred to independently fabricated Si bottom cells. Hence, both the Si and III-V cells are manufactured using the optimum processes. Subsequent integration by wafer-bonding requires extremely clean, mirror-polished surfaces to allow atomic bond formation across the heterogeneous semiconductor interface.119 Improved current-matching of the subcells has recently raised the record efficiency of wafer-bonded GaInP/GaInAsP/Si 3J solar cells to 35.9%.85 The comparably easier integration by mechanical stacking involves typically an adhesive or glue between the III-V and Si cell that can level out interface irregularities. When applying front and rear contacts on each subcell, the tandem device can also be operated in 4-Terminal (4T) configuration. Compared to the commonly used 2-Terminal configuration, 4T offer the advantage that top and bottom cells are operated independently. Hence no current-matching is needed, a wider range of top cell bandgap energies can be used, and the energy output depends less on the temperature and solar spectrum.120 Mechanically stacked 4T devices reached record efficiencies of 32.8% for 2J (GaAs/Si; see Fig. 10) and 35.9% for 3J (GaInP/GaAs/Si) using both-sides contacted Si heterojuction cells.94 Similar 3J efficiency of 35.4% was achieved when using a rear-contacted Si bottom cell.121 Devices have also been stacked using transparent conductive adhesives or metal nanoparticles, often with the goal of forming 3T tandems, and these devices have reached efficiencies of 27.3%7 (2J,TCA) and 30.8%122 (3J, Pd nanoparticles).

Fig. 10

External Quantum Efficiency and sketch of a 4T GaAs/Si tandem solar cell with 32.8% efficiency.94

JPE_13_4_042301_f010.png

Despite the high efficiencies and good scalability of III-V/Si multijunction solar cells, they will only be successful if their costs on a US$/Watt level will approach those of commercial Si single-junction technology.92 Main cost drivers are the MOCVD deposition of the III-V top cell, and the III-V growth substrate.123 Therefore, extensive research on cost reduction strategies like the use of hydride vapor phase epitaxy (HVPE)123 instead of MOCVD and advanced lift-off124 and substrate reuse125 approaches are needed.126

5.

Perovskite PV and Tandems

5.1.

Stability of Perovskite Solar Panels with Partial Shading

Section Authors: Isaac Gould (University of Colorado Boulder) and Michael D. McGehee (University of Colorado Boulder and RASEI)

In a solar panel, partial shading prevents only some of the cells from generating photocurrent. Since all cells must pass the same amount of current for the panel to generate power and the cells are in series with each other, the cells that are illuminated put the shaded cells into whatever reverse bias is needed for current matching to occur. Several research groups have tested perovskite cells in reverse bias and observed both short-term reversible and irreversible degradation problems.127131

Several interesting observations have been made on the reverse bias behavior of perovskite solar cells. When metal electrodes are used the current typically increases very suddenly and localized heating is observed with thermal imaging.127,132 This behavior is attributed to the formation of a shunt caused by the metal penetrating through the perovskite. The extreme heating at the shunt causes permanent damage to most layers of the cell. In contrast, this behavior is not seen when transparent conducting oxides and carbon electrodes are used. In these cells there is an exponential increase in current that starts typically between -2 V and -6 V when sweeping the voltage to increasing negative voltages.127131 This current is enabled by tunneling of holes from the electron transport layer due to strong band bending associated with mobile ions redistributing to screen the applied field. After passing current through a cell in reverse bias for as little as one minute, the efficiency can be reduced by more than 50 %.129,133 Much, but not all, of this efficiency recovers if the cell is exposed to one sun at a positive voltage for approximately 30 minutes.129,133,134 The rapid degradation most likely occurs because holes oxidize iodide on the octahedral corners to form smaller neutral iodine species that can move to interstitial sites, creating a vacancy in the process.135,136 With few electrons being present in reverse bias, the reverse reaction is slow. The iodine interstitials act as recombination centers, which dramatically reduces the efficiency of the cell. During the recovery process at positive voltage, the iodine interstitials are reduced and return to the octahedral corners. The recovery only occurs, however, if the iodine species stay in the perovskite film.129,137,138 For this reason, internal barriers that confine the iodine to the perovskite are critically important.

There has been some progress towards improving reverse bias stability by using carbon electrodes.

If strategies for preventing degradation in reverse bias are not found, then engineering solutions for preventing degradation due to partial shading will be needed. Wolf et al. have analyzed different panel designs and suggested that perovskite-silicon tandems can more readily be protected with bypass diodes because the high breakdown strength of silicon enables many tandems to be protected with one bypass diode.139 It is not clear how the current could be routed from the electrodes in a monolithically integrated panel (made with scribes) to a bypass diode on the side. The voltage drop would be quite large if current had to travel the length of a panel through a transparent conducting oxide. In contrast, it is easy to run current from silicon or perovskite-silicon tandem cells with metal ribbon to bypass diodes, which are typically incorporated into the junction box.

It is likely that bypass diodes can be used to protect perovskite-silicon tandems. Research will be needed to find ways to protect other types of perovskite solar cells to prevent degradation due to partial shading.

5.2.

Perovskite/Silicon Tandem Solar Cells

Section Authors: Aditya Chaudhary (ipv, University of Stuttgart) and Stephanie Essig (ipv, University of Stuttgart)

Since crystalline silicon (c-Si) solar cells approach their theoretical efficiency limit, there is an increasing interest in c-Si-based tandem PV. Stacking a wider bandgap metal-halide perovskite (ABX3) cell onto a c-Si cell offers the potential to reach conversion efficiency over 30%, and to reduce levelized costs of electricity.140,141 There are three architectures (Fig. 11) in which a tandem cell can be realized, i.e., 2-terminal (2T), 3-terminal (3T) and 4-terminal (4T). All these architectures have their own advantages as well as limitations. Owing to the easy industrial realization and lower requirements on the balance of system, 2T architecture is touted to be the most promising.142

Fig. 11

Comparison of perovskite/c-Si tandem solar cells with 2-,3- and 4-terminal configuration.

JPE_13_4_042301_f011.png

The first results of 2T perovskite/c-Si tandem cells, in which the top and bottom cells are connected in series, were published in 2015.143 From then on research has focused on the optimum device architecture, stability improvement and the defect-free large-area deposition of the top perovskite cell on c-Si.144146

To reach efficiencies over 30% with 2T tandem cells for operation temperatures between 20 and 100°C, a perovskite top cell with bandgap energy in the range of 1.65-1.73 eV is needed,147 which can be realized by varying the ABX3 composition.148 Furthermore, maximum energy yield requires that the perovskite cell coats the textured c-Si bottom cell conformally149,150 which poses a challenge for solution-based deposition methods of the perovskite layers. In recent time, strategies have been developed to circumvent this issue, such as reducing the c-Si cell pyramid size or using innovative surface textures, increasing the thickness of perovskite film or application of self-assembled monolayers (SAM).151153

Recently, record efficiencies of 33.7% and 33.9% were reported for 2T perovskite/c-Si tandem cells by a group from KAUST, Saudi Arabia, and LONGi, China, respectively.154 For the 4T architecture, a record tandem cell efficiency of 30.3% was demonstrated in early 2023 by the group from The Australian National University.155 In contrast to the monolithic 2T devices, the top and bottom cells of 4T tandems are manufactured separately and then mechanically stacked.150 This decouples the cells electrically, and hence there is no need for current matching of the sub-cells. Efficiencies of 3T tandem cells are still below 20%, nevertheless this architecture has great potential as the third contact allows the excess photo current of a subcell to be used, which would be lost in a 2T configuration.156,157

Current world records for perovskite/c-Si tandem cells have been achieved with small cell areas (4  cm2), almost a factor of 50 less than commercial c-Si solar cells. Upscaling to an area as large as c-Si cells without losses in efficiency is the foremost challenge. Current research suggests that the most scalable production techniques for the perovskite cells are CVD and solution-based methods, such as slot-die coating, blade coating, inkjet printing and screen printing.158160

Apart from upscaling, degradation of the perovskite top cells161 is considered the major obstacle to commercialization of perovskite/c-Si tandem cells. To be competitive with the established c-Si single-junction modules, stable operation for a minimum of 20-25 years is needed. Yet, a recent study by a group from KAUST showed, that employing 2D/3D perovskite heterojunctions significantly improves the damp heat-stability and allows to pass industrial test protocols for PV modules.162

5.3.

Perovskite Silicon Tandem Cells and Modules

Section Authors: Daniel Kirk (Oxford Photovoltaics Ltd.) and Christopher Case (Oxford Photovoltaics Ltd.)

Mainstream solar cell technology (or technologies) has developed rapidly in terms of efficiency163 concomitant with a huge increase in scale; global PV installations in 2022 were above 250GW and year-on-year growth is expected for decades to come. However, the giant strides taken in silicon technology development mean that it is now approaching its efficiency limit—perovskite-silicon tandem technology opens up huge new possibilities for higher efficiency devices and have become the subject of intense academic and industrial interest.

The majority of the high efficiency devices reported in the literature have been at small area142 with very few examples of commercially viable cell formats. Oxford PV has recently measured efficiency of >28% on full area tandem cells with >250  cm2 area; the IV curve and details are shown in Fig. 12. This fundamentally proves there is no issue with transferring high performance from small area cells to large area, mass production. Moreover, the assembly of high efficiency cells into industrial scale modules has also been demonstrated, with module efficiency of >23% already shown.164 Albeit there is still clearly work to do to increase the cell and module efficiency in line with small cell records, it is already true that perovskite-based tandem technology is class-leading in this area.

Fig. 12

Light-IV (solid line) and Dark-IV (dotted line) of >28% tandem cell. Image courtesy of Oxford PV.

JPE_13_4_042301_f012.png

As mentioned, more than 250GW of solar was installed in 2022—Oxford PV capacity for tandem cells is roughly 0.01% of this; the biggest challenge towards commercialization is to move towards production volumes where real economies of scale can be realised. The biggest barrier to this movement is procurement of tools capable suited for tandem cell application with very high throughput and short lead times. Concerted global effort on equipment provision will greatly accelerate progress towards multi-GW tandem production.

In the short term, one area of challenge is around measurement and certification of PVSK-Si tandem products. The technology for accurate measurement of full-size tandems exists but is not widely available. Work to define appropriate measurement systems and standards is active and has made much progress but adoption of steady-state lamps with advanced spectral control is a pre-requisite for classifying such modules ready for sale. Furthermore, such measurement is essential for certification of the modules as compliant with international standards. The most appropriate standards for tandem modules are the current silicon-oriented ones (IEC 61215 & IEC 61730) and conformity to these requirements is currently being tested, however there is a need for the community to scrutinize such standards and determine if further or different tests are required for this new technology. To ensure these standards are appropriate also requires a growing understanding of long-term performance of tandem devices in operation; deploying significant quantities of tandem modules in appropriate climatic zones is therefore of utmost importance.

6.

Organic PV

6.1.

Nonfullerene Acceptors: Progress, Challenges, and Perspectives

Section Authors: Min Hun Jee (Korea University), Minyoung Lim (Korea University), and Han Young Woo (Korea University)

Over the past decade, significant progress has been achieved in organic photovoltaics (OPV), leading to the successful demonstration of power conversion efficiencies (PCEs) reaching nearly 20%,165167 which was benefitted from the development of the outstanding non-fullerene acceptor (NFA) of Y6 in 2019.168 Numerous derivatives of Y6 and other NFAs have been reported, employing innovative strategies such as chlorination, conjugation-extension, and symmetry-breaking, etc.169172 Compared to fullerene acceptors, NFAs offer several distinct advantages, including facilely tunable optical and electronic structures, longer exciton diffusion length, smaller energy offsets for charge separation and reduced energetic disorder, leading to lower energy losses (Fig. 13). He et al.166 reported an NFA of BTP-H2 to have a near-zero highest occupied molecular orbital (HOMO) offset with the polymer donor PM6, showing efficient hole transfer and small voltage loss to achieve PCEs of 18.5% for binary PM6:BTP-H2 and 19.2% for ternary PM6:BTP-H2:L8-BO blends. To address thermal, mechanical, and long-term stability for small molecular NFA-based OSCs, Yu et al. developed a vinylene-linker-based polymer NFA (PY-V-γ) featuring a coplanar and rigid molecular conformation with high mobility and reduced energetic disorder, achieving a 17.1% PCE with PM6 as a donor.173 To combine the advantages of small molecular and polymerized NFA systems, Sun et al. reported a dimerized Y6-derived NFA with an electron-donating benzodithiophene linker (DYBO), successfully demonstrating both high efficiency (PCE>18%) and excellent stability (t80%>6,000 hours).174

Fig. 13

Design strategies and advantages of non-fullerene acceptors.

JPE_13_4_042301_f013.png

For real industrialization of OPVs, some remaining challenges need to be carefully considered. To realize large-area OPV devices using high-throughput roll-to-roll printing processes,175 it is essential to develop organic materials that can be processed using non-spin coating methods, utilizing environmentally friendly and green solvents. In the case of semi-transparent OPVs, there is a need to improve both PV performance and average visible transmittance by incorporating near-infrared absorbing materials into the active layer.173,175177 Additionally, the use of flexible linkers can be explored to enhance both the PV efficiency and mechanical properties of the stretchable devices.178,179 Moreover, it is crucial to develop cost-effective synthesis for mass production of NFAs. It is also crucial to address the issue of stability in OPVs in order to achieve commercial viability for large-area applications. While encapsulation can help mitigate extrinsic degradation of OPVs, there is a need for a fundamental solution to address the problem of low stability resulting from intrinsic factors such as thermal degradation and photoreaction. Therefore, when developing new high-performance NFAs, it is imperative to conduct extensive research on the impact of their chemical structures on the long-term stability of OPV devices under various conditions, including thermal stress and exposure to light in ambient air. By addressing these challenges, significant progress can be made towards making OPV commercially feasible.

6.2.

Optical Incoupling in Thin Film Organic PV Devices

Section Authors: Zheng Tang (Donghua University) and Olle Inganäs (Linköping University)

Organic photovoltaic (OPV) materials have poor transport properties for charge carriers, compared to classical inorganic PV materials. Thin active layers (<< 1 micrometer) are therefore necessary for OPV devices. This leads to optical losses due to reflection and parasitic absorption. OPV devices are thus easy targets for light trapping and optical incoupling.180182 With the recent OPV materials of high performance, increasing exciton diffusion lengths183 and cost, stability is improved with thinner absorbers184; thus optical incoupling is also relevant here to enable extremely thin film devices.

Strategies for improving light incoupling in the thin OPV devices include both macro- and mesoscale geometry modifications of device architecture, as well as micro- and nanopatterning of materials and devices, using physical effects from microcavity optics in multilayer structures, diffraction or scattering of light, and modification of incoming light with plasmonics. As a transparent electrode is necessary for light incoupling, a considerable effort has gone into making transparent anodes (cathodes), with decreased power dissipation or decreased reflection in the electrode facing incoming light, but not covered in this overview.

6.2.1.

Mode structure of light in multilayer devices

Solutions to Maxwell’s equation in the one dimensional device model of a stack of materials with different dielectric functions are now a standard tool for simulation of devices.185 This transfer matrix method was used also to shape the electromagnetic power dissipation into the optical absorber in the stack, and used to design optical cavity structures.186 Thin organic layers were also used simultaneously for exciton blocking and optical cavities.187

When the active layers are thinner than optimal, cavity optimization become beneficial.188 A sizable absorption enhancement over a wide wavelength range was realized for PV devices by designing new cavities, with the enhancement more pronounced for the absorption tail;189 see Fig. 14.

Fig. 14

Optical cavity inverse design. (a) Schematic representation of the dielectric/metal (DML) layer optical cavity. (b) Left axis: Computed EQEs for devices with a standard ITO cell configuration (solid grey line), the DML cell configuration with an Ag front electrode of 7 nm (solid blue line), and 5 nm Ag (dashed blue line), an optical ergodic geometry configuration with a 99% wave-length independent reflectivity back mirror (dashed red line). Right axis: EQE percentage enhancement for DML cells with a 7 nm (solid green line) and a 5 nm (dashed green line) Ag front electrode relative to the optical ergodic geometry configuration. (c) Energy storing capacity (vertical axis) of the DML cavity configuration as a function of the wavelength and dielectric layer (TiO2) thickness. A dashed rectangle indicates the broadband energy storage region. Reproduced by permission from Liu et al., Ref. 189.

JPE_13_4_042301_f014.png

When tandem devices are built, the distribution of different wavelengths of exciting light between the active materials is critical, and with series connected devices, the balance of photocurrents from these must sometimes be found; here the microcavity structure is also of importance.190,191

While optical cavity structures are often useful for narrowband organic photodiodes,192 they have smaller relevance for PV devices where the overriding aim is power generation, and with one cavity but many wavelengths of light.188

6.2.2.

Meso- and microscopic geometry design for light incoupling

By modifying the mesoscopic and macroscopic geometry of the thin OPV films reflected light from opaque devices can generate higher photocurrent. This was demonstrated in folded single junction and tandem junction devices,193195 where a macroscopic V-shaped cell geometry was used, in which adjacent OPV devices were located at an angle, with subsequent reflection and possible absorption in the adjacent device. This deviation from the standard planar geometry comes at the price of use of more absorber material.

Microstructures using V-shaped light traps196 were demonstrated enhancing the incoupling of light. With another light trap based on microlenses covering a surface and focusing light into aligned apertures in a highly reflective metal mirror in front of the device, effective light trapping was demonstrated.197 An reflective echelle grating at the back of a semitransparent solar cell can also be used to trap transmitted light, and bring it back for absorption.198

Micro- and nanopatterning of the OPV film was first used in soft embossed optical gratings199 where light incoupling was demonstrated. Subsequent work has used periodic or nonperiodic patterning of a flexible substrate for the OPV to guide light into the active layer.186,200 This has also been done with extremely thin film polymer substrates.200

Microstructured light couplers may also be formed as external elements,201204 in the form of periodic textures, diffraction gratings or aperiodic diffractive elements.205,206 Colloidal photonic crystals have been used for large area printed OPV modules.207

6.2.3.

Plasmons and light scattering

Coupling of light to plasmons in metallic nanostructures have been demonstrated, both with metal gratings coupling light into surface polariton modes at the interface between metal electrode and optical absorber,194,202 and with metallic nanoparticles208 inserted in the different layers of the OPV devices.209 Both near field and far field effects due to metal nanoparticles could be used to modify the generation of excited states in OPV materials; for small nanoparticles (2-20 nm), theory predicts that the near field enhancement can be sizable also without plasmonic coupling210; for large particles light scattering can lead to redistribution of light into the optical absorber in devices. The topic is reviewed in,211 which concludes that the plasmonic effects are considerable, though the balance between enhanced absorption due to plasmons, light scattering and modified recombination/generation/transport can have many different results. The inevitable loss of energy of the incoming light in exciting the electrons in the metal always has to be compensated by improved exciton/charge generation.

Light scattering from inhomogenous polymeric films and dielectric nanoparticles are (almost) lossless processes. Dielectric scatterer based on titania can be used to redirect light in stacked homo-212 and hetero-213tandem devices based on semitransparent212215 OPV devices, converting the semitransparent OPV to an opaque OPV. With white paper216 or nanoporous polymers,217 as well as a microstructured retroreflector,218 light harvesting is improved.

7.

Dye-Sensitized Solar Cells: Current Progress and Challenges

Section Author: Marina Freitag (Newcastle University)

Dye-sensitized solar cells (DSSCs) represent a promising PV technology due to their economic feasibility and potential for high conversion efficiencies. (DSSC device structure and principles may be seen in Fig. 15.) Recent advancements in the sector have explored improvements in the areas of device structure, redox shuttle alternatives, solid-state hole conductors, TiO2 photoelectrodes, catalyst materials, and novel sealing techniques, all contributing to the development of new monolithic cell designs.219 Unique attributes include their high conversion efficiencies and commendable performance under diffused light and high temperatures, where traditional technologies struggle. Current research predicts future DSSCs to demonstrate up to 20% power conversion efficiencies (PCE) under sunlight and an astounding 45% for ambient light.220

Fig. 15

(a) DSSC device structure; (b) DSSC working principles.

JPE_13_4_042301_f015.png

Challenges in the path towards commercial scalability of DSSCs primarily revolve around the issue of charge recombination,221223 which currently serves as a major cause of efficiency loss in DSSCs. The intricate system dynamics involved in DSSC operation mean that modifications to one component—dye, redox shuttle, or semiconductor224—can yield impacts across the entire system, potentially boosting or diminishing the overall performance. Future developments must therefore maintain a holistic perspective when introducing new materials, ensuring that the system is adapted to accommodate such changes.

Currently, DSSCs are breaking new ground in several domains. Promising developments include use in ambient and IoT applications with PCE up to 40%,225,226 as transparent solar cells227,228 and photochromic in BIPV applications,229 and in agrivoltaics for greenhouses and vertical gardening applications.227,230 A fascinating area of exploration is “zombie solar cells” using copper coordination complexes and polymers as hole transport materials.231233 Meanwhile, innovative dyes are being developed that are non-toxic, low cost, and sustainable. Pre-adsorbing a hydroxamic acid derivative on the TiO2 surface has shown to improve dye packing and PV performance, enabling record PCE up to 15.2% under standard sunlight.234 With the prospect of PCEs reaching up to 20% under sunlight and 45% under ambient light conditions, the future of DSSCs appears promising. However, the complexity of DSSC devices necessitates a nuanced approach to their improvement, considering the potential impacts of any modifications on the overall system performance.219

Furthermore, the field of photocapacitors, which store energy photochemically in a reversible manner, also offer potential synergies with DSSCs. The integration of DSSCs with photocapacitors can result in a hybrid system that harnesses solar energy and efficiently stores it, offering the potential to meet peak demand times or function during periods of low solar irradiation.235 This would further extend the utility of DSSCs, enabling more consistent energy supply and overcoming one of the key limitations of solar power.

8.

Applications and Commercialization of Emerging PV

8.1.

Indoor PV

Section Author: Lethy Krishnan Jagadamma (University of St Andrews)

One of the most emerging and thriving areas of PV is indoor light harvesting.236238 This advancement of indoor PV is spurred by the rapid development of the internet of things (IoT) and the urge to make this technology more sustainable and robust.239 The main difference between indoor PV and outdoor solar cells lies in the properties of the incident illumination spectra. The illumination spectra of indoor artificial light sources are in the visible spectra range of 400-700 nm with an irradiance of 0.05 to 0.35  mW/cm2 whereas the sun’s spectra range from UV to infrared with an irradiance of 100  mW/cm2. The extremely low light intensity and higher photon energy for indoor PV demand new design guidelines in terms of the material chemistry (not only in the photoactive layer but also the functional charge extraction layers) and the device physics. The optimum bandgap is 1.9  eV, and the theoretical power conversion efficiency (PCE) can be as high as 50-57% for indoor PV depending on whether the illumination source is a fluorescent lamp or a white LED.240,241

The most promising PV materials so far for efficient indoor light harvesting, are amorphous silicon, dye-sensitized solar cells (DSSCs), organic PV, and halide perovskites. Compared to DSSCs, the latter two possess higher prospects of commercialization due to their large area solution processibility, tuneable of bandgap, flexibility, and light weight. The latest reported champion power conversion efficiency (for 900-1000 lux indoor illumination) for these PV are respectively 36 % (a-Si: H),242 25 % (DSSC),243 30 % (OPV)244 and 40.1 % (from the halide perovskites).245 Despite these impressive PCEs, the gap between the theoretical and lab based PCE is still >20%.

To further advance the indoor PV field and its commercialization, the existing challenges need to be overcome. Presently, the field lacks a standardized illumination source. Considering the different indoor illumination levels (200 to 1500 lux), and artificial lights of white LEDs and fluorescent lamps, standardisation is not straightforward. However, reporting colour temperature, irradiance spectrum and irradiance of the indoor light source will make the comparison of the power conversion efficiency values from different research labs more meaningful. Validation of the measured short circuit current density (Jsc) with that from the external quantum efficiency (EQE) measurement should also be encouraged for indoor PV. In the case of hybrid perovskite-based indoor PV, along with the J-V curves in both scan directions, reporting steady-state power conversion efficiency would enhance the reliability of measurements. To close the gap between the theoretical and experimental efficiency values, an overall improvement in all the PV performance parameters is necessary.246 However, the Voc, loss needs immediate attention. Despite the large bandgap (1.8 to 2.0 eV) of the photoactive material, most of the reported indoor PV have a Voc of 0.8V only (loss of more than 1V). Though compared to normal outdoor solar cells, indoor PV may require less stringent stability testing protocols, many are yet to be formulated. In the coming years, research on indoor PV is expected to be focused on, maximising its PCE, developing the stability protocols, integrating with the energy storage units and implementing them in real IoT applications (see Fig. 16).

Fig. 16

Sustainable IoT for buildings.

JPE_13_4_042301_f016.png

8.2.

Commercialization of Organic PV

Section Author: Stephan Kube (Heliatek GmbH)

Heliatek is the technology leader in organic PV, founded 2006 out of research in the field of organic electronics from TU Dresden and the University of Ulm. Researchers had developed a new generation of semiconductor materials built on carbon-based molecules from organic chemistry. With this new generation of materials, Heliatek was able to develop an innovative new solar technology called “organic photovoltaics” (OPV). Since 2006, Heliatek has developed new materials, new product concepts, and a proprietary “Roll-to-Roll” vacuum evaporation production process.

In 2022, Heliatek started to produce and sell its first commercial product HeliaSol. HeliaSol is an ultralight, flexible, and ultrathin solar film that can be easily glued to various building surfaces such as metal, glass, or concrete with the integrated backside adhesive. This makes HeliaSol the perfect choice for all applications where conventional solar modules reach their limits due to weight or surface restrictions (e.g., curved shapes). With a carbon footprint of less than 10 g CO2e/kWh, HeliaSol is based on the greenest of all solar technologies and is amongst the greenest of all electricity generation technologies. The reasons for this ultralow carbon footprint mainly come from the few materials required to manufacture solar films and the abundance of toxic heavy metals (such as cadmium or lead) and rare earths. The films only use 1 g of organic material per film and the films weigh in total less than 2 kg.

The solar films are produced in an innovative roll-to-roll evaporation production process, where Heliatek processes coils with a length of up to 2.6 km and a width of up to 1.3 meter. This allows us to produce different product dimensions. The first HeliaSol that Heliatek is producing has a dimension of 2 m length and 0.44 cm width, which has a compact and easy to handle format. These films have a power between 50 and 60 W.

Heliatek has already proven the versatile applications possibilities in projects around the world, where its façade solutions, building rooftop applications, or special projects like a wind turbine tower in Spain, show that with lightweight, flexible solar films, Heliatek is able to transfer surfaces into active green electricity generators.

Heliatek is funded by the Free State of Saxony, the Federal Republic of Germany, and the European Union. The company has more than 250 employees at the headquarters in Dresden and the research location in Ulm. Figure 17 shows the diverse applications of the company’s products globally.

Fig. 17

Clockwise from top left: HeliaSol solar film (rolled), production of solar films in Roll-to-Roll manufacturing process, façade installation in Korea, wind turbine tower in Spain, and biogas plant in Germany. Images courtesy of Heliatek, Samsung, Acciona, and RWE.

JPE_13_4_042301_f017.png

8.3.

Agrivoltaics: Status, Land Use Conflicts and Synergy with Organic Photovoltaics

Section Author: Harald Hoppe (Friedrich Schiller University Jena, CEEC Jena, and IOMC)

With the current expansion of renewable energy supply in large parts of the world, primarily through wind energy, energy crop cultivation, and PV, the land use required for this is also increasing inexorably.247249 Not in all regions can land that is otherwise unusable be used for this purpose, and thus resistance to such projects is also growing.250,251 In western countries, for example, this has led to increased state regulation with the associated extensive bureaucracy for the realization of new regenerative power plants, which makes further expansion considerably more difficult. On the other hand, there are now not only competitive situations with traditional agriculture, but also land use conflicts between energy crop cultivation and photovoltaics in particular have become a priority problem, for example in Germany.252

However, a solution to these land use conflicts could be found in a dual use of the available land through agrivoltaics,253 a coexisting agriculture with PV installations, providing a solution to the water-food-energy nexus.254 Typical agrivoltaic system installation parameter considerations (presented in Fig. 18) include crop type, PV module installation angle, distance, and overlap on the crops, and the transparency of the modules. At present, more than 15 GW of installed capacity have already been realized worldwide, for which subsidy programs have even been set up in some countries.255

Fig. 18

The schematic shows the most important parameters of agrivoltaic installations. The crop type and its handling impacts, e.g., on the distance and height of installations. The angle, distance, and overlap determine the shadow or rain protection, while the transparency modulates the shadowing.

JPE_13_4_042301_f018.png

The majority of agrivoltaic systems are currently being based on the use of silicon solar cells, which is accompanied by an increased installation effort compared to classical PV installations. However, there are already first projects for the use of organic solar cells, respectively organic PV.256258

Organic solar cells have distinct advantages such as light weight and ease of integration, as well as the potential for very high average visible transmittance (AVT) due to their unique property of limited absorption bands within a spectral interval,259261 which potentially eliminates competition between PV and plant light use in greenhouses.262,263 More research and development is needed,264,265 however, in order to synthesis and mass-produce novel organic semiconductors that preferentially use the infrared (IR) section of the sun spectrum.

8.4.

Perovskites for Space PV Applications

Section Authors: Duong Nguyen Minh (National Renewable Energy Laboratory) and Joseph M. Luther (National Renewable Energy Laboratory and RASEI)

While the prospects of new PV technologies are growing rapidly for terrestrial terawatt-scale electricity generation,266 PV was first marketed as technology to power spacecraft such as satellites, and emerging PV technologies could provide substantial cost reductions for the emerging space market.267 The satellite industry is rapidly expanding due to the commercialization of launch opportunities (SpaceX and others). More than 10x the historical number of satellites are planned to launch in the upcoming 5 years.268 Traditionally, III-V, multijunction solar cells make up the bulk of space PV market because of their higher efficiency, improved radiation tolerance, and improved temperature characteristics compared to Si albeit at much higher cost. Thus, the goal for new PV technologies, like perovskites would be to provide low cost options (equivalent to terrestrial Si) that may match or improve upon the radiation tolerance and temperature characteristics of III-V technologies.269,270 Perovskite fabrication advantages enable flexible roll-out solar arrays and furthermore, the potential to even manufacture solar cells in space.267,270

Perovskites for space PV has already shown promising results.271,272 Although the early demonstrations also reported the surprising high radiation tolerance of perovskites, the testing conditions are widely different.273 Part of the rapid increase in interest of perovskites globally was because they are shown to tolerate a higher density of defects (while retaining high efficiency) than other PV technologies.273 This is further beneficial because not only can they tolerate higher defect densities, the radiation produces less defects in perovskites from particles of relevant energy as shown by the slightly lower non-ionizing energy loss and slightly higher stopping range plotted in Figs. 19(a)19(b). This property is ideally suited for radiation tolerance in space where interactions with charged particles can lead to atomic displacements or vacancies within the absorber layer. The future development of perovskite PV for space applications would benefit from a consistent ground-based radiation testing convention to accelerate the optimization of perovskite PV designs, because radiation tolerance may not only be an inherent property, but it may also be a design criterion to optimize with interfacial engineering and contact decisions. However, the well-developed radiation-testing protocol for conventional space PV technology may not be equally suitable for testing perovskite solar cells due to their soft lattices, thermal properties, and markedly different device architectures. Currently, the low-energy protons of 0.05-0.15 MeV were reported to cause atomic displacements and vacancies in unencapsulated devices,273 which effective probe the radiation interactions in perovskite solar cells.

Fig. 19

Development of perovskite solar cells for space applications. a-b) Fresh protocol testing for perovskite solar cells. a) Comparison of irradiation interaction to Si, InGaP, and perovskite PV technology. b) stopping ranges of protons for various absorbers. c-d) Development of encapsulation for perovskite solar cells. c) Simulated proton interaction with perovskite devices. d) SiOx barriers block proton irradiation to perovskites solar cells. Figure adapted with permission from Refs. 271 and 274.

JPE_13_4_042301_f019.png

However, there are still many hurdles to overcome for perovskites in space regarding thermal properties.267 Some orbits require aggressive thermal cycling tests (thousands of times more aggressive than IEC standards, and at wider and faster temperature swings), while other orbits may require constant operation indefinitely (no diurnal cycle), where the cells must be built to endure the temperature without convective heat transfer.

For either case (terrestrial or space), perovskites will need to be well-protected/packaged.275 While polymers and sealants encapsulation are likely to decompose under harsh radiations because of the weakening carbon bonds, the use of traditional cover glass could also be challenged for weight and cost savings.267,275 Comparatively thin (1 μm) SiOx encapsulation (lighter than 99% of conventional barriers) was recently shown to protect perovskites against the various key critical stressors in both terrestrial and space environments, including protons [as shown in Figs. 19(c)19(d)], alpha particles, atomic oxygen, and moisture, and provides a good starting point for more complete packaging ideas.274

9.

Strategies for Exceeding the Detailed Balance Limit

9.1.

Progress towards the Hot Carrier Solar Cell

Section Authors: Ian R. Sellers (University at Buffalo), Vincent R. Whiteside (University of Oklahoma), and David K. Ferry (Arizona State University)

Since the pioneering work of Ross & Nozik,276 the hot carrier solar cell has been considered a potential protocol to circumvent the single gap limit.277 This limit suggests that practical losses due to transmission (of low energy photons) and thermalization (due to absorption of high energy photons) limit conventional semiconductor solar cells to 30% power conversion efficiency.

Fig. 20

(a) Schematic of the typical extraction energy/voltage in a single gap solar cell. (b) Protocol for hot carrier operation using IV scattering. Hot carriers are rapidly transferred to satellite valleys then extracted via an energy matched collector at V>Voc. Figure created by Stephen J. Polly (RIT).

JPE_13_4_042301_f020.png

The proposal in Ref. 276 suggests that solar cells operating in excess of 60% might be realized, without the need for multijunction architectures.278 One scenario for the operation of a hot carrier solar cell will occur when carrier-carrier interactions have generated a steady state Maxwellian “hot carrier” distribution whose temperature is in excess of the lattice. Generally, this requires inhibiting the thermalization of high energy photogenerated carriers, then extracting these carriers at their higher energy.

Since its original inception in 1982,276 the hot carrier solar has received significant attention with much of the work related to inhibited carrier-phonon interaction or the creation of a so-called hot phonon bottleneck279,280 in low-dimensional structures such as quantum wells. Alternative approaches involve the investigation of materials with large phononic band gaps281 that may inhibit hot LO phonon dissipation such as InN,282 AlAsSb,283 and/or HfN,284 amongst others. In addition to these, recent approaches seek to control carrier-phonon interactions by design in type-II superlattice structures.285,286

While many of the early works in this area focused on spectroscopic analysis in both the CW287289 and ultrafast290292 regimes, recently hot carrier effects have also been observed in devices. Recent device results have shown: the presence of PV carrier extraction,293,294 extremely high non-equilibrium carrier temperatures in quantum well devices,295 and selective hot carrier extraction in metallic absorbers.296 While these demonstrations show significant practical progress towards the realization, decoupling electron-phonon mediated heat generation remains challenging.

Recently, an alternative protocol has been proposed in which intervalley scattering is used to transfer high energy photogenerated carriers into the satellite valleys of the absorber layer of heterostructure solar cells.297,298 In such valley PV solar cells, it is proposed that intervalley phonon scattering is harnessed to inhibit hot carrier-LO phonon interactions in a semiconductor. This process allows the transfer and stabilization of hot carriers via the efficient transfer of photogenerated carriers to the upper valleys in the absorber, a process by which radiative recombination is exponentially reduced. A schematic of this process in comparison to the typical operation of a single gap solar cell is shown in Fig. 20. Initial work in this area has shown proof of principle operation and efficient carrier scattering in the InGaAs/AlInAs system,297,298 with work now focused on efficient carrier extraction across the absorber/barrier interface. Efficient upper valley extraction would support photo-voltages in excess of the absorber band gap and power conversion efficiencies299 well above that of the single gap limit.

10.

Light Management

10.1.

Luminescent Solar Concentrator PV: Performance

Section Authors: Angèle Reinders (Eindhoven University of Technology), Xitong Zhu (Eindhoven University of Technology), Wilfried van Sark (Utrecht University), and Michael G. Debije (Eindhoven University of Technology)

Luminescent solar concentrator photovoltaics (LSC-PV) act as spectral converters, optimizing spectral matching between the irradiance spectrum and peak of the response of PV cells.300,301 An LSC-PV normally consists of a transparent optical lightguide (usually glass or a polymer such as poly [methyl methacrylate] [PMMA], polycarbonate [PC], or polydimethylsiloxane [PDMS]302) with embedded luminophores which redirects absorbed light as downshifted emission toward the edge(s) where it may be used to illuminate attached PV cells;303305 see Fig. 21 (top and middle). Spectral conversion and (in larger lightguides) total internal reflection (TIR) are used to improve the LSC-PV device’s efficiency.306,307

The concept of LSC-PV was introduced four decades ago,308310 initially to reduce the high cost of PV modules,311 but now these colorful, transparent, visually appealing solar devices could contribute to the general acceptance of solar technologies in the built environment, even with modest efficiency, thanks to their greater design flexibility than conventional flat, rectangular PV modules; see Fig. 21 (bottom).

LSC-PV perform well under various irradiance conditions, particularly under diffuse light.312 Color (from purple to red hues)313 is determined by the luminophores employed and may comprise organic dyes,314317 quantum dots (QDs),305,318322 and/or rare-earth ions.323325 The optimal dye has broad spectral absorption,326,327 high photoluminescence quantum yield (PLQY),328330 large Stokes shift,331,332 good solubility in the host matrix material,333,334 and good spectral overlap of the luminophore emission s and the spectral response of the PV.335337

The performance of LSC-PV devices is limited by several loss mechanisms, including attenuation by the lightguide material, scattering, escape cone losses,338 and reabsorption during the lightguiding process.339,340 Several methods have been proposed to minimize loss mechanisms in lightguides and enhance the internal efficiency (ηint) and power conversion efficiencies (ηPCE).319,341343 Further proposed designs include using different lightguide materials, geometric shapes (e.g., rectangular,344 square,345,346 cylindrical,347,348 and bent346), and different dimensions (from 1.5cm×1.5cm×0.1  cm,331 to 100cm×150cm×0.6  cm,349) to numerous edge-mounted and or bottom-mounted PV cell technologies, including, GaInP,350 GaAs,321 CdTe,350 CIGS,350 mc-Si,351 c-Si,352 a-Si,353355 p-Si,356 DSSC,357 and perovskite.358

Reported LSC-PV devices have shown various efficiencies.340,343 The most frequently reported world record of ηLSCPV=7.1% applies to a small-sized device (5cm×5cm×0.5  cm) with high efficiency GaAs solar cells edge attached to a PMMA containing a mixture of two organic dyes.359 However, most LSC-PV devices have modest efficiencies in the range of 1-5%. The average efficiency of LSC-PV devices has increased from 2.5% (1980–1985) to 5% (2005–2021); Fig. 22 summarizes LSC-PV devices with variety in host matrix, luminescent materials, solar cell technology, lightguide dimension, PV cell size, geometric gain (Gg), internal efficiency (ηint), power conversion efficiency (ηPCE), and other design configurations.

Note the efficiencies reported in Fig. 22 were not obtained using the same metrics or measurement procedures, as comprehensive standard protocols for performance reporting have only recently been introduced.360,361 As such, previously reported LSC-PV device efficiencies were overly positive, ignoring transparency of the LSC, and cannot be used for comparisons of other efficiencies obtained under STCs.362

Fig. 21

(Top) Solar irradiance spectrum with an indication of conversion of a blue photon to the red part of the spectrum, (middle) a lightguide with a solar cell attached to one of its edges, (bottom) in pink, possible locations of application of LSC-PV devices in a dwelling.

JPE_13_4_042301_f021.png

Fig. 22

Reported power conversion efficiencies of luminescent solar concentrator PV (LSC-PV) devices over the past 43 years with edge-attached (black squares), bottom-attached (red circles), and both edge- and bottom-attached solar cells (blue triangles).

JPE_13_4_042301_f022.png

10.2.

Luminescent Solar Concentrator PV: Technology Status

Section Author: Vivian E. Ferry (University of Minnesota)

Luminescent solar concentrators (LSCs) have been proposed for PV applications for more than 40 years.309 LSCs are attractive devices for integration of solar panels into the built environment: composed of luminescent materials (dyes, colloidal quantum dots) embedded in a polymer matrix, their colorful/semi-transparent nature and ability to capture and concentrate diffuse sunlight makes them excellent candidate architectural materials.341 Diffuse sunlight is absorbed and emitted by the luminophore, and directed onto an adjacent solar cell by total internal reflection. Historically, performance has been limited by significant optical losses, and especially re-absorption in large-area devices, since every absorption/emission event is an opportunity for losses from the non-unity quantum yield or the imperfect waveguide (escape cone, scattering); see Fig. 23.

Fig. 23

Estimated power conversion efficiency and light utilization efficiency for a tandem LSC (a) with various candidate top luminophores, and [(b),(c)] constant overall absorption of 10%. Figure adapted from Ref. 363.

JPE_13_4_042301_f023.png

The past 15 years have seen considerable new interest in LSCs, driven by new advances in light-emitting materials.364 For efficient operation, the luminophores must strongly absorb incident sunlight, have high quantum yields, and large Stokes shifts to avoid re-absorption of the luminescent light, and have low toxicity/earth-abundancy for large-scale application. Colloidal nanocrystals, such as II-VI particles,351,352,365,366 Si,331,367369 carbon dots,370 perovskites,371374 and ternary alloys,300,375 and metal-halide nanoclusters376 have all emerged as candidate materials,363 offering highly tunable and advantageous optical properties. However, these new materials must be dispersed in polymer matrices and waveguide materials370,377 without significant degradation of optical properties, light scattering, or absorption losses by the matrix material, and large-scale, cost-effective synthesis and fabrication methods are needed.367 Additionally, as the quality of the luminophores improves, other aspects of optical design become increasingly important to the overall efficiency, such as photonic structures that trap and direct luminescence toward the PV cell.351,377381 Recent calls to standardize efficiency metrics for LSCs should also be noted.382

There are other intriguing new developments in the LSC community, including the implementation of downconverting/upconverting materials, as opposed to the downshifting mechanism of a traditional LSC architecture. New application spaces for LSCs with PV are emerging,383 including the use of LSCs in greenhouses where the tailored transmission spectrum benefits crop growth.384387

Large-area and scalable fabrication remains an important challenge for LSCs. Techniques such as doctor blade coating367,388 or spray coating389 are important for production, but may require materials advances for luminophore compatibility, as well as understanding of the process to create high quality, low haze materials. Similarly, photonic structures that assist in light-guiding must be manufacturable. Lifetime is critical to commercial applications, requiring luminescent composites that are stable against oxygen and moisture. Sustainable luminescent materials with low toxicity but excellent optical properties in the composite continue to need development.390 Commercialization may also require different form factors or design modifications to integrate with application-specific components, i.e. architectural glazing or displays.391

Given the large number of design variables (shape, luminophore, concentration, matrix material, use of photonic structures, environment for implementation, and others), there is an opportunity to understand the design rules for LSCs from the perspective of the commercial application, i.e., accounting for sustainability, scalability, and the operating environment from the start, rather than designing for small-scale laboratory tests. Computational methods that accelerate this design process will likely play an essential role in both understanding and optimizing these future generations of LSCs.

10.3.

Spectrum Splitting PV Systems

Section Authors: Raymond K. Kostuk (University of Arizona) and Benjamin Chrysler (Amazon Lab)

High efficiency PV systems have many operational and financial advantages. High efficiency allows less land usage for the same electrical power generation and the ability to generate significant power in constrained areas. According to the detailed balance condition, high efficiency PV conversion can only be achieved by using multiple energy bandgap PV cells within the range of the solar energy illumination spectrum.277,392 The most common way of doing this has been to stack multijunction cells in tandem with the topmost cell converting shorter wavelengths and the bottom most cell the longer wavelengths. This approach requires precise lattice matching between different bandgap materials, current matching, and is restricted to compatible materials. These factors increase cost and require high concentration optics for economic viability. An alternative approach is spectrum splitting in which an optical element separates the spectrum of the incident solar illumination and directs spectral components to an array of single junction cells that have high responsivity to different spectral bands [Fig. 24(b)]. In spectrum splitting systems the emphasis of the design shifts from the stacked multijunction cell to the design and fabrication of the optics that perform the spectral separation of the incident solar illumination. The concept of spectrum splitting PV systems is not new (dating back to 1955)393 and comprehensive reviews have been given by Imenes and Mills394 and Mojiri et al.395

During the past ten years, a variety of methods have been proposed for implementing spectrum splitting optics. These approaches include: cascaded dichroic reflection filters,396,397 filter-mirror combinations, 398400 optimized surface relief diffraction gratings,401,402 hologram-lens combinations,403 and volume holographic lenses.404406 All spectrum splitting systems require at least single axis tracking, however, single axis tracking is being used with many more efficient types of silicon cell modules and is therefore not considered a major impediment. The most robust and compact optical approaches are surface relief diffraction gratings (SRDGs) and volume holographic lenses (VHLs). Ultimate use of either of these techniques will depend on fabrication cost and durability. Volume holographic elements have been successfully incorporated into traditional PV module packages that have passed accelerated life testing for 25 years or more and provide some support for the viability of this technique. In addition, methods for replicating volume holographic elements and lenses have also been demonstrated (Fig. 25).407,408

Recent work on spectrum splitting has led to an interesting result for optimal cell configurations.409 Figure 24 shows cell types: stacked multijunction, lateral separated single junction, and hybrid cells. A stacked multijunction PV cell is a form of absorption spectral filtering where each cell in the stack filters and converts a different spectral band of the incident sunlight to electrical power. The laterally separated cells completely rely on optics for achieving spectral separation, while the hybrid cell uses a combination of absorption and optical spatial spectral separation mechanisms. It was found that the hybrid approach incurs less overall system loss. This approach can also take advantage of recent perovskite-silicon tandem cell development.410

Fig. 24

(a) Stacked, (b) lateral, and (c) hybrid configurations for spectrum-splitting PV modules. In stacked modules, PV cells are placed on top of each other either. Upper cells in the stack convert shorter wavelength light and pass longer wavelength light to the lower cells. The stack can be implemented as a monolithic multijunction cell or a set of mechanically stacked single-junction cells. In a lateral configuration, an optical element sends different parts of the solar spectrum to single-junction PV cells located in different areas of the module. In a hybrid configuration, an optical element sends different parts of the solar spectrum to different cell stacks where the solar spectrum is further subdivided. In general, cells can be connected in series, series-parallel combinations, or electrically independent configurations.

JPE_13_4_042301_f024.png

Another emerging applications is for combined PV conversion and bio-growth (agrivoltaics411) where a volume holographic element filters part of the spectrum to match the photosynthesis active region of the spectrum (PAR)412 and the remainder for electrical conversion.413 The irradiance at the crop area can also be modified to further enhance plant growth. This approach can be applied to both open fields and greenhouse crops and provides dual use of much more land area with synergistic benefits for crop production.

Fig. 25

(a) An array of volume holographic lenses fabricated using a replication system.414 (b) Spectral band separation using the VHL array shown in (a).

JPE_13_4_042301_f025.png

11.

PV Sustainability and Environmental Impact

Section Authors: Annick Anctil (Michigan State University) and Preeti Nain (Michigan State University)

To evaluate the benefit or identify potential concerns of solar PV, one must consider all the stages of the PV industry, from raw material extraction to end-of-life. With the forecasted rapid increase in new PV installations and dismantling of older modules in the upcoming decade, research and development efforts need to 1) minimize the environmental footprint of PV manufacturing and 2) plan for PV recycling and waste management globally.

This update focuses on two areas: PV manufacturing and end-of-life. Over the last decade, the global PV supply chain has moved from Europe, Japan, and the US, to China, which accounts for more than 80% of worldwide production.415 The combination of policy incentives to support domestic production in North America and restrictions or tariffs on Chinese PV modules and components are responsible for many recent announcements of US PV manufacturing.416 Carbon footprint is a criterion for project selection in France and South Korea417 and other countries will likely follow. Changing the manufacturing location along the supply chain affects the carbon footprint. Manufacturing PV in the US should have a lower carbon footprint than in China, but the difference is decreasing due to grid decarbonization in both countries.418 In addition to electricity, manufacturing improvements and reduced material usage have led to a lower carbon footprint of PV modules over time.419,420

One other concern for PV modules is end-of-life management. Recycling methods for CdTe were established many years ago due to the scarcity of tellurium and toxicity of cadmium421 but are not as commonly used for silicon. Standard waste characterization methods (i.e., Toxicity Characteristic Leaching Procedure, Waste Extraction Test, etc.) are used to evaluate the toxicity of PV. Copper, iron, aluminum, and lead can leach from PV modules when tested, but their concentration would not exceed the hazardous waste limit.422 PV end-of-life management is required in Europe and a few states in the US. Still, additional efforts are needed globally to ensure that PV waste is recycled, and environmental impacts are minimized if landfilling is the only option.

A circular economy for PV has become one primary objective for long-term sustainability.423,424 For manufacturers, low-carbon solar modules are becoming a competitive advantage, and being able to conduct LCA rapidly would facilitate product certification. Changes in material intensity and manufacturing process need to be incorporated in LCA databases to ensure a reduction of the environmental footprint of PV manufacturing with new module technologies.

For recycling, cost remains a challenge for closed-loop recycling of PV modules. Until sufficient modules wastes are produced, a centralized PV recycling scheme would ship wastes over long distances unless waste can be incorporated into the existing waste stream.421 Another issue is forecasting the volume and value of future waste stream to plan for recycling facilities.425 Regulations and incentives for recycling are required in addition to inexpensive and effective recycling methods. Figure 26 summarizes the areas of research needs for sustainable PV management. Overall, manufacturers could improve PV life cycle assessment and end-of-life management with a bill of material disclosure or through third-party certification. Solar technology is considered a green technology, and a holistic approach considering the environmental impact throughout its life-cycle is required to minimize adverse environmental impacts.

Fig. 26

Framework for sustainable PV management.

JPE_13_4_042301_f026.png

12.

Conclusions and Perspectives

12.1.

Common Themes and Challenges

The diverse contributions to this report demonstrate the remarkable range of emerging PV technologies as well as developments in their applications. They also describe some of the challenges to widespread deployment. Some of these challenges are specific to particular technologies—for example, finding materials for lattice matching for III-V solar cells. However, it is interesting to see that there are also challenges that arise across many (or all) solar technologies. One of these recurring themes is translating lab solar cell performance to modules and on to the production line. This includes challenges of scaling up of coating, module design, and developing appropriate manufacturing tools and processes. It applies not only to solar cells themselves but also to ways of enhancing their efficiency—for example, through light management and photonic structures.

Manufacture is also closely associated with related issues of cost, energy of manufacture and environmental impact. Whilst cost has been recognised for a long time as crucial for determining extent of deployment, the articles show increasing attention being paid to energy payback time, i.e., the time for a solar cell to generate energy equal to that required for its manufacture. There is also growing concern to develop clean manufacturing processes—for example, by using “green” solvents, and to consider the environmental impact of both the process and the materials used. There is increasing awareness of the importance of life cycle analysis. Indeed, to become truly sustainable it will be necessary to develop a circular economy for PV.

Progress in stability is needed for many solar technologies, especially perovskites, organics, and dye-sensitized solar cells, with a 25-year lifetime in the field being desirable. There are grounds for optimism as in related fields, such as organic light-emitting diodes (OLEDs), tremendous progress in stability has been made and the best lifetimes of OLED pixels are much longer than 25 years. Aspects of advancing stability include understanding degradation mechanisms and developing appropriate encapsulation and packaging. The impact of shading also needs to be considered as a shaded solar cell can experience a substantial reverse bias due to its neighbours.

The exciting applications emerging bring their own challenges that need to be addressed. One such issue is developing appropriate testing protocols as the intensity, spectrum and temperature may differ greatly from the standard test conditions developed for silicon solar cells. For example, indoor applications use light that is at shorter wavelength and much less intense than most of the solar spectrum. Stability testing protocols will also need tailoring to the application. Particular applications place their own requirements on solar cells—for example, some large warehouses have flimsy roofs that would only be able to support lightweight solar cells. For agriphotovoltaics and building integration, the transmission (and reflection) spectrum of the solar cells is important.

Finally, the authors share a spirit of optimism regarding the future of PV technology, despite varying perspectives on the priorities and challenges involved (see Sec. 12.2). Overall, this report provides an overview of emerging PV approaches that show potential in helping to achieve the development of new materials, device concepts, and light management strategies to enable higher efficiencies and more scalable and sustainable manufacturing.

12.2.

Author Survey and Perspectives

As a way to gain a broader perspective on the field of emerging PV, the authors of this article were asked to complete an anonymous informal survey regarding their projections for the future directions of the field. 19 responses were received. Below is a list of the questions and summary of responses:

Q1: What do you think are the most near-term promising applications of emerging PV technology?

Responses (rank-ordered %): Agriculture (22%), PV-integrated parking structures (14%), Non-III-V in space (12%), IOT (12%), Automotive/mobile (12%), Semitransparent PV windows (10%), Indoor (8%), Other (8%)

Q2: When do you predict perovskites (all perovskite tandem or tandem with other materials) will reach commercialization at the level of 1 MW installation?

Responses (%): 1 - 3 years (26%), 3 - 10 years (47%), 10+ year (0%), Never (11%), No idea (16%)

Q3: When do you predict OPVs of any form will reach commercialization at the level of 1 MW installation?

Responses (%): 1 - 3 years (5%), 3 - 10 years (42%), 10+ year (32%), Never (11%), No idea (11%)

Q4: When do you predict DSSCs will reach commercialization at the level of 1 MW installation?

Responses (%): 1 - 3 years (5%), 3 - 10 years (11%), 10+ year (0%), Never (68%), No idea (16%)

Q5: Which of the “exotic” mechanisms for breaking the detailed balance limit will most likely be commercialized someday?

Responses (rank-ordered %): Photon up/down conversion (32%), Spectrum splitting (24%), Multiexciton generation (11%), None (11%), Hot carrier solar cells (8%), Intermediate band solar cells (5%), 2D absorbers (5%), Singlet fission (3%))

Q6: What PV materials are understudied and deserve more attention?

Responses (alphabetical order): Chalcogenides, Dilute nitrides, Inorganic semiconductors, Lead- and tin-free perovskites, Low bandgap materials, Mismatched materials, Non-toxic earth-abundant Inorganic materials, OPV, Solar concentrators, Thermophotovoltaics, Quantum dots

Q7: Which are the most pressing environmental and sustainability concerns of solar PV, which the community should focus on as technologies are being developed? Please rank the following 5 concerns.

Responses (rank, number): Recyclable solar cells (1st, 4; 2nd, 5; 3rd, 3; 4th, 5; 5th, 2), Lifecycle analysis (1st, 4; 2nd, 1; 3rd, 9; 4th, 4; 1), Elemental availability (1st, 5; 2nd, 3; 3rd, 2; 4th, 5; 5th, 3), Green chemistry for fabrication (1st, 2; 2nd, 7; 3rd, 4; 4th, 3; 5th, 2), Land usage for sustainability (1st, 4; 2nd, 2; 3rd, 1; 4th, 1; 5th, 11)

12.2.1.

Write-in responses

The authors were also asked to share any final thoughts about the future of PV technologies as write-in responses. Several key responses are quoted below, anonymously.

“It is difficult to predict the future however I dare to say that PV technologies will be widely applied, in various environments, at a terrawatt level, meaning that PV waste management will be one of the greatest challenges for truly sustainable terrestrial and space PV applications.”

“I imagine a broadening application space, solar integration and power transfer beyond today’s technology.”

“Perovskites will be important only if their lifetimes can be made comparable to silicon, CdTe, or CIGS.”

“Spectrum splitting will become important if a low-cost wide bandgap material become available.”

“Even though I ranked it last, land use is definitely important. I worry when I see articles about different environmental groups on opposite sides of a solar debate.”

“PV technologies share in our power generation infrastructure will continue to grow. PV will always be part of the solution for renewable clean energy generation, however, not necessarily the only solution.”

12.2.2.

Survey summary and conclusions

Several speculative conclusions can be drawn from the survey results. In response to the first question, “What do you think are the most near-term promising applications of emerging PV technology,” agriculture was the clear winner. This speaks to potential advantages and effectiveness of applying emerging technologies to the agricultural space as well as the importance of the food-energy nexus in global systems. Additionally, one author submitted as a write-in response “Extension of PV installations geographically further north” as a promising near-term application. This statement illustrates how much the PV industry has grown in recent years such that development in regions of the world with sub-optimal solar insolation should be expanded. It also hints at the opportunities for “geographically tailored” PV whose spectral and temperature responses, stability characteristics, and testing protocols are designed for more polar conditions.

In response to questions 2–4, regarding at which time in the future particular technologies will be commercialized at the level of 1 MW installation (an arbitrary value chosen for the purpose of this informal survey), authors were generally optimistic about the development of perovskites (all perovskite or tandems with other materials), with nearly 75% of authors predicting it will occur in less than ten years. OPV were also viewed optimistically with 47% predicting the target will be reached in that timeframe. Dye-sensitized solar cells were generally not viewed optimistically with 68% concluding that they will never reach 1 MW commercialization. However, it is important to note that this was an informal survey given to a statistically small number of researchers in the field. The results should not be considered as well-researched technoeconomic analysis, nor are they intended as guidance for researchers or policy makers in the field.

For question 5, on what mechanisms for breaking the detailed balance limit will most likely be commercialized someday (with no timeframe provided), photon up/down conversion and spectrum splitting received the most responses, with 32% and 12%, respectively. Multiple exciton generation and hot carrier solar cells garnered the third and fourth-highest number of votes. Four authors responded that none of the mechanisms listed would ever be commercialized. The same caveat as mentioned in the previous question also applies to this topic.

In the final question in the survey, on the question about the most pressing environmental and sustainability concerns of solar PV, the authors chose nearly evenly among the listed topics of recyclable solar cells, lifecycle analysis, elemental availability, green chemistry, and land usage. That no clear winner emerged from this question suggests that all these issues require deeper investigation and are critical for future PV implementation at global scale. Overall, the results suggest that researchers and educators in the physical sciences and engineering fields would do well to broaden their knowledge base and research efforts to encompass more aspects of sustainability.

Code and Data Availability

Data sharing is not applicable to this article, as no new data were created or analyzed.

Acknowledgments

This report was edited by Fatima Toor, Ifor D.W. Samuel, and Sean E. Shaheen. Annick Anctil acknowledges grant support from the US National Science Foundation (NSF) Division of Chemical, Bioengineering, Environmental and Transport Systems (CBET): Environmental Sustainability of Photovoltaics in the US (CAREER award no. 2044886). Raymond K. Kostuk and Benjamin D. Chrysler would like to acknowledge that the US NSF and DOE provided funding for the QESST Engineering Research Center (award no. EEC-1041895) for support of their research. Benjamin D. Chrysler would like to acknowledge support from the NSF Graduate Research Fellowship Program (award no. DGE-1143953). Hele Savin and Ville Vähänissi would like to acknowledge the provision of facilities and technical support by Micronova Nanofabrication Centre in Espoo, Finland, within the OtaNano research infrastructure at Aalto University. Hele Savin and Ville Vähänissi would also like to acknowledge funding provided by the Flagship on Photonics Research and Innovation “PREIN” (#346529) funded by the Research Council of Finland. H. Y. Woo acknowledges financial support from the National Research Foundation (NRF) of Korea (2019R1A6A1A11044070). This work was authored in part by the National Renewable Energy Laboratory (NREL), operated by Alliance for Sustainable Energy, LLC, for the US DOE under contract no. DE-AC36-08GO28308. Funding provided by the US DOE Office of Energy Efficiency and the Renewable Energy Solar Energy Technologies Office under award nos. 38257 and 38266. NREL authors Duong Nguyen Minh and Joseph M. Luther also acknowledge support from the Operational Energy Capability Improvement Fund of the US Department of Defense. The views expressed in the article do not necessarily represent the views of the DOE or the US Government (USG). The USG retains and the publisher, by accepting the article for publication, acknowledges that the USG retains a nonexclusive, paid-up, irrevocable, worldwide license to publish or reproduce the published form of this work, or allow others to do so, for USG purposes.

References

1. 

L. Fernández, “Share of electricity generation from solar energy worldwide from 2010 to 2022,” (July 2023). Google Scholar

2. 

E. J. Mitchel, “Photovoltaic power systems programme annual report,” (2022). https://iea-pvps.org/wp-content/uploads/2023/04/PVPS_Annual_Report_2022_v7-1.pdf Google Scholar

3. 

N. M. Haegel et al., “Photovoltaics at multi-terawatt scale: waiting is not an option,” Science, 380 (6640), 39 –42 https://doi.org/10.1126/science.adf6957 SCIEAS 0036-8075 (2023). Google Scholar

4. 

T. S. Liang et al., “A review of crystalline silicon bifacial photovoltaic performance characterisation and simulation,” Energy Environ. Sci., 12 (1), 116 –148 https://doi.org/10.1039/C8EE02184H (2019). Google Scholar

5. 

J. Trube and P. Baliozian, International Technology Roadmap for Photovoltaic (ITRPV) 2022 Results, 14th ed.VDMA e.V.( (2023). Google Scholar

6. 

A. Blakers, “Development of the PERC solar cell,” IEEE J. Photovoltaics, 9 (3), 629 –635 https://doi.org/10.1109/JPHOTOV.2019.2899460 IJPEG8 2156-3381 (2019). Google Scholar

7. 

J. Zhao et al., “24% efficient perl silicon solar cell: recent improvements in high efficiency silicon cell research,” Sol. Energy Mater. Sol. Cells, 41 87 –99 https://doi.org/10.1016/0927-0248(95)00117-4 SEMCEQ 0927-0248 (1996). Google Scholar

8. 

J. Zhao, A. Wang and M. A. Green, “24.5% efficiency PERT silicon solar cells on SEH MCZ substrates and cell performance on other SEH CZ and FZ substrates,” Sol. Energy Mater. Sol. Cells, 66 (1), 27 –36 https://doi.org/10.1016/S0927-0248(00)00155-0 SEMCEQ 0927-0248 (2001). Google Scholar

9. 

D. K. Ghosh et al., “Fundamentals, present status and future perspective of TOPCon solar cells: a comprehensive review,” Surf. Interfaces, 30 101917 https://doi.org/10.1016/j.surfin.2022.101917 (2022). Google Scholar

10. 

J. S. Stein, “Bifacial PV Project - PV Performance Modeling Collaborative (PVPMC),” (2023). https://pvpmc.sandia.gov/pv-research/bifacial-pv-project/ (July 2023). Google Scholar

11. 

G. J. Ward, “The RADIANCE lighting simulation and rendering system,” in SIGGRAPH ’94: Proc. 21st Annu. Conf. Comput. Graphics and Interact. Tech., 459 –472 (1994). Google Scholar

12. 

A. Asgharzadeh et al., “A sensitivity study of the impact of installation parameters and system configuration on the performance of bifacial PV arrays,” IEEE J. Photovoltaics, 8 (3), 798 –805 https://doi.org/10.1109/JPHOTOV.2018.2819676 IJPEG8 2156-3381 (2018). Google Scholar

13. 

A. Asgharzadeh et al., “A comparison study of the performance of south/north-facing vs east/west-facing bifacial modules under shading condition,” in IEEE 7th World Conf. Photovoltaics Energy Convers. (WCPEC) (Joint Conf. of 45th IEEE PVSC, 28th PVSEC & 34th EU PVSEC), 1730 –1734 (2018). Google Scholar

14. 

J. S. Stein et al., “Outdoor field performance from bifacial photovoltaic modules and systems,” in IEEE 44th Photovoltaics Spec. Conf. (PVSC), 3184 –3189 (2017). https://doi.org/10.1109/PVSC.2017.8366042 Google Scholar

15. 

D. Riley et al., “A performance model for bifacial PV modules,” in IEEE 44th Photovoltaics Spec. Conf. (PVSC), 3348 –3353 (2017). https://doi.org/10.1109/PVSC.2017.8366045 Google Scholar

16. 

B. Marion et al., “A practical irradiance model for bifacial PV modules,” in IEEE 44th Photovoltaics Spec. Conf. (PVSC), 1537 –1542 (2017). https://doi.org/10.1109/PVSC.2017.8366263 Google Scholar

17. 

C. W. Hansen et al., “Analysis of irradiance models for bifacial PV modules,” in IEEE 43rd Photovoltaics Spec. Conf. (PVSC), 0138 –0143 (2016). https://doi.org/10.1109/PVSC.2016.7749564 Google Scholar

18. 

C. W. Hansen et al., “A detailed model of rear-side irradiance for bifacial PV modules,” in IEEE 44th Photovoltaics Spec. Conf. (PVSC), 1543 –1548 (2017). https://doi.org/10.1109/PVSC.2017.8366707 Google Scholar

19. 

C. Deline et al., “Evaluation and field assessment of bifacial photovoltaic module power rating methodologies,” in IEEE 44th Photovoltaics Spec. Conf. (PVSC), 3698 –3703 (2017). https://doi.org/10.1109/PVSC.2017.8366887 Google Scholar

20. 

C. Deline et al., “Assessment of bifacial photovoltaic module power rating methodologies—inside and out,” IEEE J. Photovoltaics, 7 (2), 575 –580 https://doi.org/10.1109/JPHOTOV.2017.2650565 IJPEG8 2156-3381 (2017). Google Scholar

21. 

“IEC 60904: photovoltaic devices – Part 1-2: measurement of current-voltage characteristics of bifacial photovoltaic (PV) devices (Draft A),” https://pvpmc.sandia.gov/app/uploads/sites/243/2022/11/IEC_60904-1-2_IV_Bifacial_draft_A_ver7_submitted.pdf (2023). Google Scholar

22. 

T. Pasanen et al., “Industrial applicability of antireflection-coating-free black silicon on PERC solar cells and modules,” in 35th Eur. Photovoltaics Solar Energy Conf. and Exhibit., 552 –556 (2018). https://doi.org/10.4229/35thEUPVSEC20182018-2AV.2.5 Google Scholar

23. 

L. L. Ma et al., “Wide-band “black silicon” based on porous silicon,” Appl. Phys. Lett., 88 (17), 171907 https://doi.org/10.1063/1.2199593 APPLAB 0003-6951 (2006). Google Scholar

24. 

C. C. Striemer and P. M. Fauchet, “Dynamic etching of silicon for broadband antireflection applications,” Appl. Phys. Lett., 81 (16), 2980 –2982 https://doi.org/10.1063/1.1514832 APPLAB 0003-6951 (2002). Google Scholar

25. 

H. K. Raut et al., “Anti-reflective coatings: a critical, in-depth review,” Energy Environ. Sci., 4 (10), 3779 https://doi.org/10.1039/c1ee01297e (2011). Google Scholar

26. 

M. Otto et al., “Black silicon photovoltaics,” Adv. Opt. Mater., 3 (2), 147 –164 https://doi.org/10.1002/adom.201400395 2195-1071 (2015). Google Scholar

27. 

A. J. Bett et al., “Wave optical simulation of the light trapping properties of black silicon surface textures,” Opt. Express, 24 (6), A434 https://doi.org/10.1364/OE.24.00A434 OPEXFF 1094-4087 (2016). Google Scholar

28. 

A. Ingenito, O. Isabella and M. Zeman, “Experimental demonstration of 4n2 classical absorption limit in nanotextured ultrathin solar cells with dielectric omnidirectional back reflector,” ACS Photonics, 1 (3), 270 –278 https://doi.org/10.1021/ph4001586 (2014). Google Scholar

29. 

M. Garín et al., “Black ultra-thin crystalline silicon wafers reach the 4n2 absorption limit–Application to IBC Solar Cells,” Small, 19 (39), 2302250 https://doi.org/10.1002/smll.202302250 (2023). Google Scholar

30. 

G. Dingemans and W. M. M. Kessels, “(Invited) Aluminum oxide and other ALD materials for Si surface passivation,” ECS Trans., 41 (2), 293 –301 https://doi.org/10.1149/1.3633680 1938-5862 (2011). Google Scholar

31. 

G. Dingemans and W. M. M. Kessels, “Status and prospects of Al2O3-based surface passivation schemes for silicon solar cells,” J. Vac. Sci. Technol. A: Vac. Surf. Films, 30 (4), 040802 https://doi.org/10.1116/1.4728205 (2012). Google Scholar

32. 

M. Otto et al., “Extremely low surface recombination velocities in black silicon passivated by atomic layer deposition,” Appl. Phys. Lett., 100 (19), 191603 https://doi.org/10.1063/1.4714546 APPLAB 0003-6951 (2012). Google Scholar

33. 

P. Repo et al., “Effective passivation of black silicon surfaces by atomic layer deposition,” IEEE J. Photovoltaics, 3 (1), 90 –94 https://doi.org/10.1109/JPHOTOV.2012.2210031 IJPEG8 2156-3381 (2013). Google Scholar

34. 

H. Savin et al., “Black silicon solar cells with interdigitated back-contacts achieve 22.1% efficiency,” Nat. Nanotechnol., 10 (7), 624 –628 https://doi.org/10.1038/nnano.2015.89 NNAABX 1748-3387 (2015). Google Scholar

35. 

J. Oh, H.-C. Yuan and H. M. Branz, “An 18.2%-efficient black-silicon solar cell achieved through control of carrier recombination in nanostructures,” Nat. Nanotechnol., 7 (11), 743 –748 https://doi.org/10.1038/nnano.2012.166 NNAABX 1748-3387 (2012). Google Scholar

36. 

S. Zhong et al., “Influence of the texturing structure on the properties of black silicon solar cell,” Sol. Energy Mater. Sol. Cells, 108 200 –204 https://doi.org/10.1016/j.solmat.2012.10.001 SEMCEQ 0927-0248 (2013). Google Scholar

37. 

P. Li et al., “Effective optimization of emitters and surface passivation for nanostructured silicon solar cells,” RSC Adv., 6 (106), 104073 –104081 https://doi.org/10.1039/C6RA20945A (2016). Google Scholar

38. 

B. Kafle et al., “On the emitter formation in nanotextured silicon solar cells to achieve improved electrical performances,” Sol. Energy Mater. Sol. Cells, 152 94 –102 https://doi.org/10.1016/j.solmat.2016.03.031 SEMCEQ 0927-0248 (2016). Google Scholar

39. 

J. Benick et al., “High-efficiency n-Type HP mc silicon solar cells,” IEEE J. Photovoltaics, 7 (5), 1171 –1175 https://doi.org/10.1109/JPHOTOV.2017.2714139 IJPEG8 2156-3381 (2017). Google Scholar

40. 

X. Dai et al., “The influence of surface structure on diffusion and passivation in multicrystalline silicon solar cells textured by metal assisted chemical etching (MACE) method,” Sol. Energy Mater. Sol. Cells, 186 42 –49 https://doi.org/10.1016/j.solmat.2018.06.011 SEMCEQ 0927-0248 (2018). Google Scholar

41. 

T. H. Fung et al., “Improved emitter performance of RIE black silicon through the application of in-situ oxidation during POCl3 diffusion,” Sol. Energy Mater. Sol. Cells, 210 110480 https://doi.org/10.1016/j.solmat.2020.110480 SEMCEQ 0927-0248 (2020). Google Scholar

42. 

G. Scardera et al., “On the enhanced phosphorus doping of nanotextured black silicon,” IEEE J. Photovoltaics, 11 (2), 298 –305 https://doi.org/10.1109/JPHOTOV.2020.3047420 IJPEG8 2156-3381 (2021). Google Scholar

43. 

G. Von Gastrow et al., “Recombination processes in passivated boron-implanted black silicon emitters,” J. Appl. Phys., 121 (18), 185706 https://doi.org/10.1063/1.4983297 JAPIAU 0021-8979 (2017). Google Scholar

44. 

K. Chen et al., “Harnessing carrier multiplication in silicon solar cells using UV photons,” IEEE Photonics Technol. Lett., 33 (24), 1415 –1418 https://doi.org/10.1109/LPT.2021.3124307 IPTLEL 1041-1135 (2021). Google Scholar

45. 

L. Sainiemi et al., “Non-reflecting silicon and polymer surfaces by plasma etching and replication,” Adv. Mater., 23 (1), 122 –126 https://doi.org/10.1002/adma.201001810 ADVMEW 0935-9648 (2011). Google Scholar

46. 

M. M. Plakhotnyuk et al., “Low surface damage dry etched black silicon,” J. Appl. Phys., 122 (14), 143101 https://doi.org/10.1063/1.4993425 JAPIAU 0021-8979 (2017). Google Scholar

47. 

B. Iandolo et al., “Black silicon with ultra-low surface recombination velocity fabricated by inductively coupled power plasma,” Physica Status Solidi (RRL) – Rapid Res. Lett., 13 (2), 1800477 https://doi.org/10.1002/pssr.201800477 (2019). Google Scholar

48. 

X. Li and P. W. Bohn, “Metal-assisted chemical etching in HF/H2O2 produces porous silicon,” Appl. Phys. Lett., 77 (16), 2572 –2574 https://doi.org/10.1063/1.1319191 APPLAB 0003-6951 (2000). Google Scholar

49. 

Z. Huang et al., “Metal-assisted chemical etching of silicon: a review: in memory of Prof. Ulrich Gösele,” Adv. Mater., 23 (2), 285 –308 https://doi.org/10.1002/adma.201001784 ADVMEW 0935-9648 (2011). Google Scholar

50. 

F. Toor et al., “Nanostructured silicon via metal assisted catalyzed etch (MACE): chemistry fundamentals and pattern engineering,” Nanotechnology, 27 (41), 412003 https://doi.org/10.1088/0957-4484/27/41/412003 NNOTER 0957-4484 (2016). Google Scholar

51. 

F. Toor et al., “Metal assisted catalyzed etched (MACE) black Si: optics and device physics,” Nanoscale, 8 (34), 15448 –15466 https://doi.org/10.1039/C6NR04506E NANOHL 2040-3364 (2016). Google Scholar

52. 

Y.-T. Lu and A. R. Barron, “Nanopore-type black silicon anti-reflection layers fabricated by a one-step silver-assisted chemical etching,” Phys. Chem. Chem. Phys., 15 (24), 9862 https://doi.org/10.1039/c3cp51835c PPCPFQ 1463-9076 (2013). Google Scholar

53. 

Y. Matsui and S. Adachi, “Optical properties of “black silicon” formed by catalytic etching of Au/Si(100) wafers,” J. Appl. Phys., 113 (17), 173502 https://doi.org/10.1063/1.4803152 JAPIAU 0021-8979 (2013). Google Scholar

54. 

D. T. Cao et al., “Effect of AgNO3 concentration on structure of aligned silicon nanowire arrays fabricated via silver-assisted chemical etching,” Int. J. Nanotechnol., 10 (3/4), 343 https://doi.org/10.1504/IJNT.2013.053147 1475-7435 (2013). Google Scholar

55. 

K. Chen et al., “MACE nano-texture process applicable for both single- and multi-crystalline diamond-wire sawn Si solar cells,” Sol. Energy Mater. Sol. Cells, 191 1 –8 https://doi.org/10.1016/j.solmat.2018.10.015 SEMCEQ 0927-0248 (2019). Google Scholar

56. 

S. K. Sundaram and E. Mazur, “Inducing and probing non-thermal transitions in semiconductors using femtosecond laser pulses,” Nat. Mater., 1 (4), 217 –224 https://doi.org/10.1038/nmat767 NMAACR 1476-1122 (2002). Google Scholar

57. 

M. Huang et al., “The morphological and optical characteristics of femtosecond laser-induced large-area micro/nanostructures on GaAs, Si, and brass,” Opt. Express, 18 (S4), A600 https://doi.org/10.1364/OE.18.00A600 OPEXFF 1094-4087 (2010). Google Scholar

58. 

C. H. Crouch et al., “Infrared absorption by sulfur-doped silicon formed by femtosecond laser irradiation,” Appl. Phys. A, 79 (7), 1635 –1641 https://doi.org/10.1007/s00339-004-2676-0 (2004). Google Scholar

59. 

Y. Peng et al., “Optimal proportional relation between laser power and pulse number for the fabrication of surface-microstructured silicon,” Appl. Opt., 50 (24), 4765 https://doi.org/10.1364/AO.50.004765 APOPAI 0003-6935 (2011). Google Scholar

60. 

B. Radfar, F. Es and R. Turan, “Effects of different laser modified surface morphologies and post-texturing cleanings on c-Si solar cell performance,” Renew. Energy, 145 2707 –2714 https://doi.org/10.1016/j.renene.2019.08.031 RNENE3 0960-1481 (2020). Google Scholar

61. 

B. Radfar et al., “Optoelectronic properties of black silicon fabricated by femtosecond laser in ambient air: exploring a large parameter space,” Opt. Lett., 48 (5), 1224 https://doi.org/10.1364/OL.481890 OPLEDP 0146-9592 (2023). Google Scholar

62. 

T. P. Pasanen et al., “Impact of black silicon on light- and elevated temperature-induced degradation in industrial passivated emitter and rear cells,” Prog. Photovoltaics Res. Appl., 27 (11), 918 –925 https://doi.org/10.1002/pip.3088 PPHOED 1062-7995 (2019). Google Scholar

63. 

T. P. Pasanen et al., “Black silicon significantly enhances phosphorus diffusion gettering,” Sci. Rep., 8 (1), 1991 https://doi.org/10.1038/s41598-018-20494-y (2018). Google Scholar

64. 

M. U. Khan et al., “The role of metal-catalyzed chemical etching black silicon in the reduction of light- and elevated temperature-induced degradation in P-type multicrystalline wafers,” IEEE J. Photovoltaics, 11 (3), 627 –633 https://doi.org/10.1109/JPHOTOV.2021.3059426 IJPEG8 2156-3381 (2021). Google Scholar

65. 

P. Ortega et al., “Black silicon back-contact module with wide light acceptance angle,” Prog. Photovoltaics Res. Appl., 28 (3), 210 –216 https://doi.org/10.1002/pip.3231 PPHOED 1062-7995 (2020). Google Scholar

66. 

“Best research-cell efficiency chart,” https://www.nrel.gov/pv/cell-efficiency.html (2023). Google Scholar

67. 

H. M. Wikoff, S. B. Reese and M. O. Reese, “Embodied energy and carbon from the manufacture of cadmium telluride and silicon photovoltaics,” Joule, 6 (7), 1710 –1725 https://doi.org/10.1016/j.joule.2022.06.006 (2022). Google Scholar

68. 

V. Bermudez and A. Perez-Rodriguez, “Understanding the cell-to-module efficiency gap in Cu(In,Ga)(S,Se)2 photovoltaics scale-up,” Nat. Energy, 3 (6), 466 –475 https://doi.org/10.1038/s41560-018-0177-1 (2018). Google Scholar

69. 

T. Song, A. Kanevce and J. R. Sites, “Emitter/absorber interface of CdTe solar cells,” J. Appl. Phys., 119 (23), 233104 https://doi.org/10.1063/1.4953820 JAPIAU 0021-8979 (2016). Google Scholar

70. 

M. A. Green et al., “Solar cell efficiency tables (Version 61),” Prog. Photovoltaics Res. Appl., 31 (1), 3 –16 https://doi.org/10.1002/pip.3646 PPHOED 1062-7995 (2023). Google Scholar

71. 

T. Ablekim et al., “Exceeding 200 ns lifetimes in polycrystalline CdTe solar cells,” Solar RRL, 5 (8), 2100173 https://doi.org/10.1002/solr.202100173 (2021). Google Scholar

72. 

J. M. Kephart et al., “Sputter-deposited oxides for interface passivation of CdTe photovoltaics,” IEEE J. Photovoltaics, 8 (2), 587 –593 https://doi.org/10.1109/JPHOTOV.2017.2787021 IJPEG8 2156-3381 (2018). Google Scholar

73. 

J. M. Burst et al., “CdTe solar cells with open-circuit voltage breaking the 1 V barrier,” Nat. Energy, 1 (3), 16015 https://doi.org/10.1038/nenergy.2016.15 (2016). Google Scholar

74. 

J. M. Burst et al., “Carrier density and lifetime for different dopants in single-crystal and polycrystalline CdTe,” APL Mater., 4 (11), 116102 https://doi.org/10.1063/1.4966209 (2016). Google Scholar

75. 

D. Krasikov et al., “Comparative study of As and Cu doping stability in CdSeTe absorbers,” Sol. Energy Mater. Sol. Cells, 224 111012 https://doi.org/10.1016/j.solmat.2021.111012 SEMCEQ 0927-0248 (2021). Google Scholar

76. 

A. Kanevce et al., “The roles of carrier concentration and interface, bulk, and grain-boundary recombination for 25% efficient CdTe solar cells,” J. Appl. Phys., 121 (21), https://doi.org/10.1063/1.4984320 JAPIAU 0021-8979 (2017). Google Scholar

77. 

J. Moseley et al., “Impact of dopant-induced optoelectronic tails on open-circuit voltage in arsenic-doped Cd(Se)Te solar cells,” J. Appl. Phys., 128 (10), 103105 https://doi.org/10.1063/5.0018955 JAPIAU 0021-8979 (2020). Google Scholar

78. 

R. Mallick et al., “Arsenic-doped CdSeTe solar cells achieve world record 22.3% efficiency,” IEEE J. Photovoltaics, 13 510 –515 https://doi.org/10.1109/JPHOTOV.2023.3282581 IJPEG8 2156-3381 (2023). Google Scholar

79. 

G. Xiong and W. Metzger, Thin Film Cadmium Telluride Photovoltaics, Elsevier( (2022). Google Scholar

80. 

W. K. Metzger et al., “Exceeding 20% efficiency with in situ group V doping in polycrystalline CdTe solar cells,” Nat. Energy, 4 (10), 837 –845 https://doi.org/10.1038/s41560-019-0446-7 (2019). Google Scholar

81. 

W. K. Metzger et al., “As-doped CdSeTe solar cells achieving 22% efficiency with −0.23%/°C temperature coefficient,” IEEE J. Photovoltaics, 12 (6), 1435 –1438 https://doi.org/10.1109/JPHOTOV.2022.3201479 IJPEG8 2156-3381 (2022). Google Scholar

82. 

K. K. Subedi et al., “Enabling bifacial thin film devices by developing a back surface field using CuxAlOy,” Nano Energy, 83 105827 https://doi.org/10.1016/j.nanoen.2021.105827 (2021). Google Scholar

83. 

A. W. Walker et al., “The effects of absorption and recombination on quantum dot multijunction solar cell efficiency,” IEEE J. Photovoltaics, 3 (3), 1118 –1124 https://doi.org/10.1109/JPHOTOV.2013.2257920 IJPEG8 2156-3381 (2013). Google Scholar

84. 

R. M. France et al., “Triple-junction solar cells with 39.5% terrestrial and 34.2% space efficiency enabled by thick quantum well superlattices,” Joule, 6 (5), 1121 –1135 https://doi.org/10.1016/j.joule.2022.04.024 (2022). Google Scholar

85. 

P. Schygulla et al., “Wafer-bonded two-terminal III–V//Si triple-junction solar cell with power conversion efficiency of 36.1% at AM1.5g,” Prog. Photovoltaics Res. Appl., (). Google Scholar

86. 

F. Dimroth et al., “Four-junction wafer-bonded concentrator solar cells,” IEEE J. Photovoltaics, 6 (1), 343 –349 https://doi.org/10.1109/JPHOTOV.2015.2501729 IJPEG8 2156-3381 (2016). Google Scholar

87. 

R. Cariou et al., “III–V-on-silicon solar cells reaching 33% photoconversion efficiency in two-terminal configuration,” Nat. Energy, 3 (4), 326 –333 https://doi.org/10.1038/s41560-018-0125-0 (2018). Google Scholar

88. 

T. J. Grassman et al., “GaAs0.75P0.25/Si dual-junction solar cells grown by MBE and MOCVD,” IEEE J. Photovoltaics, 6 (1), 326 –331 https://doi.org/10.1109/JPHOTOV.2015.2493365 IJPEG8 2156-3381 (2016). Google Scholar

89. 

J. F. Geisz et al., “Building a six-junction inverted metamorphic concentrator solar cell,” IEEE J. Photovoltaics, 8 (2), 626 –632 https://doi.org/10.1109/JPHOTOV.2017.2778567 IJPEG8 2156-3381 (2018). Google Scholar

90. 

R. M. France et al., “Metamorphic epitaxy for multijunction solar cells,” MRS Bull., 41 (3), 202 –209 https://doi.org/10.1557/mrs.2016.25 MRSBEA 0883-7694 (2016). Google Scholar

91. 

M. A. Green et al., “Solar cell efficiency tables (version 54),” Prog. Photovoltaics Res. Appl., 27 (7), 565 –575 https://doi.org/10.1002/pip.3171 PPHOED 1062-7995 (2019). Google Scholar

92. 

K. T. VanSant, A. C. Tamboli and E. L. Warren, “III-V-on-Si tandem solar cells,” Joule, 5 (3), 514 –518 https://doi.org/10.1016/j.joule.2021.01.010 (2021). Google Scholar

93. 

K. Makita et al., “III-V//CuxIn1yGaySe2 multijunction solar cells with 27.2% efficiency fabricated using modified smart stack technology with Pd nanoparticle array and adhesive material,” Prog. Photovoltaics Res. Appl., 29 (8), 887 –898 https://doi.org/10.1002/pip.3398 PPHOED 1062-7995 (2021). Google Scholar

94. 

S. Essig et al., “Raising the one-sun conversion efficiency of III–V/Si solar cells to 32.8% for two junctions and 35.9% for three junctions,” Nat. Energy, 2 (9), 17144 https://doi.org/10.1038/nenergy.2017.144 (2017). Google Scholar

95. 

M. A. Green et al., “Solar cell efficiency tables (version 50),” Prog. Photovoltaics Res. Appl., 25 (7), 668 –676 https://doi.org/10.1002/pip.2909 PPHOED 1062-7995 (2017). Google Scholar

96. 

C. E. Valdivia and K. Hinzer, “Subcell segmentation for current matching and design flexibility in multijunction solar cells,” IEEE J. Photovoltaics, 10 (5), 1329 –1339 https://doi.org/10.1109/JPHOTOV.2020.3005630 IJPEG8 2156-3381 (2020). Google Scholar

97. 

M. M. Wilkins et al., “Ripple-free boost-mode power supply using photonic power conversion,” IEEE Trans. Power Electron., 34 (2), 1054 –1064 https://doi.org/10.1109/TPEL.2018.2843158 ITPEE8 0885-8993 (2019). Google Scholar

98. 

S. Fafard et al., “Ultrahigh efficiencies in vertical epitaxial heterostructure architectures,” Appl. Phys. Lett., 108 (7), 071101 https://doi.org/10.1063/1.4941240 APPLAB 0003-6951 (2016). Google Scholar

99. 

S. Fafard et al., “High-photovoltage GaAs vertical epitaxial monolithic heterostructures with 20 thin p/n junctions and a conversion efficiency of 60%,” Appl. Phys. Lett., 109 (13), 131107 https://doi.org/10.1063/1.4964120 APPLAB 0003-6951 (2016). Google Scholar

100. 

S. Fafard and D. P. Masson, “High-efficiency and high-power multijunction InGaAs/InP photovoltaic laser power converters for 1470 nm,” Photonics, 9 (7), 438 https://doi.org/10.3390/photonics9070438 (2022). Google Scholar

101. 

M. Feifel et al., “Epitaxial GaInP/GaAs/Si triple-junction solar cell with 25.9% AM1.5g efficiency enabled by transparent metamorphic AlxGa1xAsyP1y step-graded buffer structures,” Solar RRL, 5 2000763 https://doi.org/10.1002/solr.202000763 (2021). Google Scholar

102. 

M. A. Green et al., “Solar cell efficiency tables (version 51),” Prog. Photovoltaics Res. Appl., 26 (1), 3 –12 https://doi.org/10.1002/pip.2978 PPHOED 1062-7995 (2018). Google Scholar

103. 

M. A. Green et al., “Solar cell efficiency tables (version 56),” Prog. Photovoltaics Res. Appl., 28 (7), 629 –638 https://doi.org/10.1002/pip.3303 PPHOED 1062-7995 (2020). Google Scholar

104. 

M. A. Green et al., “Solar cell efficiency tables (Version 53),” Prog. Photovoltaics Res. Appl., 27 (1), 3 –12 https://doi.org/10.1002/pip.3102 PPHOED 1062-7995 (2019). Google Scholar

105. 

, “Press release: sharp develops concentrator solar cell with world’s highest conversion efficiency of 43.5%,” http://sharp-world.com/corporate/news/120531.html (2012). Google Scholar

106. 

, “News release: NREL demonstrates 45.7% efficiency for concentrator solar cell,” https://www.nrel.gov/news/press/2014/15436.html (2014). Google Scholar

107. 

I. Almansouri et al., “Supercharging silicon solar cell performance by means of multijunction concept,” IEEE J. Photovoltaics, 5 (3), 968 –976 https://doi.org/10.1109/JPHOTOV.2015.2395140 IJPEG8 2156-3381 (2015). Google Scholar

108. 

M. Levinshtein, S. Rumyantsev and M. Shur, Handbook Series on Semiconductor Parameters: Volume 2: Ternary And Quaternary III-V Compounds, World Scientific( (1996). Google Scholar

109. 

J. M. Gee and G. F. Virshup, “A 31%-efficient GaAs/silicon mechanically stacked, multijunction concentrator solar cell,” in Conf. Rec. of the Twentieth IEEE Photovoltaics Spec. Conf., (1988). https://doi.org/10.1109/PVSC.1988.105803 Google Scholar

110. 

S. Essig et al., “Boosting the efficiency of III-V/Si tandem solar cells,” in IEEE 43rd Photovoltaics Spec. Conf. (PVSC), (2016). https://doi.org/10.1109/PVSC.2016.7749987 Google Scholar

111. 

M. Schnabel et al., “Three-terminal III–V/Si tandem solar cells enabled by a transparent conductive adhesive,” Sustain. Energy Fuels, 4 (2), 549 –558 https://doi.org/10.1039/C9SE00893D (2020). Google Scholar

112. 

H. Mizuno et al., “Integration of Si heterojunction solar cells with III–V solar cells by the Pd nanoparticle array-mediated “Smart Stack” approach,” ACS Appl. Mater. Interfaces, 14 (9), 11322 –11329 https://doi.org/10.1021/acsami.1c22458 (2022). Google Scholar

113. 

K. Tanabe, K. Watanabe and Y. Arakawa, “III-V/Si hybrid photonic devices by direct fusion bonding,” Sci. Rep., 2 (1), 349 https://doi.org/10.1038/srep00349 (2012). Google Scholar

114. 

K. Derendorf et al., “Fabrication of GaInP/GaAs//Si solar cells by surface activated direct wafer bonding,” IEEE J. Photovoltaics, 3 (4), 1423 –1428 https://doi.org/10.1109/JPHOTOV.2013.2273097 IJPEG8 2156-3381 (2013). Google Scholar

115. 

S. Fan et al., “Current-matched III–V/Si epitaxial tandem solar cells with 25.0% efficiency,” Cell Rep. Phys. Sci., 1 (9), 100208 https://doi.org/10.1016/j.xcrp.2020.100208 (2020). Google Scholar

116. 

B. Kunert et al., “How to control defect formation in monolithic III/V hetero-epitaxy on (100) Si? A critical review on current approaches,” Semicond. Sci. Technol., 33 (9), 093002 https://doi.org/10.1088/1361-6641/aad655 SSTEET 0268-1242 (2018). Google Scholar

117. 

J. T. Boyer et al., “Correlation of early-stage growth process conditions with dislocation evolution in MOCVD-based GaP/Si heteroepitaxy,” J. Cryst. Growth, 571 126251 https://doi.org/10.1016/j.jcrysgro.2021.126251 JCRGAE 0022-0248 (2021). Google Scholar

118. 

R. D. Hool et al., “Challenges of relaxed n-type GaP on Si and strategies to enable low threading dislocation density,” J. Appl. Phys., 130 (24), 243104 https://doi.org/10.1063/5.0073525 JAPIAU 0021-8979 (2021). Google Scholar

119. 

S. Essig et al., “Fast atom beam-activated n-Si/n-GaAs wafer bonding with high interfacial transparency and electrical conductivity,” J. Appl. Phys., 113 (20), 203512 https://doi.org/10.1063/1.4807905 JAPIAU 0021-8979 (2013). Google Scholar

120. 

H. Liu et al., “A worldwide theoretical comparison of outdoor potential for various silicon-based tandem module architecture,” Cell Rep. Phys. Sci., 1 (4), 100037 https://doi.org/10.1016/j.xcrp.2020.100037 (2020). Google Scholar

121. 

M. Rienäcker et al., “Mechanically stacked dual-junction and triple-junction III-V/Si-IBC cells with efficiencies of 31.5 % and 35.4,” in 33rd Eur. Photovoltaics Solar Energy Conf. and Exhibit., (2017). https://doi.org/10.4229/EUPVSEC20172017-1AP.1.2 Google Scholar

122. 

K. Makita et al., “III-V//Si multijunction solar cells with 30% efficiency using smart stack technology with Pd nanoparticle array,” Prog. Photovoltaics Res. Appl., 28 (1), 16 –24 https://doi.org/10.1002/pip.3200 PPHOED 1062-7995 (2020). Google Scholar

123. 

J. Simon et al., “III-V-based optoelectronics with low-cost dynamic hydride vapor phase epitaxy,” Crystals, 9 (1), 3 https://doi.org/10.3390/cryst9010003 CRYSBC 2073-4352 (2018). Google Scholar

124. 

J. Chen and C. E. Packard, “Controlled spalling-based mechanical substrate exfoliation for III-V solar cells: a review,” Sol. Energy Mater. Sol. Cells, 225 111018 https://doi.org/10.1016/j.solmat.2021.111018 SEMCEQ 0927-0248 (2021). Google Scholar

125. 

K. Schulte, High-Efficiency, Low-Cost III-V Solar Cells by Dynamic Hydride Vapor Phase Epitaxy Coupled with Rapid, Polishing-Free Wafer Reuse through Orientation-Optimized (110) Spalling, National Renewable Energy Lab. (NREL), Golden, Colorado (2021). Google Scholar

126. 

K. A. Horowitz et al., A Techno-Economic Analysis and Cost Reduction Roadmap for III-V Solar Cells, (2018). Google Scholar

127. 

A. R. Bowring et al., “Reverse bias behavior of halide perovskite solar cells,” Adv. Energy Mater., 8 (8), 1702365 https://doi.org/10.1002/aenm.201702365 ADEMBC 1614-6840 (2018). Google Scholar

128. 

D. Bogachuk et al., “Perovskite photovoltaic devices with carbon-based electrodes withstanding reverse-bias voltages up to –9 V and surpassing IEC 61215:2016 International Standard,” Solar RRL, 6 (3), 2100527 https://doi.org/10.1002/solr.202100527 (2022). Google Scholar

129. 

L. Bertoluzzi et al., “Incorporating electrochemical halide oxidation into drift-diffusion models to explain performance losses in perovskite solar cells under prolonged reverse bias,” Adv. Energy Mater., 11 (10), 2002614 https://doi.org/10.1002/aenm.202002614 ADEMBC 1614-6840 (2021). Google Scholar

130. 

I. E. Gould et al., “In-operando characterization of P-I-N perovskite solar cells under reverse bias,” in IEEE 48th Photovoltaics Spec. Conf. (PVSC), 1365 –1367 (2021). https://doi.org/10.1109/PVSC43889.2021.9518723 Google Scholar

131. 

R. A. Z. Razera et al., “Instability of p–i–n perovskite solar cells under reverse bias,” J. Mater. Chem. A, 8 (1), 242 –250 https://doi.org/10.1039/C9TA12032G (2020). Google Scholar

132. 

W. Li et al., “Sparkling hot spots in perovskite solar cells under reverse bias,” ChemPhysMater, 1 (1), 71 –76 https://doi.org/10.1016/j.chphma.2021.10.001 (2022). Google Scholar

133. 

Z. Ni et al., “Evolution of defects during the degradation of metal halide perovskite solar cells under reverse bias and illumination,” Nat. Energy, 7 (1), 65 –73 https://doi.org/10.1038/s41560-021-00949-9 (2021). Google Scholar

134. 

J. Qian et al., “Destructive reverse bias pinning in perovskite/silicon tandem solar modules caused by perovskite hysteresis under dynamic shading,” Sustain. Energy Fuels, 4 (8), 4067 –4075 https://doi.org/10.1039/C9SE01246J (2020). Google Scholar

135. 

S. G. Motti et al., “Controlling competing photochemical reactions stabilizes perovskite solar cells,” Nat. Photonics, 13 (8), 532 –539 https://doi.org/10.1038/s41566-019-0435-1 NPAHBY 1749-4885 (2019). Google Scholar

136. 

G. Y. Kim et al., “Photo-effect on ion transport in mixed cation and halide perovskites and implications for photo-demixing,” Angew. Chem., 133 (2), 833 –839 https://doi.org/10.1002/ange.202005853 ANCEAD 0044-8249 (2021). Google Scholar

137. 

S. Bitton and N. Tessler, “Perovskite ionics – elucidating degradation mechanisms in perovskite solar cells via device modelling and iodine chemistry,” Energy Environ. Sci., 16 (6), 2621 –2628 https://doi.org/10.1039/D3EE00881A (2023). Google Scholar

138. 

Z. Xu et al., “Halogen redox shuttle explains voltage-induced halide redistribution in mixed-halide perovskite devices,” ACS Energy Lett., 8 (1), 513 –520 https://doi.org/10.1021/acsenergylett.2c02385 (2023). Google Scholar

139. 

E. J. Wolf et al., “Designing modules to prevent reverse bias degradation in perovskite solar cells when partial shading occurs,” Solar RRL, 6 (3), 2100239 https://doi.org/10.1002/solr.202100239 (2022). Google Scholar

140. 

B. Chen et al., “Insights into the development of monolithic perovskite/silicon tandem solar cells,” Adv. Energy Mater., 12 (4), 2003628 https://doi.org/10.1002/aenm.202003628 ADEMBC 1614-6840 (2022). Google Scholar

141. 

L. A. Zafoschnig, S. Nold and J. C. Goldschmidt, “The race for lowest costs of electricity production: techno-economic analysis of silicon, perovskite and tandem solar cells,” IEEE J. Photovoltaics, 10 (6), 1632 –1641 https://doi.org/10.1109/JPHOTOV.2020.3024739 IJPEG8 2156-3381 (2020). Google Scholar

142. 

F. Fu et al., “Monolithic perovskite-silicon tandem solar cells: from the lab to fab?,” Adv. Mater., 34 (24), 2106540 https://doi.org/10.1002/adma.202106540 ADVMEW 0935-9648 (2022). Google Scholar

143. 

J. P. Mailoa et al., “A 2-terminal perovskite/silicon multijunction solar cell enabled by a silicon tunnel junction,” Appl. Phys. Lett., 106 (12), 121105 https://doi.org/10.1063/1.4914179 APPLAB 0003-6951 (2015). Google Scholar

144. 

C. D. Bailie et al., “Mechanically stacked and monolithically integrated perovskite/silicon tandems and the challenges for high efficiency,” in IEEE 42nd Photovoltaics Spec. Conf. (PVSC), (2015). Google Scholar

145. 

J. Liu et al., “Efficient and stable perovskite-silicon tandem solar cells through contact displacement by MgFx,” Science, 377 (6603), 302 –306 https://doi.org/10.1126/science.abn8910 SCIEAS 0036-8075 (2022). Google Scholar

146. 

F. Sahli et al., “Fully textured monolithic perovskite/silicon tandem solar cells with 25.2% power conversion efficiency,” Nat. Mater., 17 (9), 820 –826 https://doi.org/10.1038/s41563-018-0115-4 NMAACR 1476-1122 (2018). Google Scholar

147. 

E. Aydin et al., “Interplay between temperature and bandgap energies on the outdoor performance of perovskite/silicon tandem solar cells,” Nat. Energy, 5 (11), 851 –859 https://doi.org/10.1038/s41560-020-00687-4 (2020). Google Scholar

148. 

Z. Song et al., “Wide-bandgap, low-bandgap, and tandem perovskite solar cells,” Semicond. Sci. Technol., 34 (9), 093001 https://doi.org/10.1088/1361-6641/ab27f7 SSTEET 0268-1242 (2019). Google Scholar

149. 

F. Gota et al., “Energy yield modelling of textured perovskite/silicon tandem photovoltaics with thick perovskite top cells,” Opt. Express, 30 (9), 14172 https://doi.org/10.1364/OE.447069 OPEXFF 1094-4087 (2022). Google Scholar

150. 

S. Akhil et al., “Review on perovskite silicon tandem solar cells: status and prospects 2T, 3T and 4T for real world conditions,” Mater. Des., 211 110138 https://doi.org/10.1016/j.matdes.2021.110138 MADSD2 0264-1275 (2021). Google Scholar

151. 

Y. Huang et al., “Limitations and solutions for achieving high-performance perovskite tandem photovoltaics,” Nano Energy, 88 106219 https://doi.org/10.1016/j.nanoen.2021.106219 (2021). Google Scholar

152. 

Y. Hou et al., “Efficient tandem solar cells with solution-processed perovskite on textured crystalline silicon,” Science, 367 (6482), 1135 –1140 https://doi.org/10.1126/science.aaz3691 SCIEAS 0036-8075 (2020). Google Scholar

153. 

P. Tockhorn et al., “Nano-optical designs for high-efficiency monolithic perovskite–silicon tandem solar cells,” Nat. Nanotechnol., 17 1214 –1221 https://doi.org/10.1038/s41565-022-01228-8 (2022). Google Scholar

154. 

Interactive best research-cell efficiency chart, https://www.nrel.gov/pv/interactive-cell-efficiency.html (30 November 2023). Google Scholar

155. 

T. Duong et al., “Bulk incorporation with 4-methylphenethylammonium chloride for efficient and stable methylammonium-free perovskite and perovskite-silicon tandem solar cells,” Adv. Energy Mater., 13 (9), 2203607 https://doi.org/10.1002/aenm.202203607 ADEMBC 1614-6840 (2023). Google Scholar

156. 

H. Kanda et al., “Three-terminal perovskite/integrated back contact silicon tandem solar cells under low light intensity conditions,” Interdiscipl. Mater., 1 (1), 148 –156 https://doi.org/10.1002/idm2.12006 (2022). Google Scholar

157. 

P. Tockhorn et al., “Three-terminal perovskite/silicon tandem solar cells with top and interdigitated rear contacts,” ACS Appl. Energy Mater., 3 (2), 1381 –1392 https://doi.org/10.1021/acsaem.9b01800 (2020). Google Scholar

158. 

C. Chen et al., “Perovskite solar cells based on screen-printed thin films,” Nature, 612 (7939), 266 –271 https://doi.org/10.1038/s41586-022-05346-0 (2022). Google Scholar

159. 

Z. Li et al., “Scalable fabrication of perovskite solar cells,” Nat. Rev. Mater., 3 (4), 18017 https://doi.org/10.1038/natrevmats.2018.17 (2018). Google Scholar

160. 

Y. Rong et al., “Toward industrial-scale production of perovskite solar cells: screen printing, slot-die coating, and emerging techniques,” J. Phys. Chem. Lett., 9 (10), 2707 –2713 https://doi.org/10.1021/acs.jpclett.8b00912 (2018). Google Scholar

161. 

S. P. Dunfield et al., “From defects to degradation: a mechanistic understanding of degradation in perovskite solar cell devices and modules,” Adv. Energy Mater., 10 (26), 1904054 https://doi.org/10.1002/aenm.201904054 ADEMBC 1614-6840 (2020). Google Scholar

162. 

R. Azm et al., “Damp heat–stable perovskite solar cells with tailored-dimensionality 2D/3D heterojunctions,” Science, 376 (6588), 73 –77 https://doi.org/10.1126/science.abm5784 SCIEAS 0036-8075 (2022). Google Scholar

163. 

H. Lin et al., “Silicon heterojunction solar cells with up to 26.81% efficiency achieved by electrically optimized nanocrystalline-silicon hole contact layers,” Nat. Energy, 8 789 –799 https://doi.org/10.1038/s41560-023-01255-2 (2023). Google Scholar

164. 

T. Rößler et al., “Production of Full-size Perovskite Silicon Heterojunction Tandem PV Modules with more than 23% Efficiency and more than 430 Wp,” (2022). (2022). Google Scholar

165. 

K. Chong et al., “Realizing 19.05% efficiency polymer solar cells by progressively improving charge extraction and suppressing charge recombination,” Adv. Mater., 34 (13), 2109516 https://doi.org/10.1002/adma.202109516 ADVMEW 0935-9648 (2022). Google Scholar

166. 

C. He et al., “Manipulating the D:A interfacial energetics and intermolecular packing for 19.2% efficiency organic photovoltaics,” Energy Environ. Sci., 15 (6), 2537 –2544 https://doi.org/10.1039/D2EE00595F (2022). Google Scholar

167. 

R. Sun et al., “Single-junction organic solar cells with 19.17% efficiency enabled by introducing one asymmetric guest acceptor,” Adv. Mater., 34 (26), 2110147 https://doi.org/10.1002/adma.202110147 ADVMEW 0935-9648 (2022). Google Scholar

168. 

J. Yuan et al., “Single-junction organic solar cell with over 15% efficiency using fused-ring acceptor with electron-deficient core,” Joule, 3 (4), 1140 –1151 https://doi.org/10.1016/j.joule.2019.01.004 (2019). Google Scholar

169. 

Y. Cui et al., “Over 16% efficiency organic photovoltaic cells enabled by a chlorinated acceptor with increased open-circuit voltages,” Nat. Commun., 10 (1), 2515 https://doi.org/10.1038/s41467-019-10351-5 NCAOBW 2041-1723 (2019). Google Scholar

170. 

Z. Luo et al., “Fine-tuning energy levels via asymmetric end groups enables polymer solar cells with efficiencies over 17%,” Joule, 4 (6), 1236 –1247 https://doi.org/10.1016/j.joule.2020.03.023 (2020). Google Scholar

171. 

L. Yan et al., “Regioisomer-free difluoro-monochloro terminal-based hexa-halogenated acceptor with optimized crystal packing for efficient binary organic solar cells,” Angew. Chem. Int. Ed., 134 (46), e202209454 https://doi.org/10.1002/anie.202209454 ACIEAY 0570-0833 (2022). Google Scholar

172. 

D. Yuk et al., “Simplified Y6-based nonfullerene acceptors: in-depth study on molecular structure–property relation, molecular dynamics simulation, and charge dynamics,” Small, 19 (10), 2206547 https://doi.org/10.1002/smll.202206547 SMALBC 1613-6810 (2023). Google Scholar

173. 

H. Yu et al., “A vinylene-linker-based polymer acceptor featuring a coplanar and rigid molecular conformation enables high-performance all-polymer solar cells with over 17% efficiency,” Adv. Mater., 34 (27), 2200361 https://doi.org/10.1002/adma.202200361 ADVMEW 0935-9648 (2022). Google Scholar

174. 

C. Sun et al., “Dimerized small-molecule acceptors enable efficient and stable organic solar cells,” Joule, 7 (2), 416 –430 https://doi.org/10.1016/j.joule.2023.01.009 (2023). Google Scholar

175. 

G. Wang et al., “Large-area organic solar cells: material requirements, modular designs, and printing methods,” Adv. Mater., 31 (45), 1805089 https://doi.org/10.1002/adma.201805089 ADVMEW 0935-9648 (2019). Google Scholar

176. 

J. S. Park et al., “Material design and device fabrication strategies for stretchable organic solar cells,” Adv. Mater., 34 (31), 2201623 https://doi.org/10.1002/adma.202201623 ADVMEW 0935-9648 (2022). Google Scholar

177. 

Y. Li et al., “Semitransparent organic photovoltaics for building-integrated photovoltaic applications,” Nat. Rev. Mater., 8 (3), 186 –201 https://doi.org/10.1038/s41578-022-00514-0 (2022). Google Scholar

178. 

J. W. Lee et al., “Intrinsically stretchable and non-halogenated solvent processed polymer solar cells enabled by hydrophilic spacer-incorporated polymers,” Adv. Energy Mater., 12 (46), 2202224 https://doi.org/10.1002/aenm.202202224 ADEMBC 1614-6840 (2022). Google Scholar

179. 

F. Qi et al., “Dimer acceptor adopting a flexible linker for efficient and durable organic solar cells,” Angew. Chem. Int. Ed., 62 (21), e202303066 https://doi.org/10.1002/anie.202303066 ACIEAY 0570-0833 (2023). Google Scholar

180. 

M. Niggemann et al., “Light trapping in organic solar cells: light trapping in organic solar cells,” Physica Status Solidi (A), 205 (12), 2862 –2874 https://doi.org/10.1002/pssa.200880461 PSSABA 0031-8965 (2008). Google Scholar

181. 

Z. Tang, W. Tress and O. Inganäs, “Light trapping in thin film organic solar cells,” Mater. Today, 17 (8), 389 –396 https://doi.org/10.1016/j.mattod.2014.05.008 MATOBY 1369-7021 (2014). Google Scholar

182. 

O. Inganäs et al., Organic Solar Cells: Fundamentals, Devices, and Upscaling, CRC Press, Taylor and Francis Group, Boca Raton (2014). Google Scholar

183. 

Y. Firdaus et al., “Long-range exciton diffusion in molecular non-fullerene acceptors,” Nat. Commun., 11 (1), 5220 https://doi.org/10.1038/s41467-020-19029-9 NCAOBW 2041-1723 (2020). Google Scholar

184. 

W. Yang et al., “Balancing the efficiency, stability, and cost potential for organic solar cells via a new figure of merit,” Joule, 5 (5), 1209 –1230 https://doi.org/10.1016/j.joule.2021.03.014 (2021). Google Scholar

185. 

L. A. A. Pettersson, L. S. Roman and O. Inganäs, “Modeling photocurrent action spectra of photovoltaic devices based on organic thin films,” J. Appl. Phys., 86 (1), 487 –496 https://doi.org/10.1063/1.370757 JAPIAU 0021-8979 (1999). Google Scholar

186. 

L. S. Roman et al., “High quantum efficiency polythiophene,” Adv. Mater., 10 (10), 774 –777 https://doi.org/10.1002/(SICI)1521-4095(199807)10:10<774::AID-ADMA774>3.0.CO;2-J ADVMEW 0935-9648 (1998). Google Scholar

187. 

P. Peumans, V. Bulović and S. R. Forrest, “Efficient photon harvesting at high optical intensities in ultrathin organic double-heterostructure photovoltaic diodes,” Appl. Phys. Lett., 76 (19), 2650 –2652 https://doi.org/10.1063/1.126433 APPLAB 0003-6951 (2000). Google Scholar

188. 

B. V. Andersson et al., “An optical spacer is no panacea for light collection in organic solar cells,” Appl. Phys. Lett., 94 (4), 043302 https://doi.org/10.1063/1.3073710 APPLAB 0003-6951 (2009). Google Scholar

189. 

Q. Liu et al., “Inverse optical cavity design for ultrabroadband light absorption beyond the conventional limit in low-bandgap nonfullerene acceptor–based solar cells,” Adv. Energy Mater., 9 (20), 1900463 https://doi.org/10.1002/aenm.201900463 ADEMBC 1614-6840 (2019). Google Scholar

190. 

J. Wang et al., “A tandem organic photovoltaic cell with 19.6% efficiency enabled by light distribution control,” Adv. Mater., 33 (39), 2102787 https://doi.org/10.1002/adma.202102787 ADVMEW 0935-9648 (2021). Google Scholar

191. 

L. Zuo et al., “Microcavity-enhanced light-trapping for highly efficient organic parallel tandem solar cells,” Adv. Mater., 26 (39), 6778 –6784 https://doi.org/10.1002/adma.201402782 ADVMEW 0935-9648 (2014). Google Scholar

192. 

B. Siegmund et al., “Organic narrowband near-infrared photodetectors based on intermolecular charge-transfer absorption,” Nat. Commun., 8 (1), 15421 https://doi.org/10.1038/ncomms15421 NCAOBW 2041-1723 (2017). Google Scholar

193. 

B. V. Andersson, N.-K. Persson and O. Inganäs, “Comparative study of organic thin film tandem solar cells in alternative geometries,” J. Appl. Phys., 104 (12), 124508 https://doi.org/10.1063/1.3050346 JAPIAU 0021-8979 (2008). Google Scholar

194. 

K. Tvingstedt et al., “Surface plasmon increase absorption in polymer photovoltaic cells,” Appl. Phys. Lett., 91 (11), 113514 https://doi.org/10.1063/1.2782910 APPLAB 0003-6951 (2007). Google Scholar

195. 

Y. Zhou et al., “Multifolded polymer solar cells on flexible substrates,” Appl. Phys. Lett., 93 (3), 033302 https://doi.org/10.1063/1.2957995 APPLAB 0003-6951 (2008). Google Scholar

196. 

S.-B. Rim et al., “An effective light trapping configuration for thin-film solar cells,” Appl. Phys. Lett., 91 (24), 243501 https://doi.org/10.1063/1.2789677 APPLAB 0003-6951 (2007). Google Scholar

197. 

K. Tvingstedt et al., “Trapping light with micro lenses in thin film organic photovoltaic cells,” Opt. Express, 16 (26), 21608 https://doi.org/10.1364/OE.16.021608 OPEXFF 1094-4087 (2008). Google Scholar

198. 

K. Tvingstedt, Z. Tang and O. Inganäs, “Light trapping with total internal reflection and transparent electrodes in organic photovoltaic devices,” Appl. Phys. Lett., 101 (16), 163902 https://doi.org/10.1063/1.4759125 APPLAB 0003-6951 (2012). Google Scholar

199. 

L. Stolz Roman et al., “Trapping light in polymer photodiodes with soft embossed gratings,” Adv. Mater., 12 (3), 189 –195 https://doi.org/10.1002/(SICI)1521-4095(200002)12:3<189::AID-ADMA189>3.0.CO;2-2 ADVMEW 0935-9648 (2000). Google Scholar

200. 

M. Takakuwa et al., “Nanograting structured ultrathin substrate for ultraflexible organic photovoltaics,” Small Methods, 4 (3), 1900762 https://doi.org/10.1002/smtd.201900762 (2020). Google Scholar

201. 

B. Lipovsek, J. Krc and M. Topic, “Optimization of microtextured light-management films for enhanced light trapping in organic solar cells under perpendicular and oblique illumination conditions,” IEEE J. Photovoltaics, 4 (2), 639 –646 https://doi.org/10.1109/JPHOTOV.2013.2293875 IJPEG8 2156-3381 (2014). Google Scholar

202. 

B. Lipovsek, J. Krc and M. Topic, “Microtextured light-management foils and their optimization for planar organic and perovskite solar cells,” IEEE J. Photovoltaics, 8 (3), 783 –792 https://doi.org/10.1109/JPHOTOV.2018.2810844 IJPEG8 2156-3381 (2018). Google Scholar

203. 

J. A. Mayer et al., “Self-contained optical enhancement film for printed photovoltaics,” Sol. Energy Mater. Sol. Cells, 163 51 –57 https://doi.org/10.1016/j.solmat.2017.01.015 SEMCEQ 0927-0248 (2017). Google Scholar

204. 

J. A. Mayer et al., “Optical enhancement of a printed organic tandem solar cell using diffractive nanostructures,” Opt. Express, 26 (6), A240 https://doi.org/10.1364/OE.26.00A240 OPEXFF 1094-4087 (2018). Google Scholar

205. 

M. Merkel, J. Imbrock and C. Denz, “Diffraction-optimized aperiodic surface structures for enhanced current density in organic solar cells,” Opt. Express, 30 (20), 36678 https://doi.org/10.1364/OE.465177 OPEXFF 1094-4087 (2022). Google Scholar

206. 

M. Merkel, T. Schemme and C. Denz, “Aperiodic biomimetic Vogel spirals as diffractive optical elements for tailored light distribution in functional polymer layers,” J. Opt., 23 (6), 065401 https://doi.org/10.1088/2040-8986/abf8cc JOOPDB 0150-536X (2021). Google Scholar

207. 

M. Parchine et al., “Large area colloidal photonic crystals for light trapping in flexible organic photovoltaic modules applied using a roll-to-roll Langmuir-Blodgett method,” Sol. Energy Mater. Sol. Cells, 185 158 –165 https://doi.org/10.1016/j.solmat.2018.05.026 SEMCEQ 0927-0248 (2018). Google Scholar

208. 

B. P. Rand, P. Peumans and S. R. Forrest, “Long-range absorption enhancement in organic tandem thin-film solar cells containing silver nanoclusters,” J. Appl. Phys., 96 (12), 7519 –7526 https://doi.org/10.1063/1.1812589 JAPIAU 0021-8979 (2004). Google Scholar

209. 

Y.-F. Li et al., “Plasmon-enhanced organic and perovskite solar cells with metal nanoparticles,” Nanophotonics, 9 (10), 3111 –3133 https://doi.org/10.1515/nanoph-2020-0099 (2020). Google Scholar

210. 

J.-Y. Lee and P. Peumans, “The origin of enhanced optical absorption in solar cells with metal nanoparticles embedded in the active layer,” Opt. Express, 18 (10), 10078 https://doi.org/10.1364/OE.18.010078 OPEXFF 1094-4087 (2010). Google Scholar

211. 

S. Liu et al., “A review on plasmonic nanostructures for efficiency enhancement of organic solar cells,” Mater. Today Phys., 24 100680 https://doi.org/10.1016/j.mtphys.2022.100680 (2022). Google Scholar

212. 

Z. Tang et al., “Semi-transparent tandem organic solar cells with 90% internal quantum efficiency,” Adv. Energy Mater., 2 (12), 1467 –1476 https://doi.org/10.1002/aenm.201200204 ADEMBC 1614-6840 (2012). Google Scholar

213. 

Z. Tang et al., “Light trapping with dielectric scatterers in single- and tandem-junction organic solar cells,” Adv. Energy Mater., 3 (12), 1606 –1613 https://doi.org/10.1002/aenm.201300524 ADEMBC 1614-6840 (2013). Google Scholar

214. 

A. Gadisa et al., “Transparent polymer cathode for organic photovoltaic devices,” Synth. Met., 156 (16–17), 1102 –1107 https://doi.org/10.1016/j.synthmet.2006.07.006 SYMEDZ 0379-6779 (2006). Google Scholar

215. 

J.-Y. Lee et al., “Semitransparent organic photovoltaic cells with laminated top electrode,” Nano Lett., 10 (4), 1276 –1279 https://doi.org/10.1021/nl903892x NALEFD 1530-6984 (2010). Google Scholar

216. 

Z. Tang et al., “Fully-solution-processed organic solar cells with a highly efficient paper-based light trapping element,” J. Mater. Chem. A, 3 (48), 24289 –24296 https://doi.org/10.1039/C5TA07154B (2015). Google Scholar

217. 

S. Yu et al., “Nanoporous polymer reflectors for organic solar cells,” Energy Technol., 10 (2), 2100676 https://doi.org/10.1002/ente.202100676 ENTED9 (2022). Google Scholar

218. 

S. Esiner et al., “Quantification and validation of the efficiency enhancement reached by application of a retroreflective light trapping texture on a polymer solar cell,” Adv. Energy Mater., 3 (8), 1013 –1017 https://doi.org/10.1002/aenm.201300227 ADEMBC 1614-6840 (2013). Google Scholar

219. 

A. B. Muñoz-García et al., “Dye-sensitized solar cells strike back,” Chem. Soc. Rev., 50 (22), 12450 –12550 https://doi.org/10.1039/D0CS01336F CSRVBR 0306-0012 (2021). Google Scholar

220. 

J. Briscoe and S. Dunn, “Dye-sensitized solar cells: the future of using earth-abundant elements in counter electrodes for dye-sensitized solar cells (Adv. Mater. 20/2016),” Adv. Mater., 28 (20), 3976 –3976 https://doi.org/10.1002/adma.201670140 ADVMEW 0935-9648 (2016). Google Scholar

221. 

C. Dong et al., “Controlling interfacial recombination in aqueous dye-sensitized solar cells by octadecyltrichlorosilane surface treatment,” Angew. Chem. Int. Ed., 53 (27), 6933 –6937 https://doi.org/10.1002/anie.201400723 ACIEAY 0570-0833 (2014). Google Scholar

222. 

F. Fabregat-Santiago et al., “Electron transport and recombination in solid-state dye solar cell with spiro-OMeTAD as hole conductor,” J. Am. Chem. Soc., 131 (2), 558 –562 https://doi.org/10.1021/ja805850q JACSAT 0002-7863 (2009). Google Scholar

223. 

J. W. Ondersma and T. W. Hamann, “Recombination and redox couples in dye-sensitized solar cells,” Coord. Chem. Rev., 257 (9–10), 1533 –1543 https://doi.org/10.1016/j.ccr.2012.09.010 CCHRAM 0010-8545 (2013). Google Scholar

224. 

I. Benesperi et al., “Dynamic dimer copper coordination redox shuttles,” Chem, 8 (2), 439 –449 https://doi.org/10.1016/j.chempr.2021.10.017 (2022). Google Scholar

225. 

M. Freitag et al., “Dye-sensitized solar cells for efficient power generation under ambient lighting,” Nat. Photonics, 11 (6), 372 –378 https://doi.org/10.1038/nphoton.2017.60 NPAHBY 1749-4885 (2017). Google Scholar

226. 

H. Michaels et al., “Emerging indoor photovoltaics for self-powered and self-aware IoT towards sustainable energy management,” Chem. Sci., 14 (20), 5350 –5360 https://doi.org/10.1039/D3SC00659J 1478-6524 (2023). Google Scholar

227. 

A. Dessì et al., “Thiazolo[5,4-d]thiazole-based organic sensitizers with improved spectral properties for application in greenhouse-integrated dye-sensitized solar cells,” Sustain. Energy Fuels, 4 (5), 2309 –2321 https://doi.org/10.1039/D0SE00124D (2020). Google Scholar

228. 

F. Grifoni et al., “Toward sustainable, colorless, and transparent photovoltaics: state of the art and perspectives for the development of selective near-infrared dye-sensitized solar cells,” Adv. Energy Mater., 11 (43), 2101598 https://doi.org/10.1002/aenm.202101598 ADEMBC 1614-6840 (2021). Google Scholar

229. 

Q. Huaulmé et al., “Photochromic dye-sensitized solar cells with light-driven adjustable optical transmission and power conversion efficiency,” Nat. Energy, 5 (6), 468 –477 https://doi.org/10.1038/s41560-020-0624-7 (2020). Google Scholar

230. 

E. F. Fernández et al., “Global energy assessment of the potential of photovoltaics for greenhouse farming,” Appl. Energy, 309 118474 https://doi.org/10.1016/j.apenergy.2021.118474 APENDX 0306-2619 (2022). Google Scholar

231. 

H. Michaels et al., “Dye-sensitized solar cells under ambient light powering machine learning: towards autonomous smart sensors for the internet of things,” Chem. Sci., 11 (11), 2895 –2906 https://doi.org/10.1039/C9SC06145B 1478-6524 (2020). Google Scholar

232. 

G. H. Morritt, H. Michaels and M. Freitag, “Coordination polymers for emerging molecular devices,” Chem. Phys. Rev., 3 (1), 011306 https://doi.org/10.1063/5.0075283 (2022). Google Scholar

233. 

E. Tanaka et al., “Synergy of co-sensitizers in a copper bipyridyl redox system for efficient and cost-effective dye-sensitized solar cells in solar and ambient light,” J. Mater. Chem. A, 8 (3), 1279 –1287 https://doi.org/10.1039/C9TA10779G (2020). Google Scholar

234. 

Y. Ren et al., “Hydroxamic acid pre-adsorption raises the efficiency of cosensitized solar cells,” Nature, 613 (7942), 60 –65 https://doi.org/10.1038/s41586-022-05460-z (2023). Google Scholar

235. 

N. Flores-Diaz et al., “Progress of photocapacitors,” Chem. Rev., 123 (15), 9327 –9355 https://doi.org/10.1021/acs.chemrev.2c00773 CHREAY 0009-2665 (2023). Google Scholar

236. 

B. Li, B. Hou and G. A. J. Amaratunga, “Indoor photovoltaics, The Next Big Trend in solution-processed solar cells,” InfoMat, 3 (5), 445 –459 https://doi.org/10.1002/inf2.12180 (2021). Google Scholar

237. 

B. T. Muhammad et al., “Halide perovskite-based indoor photovoltaics: recent development and challenges,” Mater. Today Energy, 23 100907 https://doi.org/10.1016/j.mtener.2021.100907 (2022). Google Scholar

238. 

B. Yan et al., “Indoor photovoltaics awaken the world’s first solar cells,” Sci. Adv., 8 (49), eadc9923 https://doi.org/10.1126/sciadv.adc9923 (2022). Google Scholar

239. 

V. Pecunia, L. G. Occhipinti and R. L. Z. Hoye, “Emerging indoor photovoltaic technologies for sustainable internet of things,” Adv. Energy Mater., 11 (29), 2100698 https://doi.org/10.1002/aenm.202100698 ADEMBC 1614-6840 (2021). Google Scholar

240. 

M. FreunekM. Freunek and L. M. Reindl, “Maximum efficiencies of indoor photovoltaic devices,” IEEE J. Photovoltaics, 3 (1), 59 –64 https://doi.org/10.1109/JPHOTOV.2012.2225023 IJPEG8 2156-3381 (2013). Google Scholar

241. 

J. K. W. Ho, H. Yin and S. K. So, “From 33% to 57% – an elevated potential of efficiency limit for indoor photovoltaics,” J. Mater. Chem. A, 8 (4), 1717 –1723 https://doi.org/10.1039/C9TA11894B (2020). Google Scholar

242. 

G. Kim et al., “Transparent thin-film silicon solar cells for indoor light harvesting with conversion efficiencies of 36% without photodegradation,” ACS Appl. Mater. Interfaces, 12 (24), 27122 –27130 https://doi.org/10.1021/acsami.0c04517 (2020). Google Scholar

243. 

C. C. Chen et al., “Anthracene-bridged sensitizers for dye-sensitized solar cells with 37% efficiency under dim light (Adv. Energy Mater. 20/2022),” Adv. Energy Mater., 12 (20), 2270080 https://doi.org/10.1002/aenm.202270080 ADEMBC 1614-6840 (2022). Google Scholar

244. 

C. Lee et al., “Over 30% efficient indoor organic photovoltaics enabled by morphological modification using two compatible non-fullerene acceptors,” Adv. Energy Mater., 12 (22), 2200275 https://doi.org/10.1002/aenm.202200275 ADEMBC 1614-6840 (2022). Google Scholar

245. 

X. He et al., “40.1% record low-light solar-cell efficiency by holistic trap-passivation using micrometer-thick perovskite film,” Adv. Mater., 33 (27), 2100770 https://doi.org/10.1002/adma.202100770 ADVMEW 0935-9648 (2021). Google Scholar

246. 

B. H. Hamadani, “Understanding photovoltaic energy losses under indoor lighting conditions,” Appl. Phys. Lett., 117 (4), 043904 https://doi.org/10.1063/5.0017890 APPLAB 0003-6951 (2020). Google Scholar

247. 

J. Lovering et al., “Land-use intensity of electricity production and tomorrow’s energy landscape,” PLoS One, 17 (7), e0270155 https://doi.org/10.1371/journal.pone.0270155 POLNCL 1932-6203 (2022). Google Scholar

248. 

S. Nonhebel, “Renewable energy and food supply: will there be enough land?,” Renew. Sustain. Energy Rev., 9 (2), 191 –201 https://doi.org/10.1016/j.rser.2004.02.003 (2005). Google Scholar

249. 

M. Pogson, A. Hastings and P. Smith, “How does bioenergy compare with other land-based renewable energy sources globally?,” GCB Bioenergy, 5 (5), 513 –524 https://doi.org/10.1111/gcbb.12013 (2013). Google Scholar

250. 

H. Stevens, “We need an area the size of Texas for wind and solar. Here’s how to halve it,” The Washington Post, (October 2023). https://www.washingtonpost.com/climate-environment/interactive/2023/renewable-energy-land-use-wind-solar/ Google Scholar

251. 

E. Rosen, “As demand for green energy grows, solar farms face local resistance,” The New York Times, https://www.nytimes.com/2021/11/02/business/solar-farms-resistance.html (October 2023). Google Scholar

252. 

M. Unger and T. Lakes, “Land use conflicts and synergies on agricultural land in Brandenburg, Germany,” Sustainability, 15 (5), 4546 https://doi.org/10.3390/su15054546 (2023). Google Scholar

253. 

A. Goetzberger, “Kartoffeln unter dem Kollektor,” Sonnenenergie, 3 19 –22 (1981). Google Scholar

254. 

M. Wagner et al., “Agrivoltaics: the environmental impacts of combining food crop cultivation and solar energy generation,” Agronomy, 13 (2), 299 https://doi.org/10.3390/agronomy13020299 AGRYAV 0065-4663 (2023). Google Scholar

255. 

A. Klokov et al., “A mini-review of current activities and future trends in agrivoltaics,” Energies, 16 (7), 3009 https://doi.org/10.3390/en16073009 NRGSDB 0165-2117 (2023). Google Scholar

256. 

L. La Notte et al., “Hybrid and organic photovoltaics for greenhouse applications,” Appl. Energy, 278 115582 https://doi.org/10.1016/j.apenergy.2020.115582 APENDX 0306-2619 (2020). Google Scholar

257. 

L. Lu et al., “Comprehensive review on the application of inorganic and organic photovoltaics as greenhouse shading materials,” Sustain. Energy Technol. Assess., 52 102077 https://doi.org/10.1016/j.seta.2022.102077 (2022). Google Scholar

258. 

R. Waller et al., “Evaluating the performance of flexible, semi-transparent large-area organic photovoltaic arrays deployed on a greenhouse,” AgriEngineering, 4 (4), 969 –992 https://doi.org/10.3390/agriengineering4040062 (2022). Google Scholar

259. 

M. Charles et al., “Emergent molecular traits of lettuce and tomato grown under wavelength-selective solar cells,” Front. Plant. Sci., 14 1087707 https://doi.org/10.3389/fpls.2023.1087707 (2023). Google Scholar

260. 

B. Park et al., “Synthesis of a halogenated low bandgap polymeric donor for semi-transparent and near-infrared organic solar cells,” Org. Electron., 113 106717 https://doi.org/10.1016/j.orgel.2022.106717 OERLAU 1566-1199 (2023). Google Scholar

261. 

S. S. Dipta et al., “Estimating the potential for semitransparent organic solar cells in agrophotovoltaic greenhouses,” Appl. Energy, 328 120208 https://doi.org/10.1016/j.apenergy.2022.120208 APENDX 0306-2619 (2022). Google Scholar

262. 

R. Meitzner, U. S. Schubert and H. Hoppe, “Agrivoltaics—the perfect fit for the future of organic photovoltaics,” Adv. Energy Mater., 11 (1), 2002551 https://doi.org/10.1002/aenm.202002551 ADEMBC 1614-6840 (2021). Google Scholar

263. 

E. Ravishankar et al., “Organic solar powered greenhouse performance optimization and global economic opportunity,” Energy Environ. Sci., 15 (4), 1659 –1671 https://doi.org/10.1039/D1EE03474J (2022). Google Scholar

264. 

C. J. M. Emmott et al., “Organic photovoltaic greenhouses: a unique application for semi-transparent PV?,” Energy Environ. Sci., 8 (4), 1317 –1328 https://doi.org/10.1039/C4EE03132F (2015). Google Scholar

265. 

X. Rodríguez-Martínez et al., “Laminated organic photovoltaic modules for agrivoltaics and beyond: an outdoor stability study of all-polymer and polymer: small molecule blends,” Adv. Funct. Mater., 33 (10), 2213220 https://doi.org/10.1002/adfm.202213220 AFMDC6 1616-301X (2023). Google Scholar

266. 

R. E. Smalley, “Future global energy prosperity: the terawatt challenge,” MRS Bull., 30 (6), 412 –417 https://doi.org/10.1557/mrs2005.124 MRSBEA 0883-7694 (2005). Google Scholar

267. 

V. Romano et al., “Advances in perovskites for photovoltaic applications in space,” ACS Energy Lett., 7 (8), 2490 –2514 https://doi.org/10.1021/acsenergylett.2c01099 (2022). Google Scholar

268. 

A. W. Y. Ho-Baillie et al., “Deployment opportunities for space photovoltaics and the prospects for perovskite solar cells,” Adv. Mater. Technol., 7 (3), 2101059 https://doi.org/10.1002/admt.202101059 (2022). Google Scholar

269. 

J. J. Berry et al., “Perovskite photovoltaics: the path to a printable terawatt-scale technology,” ACS Energy Lett., 2 (11), 2540 –2544 https://doi.org/10.1021/acsenergylett.7b00964 (2017). Google Scholar

270. 

L. McMillon-Brown, J. M. Luther and T. J. Peshek, “What would it take to manufacture perovskite solar cells in space?,” ACS Energy Lett., 7 (3), 1040 –1042 https://doi.org/10.1021/acsenergylett.2c00276 (2022). Google Scholar

271. 

I. Cardinaletti et al., “Organic and perovskite solar cells for space applications,” Sol. Energy Mater. Sol. Cells, 182 121 –127 https://doi.org/10.1016/j.solmat.2018.03.024 SEMCEQ 0927-0248 (2018). Google Scholar

272. 

L. K. Reb et al., “Perovskite and organic solar cells on a rocket flight,” Joule, 4 (9), 1880 –1892 https://doi.org/10.1016/j.joule.2020.07.004 (2020). Google Scholar

273. 

A. R. Kirmani et al., “Countdown to perovskite space launch: guidelines to performing relevant radiation-hardness experiments,” Joule, 6 (5), 1015 –1031 https://doi.org/10.1016/j.joule.2022.03.004 (2022). Google Scholar

274. 

A. R. Kirmani et al., “Metal oxide barrier layers for terrestrial and space perovskite photovoltaics,” Nat. Energy, 8 (2), 191 –202 https://doi.org/10.1038/s41560-022-01189-1 (2023). Google Scholar

275. 

N. Y. Nia, “Barrier layer for space,” Nat. Energy, 8 (2), 117 –118 https://doi.org/10.1038/s41560-023-01197-9 (2023). Google Scholar

276. 

R. T. Ross and A. J. Nozik, “Efficiency of hot-carrier solar energy converters,” J. Appl. Phys., 53 (5), 3813 –3818 https://doi.org/10.1063/1.331124 JAPIAU 0021-8979 (1982). Google Scholar

277. 

W. Shockley and H. J. Queisser, “Detailed balance limit of efficiency of p-n junction solar cells,” J. Appl. Phys., 32 (3), 510 –519 https://doi.org/10.1063/1.1736034 JAPIAU 0021-8979 (1961). Google Scholar

278. 

H. Cotal et al., “III–V multijunction solar cells for concentrating photovoltaics,” Energy Environ. Sci., 2 (2), 174 –192 https://doi.org/10.1039/B809257E EESNBY 1754-5692 (2009). Google Scholar

279. 

G. Conibeer et al., “Progress on hot carrier cells,” Sol. Energy Mater. Sol. Cells, 93 (6–7), 713 –719 https://doi.org/10.1016/j.solmat.2008.09.034 SEMCEQ 0927-0248 (2009). Google Scholar

280. 

P. Lugli and S. M. Goodnick, “Nonequilibrium longitudinal-optical phonon effects in GaAs-AlGaAs quantum wells,” Phys. Rev. Lett., 59 (6), 716 –719 https://doi.org/10.1103/PhysRevLett.59.716 PRLTAO 0031-9007 (1987). Google Scholar

281. 

Y. Yao and D. König, “Comparison of bulk material candidates for hot carrier absorber,” Sol. Energy Mater. Sol. Cells, 140 422 –427 https://doi.org/10.1016/j.solmat.2015.05.002 SEMCEQ 0927-0248 (2015). Google Scholar

282. 

P. Aliberti et al., “Investigation of theoretical efficiency limit of hot carriers solar cells with a bulk indium nitride absorber,” J. Appl. Phys., 108 (9), 094507 https://doi.org/10.1063/1.3494047 JAPIAU 0021-8979 (2010). Google Scholar

283. 

L. Lindsay, D. A. Broido and T. L. Reinecke, “Ab initio thermal transport in compound semiconductors,” Phys. Rev. B, 87 (16), 165201 https://doi.org/10.1103/PhysRevB.87.165201 PRBMDO 1098-0121 (2013). Google Scholar

284. 

S. Chung et al., “Nanosecond long excited state lifetimes observed in hafnium nitride,” Sol. Energy Mater. Sol. Cells, 169 13 –18 https://doi.org/10.1016/j.solmat.2017.04.037 SEMCEQ 0927-0248 (2017). Google Scholar

285. 

H. Esmaielpour et al., “Hot carrier relaxation and inhibited thermalization in superlattice heterostructures: the potential for phonon management,” Appl. Phys. Lett., 118 (21), 213902 https://doi.org/10.1063/5.0052600 APPLAB 0003-6951 (2021). Google Scholar

286. 

J. Garg and I. R. Sellers, “Phonon linewidths in InAs/AlSb superlattices derived from first-principles—application towards quantum well hot carrier solar cells,” Semicond. Sci. Technol., 35 (4), 044001 https://doi.org/10.1088/1361-6641/ab73f0 SSTEET 0268-1242 (2020). Google Scholar

287. 

L. C. Hirst et al., “Hot carriers in quantum wells for photovoltaic efficiency enhancement,” IEEE J. Photovoltaics, 4 (1), 244 –252 https://doi.org/10.1109/JPHOTOV.2013.2289321 IJPEG8 2156-3381 (2014). Google Scholar

288. 

A. Le Bris et al., “Thermalisation rate study of GaSb-based heterostructures by continuous wave photoluminescence and their potential as hot carrier solar cell absorbers,” Energy Environ. Sci., 5 (3), 6225 https://doi.org/10.1039/c2ee02843c (2012). Google Scholar

289. 

J. Tang et al., “Effects of localization on hot carriers in InAs/AlAsxSb1x quantum wells,” Appl. Phys. Lett., 106 (6), 061902 https://doi.org/10.1063/1.4907630 APPLAB 0003-6951 (2015). Google Scholar

290. 

Y. Rosenwaks et al., “Hot-carrier cooling in GaAs: quantum wells versus bulk,” Phys. Rev. B, 48 (19), 14675 –14678 https://doi.org/10.1103/PhysRevB.48.14675 PRBMDO 1098-0121 (1993). Google Scholar

291. 

J. Shah, “Hot carriers in quasi-2-D polar semiconductors,” IEEE J. Quantum Electron., 22 (9), 1728 –1743 https://doi.org/10.1109/JQE.1986.1073164 IEJQA7 0018-9197 (1986). Google Scholar

292. 

J. Shah et al., “Energy-loss rates for hot electrons and holes in GaAs quantum wells,” Phys. Rev. Lett., 54 (18), 2045 –2048 https://doi.org/10.1103/PhysRevLett.54.2045 PRLTAO 0031-9007 (1985). Google Scholar

293. 

L. C. Hirst et al., “Experimental demonstration of hot-carrier photo-current in an InGaAs quantum well solar cell,” Appl. Phys. Lett., 104 (23), 231115 https://doi.org/10.1063/1.4883648 APPLAB 0003-6951 (2014). Google Scholar

294. 

J. Rodière et al., “Experimental evidence of hot carriers solar cell operation in multi-quantum wells heterostructures,” Appl. Phys. Lett., 106 (18), 183901 https://doi.org/10.1063/1.4919901 APPLAB 0003-6951 (2015). Google Scholar

295. 

D.-T. Nguyen et al., “Quantitative experimental assessment of hot carrier-enhanced solar cells at room temperature,” Nat. Energy, 3 (3), 236 –242 https://doi.org/10.1038/s41560-018-0106-3 (2018). Google Scholar

296. 

J. A. R. Dimmock et al., “Demonstration of a hot-carrier photovoltaic cell: demonstration of a hot-carrier photovoltaic cell,” Prog. Photovoltaics Res. Appl., 22 (2), 151 –160 https://doi.org/10.1002/pip.2444 PPHOED 1062-7995 (2014). Google Scholar

297. 

H. Esmaielpour et al., “Exploiting intervalley scattering to harness hot carriers in III–V solar cells,” Nat. Energy, 5 (4), 336 –343 https://doi.org/10.1038/s41560-020-0602-0 (2020). Google Scholar

298. 

V. R. Whiteside et al., “The role of intervalley phonons in hot carrier transfer and extraction in type-II InAs/AlAsSb quantum-well solar cells,” Semicond. Sci. Technol., 34 (9), 094001 https://doi.org/10.1088/1361-6641/ab312b SSTEET 0268-1242 (2019). Google Scholar

299. 

D. K. Ferry, “In search of a true hot carrier solar cell,” Semicond. Sci. Technol., 34 (4), 044001 https://doi.org/10.1088/1361-6641/ab0bc3 SSTEET 0268-1242 (2019). Google Scholar

300. 

F. Meinardi et al., “Highly efficient large-area colourless luminescent solar concentrators using heavy-metal-free colloidal quantum dots,” Nat. Nanotechnol., 10 (10), 878 –885 https://doi.org/10.1038/nnano.2015.178 NNAABX 1748-3387 (2015). Google Scholar

301. 

W. G. J. H. M. Van Sark, “Luminescent solar concentrators – a low cost photovoltaics alternative,” Renew. Energy, 49 207 –210 https://doi.org/10.1016/j.renene.2012.01.030 RNENE3 0960-1481 (2013). Google Scholar

302. 

M. Buffa et al., “Dye-doped polysiloxane rubbers for luminescent solar concentrator systems,” Sol. Energy Mater. Sol. Cells, 103 114 –118 https://doi.org/10.1016/j.solmat.2012.04.019 SEMCEQ 0927-0248 (2012). Google Scholar

303. 

S. M. El-Bashir, “Enhanced fluorescence polarization of fluorescent polycarbonate/zirconia nanocomposites for second generation luminescent solar concentrators,” Renew. Energy, 115 269 –275 https://doi.org/10.1016/j.renene.2017.08.016 RNENE3 0960-1481 (2018). Google Scholar

304. 

A. Reinders et al., “Luminescent solar concentrator photovoltaic designs,” Jpn. J. Appl. Phys., 57 (8S3), 08RD10 https://doi.org/10.7567/JJAP.57.08RD10 JJAPA5 0021-4922 (2018). Google Scholar

305. 

B. C. Rowan, L. R. Wilson and B. S. Richards, “Advanced material concepts for luminescent solar concentrators,” IEEE J. Sel. Top. Quantum Electron., 14 (5), 1312 –1322 https://doi.org/10.1109/JSTQE.2008.920282 IJSQEN 1077-260X (2008). Google Scholar

306. 

Third Generation Photovoltaics, Springer Berlin Heidelberg( (2006). Google Scholar

307. 

A. A Martí, Next Generation Photovoltaics: High Efficiency through Full Spectrum Utilization, CRC Press( (2003). Google Scholar

308. 

A. Goetzberger, “Fluorescent solar energy collectors: operating conditions with diffuse light,” Appl. Phys., 16 (4), 399 –404 https://doi.org/10.1007/BF00885865 (1978). Google Scholar

309. 

W. H. Weber and J. Lambe, “Luminescent greenhouse collector for solar radiation,” Appl. Opt., 15 (10), 2299 https://doi.org/10.1364/AO.15.002299 APOPAI 0003-6935 (1976). Google Scholar

310. 

E. Yablonovitch, “Thermodynamics of the fluorescent planar concentrator,” J. Opt. Soc. Am., 70 (11), 1362 https://doi.org/10.1364/JOSA.70.001362 JOSAAH 0030-3941 (1980). Google Scholar

311. 

X. Wang and A. Barnett, “The evolving value of photovoltaic module efficiency,” Appl. Sci., 9 (6), 1227 https://doi.org/10.3390/app9061227 (2019). Google Scholar

312. 

M. G. Debije and V. A. Rajkumar, “Direct versus indirect illumination of a prototype luminescent solar concentrator,” Sol. Energy, 122 334 –340 https://doi.org/10.1016/j.solener.2015.08.036 SRENA4 0038-092X (2015). Google Scholar

313. 

A. Reinders, Designing with Photovoltaics, CRC Press( (2020). Google Scholar

314. 

P. Della Sala et al., “First demonstration of the use of very large Stokes shift cycloparaphenylenes as promising organic luminophores for transparent luminescent solar concentrators,” Chem. Commun., 55 (21), 3160 –3163 https://doi.org/10.1039/C8CC09859J (2019). Google Scholar

315. 

Y. El Mouedden et al., “A cost-effective, long-lifetime efficient organic luminescent solar concentrator,” J. Appl. Phys., 118 (1), 015502 https://doi.org/10.1063/1.4923389 JAPIAU 0021-8979 (2015). Google Scholar

316. 

Y. Li et al., “A structurally modified perylene dye for efficient luminescent solar concentrators,” Sol. Energy, 136 668 –674 https://doi.org/10.1016/j.solener.2016.07.051 SRENA4 0038-092X (2016). Google Scholar

317. 

A. Sanguineti et al., “High stokes shift perylene dyes for luminescent solar concentrators,” Chem. Commun., 49 (16), 1618 https://doi.org/10.1039/c2cc38708e (2013). Google Scholar

318. 

V. I. Klimov et al., “Quality factor of luminescent solar concentrators and practical concentration limits attainable with semiconductor quantum dots,” ACS Photonics, 3 (6), 1138 –1148 https://doi.org/10.1021/acsphotonics.6b00307 (2016). Google Scholar

319. 

P. Moraitis, R. E. I. Schropp and W. G. J. H. M. Van Sark, “Nanoparticles for luminescent solar concentrators - a review,” Opt. Mater., 84 636 –645 https://doi.org/10.1016/j.optmat.2018.07.034 (2018). Google Scholar

320. 

F. Purcell-Milton and Y. K. Gun’ko, “Quantum dots for luminescent solar concentrators,” J. Mater. Chem., 22 (33), 16687 https://doi.org/10.1039/c2jm32366d JMACEP 0959-9428 (2012). Google Scholar

321. 

K. Wu, H. Li and V. I. Klimov, “Tandem luminescent solar concentrators based on engineered quantum dots,” Nat. Photonics, 12 (2), 105 –110 https://doi.org/10.1038/s41566-017-0070-7 NPAHBY 1749-4885 (2018). Google Scholar

322. 

Y. Zhou et al., “Harnessing the properties of colloidal quantum dots in luminescent solar concentrators,” Chem. Soc. Rev., 47 (15), 5866 –5890 https://doi.org/10.1039/C7CS00701A CSRVBR 0306-0012 (2018). Google Scholar

323. 

V. T. Freitas et al., “Eu3+ -based bridged silsesquioxanes for transparent luminescent solar concentrators,” ACS Appl. Mater. Interfaces, 7 (16), 8770 –8778 https://doi.org/10.1021/acsami.5b01281 (2015). Google Scholar

324. 

A. Frias et al., “Transparent luminescent solar concentrators using Ln3+-based ionosilicas towards photovoltaic windows,” Energies, 12 (3), 451 https://doi.org/10.3390/en12030451 NRGSDB 0165-2117 (2019). Google Scholar

325. 

J. Liu et al., “Preparation of cerium modified titanium dioxide nanoparticles and investigation of their visible light photocatalytic performance,” Int. J. Photoenergy, 2014 695679 https://doi.org/10.1155/2014/695679 (2014). Google Scholar

326. 

G. V. Shcherbatyuk et al., “Viability of using near infrared PbS quantum dots as active materials in luminescent solar concentrators,” Appl. Phys. Lett., 96 (19), 191901 https://doi.org/10.1063/1.3422485 APPLAB 0003-6951 (2010). Google Scholar

327. 

Y. Zhou et al., “Near infrared, highly efficient luminescent solar concentrators,” Adv. Energy Mater., 6 (11), 1501913 https://doi.org/10.1002/aenm.201501913 ADEMBC 1614-6840 (2016). Google Scholar

328. 

M. R. Bergren et al., “High-performance CuInS2 quantum dot laminated glass luminescent solar concentrators for windows,” ACS Energy Lett., 3 (3), 520 –525 https://doi.org/10.1021/acsenergylett.7b01346 (2018). Google Scholar

329. 

K. E. Knowles et al., “Bright CuInS2/CdS nanocrystal phosphors for high-gain full-spectrum luminescent solar concentrators,” Chem. Commun., 51 (44), 9129 –9132 https://doi.org/10.1039/C5CC02007G (2015). Google Scholar

330. 

B. Zhang et al., “Aggregation-induced emission-mediated spectral downconversion in luminescent solar concentrators,” Mater. Chem. Front., 2 (3), 615 –619 https://doi.org/10.1039/C7QM00598A (2018). Google Scholar

331. 

F. Meinardi et al., “Doped halide perovskite nanocrystals for reabsorption-free luminescent solar concentrators,” ACS Energy Lett., 2 (10), 2368 –2377 https://doi.org/10.1021/acsenergylett.7b00701 (2017). Google Scholar

332. 

O. M. Ten Kate, K. W. Krämer and E. Van Der Kolk, “Efficient luminescent solar concentrators based on self-absorption free, Tm2+ doped halides,” Sol. Energy Mater. Sol. Cells, 140 115 –120 https://doi.org/10.1016/j.solmat.2015.04.002 SEMCEQ 0927-0248 (2015). Google Scholar

333. 

G. Griffini, “Host matrix materials for luminescent solar concentrators: recent achievements and forthcoming challenges,” Front. Mater., 6 29 https://doi.org/10.3389/fmats.2019.00029 (2019). Google Scholar

334. 

M. J. Kastelijn, C. W. M. Bastiaansen and M. G. Debije, “Influence of waveguide material on light emission in luminescent solar concentrators,” Opt. Mater., 31 (11), 1720 –1722 https://doi.org/10.1016/j.optmat.2009.05.003 (2009). Google Scholar

335. 

W. Chen et al., “Heavy metal free nanocrystals with near infrared emission applying in luminescent solar concentrator,” Solar RRL, 1 (6), 1700041 https://doi.org/10.1002/solr.201700041 (2017). Google Scholar

336. 

A. Reinders, M. G. Debije and A. Rosemann, “Measured efficiency of a luminescent solar concentrator PV module called leaf roof,” IEEE J. Photovoltaics, 7 (6), 1663 –1666 https://doi.org/10.1109/JPHOTOV.2017.2751513 IJPEG8 2156-3381 (2017). Google Scholar

337. 

A. Sanguineti et al., “NIR emitting ytterbium chelates for colourless luminescent solar concentrators,” Phys. Chem. Chem. Phys., 14 (18), 6452 https://doi.org/10.1039/c2cp40791d PPCPFQ 1463-9076 (2012). Google Scholar

338. 

M. G. Debije et al., “Measured surface loss from luminescent solar concentrator waveguides,” Appl. Opt., 47 (36), 6763 https://doi.org/10.1364/AO.47.006763 APOPAI 0003-6935 (2008). Google Scholar

339. 

M. Kennedy et al., “Large Stokes shift downshifting Eu(III) films as efficiency enhancing UV blocking layers for dye sensitized solar cells: Eu(III) films as efficiency enhancing UV blocking layers for dye sensitized solar cells,” Physica Status Solidi (A), 212 (1), 203 –210 https://doi.org/10.1002/pssa.201431683 PSSABA 0031-8965 (2015). Google Scholar

340. 

M. Rafiee et al., “An overview of various configurations of Luminescent Solar Concentrators for photovoltaic applications,” Opt. Mater., 91 212 –227 https://doi.org/10.1016/j.optmat.2019.01.007 (2019). Google Scholar

341. 

M. G. Debije and P. P. C. Verbunt, “Thirty years of luminescent solar concentrator research: solar energy for the built environment,” Adv. Energy Mater., 2 (1), 12 –35 https://doi.org/10.1002/aenm.201100554 ADEMBC 1614-6840 (2012). Google Scholar

342. 

J. C. Goldschmidt et al., “Increasing the efficiency of fluorescent concentrator systems,” Sol. Energy Mater. Sol. Cells, 93 (2), 176 –182 https://doi.org/10.1016/j.solmat.2008.09.048 SEMCEQ 0927-0248 (2009). Google Scholar

343. 

W. G. J. H. M. Van Sark et al., “Luminescent solar concentrators - a review of recent results,” Opt. Express, 16 (26), 21773 https://doi.org/10.1364/OE.16.021773 OPEXFF 1094-4087 (2008). Google Scholar

344. 

M. Kanellis et al., “The solar noise barrier project: 1. Effect of incident light orientation on the performance of a large-scale luminescent solar concentrator noise barrier,” Renew. Energy, 103 647 –652 https://doi.org/10.1016/j.renene.2016.10.078 RNENE3 0960-1481 (2017). Google Scholar

345. 

R. López-Martín and L. Valenzuela, “Optical efficiency measurement of solar receiver tubes: a testbed and case studies,” Case Stud. Thermal Eng., 12 414 –422 https://doi.org/10.1016/j.csite.2018.06.002 (2018). Google Scholar

346. 

B. Vishwanathan et al., “A comparison of performance of flat and bent photovoltaic luminescent solar concentrators,” Sol. Energy, 112 120 –127 https://doi.org/10.1016/j.solener.2014.12.001 SRENA4 0038-092X (2015). Google Scholar

347. 

R. H. Inman et al., “Cylindrical luminescent solar concentrators with near-infrared quantum dots,” Opt. Express, 19 (24), 24308 https://doi.org/10.1364/OE.19.024308 OPEXFF 1094-4087 (2011). Google Scholar

348. 

J. J. H. Videira, E. Bilotti and A. J. Chatten, “Cylindrical array luminescent solar concentrators: performance boosts by geometric effects,” Opt. Express, 24 (14), A1188 https://doi.org/10.1364/OE.24.0A1188 OPEXFF 1094-4087 (2016). Google Scholar

349. 

N. Aste et al., “Performance analysis of a large-area luminescent solar concentrator module,” Renew. Energy, 76 330 –337 https://doi.org/10.1016/j.renene.2014.11.026 RNENE3 0960-1481 (2015). Google Scholar

350. 

M. J. Currie et al., “High-efficiency organic solar concentrators for photovoltaics,” Science, 321 (5886), 226 –228 https://doi.org/10.1126/science.1158342 SCIEAS 0036-8075 (2008). Google Scholar

351. 

N. D. Bronstein et al., “Quantum dot luminescent concentrator cavity exhibiting 30-fold concentration,” ACS Photonics, 2 (11), 1576 –1583 https://doi.org/10.1021/acsphotonics.5b00334 (2015). Google Scholar

352. 

I. Coropceanu and M. G. Bawendi, “Core/shell quantum dot based luminescent solar concentrators with reduced reabsorption and enhanced efficiency,” Nano Lett., 14 (7), 4097 –4101 https://doi.org/10.1021/nl501627e NALEFD 1530-6984 (2014). Google Scholar

353. 

S. M. El-Bashir, F. M. Barakat and M. S. AlSalhi, “Metal-enhanced fluorescence of mixed Coumarin dyes by silver and gold nanoparticles: towards plasmonic thin-film luminescent solar concentrator,” J. Lumin., 143 43 –49 https://doi.org/10.1016/j.jlumin.2013.04.029 JLUMA8 0022-2313 (2013). Google Scholar

354. 

S. M. El-Bashir, F. M. Barakat and M. S. AlSalhi, “Double layered plasmonic thin-film luminescent solar concentrators based on polycarbonate supports,” Renew. Energy, 63 642 –649 https://doi.org/10.1016/j.renene.2013.10.014 RNENE3 0960-1481 (2014). Google Scholar

355. 

D. J. Farrell et al., “Using amorphous silicon solar cells to boost the viability of luminescent solar concentrators,” Physica Status Solidi C, 7 (3–4), 1045 –1048 https://doi.org/10.1002/pssc.200982866 PSSCGL 1862-6351 (2010). Google Scholar

356. 

C. Tummeltshammer et al., “On the ability of Förster resonance energy transfer to enhance luminescent solar concentrator efficiency,” Nano Energy, 32 263 –270 https://doi.org/10.1016/j.nanoen.2016.11.058 (2017). Google Scholar

357. 

S.-J. Ha et al., “Upconversion-assisted dual-band luminescent solar concentrator coupled for high power conversion efficiency photovoltaic systems,” ACS Photonics, 5 (9), 3621 –3627 https://doi.org/10.1021/acsphotonics.8b00498 (2018). Google Scholar

358. 

K. Kim et al., “Photon upconversion-assisted dual-band luminescence solar concentrators coupled with perovskite solar cells for highly efficient semi-transparent photovoltaic systems,” Nanoscale, 12 (23), 12426 –12431 https://doi.org/10.1039/D0NR02106G NANOHL 2040-3364 (2020). Google Scholar

359. 

L. H. Slooff et al., “A luminescent solar concentrator with 7.1% power conversion efficiency: carbon nanotube/epoxy resin composites using a block copolymer as a dispersing agent,” Physica Status Solidi (RRL) - Rapid Res. Lett., 2 (6), 257 –259 https://doi.org/10.1002/pssr.200802186 (2008). Google Scholar

360. 

M. G. Debije, R. C. Evans and G. Griffini, “Laboratory protocols for measuring and reporting the performance of luminescent solar concentrators,” Energy Environ. Sci., 14 (1), 293 –301 https://doi.org/10.1039/D0EE02967J (2021). Google Scholar

361. 

C. Yang, D. Liu and R. R. Lunt, “How to accurately report transparent luminescent solar concentrators,” Joule, 3 (12), 2871 –2876 https://doi.org/10.1016/j.joule.2019.10.009 (2019). Google Scholar

362. 

T. Warner, K. P. Ghiggino and G. Rosengarten, “A critical analysis of luminescent solar concentrator terminology and efficiency results,” Sol. Energy, 246 119 –140 https://doi.org/10.1016/j.solener.2022.09.011 SRENA4 0038-092X (2022). Google Scholar

363. 

J. Keil et al., “Evaluating tandem luminscent solar concentrator performance based on luminophore selection,” in IEEE 48th Photovoltaics Spec. Conf. (PVSC), (2021). https://doi.org/10.1109/PVSC43889.2021.9518707 Google Scholar

364. 

R. Mazzaro and A. Vomiero, “The renaissance of luminescent solar concentrators: the role of inorganic nanomaterials,” Adv. Energy Mater., 8 (33), 1801903 https://doi.org/10.1002/aenm.201801903 ADEMBC 1614-6840 (2018). Google Scholar

365. 

L. R. Bradshaw et al., “Nanocrystals for luminescent solar concentrators,” Nano Lett., 15 (2), 1315 –1323 https://doi.org/10.1021/nl504510t NALEFD 1530-6984 (2015). Google Scholar

366. 

N. D. Bronstein et al., “Luminescent solar concentration with semiconductor nanorods and transfer-printed micro-silicon solar cells,” ACS Nano, 8 (1), 44 –53 https://doi.org/10.1021/nn404418h ANCAC3 1936-0851 (2014). Google Scholar

367. 

S. K. E. Hill et al., “Poly(methyl methacrylate) films with high concentrations of silicon quantum dots for visibly transparent luminescent solar concentrators,” ACS Appl. Mater. Interfaces, 12 (4), 4572 –4578 https://doi.org/10.1021/acsami.9b22903 (2020). Google Scholar

368. 

S. K. E. Hill et al., “Silicon quantum dot–poly(methyl methacrylate) nanocomposites with reduced light scattering for luminescent solar concentrators,” ACS Photonics, 6 (1), 170 –180 https://doi.org/10.1021/acsphotonics.8b01346 (2019). Google Scholar

369. 

R. Mazzaro et al., “Hybrid Silicon nanocrystals for color-neutral and transparent luminescent solar concentrators,” ACS Photonics, 6 (9), 2303 –2311 https://doi.org/10.1021/acsphotonics.9b00802 (2019). Google Scholar

370. 

Y. Li et al., “N-doped carbon-dots for luminescent solar concentrators,” J. Mater. Chem. A, 5 (40), 21452 –21459 https://doi.org/10.1039/C7TA05220K (2017). Google Scholar

371. 

F. Meinardi et al., “Highly efficient luminescent solar concentrators based on earth-abundant indirect-bandgap silicon quantum dots,” Nat. Photonics, 11 (3), 177 –185 https://doi.org/10.1038/nphoton.2017.5 NPAHBY 1749-4885 (2017). Google Scholar

372. 

S. Mirershadi and S. Ahmadi-Kandjani, “Efficient thin luminescent solar concentrator based on organometal halide perovskite,” Dyes Pigm., 120 15 –21 https://doi.org/10.1016/j.dyepig.2015.03.035 DYPIDX 0143-7208 (2015). Google Scholar

373. 

K. Nikolaidou et al., “Hybrid perovskite thin films as highly efficient luminescent solar concentrators,” Adv. Opt. Mater., 4 (12), 2126 –2132 https://doi.org/10.1002/adom.201600634 2195-1071 (2016). Google Scholar

374. 

H. Zhao et al., “Perovskite quantum dots integrated in large-area luminescent solar concentrators,” Nano Energy, 37 214 –223 https://doi.org/10.1016/j.nanoen.2017.05.030 (2017). Google Scholar

375. 

C. Li et al., “Large Stokes shift and high efficiency luminescent solar concentrator incorporated with CuInS2/ZnS quantum dots,” Sci. Rep., 5 (1), 17777 https://doi.org/10.1038/srep17777 (2015). Google Scholar

376. 

Y. Zhao et al., “Near-infrared harvesting transparent luminescent solar concentrators,” Adv. Opt. Mater., 2 (7), 606 –611 https://doi.org/10.1002/adom.201400103 2195-1071 (2014). Google Scholar

377. 

R. Connell et al., “CdSe/CdS–poly(cyclohexylethylene) thin film luminescent solar concentrators,” APL Mater., 7 (10), 101123 https://doi.org/10.1063/1.5121441 (2019). Google Scholar

378. 

M. G. Debije et al., “Effect on the output of a luminescent solar concentrator on application of organic wavelength-selective mirrors,” Appl. Opt., 49 (4), 745 https://doi.org/10.1364/AO.49.000745 APOPAI 0003-6935 (2010). Google Scholar

379. 

U. Rau, F. Einsele and G. C. Glaeser, “Efficiency limits of photovoltaic fluorescent collectors,” Appl. Phys. Lett., 87 (17), 171101 https://doi.org/10.1063/1.2112196 APPLAB 0003-6951 (2005). Google Scholar

380. 

R. Connell, C. Pinnell and V. E. Ferry, “Designing spectrally-selective mirrors for use in luminescent solar concentrators,” J. Opt., 20 (2), 024009 https://doi.org/10.1088/2040-8986/aaa074 JOOPDB 0150-536X (2018). Google Scholar

381. 

P. P. C. Verbunt et al., “Increased efficiency of luminescent solar concentrators after application of organic wavelength selective mirrors,” Opt. Express, 20 (S5), A655 https://doi.org/10.1364/OE.20.00A655 OPEXFF 1094-4087 (2012). Google Scholar

382. 

C. Yang et al., “Consensus statement: standardized reporting of power-producing luminescent solar concentrator performance,” Joule, 6 (1), 8 –15 https://doi.org/10.1016/j.joule.2021.12.004 (2022). Google Scholar

383. 

I. Papakonstantinou, M. Portnoi and M. G. Debije, “The hidden potential of luminescent solar concentrators,” Adv. Energy Mater., 11 (3), 2002883 https://doi.org/10.1002/aenm.202002883 ADEMBC 1614-6840 (2021). Google Scholar

384. 

C. Corrado et al., “Power generation study of luminescent solar concentrator greenhouse,” J. Renew. Sustain. Energy, 8 (4), 043502 https://doi.org/10.1063/1.4958735 (2016). Google Scholar

385. 

D. Hebert et al., “Luminescent quantum dot films improve light use efficiency and crop quality in greenhouse horticulture,” Front. Chem., 10 988227 https://doi.org/10.3389/fchem.2022.988227 (2022). Google Scholar

386. 

J. Keil et al., “Bilayer luminescent solar concentrators with enhanced absorption and efficiency for agrivoltaic applications,” ACS Appl. Energy Mater., 4 (12), 14102 –14110 https://doi.org/10.1021/acsaem.1c02860 (2021). Google Scholar

387. 

L. Shen et al., “Increasing greenhouse production by spectral-shifting and unidirectional light-extracting photonics,” Nat. Food, 2 (6), 434 –441 https://doi.org/10.1038/s43016-021-00307-8 (2021). Google Scholar

388. 

H. Li et al., “Doctor-blade deposition of quantum dots onto standard window glass for low-loss large-area luminescent solar concentrators,” Nat. Energy, 1 (12), 16157 https://doi.org/10.1038/nenergy.2016.157 (2016). Google Scholar

389. 

J. Wang et al., “Quantum dot-based luminescent solar concentrators fabricated through the ultrasonic spray-coating method,” ACS Appl. Mater. Interfaces, 14 (36), 41013 –41021 https://doi.org/10.1021/acsami.2c11205 (2022). Google Scholar

390. 

M. A. Hernández-Rodríguez et al., “A perspective on sustainable luminescent solar concentrators,” J. Appl. Phys., 131 (14), 140901 https://doi.org/10.1063/5.0084182 JAPIAU 0021-8979 (2022). Google Scholar

391. 

C. Yang et al., “Integration of near-infrared harvesting transparent luminescent solar concentrators onto arbitrary surfaces,” J. Lumin., 210 239 –246 https://doi.org/10.1016/j.jlumin.2019.02.042 JLUMA8 0022-2313 (2019). Google Scholar

392. 

C. H. Henry, “Limiting efficiencies of ideal single and multiple energy gap terrestrial solar cells,” J. Appl. Phys., 51 (8), 4494 –4500 https://doi.org/10.1063/1.328272 JAPIAU 0021-8979 (1980). Google Scholar

393. 

E. D. Jackson, “Areas for improvement of the semiconductor solar energy converter,” in Trans. Conf. Use of Solar Energy, 122 (1955). Google Scholar

394. 

A. G. Imenes and D. R. Mills, “Spectral beam splitting technology for increased conversion efficiency in solar concentrating systems: a review,” Sol. Energy Mater. Sol. Cells, 84 (1–4), 19 –69 https://doi.org/10.1016/j.solmat.2004.01.038 SEMCEQ 0927-0248 (2004). Google Scholar

395. 

A. Mojiri et al., “Spectral beam splitting for efficient conversion of solar energy—a review,” Renew. Sustain. Energy Rev., 28 654 –663 https://doi.org/10.1016/j.rser.2013.08.026 (2013). Google Scholar

396. 

J. D. McCambridge et al., “Compact spectrum splitting photovoltaic module with high efficiency,” Prog. Photovoltaics Res. Appl., 19 (3), 352 –360 https://doi.org/10.1002/pip.1030 PPHOED 1062-7995 (2011). Google Scholar

397. 

A. Polman and H. A. Atwater, “Photonic design principles for ultrahigh-efficiency photovoltaics,” Nat. Mater., 11 (3), 174 –177 https://doi.org/10.1038/nmat3263 NMAACR 1476-1122 (2012). Google Scholar

398. 

C. G. Stojanoff, “Engineering applications of HOEs manufactured with enhanced performance DCG films,” Proc. SPIE, 6316 613601 https://doi.org/10.1117/12.642986 PSISDG 0277-786X (2006). Google Scholar

399. 

D. Vincenzi et al., “Concentrating PV system based on spectral separation of solar radiation,” Physica Status Solidi (A), 206 (2), 375 –378 https://doi.org/10.1002/pssa.200824468 PSSABA 0031-8965 (2009). Google Scholar

400. 

Y. Wu et al., “Design of folded holographic spectrum-splitting photovoltaic system for direct and diffuse illumination conditions,” Proc. SPIE, 9175 91750G https://doi.org/10.1117/12.2060882 PSISDG 0277-786X (2014). Google Scholar

401. 

N. Mohammad et al., “Enhancing photovoltaic output power by 3-band spectrum-splitting and concentration using a diffractive micro-optic,” Opt. Express, 22 (S6), A1519 https://doi.org/10.1364/OE.22.0A1519 OPEXFF 1094-4087 (2014). Google Scholar

402. 

S. Vorndran et al., “Broadband Gerchberg-Saxton algorithm for freeform diffractive spectral filter design,” Opt. Express, 23 (24), A1512 https://doi.org/10.1364/OE.23.0A1512 OPEXFF 1094-4087 (2015). Google Scholar

403. 

J. M. Russo et al., “Spectrum splitting metrics and effect of filter characteristics on photovoltaic system performance,” Opt. Express, 22 (S2), A528 https://doi.org/10.1364/OE.22.00A528 OPEXFF 1094-4087 (2014). Google Scholar

404. 

B. D. Chrysler, S. E. Shaheen and R. K. Kostuk, “Lateral spectrum splitting system with perovskite photovoltaic cells,” J. Photonics Energy, 12 (2), 022206 https://doi.org/10.1117/1.JPE.12.022206 (2022). Google Scholar

405. 

S. D. Vorndran et al., “Off-axis holographic lens spectrum-splitting photovoltaic system for direct and diffuse solar energy conversion,” Appl. Opt., 55 (27), 7522 https://doi.org/10.1364/AO.55.007522 APOPAI 0003-6935 (2016). Google Scholar

406. 

D. Zhang et al., “Spectrum-splitting photovoltaic system using transmission holographic lenses,” J. Photonics Energy, 3 (1), 034597 https://doi.org/10.1117/1.JPE.3.034597 (2013). Google Scholar

407. 

B. D. Chrysler and R. K. Kostuk, “Volume hologram replication system for spectrum-splitting photovoltaic applications,” Appl. Opt., 57 (30), 8887 https://doi.org/10.1364/AO.57.008887 APOPAI 0003-6935 (2018). Google Scholar

408. 

A. D. Cronin et al., “Holographic CPV field tests at the Tucson Electric Power solar test yard,” in 37th IEEE Photovoltaics Spec. Conf., (2011). https://doi.org/10.1109/PVSC.2011.6186423 Google Scholar

409. 

B. D. Chrysler and R. K. Kostuk, “A comparison of spectrum-splitting configurations for high-efficiency photovoltaic systems with perovskite cells,” IEEE J. Photovoltaics, 12 (6), 1477 –1486 https://doi.org/10.1109/JPHOTOV.2022.3195688 IJPEG8 2156-3381 (2022). Google Scholar

410. 

J. Werner, B. Niesen and C. Ballif, “Perovskite/silicon tandem solar cells: marriage of convenience or true love story? – An overview,” Adv. Mater. Interfaces, 5 (1), 1700731 https://doi.org/10.1002/admi.201700731 (2018). Google Scholar

411. 

G. A. Barron-Gafford et al., “Agrivoltaics provide mutual benefits across the food–energy–water nexus in drylands,” Nat. Sustain., 2 (9), 848 –855 https://doi.org/10.1038/s41893-019-0364-5 (2019). Google Scholar

412. 

B. B. Jorgensen, Y. Cohen and D. J. Des Marais, “Photosynthetic action spectra and adaptation to spectral light distribution in a benthic cyanobacterial mat,” Appl. Environ. Microbiol., 53 (4), 879 –886 https://doi.org/10.1128/aem.53.4.879-886.1987 AEMIDF 0099-2240 (1987). Google Scholar

413. 

S. Vorndran et al., “Holographic diffraction-through-aperture spectrum splitting for increased hybrid solar energy conversion efficiency: holographic diffraction-through-aperture spectrum-splitting system,” Int. J. Energy Res., 39 (3), 326 –335 https://doi.org/10.1002/er.3245 IJERDN 0363-907X (2015). Google Scholar

414. 

, “Holograms increase solar energy yield,” https://spie.org/news/holograms-increase-solar-energy-yield?SSO=1 (2021). Google Scholar

415. 

, “The world needs more diverse solar panel supply chains to ensure a secure transition to net zero emissions,” https://www.iea.org/news/the-world-needs-more-diverse-solar-panel-supply-chains-to-ensure-a-secure-transition-to-net-zero-emissions (2022). Google Scholar

416. 

DOE Solar Energy Technologies Office, Federal Tax Credits for Solar Manufacturers, (2022). Google Scholar

417. 

Ultra Low-Carbon Solar Alliance, “South Korea is implementing carbon footprint assessment regulations for the PV market,” (2021). https://ultralowcarbonsolar.org/blog/south-korea-implementing-carbon-footprint-assessment-regulations/ Google Scholar

418. 

A. Anctil, “Comparing the carbon footprint of monocrystalline silicon solar modules manufactured in China and the United States,” in IEEE 48th Photovoltaics Spec. Conf. (PVSC), (2021). https://doi.org/10.1109/PVSC43889.2021.9518632 Google Scholar

419. 

V. Fthenakis and E. Leccisi, “Updated sustainability status of crystalline silicon-based photovoltaic systems: life-cycle energy and environmental impact reduction trends,” Prog. Photovoltaics Res. Appl., 29 (10), 1068 –1077 https://doi.org/10.1002/pip.3441 PPHOED 1062-7995 (2021). Google Scholar

420. 

L. Yuan and A. Anctil, “Material use and life cycle impact of crystalline silicon PV modules over time,” in IEEE 49th Photovoltaics Spec. Conf. (PVSC), (2022). https://doi.org/10.1109/PVSC48317.2022.9938923 Google Scholar

421. 

H. Zhang et al., “Green or not? Environmental challenges from photovoltaic technology,” Environ. Pollut., 320 121066 https://doi.org/10.1016/j.envpol.2023.121066 (2023). Google Scholar

422. 

P. Nain and A. Kumar, “Understanding metal dissolution from solar photovoltaics in MSW leachate under standard waste characterization conditions for informing end-of-life photovoltaic waste management,” Waste Manage., 123 97 –110 https://doi.org/10.1016/j.wasman.2021.01.013 WAMAE2 0956-053X (2021). Google Scholar

423. 

D. Sica et al., “Management of end-of-life photovoltaic panels as a step towards a circular economy,” Renew. Sustain. Energy Rev., 82 2934 –2945 https://doi.org/10.1016/j.rser.2017.10.039 (2018). Google Scholar

424. 

S. Mahmoudi, N. Huda and M. Behnia, “Multi-levels of photovoltaic waste management: a holistic framework,” J. Cleaner Prod., 294 126252 https://doi.org/10.1016/j.jclepro.2021.126252 (2021). Google Scholar

425. 

L. Hocine and K. M. Samira, “Optimal PV panel’s end-life assessment based on the supervision of their own aging evolution and waste management forecasting,” Sol. Energy, 191 227 –234 https://doi.org/10.1016/j.solener.2019.08.058 SRENA4 0038-092X (2019). Google Scholar

Biography

Annick Anctil is an associate professor in Civil and Environmental Engineering at Michigan State University (MSU), where she leads research on sustainable energy systems. She uses anticipatory lifecycle assessment to reduce the environmental impact of new technologies. She received an NSF CAREER award in 2021 to work on the impact of the solar photovoltaics industry in the US. At MSU, she teaches classes on sustainability and life cycle assessment of energy.

Meghan Beattie is a postdoctoral fellow with the SUNLAB in the School of Electrical Engineering and Computer Science at the University of Ottawa. She holds a PhD in Physics from the University of Ottawa (2021) and a BSc in Physics from Queen’s University (2015). Her research interests include materials and devices for advanced photovoltaic energy conversion and the development of photovoltaic receivers for optical wireless and fiber power transmission.

Christopher Case began his career as Assistant Professor of Engineering at Brown University, followed by ten years at AT&T; Bell Laboratories. He also had ten years as Chief Technology and Scientific Officer with BOC, part of Linde, where he oversaw the global technology strategy and R&D; for the €8 billion business. He has helped companies large and small to manage their IP portfolios, broaden their scope of influence, and build successful and profitable businesses. Chris actively contributes to advancing the solar industry including serving as a board member of the European Solar Manufacturing Council and a Steering Committee member of the European Technology and Innovation Platform for Photovoltaics (ETIP-PV). Since 2021, he has been the President of the International Thin-Film Solar Industry Association (PVthin).

Aditya Chaudhary studied power engineering at National Power Training Institute, New Delhi, India, and graduated in 2014. Afterwards, he completed a master’s degree in electrical sustainable energy from TU Delft, Delft, Netherlands, in 2017. His master thesis involved development of four terminal mechanically stacked a-SiOx/c-Si solar cells. From January 2018 to April 2022, he worked at ISC Konstanz, Konstanz, Germany, as a PhD researcher. The topic of his work was screen-printed metallization of polysilicon/SiOx passivated contacts. After the completion of the project at ISC Konstanz he moved to ipv (University of Stuttgart), to work as a postdoc in the group headed by Dr. Stephanie Essig, developing perovskite/Silicon tandem solar cells.

Michael G. Debije is an associate professor in the Stimuli-responsive Functional Materials and Devices group at the TU Eindhoven. He has been working on the luminescent concentrator device for fifteen years as part of his research line of light control in the built environment. His other projects relate to the fabrication and study of environmentally responsive liquid crystal-based actuators for soft robotics.

Stephanie Essig is heading the high-efficiency solar cells group of Institute for Photovoltaics at the University of Stuttgart. Previously, she performed research on III-V/Si multi-junction, GaAs nanowire and CIGS thin-film solar cells at NREL in Colorado, EPFL in Switzerland, Sol Voltaics AB in Sweden, and ZSW in Germany. She holds an MPhys degree from Heriot-Watt University Edinburgh, a diploma in physics from Karlsruhe Institute of Technology, and a PhD from the University of Konstanz.

Vivian E. Ferry is an associate professor at the University of Minnesota in the Department of Chemical Engineering and Materials Science. She received her SB in chemistry from the University of Chicago, and her PhD from the California Institute of Technology, working with Prof. Harry Atwater. She was a postdoctoral fellow with Prof. Paul Alivisatos at Lawrence Berkeley National Laboratory. Her research interests include light management in photovoltaics, tunable metamaterials, and nanoscale chirality.

Karin Hinzer is a professor at the School of Electrical Engineering and Computer Science at the University of Ottawa and the University Research Chair in Photonic Devices for Energy. She has published >210 refereed papers, trained >190 highly qualified personnel, and her laboratory has spun-off three Canadian companies in the energy sector. Her research interests include new materials, high efficiency light sources and detectors, solar cells and modules, new electrical grid architectures and voltage converters.

Raymond K. Kostuk is a professor in the Department of Electrical and Computer Engineering and the Wyant College of Optical Sciences at the University of Arizona. He has published extensively on the application of holography and optics in PV systems and summarized many of these concepts in the SPIE Spotlight e-book Holographic Applications in Photovoltaic Energy-Conversion Processes. He is a fellow of Optica and SPIE.

Daniel Kirk completed his doctoral work in Oxford and working in industrial research at Dupont (metallization development) and Nexeon (Li-ion battery technology). He joined Oxford PV in 2014. Since 2015 he has been heavily involved in optimization of both perovskite and silicon cells for tandem applications. More recently he has focused on supporting the scaling up the technology to full-wafer scale at Oxford PV, as well as on optimization of the product for commercial realization in various markets.

Stephan Kube is Head of Marketing at Heliatek since 2019. He has more than ten years of experience in B2B marketing in the emerging international solar industry. Stephan has worked for the technology leaders Solar Frontier (CIGS) and now Heliatek (OPV), driving their brand awareness and business success.

Preeti Nain is a postdoctoral researcher in the Department of Civil and Environmental Engineering at Michigan State University. She received her PhD in Environmental Engineering from Indian Institute of Technology, Delhi, India. Her research focuses on different aspects of end-of-life solar photovoltaics. Her current interests include PFAS analysis in solar photovoltaics and life cycle assessment for PV repowering in the United States.

Angèle Reinders is a full professor in Design of Sustainable Energy Systems in the Energy Technology Group at the TU Eindhoven, and the Director of Solliance. She aims at an optimal integration of solar technologies in products, buildings and local infrastructures by design-driven research on these technologies. She simulates, prototypes and experimentally characterizes, among others, luminescent solar concentrator PV, BIPV and VIPV. She is a physicist with a track record in PV performance research.

Ifor D. W. Samuel is Professor of Physics at the University of St Andrews. He received his MA and PhD from the University of Cambridge, working on optical spectroscopy of organic semiconductors. In 2000, he moved to the University of St Andrews where he founded and leads the Organic Semiconductor Centre. His current work concerns the photophysics of organic semiconductor materials and devices. He is a fellow of the Royal Society of Edinburgh and of SPIE.

Wilfried van Sark is full professor “Integration of Photovoltaics” at the Copernicus Institute of Sustainable Development at Utrecht University. He is an experimental physicist by training (MSc/PhD), and has over 40 years of experience in photovoltaics. He worked on various material systems such as crystalline and thin film silicon and III-V solar cells, both experimentally and theoretically. Currently he focuses on building and grid integration focusing on luminescent solar concentrators and bidirectional charging of electric vehicles.

Hele Savin received the DSc (Tech) degree in semiconductor technology from the Helsinki University of Technology, Espoo, Finland, in 2005. She is currently a professor with the Department of Electronics and Nanoengineering, Aalto University, Espoo. Her current research interests focus on photonic devices, especially photon detection and solar cells, nanofabrication methods, and defect engineering in group-IV materials.

Sean E. Shaheen is a professor of electrical, computer, and energy engineering; professor by courtesy of physics; and fellow of the Renewable and Sustainable Energy Institute at the University of Colorado Boulder. He obtained his BS and PhD in physics from Carnegie Mellon University (1991) and the University of Arizona (1999), respectively. His research interests include organic and hybrid photovoltaics, photonics, and neuromorphic/bio-inspired computing. He is currently the editor-in-chief of the Journal of Photonics for Energy.

Fatima Toor is an endowed chair and associate professor in the Department of Electrical and Computer Engineering at the University of Iowa. She earned her PhD in electrical engineering from Princeton University in 2009 and was a postdoctoral researcher in the silicon PV group at NREL from 2010 to 2011. Her team’s current research involves the development of semiconductor optoelectronic devices for applications in the health, environment, and energy industries. She is a senior member of SPIE.

Ville Vähänissi received the DSc (Tech) degree in semiconductor technology from Aalto University, Espoo, Finland, in 2016. He is currently working as a staff scientist at the Department of Electronics and Nanoengineering, School of Electrical Engineering, Aalto University. His research interests include defect engineering in silicon, especially photovoltaics, as well as atomic layer deposition and nanopatterning, and their various applications in semiconductor devices.

Emily L. Warren is a senior scientist and group manager at the National Renewable Energy Laboratory where she leads a core program on the development of high efficiency tandem solar cells. She received her PhD from the California Institute of Technology, an MPhil from the University of Cambridge in Engineering for Sustainable Development, and a BS in Chemical Engineering from Cornell University. Her research interests include the development and modeling of tandem solar cells and modules, photoelectrochemical production of solar fuels, and heteroepitaxy of III-V semiconductors.

Han Young Woo received his PhD degree in chemistry from the Korea Advanced Institute of Science and Technology (KAIST), Republic of Korea, in 1999. After postdoctoral training at the University of California, Santa Barbara (UCSB), USA, he joined the Pusan National University as an assistant professor. In 2015, he moved to Korea University where he is currently a professor in the Department of Chemistry. His current research focuses on conjugated polymers and polyelectrolytes for applications in organic optoelectronic devices.

Gang Xiong currently is the Director of First Solar California Technology Center, managing the site operation and research programs such as CdTe, perovskite, and tandem. He joined First Solar in 2007 and has since worked on cell/module research and manufacturing implementation. Representing First Solar, Dr. Xiong also serves at a few industrial advisory boards (IABs), including as the IAB chair of US Manufacturing of Advanced Cadmium Telluride Photovoltaics, the former IAB chair of Solar Powered Future 2050 under the National Science Foundation, and the IAB member of US Manufacturing of Advanced Perovskite.

Xitong Zhu is a PhD candidate in the Energy Technology Group at TU Eindhoven. He has been focused on ray-tracing simulation of luminescent solar concentrator photovoltaics (LSC PV). Currently, he focuses on different geometry LSC PV devices and the application feasibility of LSCs for BIPV. He is devoted to research on the applications of sustainable energy.

Biographies of the other authors are not available.

© 2023 Society of Photo-Optical Instrumentation Engineers (SPIE)
Annick Anctil, Meghan N. Beattie, Christopher Case, Aditya Chaudhary, Benjamin D. Chrysler, Michael G. Debije, Stephanie Essig, David K. Ferry, Vivian E. Ferry, Marina Freitag, Isaac Gould, Karin Hinzer, Harald Hoppe, Olle Inganäs, Lethy Krishnan Jagadamma, Min Hun Jee, Raymond K. Kostuk, Daniel Kirk, Stephan Kube, Minyoung Lim, Joseph M. Luther, Lorelle Mansfield, Michael D. McGehee, Duong Nguyen Minh, Preeti Nain, Matthew O. Reese, Angèle Reinders, Ifor D. W. Samuel, Wilfried van Sark, Hele Savin, Ian R. Sellers, Sean E. Shaheen, Zheng Tang, Fatima Toor, Ville Vähänissi, Ella Wassweiler, Emily L. Warren, Vincent R. Whiteside, Han Young Woo, Gang Xiong, and Xitong Zhu "Status report on emerging photovoltaics," Journal of Photonics for Energy 13(4), 042301 (14 December 2023). https://doi.org/10.1117/1.JPE.13.042301
Published: 14 December 2023
Lens.org Logo
CITATIONS
Cited by 2 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Solar cells

Photovoltaics

Solar energy

Perovskite

Silicon

Manufacturing

Design

Back to Top