L. Páramoab,
M. Pains Duarteab,
G. Fuocoab,
A. Adedapoab,
D. P. Singhc and
R. Naccache
*ab
aDepartment of Chemistry and Biochemistry and the Centre for NanoScience Research, Concordia University, Montreal, Quebec H3G 1M8, Canada. E-mail: rafik.naccache@concordia.ca
bQuebec Centre for Advanced Materials, Concordia University, Montreal, Quebec H3G 1M8, Canada
cUniversité du Littoral Côte d'Opale, UR 4476, UDSMM, Unité de Dynamique et Structure des Matériaux Moléculaires, F-62228 Calais, France
First published on 11th August 2025
CuInX2 (where X = S, Se) are among group I–III–VI's most recognized semiconductors in the ternary chalcogenides. These materials have come to the forefront of solar energy research owing to their optical and electronic properties, such as high absorption coefficients, tunable band gaps, and excellent thermal stability. This positions them as sustainable and low-cost alternatives for solar energy conversion devices. However, their practical implementation remains hindered due to several challenges, including low efficiencies associated with insufficient control in defect formation, variable stoichiometries, ligand selection, and binary or ternary compositional mixtures. A particularly limiting barrier is the current lack of understanding of the emission processes in these materials. Bringing these materials to light and their implementation in fabricating energy conversion devices will crystallize through overcoming these challenges. This review critically examines recent advances in the synthesis, characterization, and integration of CuInX2 systems in solar energy-converting devices, highlighting current challenges and opportunities and offering insights that will not only accelerate research in this field but also contribute to the advancement of other ternary or quaternary copper-based chalcogenides.
Solar energy has emerged as one of the most promising renewable alternatives due to its abundance, inexhaustibility, and environmentally friendly features. The sun provides an amount of energy of ∼170000 TWh.5,6 At the same time, according to IEA, in 2023, the global electricity demand was 26
000 TWh per year,2 which supports the idea that the sun can meet global energy demand. As such, photovoltaics (PV), which convert sunlight directly into electricity, have emerged as a key solution to use this widely available energy7 PV cells can be classified into three significant generations: (i) mono- or multi-crystalline silicon, which is the dominant technology utilized globally, (ii) thin-film cells, which includes amorphous silicon, cadmium telluride (CdTe), cadmium sulfide (CdS), as well as copper indium gallium selenide (CIGS) cells, and (iii) the emerging technologies, which includes dye-sensitized, organic, perovskite and quantum dots (QDs) solar cells.6–9 According to the Renewables 2024 report by the IEA,4 PV cells are expected to contribute to 80% of the global expansion of the renewable energy share, making solar power the largest renewable energy source. Although silicon-based solar cells currently dominate the market, they face several limitations related mainly to material and processing cost.7,9 The second and third PV cell generations were developed to reduce costs in this context. Although the second generation showed cost-effectiveness compared to the silicon counterparts, it presented some drawbacks related mainly to low efficiency and toxicity, which hindered large scale applications. Thus, the third generation emerged to overcome production costs and materials inefficiency. These solar cells have attracted significant attention due to their high efficiency, lower costs, and excellent material flexibility.8,11,12 Among them, QD solar cells stand out owing to their unique optical and electronic properties, such as size-tunable bandgap, endowing them with a broader absorption range of the solar spectrum.12–14 Nonetheless, one of the challenges QD-based cells face is their reliance on toxic metals, as the most commonly used QDs comprise Cd and Pb and include CdS, CdTe, PbS, and PbSe.14,15 In an attempt to address these shortcomings, recent research has focused on the design of more sustainable and greener QDs with solar cells based on ternary chalcogenides, with CuInX2 nanocrystals becoming a popular alternative.16,17 This is not only due to the fact that CuInX2 nanomaterials possess appropriate optical and electronic properties, but also their constituent elements are considered to be non-toxic to human health and the environment. Moreover, copper, sulfur, and selenium are relatively abundant. In contrast, while indium is regarded as a scarce material compared to the others, its environmental impact remains low.18–20
CuInS2 and CuInSe2 nanomaterials have been among the most explored compounds for solar energy in the I–III–VI2 ternary chalcopyrite semiconductors group, which belongs to the broad family of Cu-chalcogenides.18 These compounds exhibit a chalcopyrite crystalline structure, where the atoms Cu, In, S, and/or Se are arranged tetrahedrally. They also have an abundance of intrinsic defects, which guarantee them diverse optoelectronic properties.21,22 These compounds have been widely investigated for use in QDs-solar cells due to their unique optoelectronic properties, such as high energy conversion, significant Stokes shifts, tunable band gaps, and high absorption coefficients.13,18–20,22 The ability to tune their direct bandgaps based on the size or elemental composition renders them efficient sunlight absorbers across a broad spectrum of wavelengths, from the visible to near-infrared (NIR).15,18,23 The band gap of CuInS2 is ∼1.5 eV, which matches the solar irradiance spectrum peak. CuInSe2 has a smaller bandgap of ∼1.04 eV, so it absorbs more infrared light. In this context, these materials have proven to be appropriate for photovoltaic applications.20,24,25
Furthermore, CuInX2 materials offer an essential advantage relative to traditional materials used in other PV cells, such as silicon, given their high absorption coefficients often exceeding 105 cm−1.19,24 This enables them to efficiently absorb sunlight in layers typically less than several micrometers thick, compared to the 100 mm layers required in silicon-based PVs.18 As a result, CuInX2 solar cells reduce material usage and production costs, making them a promising option for PV cell technology.8 Given their aforementioned advantages, CuInX2 nanomaterials have not only been applied in solar cells but also explored in biomedicine,26 tandem devices,27 photocatalysis,28 LEDs,29 and luminescent solar concentrators (LSC).30 LSCs have proven to be one of the most significant devices developed in the field. LSCs can be incorporated or doped with different photoluminescent materials, such as organic dyes and QDs, to enhance their properties. Organic dyes were initially used, but their application was limited owing to narrow absorption bands and significant reabsorption losses. On the other hand, QDs offer advantages over organic dyes, such as a larger Stokes shift, broader absorption spectra, and improved power conversion efficiency (PCE), providing a simple and low-cost alternative to converting sunlight into electricity.31,32
Despite their attractive and promising features, applying these nanomaterials in QD solar cells still faces several challenges. Recent research has been focused on addressing these issues to improve performance and efficiency. One major challenge is the incomplete understanding of their emission mechanisms, which hinders the development of materials with enhanced properties, such as high quantum yields (QY).24,33,34 Another challenge is associated with the synthesis, including the selection of suitable ligands, which play a crucial role in charge transfer efficiency,15,18 and the insertion of selenium into the structure due to the limited availability of Se precursors compared to sulfur.35 Additionally, strategies to improve the PCE of QD-based solar cells have been explored, with recent studies focusing on grain boundaries, which can enhance charge transfer and consequently enhance device performance.18
Thus, this review aims to provide a comprehensive overview of recent design and application strategies for CuInX2 nanomaterials for solar energy conversion. The review also explores the fundamental properties of CuInX2 materials, focusing on their structure, optical and electrical properties, recent progress, research gaps in the synthesis of these materials, their emission processes, and their role in solar energy conversion devices such as PV cells and LSCs.
In CuInX2 (X = S, Se), copper exists in the +1 oxidation state, indium in the +3 oxidation state, while S or Se are in the −2 oxidation state.41 In terms of crystalline structure, CuInX2 (X = S, Se) can exist in three phases, namely chalcopyrite,42–47 zinc blend,48–50 and wurtzite.41,50 Among these, chalcopyrite is the most reported in nanomaterials synthesis due to its low energy of formation and high stability.36,39,41,42,51 The chalcopyrite unit cell structure can be constructed from the traditional zinc blend structure in ZnS by the alternate substitution of Zn sites with In and Cu cations; the differences in bond lengths between Cu and In cause a distortion in the unit cell, leading to a tetragonal unit cell, on the other hand CuInS2 and CuInSe2 in zinc blende phase possess an structure in which the Cu and In atoms are randomly organized within the structure leading to a cationic disorder within a cubic lattice.18 Unit cells of the chalcopyrite, zinc blende, and wurtzite can be seen in Fig. 1a. The distinction between the chalcopyrite phase and CuInX2 zinc blende can be challenging since both phases share a Bragg reflection of high intensity at similar angles,19 as seen in Fig. 1a. The distinction between the two phases commonly relies on the lowest intensity reflections, although signal broadening opacifies the low-intensity reflections in nanomaterials.19 Studies have shown that the use of different ligands can lead to materials with similar crystallographic characteristics, as demonstrated by Harvey et al.44 where the use of ligands like oleylamine (OLA), diphenylphosphine, and tributyl phosphine yielded nanomaterials in the chalcopyrite phase44 (Fig. 1b). On the other hand, the wurtzite structure exhibits Bragg reflections located at angles that can be clearly distinguished from the other two crystalline structures.41 A change in anions from S to Se in the crystal structure can be observed as a shift of the Bragg reflections associated with an increase in the lattice parameter due to the incorporation of a larger ion,41 as seen in Fig. 1c.
![]() | ||
Fig. 1 (a) Unit cells of chalcopyrite, zinc blende, wurtzite, and simulated XRD spectra of the three CuInS2 crystalline phases, adapted with permission from ref. 19. Copyright 2016 American Chemical Society. (b) XRD diffractogram of chalcopyrite CuInSe2 synthesized with different surface ligands, adapted with permission from ref. 40. Copyright 2021 American Chemical Society. (c) XRD diffractogram of wurtzite CuIn(Sx/Se1−x)2 with different compositions of S/Se, adapted with permission from ref. 37. Copyright 2024 American Chemical Society. |
CuInS2 and CuInSe2 are semiconductors with structures rich in various types of defects, which are associated with their optical properties, such as PL. These properties have led to the materials being commonly described as defect-tolerant.52 Although it is recognized that defects guide the emission processes, the scientific community has not yet reached a consensus on the exact origins of these mechanisms.19 The existence of multiple stoichiometric ratios is closely related to the formation of defects such as copper vacancies and indium antisite pairs (2vCu− + InCu2+).46
Composition is also affected by the choice of ligand, which depends on their chemical properties, as well as their reactivities with metal and chalcogen precursors. Commonly, the hard and soft acid–base (HSAB) theory has been used as a guiding tool to control the reactivity of selected precursors.44 A comparison between three ligands, tributyl phosphine (TBP), diphenylphosphine (DPP), and OLA, shows that among the three, TBP promoted and increased indium content in the samples compared to OLA and DPP, where the Cu:
In ratio was higher.44 The synthesis temperature also had a profound effect on composition, as demonstrated by Houck et al.46 where samples of CuInSe2 synthesized at 180 °C were copper deficient (Cu
:
In = 0.52). The Cu
:
In ratio increased with temperature, yielding a value of 0.72 at 210 °C and reaching 0.92 at 240 °C.46
![]() | ||
Fig. 2 (a) Size dependence of molar absorption coefficient and absorption spectra of C QDs. Adapted with permission from ref. 54. Copyright 2018 American Chemical Society. (b) Band positions on CuInSxSe1−x based on chalcogen composition, adapted with permission from ref. 37. Copyright 2024 American Chemical Society, (c) emission spectra of CuInSe2 QDs with varying ratio of Cu/In adapted with permission from ref. 55. Copyright 2012 American Chemical Society. (d) Absorption and emission spectra of CuInSe2 with varying sizes, adapted with permission from ref. 56. Copyright 2013 American Chemical Society. |
Due to their tunable stoichiometry, researchers can manipulate the ratios of reagents such as Cu, In, and S/Se to obtain mixed ternary compositions. Variations in the Cu:
In ratio in CuInSe2 influence their optical and crystalline properties. A higher indium-to-copper ratio results in materials with less crystalline structures and fewer defects compared to copper-rich compositions, which are associated with a more ordered crystalline structure.54 In a study by Guoshuai Wang, CuInS2 with different Cu:In compositions was synthesized. It was observed that an increasing concentration of indium caused a blue shift in the PL spectra along with an increase in PL intensity52 (Fig. 2c). Similarly, Chen et al.57 reported an increase in PL intensity with higher indium content in copper-deficient CuInSe2; however, at high indium concentrations, while the PL spectra did not exhibit further shifts, the PL intensity decreased57 (Fig. 2c). Emission spectra can also be redshifted by the increase of the QD size as seen in Fig. 2d, which also results in a redshift of the absorption spectrum.58
Another critical factor influencing the opto-electronic properties is the selection of ligands used during the synthesis of CuInSe2.44 Ligand selection can lead to the formation of either hydrophilic or hydrophobic materials, with hydrophobic materials being the most extensively studied. High-boiling-point solvents, such as octadecene, OLA, and oleic acid (OA), are well-established for the synthesis of colloidal nanomaterials and QDs via hot injection methods.44 Early studies on the formation of nanomaterials, such as CuInS2 and CuInSe2, relied on phosphine ligands, enabling the synthesis of various structures with a wide range of tunable properties. However, in recent years, efforts to develop sustainable synthetic routes using less toxic reagents have led to a reduction in the use of phosphine ligands in favour of alternative surfactants, such as OLA and OA.59–65 Among these ligands, dodecanethiol (DDT) stands out as a versatile, high-boiling-point solvent that plays multiple roles during the synthesis process. DDT serves as the primary solvent, stabilizing agent, and sulfur source for synthesizing binary, ternary, and quaternary compositions. Its utility has also been extended to the synthesis of CuInSe2, although the incorporation of selenium creates unique challenges that will be addressed in the synthetic routes section (vide infra).38,66–71 Studies have shown that using different ligands, such as tributylphosphine, OLA, and diphenylphosphine, yields CuInS2 materials with varying emission characteristics.44 While the emission wavelengths remain similar due to comparable particle sizes, QY values differ significantly. For example, CuInS2 QDs capped with OLA exhibited the lowest QY (0.003%), while those synthesized with tributylphosphine achieved the highest values (1.2%).44 This variation in QY is typically attributed to the ligands' ability to control the defect formation and passivate surface sites, which influence non-radiative emission pathways.44
In a research article published in 2016, Leach and Macdonald19 discussed several works to explain the CuInS2 emission processes. In this work, various mechanisms were attributed to the observed photoluminescence (PL), including transitions from the conduction band (CB) to a localized state in the bandgap (BG), donor–acceptor mechanisms, transitions between a localized state and the valence band (VB), as well as the presence of Cu2+ impurities.19 The same year, Berends et al.73 analyzed CuInS2 nanocrystals and their counterparts shelled with ZnS and CdS. An analysis of CdS-shelled nanocrystals using time-resolved (TAPL) analysis yielded a negative signal observed at ∼530 nm, which was attributed to band-edge bleaching. The authors stated that a donor-to-acceptor type of mechanism requires two trap charges; thus, a band-edge bleach signal shouldn't be observed, ruling this mechanism out.73 Instead, recombination was associated with a Cu trap, where the radiative decay arises from recombining a delocalized electron to a Cu-related trap. The non-radiative decay pathway was associated with an electron trapping at the material's surface, which, upon shelling, is eliminated, thus increasing PL intensity and lifetime.73
Subsequently, Fuhr et al.33 identified reduction and oxidation signals in CuInS2 samples at 350 meV above the VB, which were attributed to the Cu1+ and Cu2+ oxidation states. In their work, both oxidation states were reported to be PL active, although their distribution depends on the copper content in the samples. Copper-deficient samples were enriched in Cu2+ due to charge compensation arising from an increased concentration of copper vacancies. In their proposed mechanism, Cu2+ PL involves the recombination of a CB electron with a pre-existing hole in the Cu2+ as it has an incomplete 3d shell ([Ar]3d9). Furthermore, the authors argued that the removal of the photogenerated hole from the VB must occur via an existing trap, such as a VCu, for this pathway to become dominant, as recombination persists for hundreds of nanoseconds. Upon recombination, the reduced Cu2+ to Cu+ requires capturing the hole trapped by the VCu to recover its +2 state. On the other hand, the mechanisms involving Cu+ centers involve a capture of a hole to oxidize the Cu+ to Cu2+, which recombines with the CB electron to form the Cu+ species.33 A schematic of the recombination process is shown in Fig. 3a.
![]() | ||
Fig. 3 (a) Proposed emission pathways for stochiometric, copper-deficient, and core–shell CuInS2, adapted with permission from ref. 34. Copyright 2017 American Chemical Society. (b) LUMO and HOMO percentage of contribution of Cu and Zn for ZnS cluster with increased content of Cu and In, adapted with permission from ref. 73. Copyright 2018 American Chemical Society. (c) Proposed mechanisms for Cu+ and Cu2+ trap states in ensemble and single CuInS2 crystals, adapted with permission from ref. 74. Copyright 2018 American Chemical Society. (d) Optical properties of individual CuInS2/CdS and schematic description of hole localization on the individual crystal. Adapted with permission from ref. 76. Copyright 2021 American Chemical Society. (e) Schematic of differences between PL decay rates based on QD size and trap localization, adapted with permission from ref. 79. Copyright 2023 American Chemical Society. |
In a work published by Nelson and Gamelin,74 a deep DFT density of states (DOS) analysis of several ZnS structures with different concentrations of Cu+ and In3+ indicated that the VB states acquire a small contribution of the Cu 3d orbital by the single addition of one Cu cation into the structure, which increases as the amount of copper is increased in the structure. On the other hand, the In contribution of the DOS at the CB was not accompanied by a drastic change as compared to copper. As the concentration gradually increased, the VB energy decreased. HOMO analysis of the simulated structure revealed a high contribution of Cu in structures simulated with just one copper atom, as seen in Fig. 3b. Further increments in the copper concentration showed strong HOMO delocalization near 2–3 adjacent copper cations with no further delocalization as the copper concentration increased. Based on the obtained data, PL mechanisms were described as the recombination of a photogenerated hole localized near a Cu atom with an electron in the CB, further supporting the idea of the recombination related to a copper defect.74
At the same time, van der Stam et al.75 investigated the effects of electron injection and extraction on trap distribution in CuInS2 through spectroelectrochemistry.75 Their findings revealed that Cu+ traps are solely responsible for radiative recombination processes, contrary to the notion that Cu+ and Cu2+ traps contribute to the emission. It was mentioned that Cu2+ traps require hole trapping to enable radiative recombination. van der Stam's experiments showed that electron injection into the material increased the PL of CuInS2. This behavior was attributed to the reduction of Cu2+, which acts as non-emissive “dark centers”, to Cu+; the latter functions as emissive “bright centers”. Conversely, oxidizing the sample caused a decrease in PL intensity.75 The schematic of the proposed mechanism of dark and bright centers is shown in Fig. 3c. The reduction of Cu+ not only enhanced PL intensity but also altered its bandwidth and peak position due to the elevation of the Fermi level. This process can be better understood by considering that the activated Cu+ can participate in radiative recombination by trapping photogenerated holes, whereas Cu2+ promotes non-radiative Auger-assisted trapping, rendering those traps non-emissive.75 The authors aimed to quantify the concentration of Cu2+ atoms in the material, but they concluded that the concentration was too low to be detected using in situ XANES techniques. As noted, the precise localization of these traps remains an open question.75
Fuhr et al.33 discussed that both oxidation states are responsible for PL, in contrast to van der Stam et al.,75 who indicated that Cu2+ states are non-emissive, as electrochemical oxidation of CuInS2 leads to a decrease in the material's emissivity.75 Fuhr et al.76 gave further insight into these matters in their subsequent work, published in 2020, where, despite copper being present in two different oxidation states, both can participate in the emission processes through various mechanisms. Cu+ with a filled 3d10 shell only requires trapping a photoexcited VB hole before recombination. On the other hand, Cu2+ with a 3d9 shell can directly participate in emission by trapping a CB electron, as it already has a hole. However, the photoexcited hole must leave the valence band (VB) state for the Cu2+ cations to trap the conduction band (CB) electron radiatively. Their study further illustrated the presence of both oxidation states through magnetic dichroism, where samples with filled 3d10 shells are nonmagnetic, while those with 3d9 configurations, which possess unpaired spins, showed magnetic behaviour. As stated in their work, both mechanisms can coexist, and the dominance of one over the other depends on the sample's stoichiometry. Furthermore, it was suggested that the location of the defect where Cu+ states are antisite CuIn–InCu,, while Cu2+ emerges as charge compensation for copper-deficient samples that are rich in copper vacancies.76
Over the last years, single-nanocrystal analysis has been essential to understand that the properties like the broad PL emission correspond to an average emission arising from a population of QDs with distinct emissive properties. As shown in the work of Hinterding et al.,77 the difference in the excitation spectra of individual CuInS2 QDs compared to their ensemble is supported by: the variation among QDs, which can arise due to differences in composition, size, and shape, as well as spectral diffusion analysis of the QDs, which indicates a recombination of a delocalized electron with a localized hole. The authors proposed that the hole can be localized on the Cu+ belonging to the crystal lattice, or on one specific Cu site near a defect, or surface charge. The latter mechanism was preferred as the PL peak position and lifetime measurements supported the idea of the hole localization in a preferential Cu site based on QDs properties like composition, surface ligands, and shape,77 The optical properties of individual nanocrystals analyzed can be observed in Fig. 3d, along with a schematic of hole localization.
Later in 2021, Xia et al.,78 through single low temperature QD spectroscopy, analyzed the differences in the spectral features of individual ZnS/CuInS2 QDs, where two types of emission features were observed. Firstly, a classic broad emission (FWHM ∼210 meV) again attributed to recombination of a delocalized CB electron with a localized copper-related hole, as well as a sharp, narrow (emission FWHM ∼1 meV), which was attributed to a band edge exciton recombination. Both types of emission were not observed in the same QD, theorizing that the position of the Cu traps defines the PL mechanism. When the Cu traps are located near the QD surface band edge, recombination dominates due to better passivation. On the other hand, when Cu traps are located at the core, a free-to-bound recombination dominates due to incomplete passivation.78
Cu2+ action in the PL mechanism was reinvestigated by Harchol et al.,79 where spin Hamiltonian simulations in CuInS2/CdS indicated a strong coupling of the photogenerated hole with a Cu2+ trap, suggesting that the recombination involves a self-trapped exciton at a photogenerated Cu2+.79 Further confirmation of a free-to-bound type of PL mechanisms was provided by Szymura et al.80 At low temperatures, the PL decay rate decreased, which, as described by the authors, wouldn't be compatible with an excitonic recombination. Additionally, the application of a magnetic field did not affect PL lifetimes as compared to binary chalcogenides, which suffer from PL lifetime reduction. The authors also showed that size and copper trap localization are essential components in defining PL dynamics, where large QDs suffer from fast decay compared to smaller counterparts. Moreover, when Cu traps are positioned near the surface, the PL decay is slower, while traps in the core lead to faster decay. In contrast, core-located traps show fast decay80 (Fig. 3e).
In summary, while the consensus points towards a free-to-bound mechanism involving Cu-based trap states, the exact role of Cu2+ versus Cu+ and the location of these traps (core vs. surface) remain as open questions.
![]() | ||
Fig. 4 (a) Chemical processes that can occur during the passivation of CuInS2 with ZnS, where mechanisms like etching, epitaxial growth, alloying, cation exchange, and ZnS nucleation can occur, adapted with permission from ref. 80. Copyright 2018 American Chemical Society. (b) Differences in PL emission from CuInSe2 samples core–shelled with ZnS and ZnSe, adapted with permission from ref. 56. Copyright 2013 American Chemical Society. (c) Infrared photographs of polymer films coated with ZnS/CuInS2 quantum dots, adapted with permission from ref. 49. Copyright 2024 American Chemical Society. (d) Effect on PL emission of the amount of ZnS layers on CuInS2 quantum dots alloyed with Zn, adapted with permission from ref. 67. Copyright 2019 The Royal Society of Chemistry. (e) Emission properties of ZnS/CuInS2 under different reaction times, adapted with permission from ref. 85. Copyright 2011 The Royal Society of Chemistry. (f) Emission properties of CuInS2 quantum dots with different increasing amounts of ZnS layers, adapted with permission from ref. 83. Copyright 2015 American Chemical Society. |
While ZnS is also often used for shelling CuInSe2, studies have proven a better effect of passivation on CuInSe2 due to a minor lattice mismatch of approximately 2% compared to ZnS (4–6%),58 offering better optical properties as seen in Fig. 4b. However, the use of ZnSe has been less extensively reported, primarily due to the greater complexity associated with selenium precursors. This is in contrast to sulfur precursors like DDT that are commonly employed in shelling processes71 alongside various zinc-based precursors (e.g., ZnAc,36,51,68,88,89 ZnCl2,71 ZnI2,81 Zn(St)2,32,81,82,90,91 and ZnEt2 (ref. 58)).
ZnS has been demonstrated to improve PL on multiple occasions, increasing the PLQY value up to ∼92.1%, as achieved by Liu et al.,36 producing optical films with strong emission in the infrared region of the spectrum as seen in Fig. 4c; other examples are shown in Fig. 4d–f, where PL properties are enhanced as a function of shell time, or the amount of ZnS layers applied, achieving materials with emission in a broad range of colours. The introduction of a shell layer has brought significant improvements to shelf life, with reports of up to 6 months at 4 °C, accompanied by negligible loss of PL.92 Additionally, optical films prepared with CuInS2/ZnS have demonstrated good stability over a period of 1 month under ambient conditions. Another LED was reported to maintain 90% brightness over 30 days if stored in a glove box.71 In Lian et al.'s.93 work, the core–shell could maintain 84.2% PL after 2 months, while the core CuInSe2 samples' PL only lasted 6 days.93
ZnS possesses a bandgap of 3.68 eV, which is wider than that of CuInS2 or CuInSe2, making it a suitable material for the growth of a type-1 core–shell. As such, it is ideal for applications like light-emitting devices36 and solar concentrators, which require a high PLQY while maintaining the advantages of CuIn(S, Se)2, namely, a large Stokes shift. However, it is commonly reported that either alloying or the shelling process results in a blue shift of the PL spectra, making it possible to produce different structures emitting from the red to green regions of the spectrum,26,58,68,82–86 although it is still possible to maintain a low degree of blueshift.36
The PL of CuIn(S, Se)2 QDs significantly improves via the epitaxial growth of a ZnS shell, as evidenced by temperature-dependent PL studies and transient absorption (TA) measurements, where the ZnS shell inhibits electron–phonon interactions and Auger recombination processes by effectively passivating particle surface defects.36 Reduction in nonradiative recombination is also mentioned in the works of Yamashita et al.88 and Xia et al.,94 where the growth of a ZnS shell protects the surface of the core, suppressing surface trap-mediated recombination.88,94 More recently, Xia et al.,78 used low-temperature single QDs spectroscopy in a sample of CuInS2/ZnS, where two subpopulations of QDs with different PL behaviour were found in their study. One subpopulation exhibited sharp, resolution-limited emission lines attributed to the zero-phonon recombination of a long-lived band-edge exciton, resulting from the ZnS shell's perfect passivation. In contrast, the other population possessed broad spectra attributed to the self-trapping of holes by defects arising from non-well-passivated QDs.78
Identifying a core–shell structure formation in multiple studies is often associated with increased size distribution of QDs.36,58,84,92,93,95,96 This size increase can be attributed to the deposition of layers of ZnS or ZnSe, whose monolayer thicknesses are approximately 0.31 and 0.33 nm, respectively.82 Elemental analysis, such as EDS coupled with HRTEM, can serve to identify the elemental distribution, which can be associated with shell formation, as reported by Liu et al.36 PL and absorption spectra can also provide insights: an increase in PLQY with minimal PL blueshift indicates successful core–shell formation rather than an alloying process.58 However, as Darren Neo et al.82 noted, PL and absorption spectra alone cannot fully distinguish between passivation via alloying and core–shell formation.82
Core–shell formation with ternary chalcogenides like CuInS2 has also been explored, as demonstrated in the work by Ning et al.38,69 However, the PLQY achieved in these studies was not as high as in other works using ZnS shells, with a maximum PLQY of 20% reported for CuInSe2/CuInZnS2−xSex.38 These findings suggest that further refinement of growth conditions and shell engineering could enhance the PLQY. Notably, their study demonstrated the successful generation of core–shell structures exhibiting quasi-type I (CuInSe2/CuInZnS2−xSex), type II (CuInSe2/CuInS2), and quasi-type II or type II (CuInSe2/CuInS2−xSex).38 Additionally, a more focused investigation on the shell growth of CuInSe2 revealed that the choice between chalcopyrite and wurtzite shell structures depends on the number of precursors supplied during the shell formation process.69
![]() | ||
Fig. 5 (a) Synthesis of CuInS2 QDs in bubble wrap, adapted with permission from ref. 96. Copyright the Royal Society of Chemistry 2019. (b) Production of CuInS2 nanocrystals through cation exchange, starting from Cu2−xS, adapted with permission from ref. 54. Copyright American Chemical Society 2018. (c) Effect of halide ions (I, Cl, and Br) in the growth dynamics of CuInS2 and their respective TEMs, adapted with permission from ref. 100. Copyright The Royal Society of Chemistry 2020. (d) Use of metal xanthanates for the synthesis of CuInS2 nanoparticles and the ligand exchange to hexanethiol, adapted with permission from ref. 101. Copyright The Royal Society of Chemistry and the Centre National De La Recherche Scientifique 2019 (e) use of diaryldiselenides as Se source for the synthesis of CuIn(SxSe1−x)2, adapted with permission from ref. 37. Copyright American Chemical Society 2024. |
In syntheses involving elemental selenium, the mechanism of nanocrystal formation depends strongly on the ligands used, as they influence the availability and reactivity of the metal precursors. In a study by Kar et al., the effects of coordinating versus non-coordinating solvents such as OLA and octadecene were compared. When OLA was used, MALDI analysis revealed multiple OLA derived species with varying carbon chain lengths (from C17 to C9), as well as a chlorinated OLA derivative. FTIR analysis further confirmed the presence of an alkylamine hydrochloride, suggesting that the metal precursors (e.g., CuCl and InCl3) generate HCl in situ, which reacts with OLA to form its corresponding ammonium salt. XRD analysis of the nanocrystal products showed that under these conditions and with elemental selenium, the formation of CuInSe2 proceeds via intermediate binary phases of CuSe and InSe. In contrast, when non-coordinating solvents such as ODE or octadecane were used, chlorinated derivatives were also observed. However, the binary phases formed under these conditions were Cu2−xSe and In2Se3, rather than CuSe and InSe.107
A deeper look into the CuInSe2 growth mechanism has not been developed as extensively as for CuInS2, although it is theorized that the formation of lamellar structures with the metal components is also involved when ligands such as dodecaneselenol are used as the selenium source.108 This was shown by Hernández-Pagán et al., whose low-angle X-ray diffraction patterns revealed periodic peaks similar to those observed at initial stages in syntheses involving Cu precursors and thiolate ligands.108
For syntheses involving the combination of oleic acid and OLA, it has been found that both ligands react to form dioleylamine, which further stabilizes the metal precursors.106 However, no specific mechanism for the formation of either CuInS2 or CuInSe2 has been proposed. Although it has been mentioned that in reactions involving metal xanthates, the formation of the semiconductor proceeds through a Chugaev reaction, the presence of ligands can alter the formation mechanism, as stated by Perner et al.106
CuInSe2 formation often requires the use of a colloidally compatible Se precursor. Some syntheses still involve the use of phosphine-related chemicals, although new strategies have been under development to facilitate the synthesis of pure-phase CuInSe2. One of the first attempts to develop a colloidally suitable precursor was the dissolution of elemental selenium in a mixture of OLA with DDT, where the latter was able to reduce the Se and induce the formation of Se–OLA complexes and didodecyl disulfide.109 No particular mechanism of formation has been proposed involving this kind of precursor, although the presence of disulfides in the reaction can still act as sulfur precursors by undergoing homolytic cleavage, as observed in syntheses involving dialkyl diselenides.108
In addition to other selenium precursors, dialkyl diselenides have also gained high interest because of their colloidal stability and their capability of unlocking crystalline phases such as wurtzite. However, this will depend on factors such as bond strength, since several diorganyl diselenides have been shown to produce either the chalcopyrite or the wurtzite phase.110 For example, Tappan et al. tested the differences between three different diorganyl diselenides and their outcomes in crystal phase formation. Among them, dimethyl diselenide and dibenzyl diselenide yielded the formation of the chalcopyrite phase, although its formation wasn't direct and involved first the formation of a Cu2−xSe phase, which then required the diffusion of In3+ ions and the out-diffusion of (1−x)Cu cations. In contrast, diphenyl diselenide yielded the formation of the wurtzite structure, which first involved the formation of umangite (Cu3Se2), followed by the diffusion of In3+ and the out-diffusion of Cu+ and Cu2+ in order to stabilize the charge of the final CuInSe2 wurtzite crystal.110 Tappan et al.'s findings indicated that the bond strength of the C–Se bond had more impact on the reactivity compared to the Se–Se bond, which showed a difference of 11.21 kcal mol−1 between the strongest and weakest diselenide bonds, diphenyl diselenide and dibenzyl diselenide, respectively. On the other hand, the C–Se bond dissociation strength had the greatest variation, being the strongest in diphenyl diselenide (64.65 kcal mol−1), followed by dimethyl diselenide (57.18 kcal mol−1), and finally (43.09 kcal mol−1). In conclusion, these findings suggest that the formation of a more thermodynamically stable phase, such as chalcopyrite, is associated with diorganyl precursors having weaker C–Se bonds.110
Hernández-Pagán et al. also tested the differences between didodecyl diselenide and dodecaneselenol in the formation of binary CuSe.108 The estimated C–Se bond energy of dodecaneselenol (67.8 kcal mol−1) and didodecyl diselenide (55.4 kcal mol−1) would suggest that the selenol reagent should induce the formation of a less thermodynamically favourable structure, while the weaker diselenide bond would favor the formation of a more thermodynamically stable structure. However, the results were contradictory, dodecaneselenol led to the formation of the cubic Cu2−xSe phase, while didodecyl diselenide yielded Cu2−xSe in the wurtzite phase.108 The proposed mechanism involving dodecaneselenol suggested the formation of lamellar structures with the copper reagent, as observed in other copper/DDT reactions.108 The formation of these structures, as well as copper withdrawing electron density from the Se atom, would then weaken the C–Se bond, facilitating the formation of the cubic phase. In contrast, although didodecyl diselenide has a weaker C–Se bond, it would still require the appropriate energy to first undergo homolytic cleavage before participating in crystal nucleation.108
Berends et al. also tested the outcome of dodecaneselenol in the synthesis of CuSe nanocrystals, yielding umangite, which was then treated through cation exchange to form CuInSe2.111 The difference in crystal phase between the two previously discussed works using dodecaneselenol can be attributed to the differences in the selected precursors and ligands involved in the synthesis: Hernández-Pagán tested the precursor along with octadecene,108 while Berends et al. included the presence of trioctylphosphine oxide,111 indicating that the chosen ligand influences the reaction mechanisms and thus the resulting crystalline phase.
A deeper look into the reactivity of dodecaneselenol with other commonly used ligands was also tested by Ho et al., where the reaction of dodecaneselenol with ligands such as octadecene, OLA, OA, stearylamine, stearic acid, and dioctyl ether yielded CuSe nanocrystals in different phases such as cubic for octadecene, OA, stearic acid, and dioctyl ether; hexagonal for stearylamine; and a mixture of hexagonal and umangite for OLA. Reaction products were also identified among the different combinations of ligands and dodecaneselenol, including the formation of species such as dialkyl selenides, dialkyl diselenides, hydrogen selenide, and selenoethers, depending on the type of ligand used.112
Although the previously reported articles focus on the formation of copper selenide binary structures, their findings can help guide the direct synthesis of ternary chalcogenides such as CuInSe2 without requiring cation exchange, when using less common selenide precursors such as selenols and diorganyl diselenides. A comprehensive list of recently reported syntheses for CuInX2 materials is presented in Table 1, specifying the type of reaction used for their synthesis, along with the reported size and crystal phase.
Material | Synthesis method | Precursors | Ligand/solvent | Max temp | Time | Crystal phase | Size | Ref. |
---|---|---|---|---|---|---|---|---|
CuInS2 | Solution-based/cation exchange | InCl3 | 1-Octadecene | 180 °C | 15 min | Wurtzite | Rods, length: 55.3 ± 3.3 nm, width: 24.6 ± 1.4 nm | 101 |
Copper(I) acetate | Tri-n-octylphosphine | |||||||
Tert-dodecanethiol | ||||||||
CuInSe2 | Solution-based/hot injection | In(OAC)3 | OLA | 200 °C | 15 min | Chalcopyrite | 15 nm | 100 |
CuI | 1-Octadecene | |||||||
Se powder | ||||||||
CuInSe2 | Solution-based/hot injection | InCl3 | OLA | 240 °C | 1 hour | Chalcopyrite | 9.6 ± 1.1 nm | 99 |
CuCl | Diphenylphosphine | |||||||
Se powder | ||||||||
CuInSe2 | Solution based | InCl3 | OLA | 240 °C | 10 min | Chalcopyrite | 6.3 ± 1.5 nm | 44 |
CuCl | ||||||||
Se powder | ||||||||
CuInS2 | Solution-based/hot injection | In(acac)3 | OLA | 310 °C | 1 hour | Wurtzite | Oval nanoparticles, length: 42.78 ± 5.55 nm, width: 27.52 ± 3.06 nm | 102 |
Cu(acac)2 | DDT | |||||||
Trioctylphosphine oxide | Tert-dodecanethiol | |||||||
CuInS2 | Solution based | In(NO3)3·3H2O | Iminodiacetic acid | 30 °C | 1 hour | Chalcopyrite | 20–100 nm | 103 |
Cu(NO3)2 | Glycine | |||||||
Na2S |
The existence of defects allows solids to have different properties. Defect chemistry enables various electronic, thermal, optical, magnetic, catalytic, and mechanical properties in materials; therefore, it is essential to understand how to control such defects.118 Li et al.119 have shown that shallow defects in CuInS2 QDs help capture charge carriers without affecting their transport, while deep defects capture charge carriers and entirely confine them, which can be consequential for chemistry related to photocatalytic hydrogen production. By growing ZnS on the surfaces of the QDs, the authors showed that the deep defects become passivated. When choosing the chemical source of a metal used in synthesizing these materials, it is essential to be meticulous. Enkhbat et al.120 have found that acetate sources resulted in high defect densities and many CuZn antisites. In contrast, using nitrate and chloride sources improved solution stability and speaks to the importance of source selection in dictating defect control.
Long-chain ligands, such as OLA, have also been found to be a source of carbonaceous impurities, which can leave a discrete residue between the absorber layer and the back contact in photovoltaic devices, thereby affecting grain growth. Ellis et al.128 have developed an exhaustive hybrid organic/inorganic ligand exchange procedure to completely strip the OLA ligand and replace it with the inorganic capping agent diammonium sulfide ((NH4)2S), improving grain growth.128 Alternatively, Hayes et al.129 have found that using N-methyl-2-pyrrolidone as an alternative to long-chain ligands such as OLA improves grain size and reduces carbonaceous residues, presumably due to the reduced carbon content.129 Table 2 presents recent works.
Material | Original ligand | New ligand | Hydrophilic/hydrophobic phase | Temperature (°C) | Time | Ref. |
---|---|---|---|---|---|---|
CuInS2 | OLA | Mercaptopropionic acid | 30% NaOH solution | Room temperature | 2 h | 52 |
CuInSe2 | ||||||
CuInS2 | DDT | Mercaptopropionic acid | Diethylene glycol/tetradecane | 160–220 °C | 1 h | 113 |
CuInS2/ZnS | DDT/OA | Mercaptopropionic acid | Ligand exchange is done directly after ZnS shell formation | 250 °C | 15 min | 89 |
Cyclohexyl 3-mercaptopropionate | ||||||
CuInSe2 | OLA/dyphenilphosphine | S2− | Formamide/hexane | Room temperature | 1 h | 99 |
CuInSe2 | OLA | 2,5-Dimercapto-1,3,4-thiadiazole | Ligand exchange is done after CuInSe film formation on the substrates | Room temperature | 20 min | 114 |
2-Amino-5-mercapto-1,3,4-thiadiazole | ||||||
CuInS2/ZnS | DDT | γ-(mercaptopropyl)trimethoxysilane | 1-Hexyl-3-methylimidazolium chloride/toluene | Room temperature | 30 min | 68 |
CuInS2/ZnS | DDT | Dodecylamine, OLA | Ligand exchange was done after ZnS formation, 11-mercapto-1-undecano was used in the ligand exchange procedure | 180 °C | 30 min | 115 |
CuInSe2 | OLA, diphenylphosphine | Trioctylphosphine, octanethiol | Ligands were exchanged with hydrophobic ligands, so the primary solvent was toluene in combination with the new ligands and iodine as an aiding agent for the exchange | Room temperature | Multiple times tested | 116 |
CuInS2 | OLA, trioctylphosphine oxide, DDT | Methylbenzenethiol (4-MBT) | Dried CuInS2 was mixed with 20 times the amount of 4-MBT | 180 °C | 72 h | 66 |
CuInS2/ZnS | DDT, OLA, OA | 6-Mercaptohexanol | Ligand exchange is done after ZnS formation | 180 °C | 15 min | 117 |
CuInSe2 | OLA | 3-Mercaptopropionic acid | Chloroform/deionized water, pH adjusted to 8–9 | Room temperature | 2 h | 25 |
Notably, some of the highest PCEs for copper chalcogenides have been achieved through costly processes like co-evaporation,144 or vapor deposition.129 Nowadays, colloidal synthesis of nanomaterials has become a popular method for fabricating CuInX2 nanomaterials, offering the versatility of incorporating these materials through various methods, such as spin coating or dip coating, which are preferred in lab-scale applications as they don't require costly equipment. However, the main drawback regarding these methods is the amount of waste material generated in producing thin films. In spin coating, part of the solution is expelled from the substrate due to the centripetal/centrifugal forces, and dip coating techniques involve preparing enough ink to fully cover the substrate.145 Although adequate, the actual spin velocity provides versatility in optimizing device performance based on layer thickness.146
In addition to deposition variables, the performance of copper chalcogenides is influenced by the synthesis process, where variables such as reagents and ligands can either enhance or limit their performance. For instance, in the colloidal synthesis of nanoparticles, ligand selection for solar energy conversion devices should be considered in the design process, as it can dramatically impact the device's performance. Particularly for solar cells, performance metrics vary depending on the functional groups and the length of the ligand.72 In QD-sensitized solar cells, it is essential to achieve full coverage of the metal oxide electrode to avoid charge recombination with the electrolyte.147 However, common synthetic routes use ligands like OAm, OA, and DDT, which are about 2.2 nm long. These ligands limit the packing density of the QDs. This makes further ligand exchange necessary to replace them with smaller ligands, which increases charge mobility,25 improves packing density and enhances the light absorption capacity of the film72 Shorter ligands like hexanethiol (HT)106 or MPA whose length is around 0.7 nm are preferred to secure a higher QDs density on the surface of selected electrodes and higher coupling to the surface.72 Inorganic ligands such as SCN−1 whose length is ∼0.22 nm are also common to further increase the packing density of QDs.147 The change in surface ligand has been shown to increase PCE from 6.2% (OAm) to 9.0% and 9.4% for MPA and HT, respectively. This is associated with higher packing and thus higher absorption capacity, elevating the Jsc to values of >26 mA cm−2.72 Dual ligand coverage has also been explored using MPA and inorganic ligands such as SCN−, Cl−, B−, I−, S2−, and S2O32−. Among these, pairing MPA with SCN− resulted in higher QD deposition on TiO2. This improved film absorption and increased the PL decay lifetime from 17.10 ns (MPA-capped) to 44.94 ns (MPA, SCN− capped). This led to the highest tested cell performance, with a certified PCE of 16.10% in a liquid-junction QD solar cell (Fig. 6a).147 Other studies have substituted OAm with shorter ligands, such as N-methyl-2-pyrrolidone, in an effort to minimize the carbon residue left after the selenization of CuInGaS2 films. This resulted in better grain formation (Fig. 6b) in conjunction with Na, as compared to OAm-capped nanoparticles, thereby increasing device performance metrics.129 Other strategies to improve packing density involve secondary depositions of the QDs, as demonstrated by Song et al.143 A pretreatment of the QDs-sensitized TiO2 with metal oxyhydroxides allows better incorporation of the second layer of QDs with no interference of the light absorption143 (Fig. 6c). Polymer heterojunctions using PCDTBT (poly[[9-(1-octylnonyl)-9H-carbazole-2,7-diyl]-2,5-thiophenediyl-2,1,3-benzothiadiazole-4,7-diyl-2,5 thiophenediyl]) in devices where the active layer is composed solely of a mixture of CuInS2 nanocrystals and the polymer, with varying ratios of polymer and CuInS2, have been demonstrated to affect the PL properties of the material. However, its PCE was relatively low, 0.23% (ref. 106) (Fig. 6d).
![]() | ||
Fig. 6 (a) J–V curves and IPCE curves of QDSCs with single (MPA) and dual (MPA/SCN) ligands, adapted with permission from ref. 152. Copyright 2022 American Chemical Society. (b) Side view SEM images of CuInS2 capped with N-methyl-2-pyrrolidone (top) and oleylamine capped (bottom), adapted from ref. 116. Copyright 2024 American Chemical Society. (c) J–V and IPCE curves of devices with single, double deposition of QDs and magnesium oxyhydroxide layer (Mg), adapted from ref. 143. Copyright 2021 American Chemical Society. (d) Effect of PCDTBT/CuInS2 ratio on PL properties, device architecture, and corresponding J–V curves, adapted from ref. 101 with permission from the Centre National de la Recherche Scientifique (CNRS) and the Royal Society of Chemistry. (e) Effect of magnesium doping concentration on the I–V curves and IPCE curves of CuInSe2 sensitized-TiO2 solar cell, adapted from ref. 144 with permission from the Royal Society of Chemistry. |
In other studies, increasing the amount of OAm in the synthetic procedure has also increased the PCE of perovskite PV prepared with CuInS2 as the hole transport layer. Higher concentrations of OAm led to bigger and less covered QDs, which are believed to have fewer defects, thus avoiding nonradiative recombination pathways.135 Although ligands like OAm and HT have shown longer electron transfer times in Zn:CuInSe2 (20 ns), care must be taken to assume both ligands would perform inferior to MPA with faster electron transfer times of 0.35–3.5 ns, as demonstrated by HT, which achieves similar internal quantum efficiency values compared to MPA-capped QDs.72
As previously mentioned, compositional variations of either Cu/In or S/Se in mixed systems significantly influence the optoelectronic properties, making the study of such variations particularly interesting for improving solar cell efficiency and understanding charge dynamics.52,153,154 Although several references suggest that copper-deficient materials achieve better PCE values, stoichiometric materials should outperform them due to their lower defect density and higher mobility.27 Furthermore, PL has recently been described as a mechanism involving Cu2+ and Cu1+, whose density depends on sample stoichiometry. Samples rich in Cu+ traps (stochiometric) benefit solar conversion devices like solar cells by extending absorption in the intragap region, improving Jsc with little effect on Voc because of fast trapping of the valence hole. On the other hand, samples rich in Cu2+ traps should outperform their counterparts in luminescent solar concentrators as the intragap absorption of Cu+ is less, thus preventing self-absorption as described by the findings of Fuhr et al.76 Suresh et al.153 accomplished this as the stoichiometric samples of CuInS2 surpassed the performance metrics of copper-deficient devices.153
Wang et al.52 investigated the impact of varying the Cu/In ratio on the formation of CuInS2 QDs passivated with ZnS. This was achieved by soaking the generated films in a Zn(Ac)2 and Na2S solution. Key performance factors, such as the open-circuit voltage (Voc), were found to be influenced by the Cu/In ratio. Copper-deficient samples exhibited higher Voc, contributing to an increased fill factor (FF) and PCE.52 Interestingly, the short-circuit current density (Jsc) did not follow a clear trend with decreasing copper content. For example, a Cu/In ratio of 1:
4 delivered a PCE of 8.54%, whereas a ratio of 1
:
5 resulted in a lower PCE of 6.95%. Further analysis, including time-resolved PL and electrochemical impedance spectroscopy, suggested that the improved performance in copper-deficient samples was due to the inhibition of charge recombination at the solar cell interface. This effect was attributed to a reduction in defect states, highlighting the importance of compositional tuning in optimizing the efficiency of CuInS2-based solar cells.52 Doping with elements like Al, Zn, Ag, or Y has also controlled defects, showing improvements in PV charge transfer.131,139,150,155
Heterogeneous atom incorporation has also become quite popular in CuInX2 devices due to its ability to modify the electronic and optical properties, thereby facilitating charge separation, collection, and injection.148 Some examples of alkaline doping include Mg doping of CuInS2 at different concentrations (0, 2, 4, 8, 10 μmol). This doping modified the absorption spectra, showing a broader range toward the visible region. The band gap was reduced from 1.81 eV (undoped CuInS2) to 1.53 eV at 8 μmol. This concentration also gave the highest increase in Voc and Jsc, leading to a PCE of 6.28% compared to 4.83% for the undoped CuInS2. The improvement was explained by a higher rate of electron injection to the TiO2.
However, higher magnesium concentrations reduced both the absorption capability of the material and the device performance, as seen in Fig. 6e.148
Alkaline metals, such as sodium and potassium, have become popular approaches to improve grain boundary quality during the formation process in search of better charge carrier mobilities, as improved performance metrics were achieved by the unexpected diffusion of Na from soda lime glass into the photovoltaic layers.144 In a study by Chugh et al.,156 simulations of several grain boundaries with structural defects such as voids and dangling bonds were carried out. These were compared to cases with incorporated alkaline metals such as Na, Li, K, and Rb. The study revealed differences in ion diffusion. Metals with a larger ionic radius, like Rb, showed low diffusion rates. In contrast, lighter atoms like Na had higher diffusion rates. However, all ions were found to concentrate in the grain boundaries. It was found that both intrinsic and extrinsic defects (induced by alkaline metals) increase device efficiency by preventing nonradiative recombination from occurring at the grain boundaries.156 Although alkali metals like Rb27 and Na have demonstrated beneficial effects in performance metrics, higher concentrations of alkali metals have also been detrimental in film formation by increasing the porosity of the layer.144 Alkali metals like sodium have also been shown to positively affect the selenization process by forming a Na–Se–O layer, which prevents the formation of surface defect sites like VSe.144
Care must be taken when selecting layers from the contact and buffer layers. Layer alignment has been demonstrated to be a key aspect in elevating PCE values. For example, replacing aluminum-doped zinc oxide (AZO) with indium tin oxide IZO as the transparent contact layer TCO front contact in CIS solar cells reduces free-carrier absorption in the NIR, improving current collection. IZO has higher electron mobility (47.3 vs. 14.7 cm2 V−1 s−1). This change boosts efficiency to 19.2% despite minor losses in the UV range due to IZO's lower bandgap.27 In another study, the morphological and optoelectronic properties of CuInSe2 films were found to be dependent on the type of TCO used (FTO, FTO/NiOx, FTO/MoO3), which affected grain size, band gap, and charge carrier mobility.157
Among the different solar cell architectures, tandem solar cells hold the most promise, as the use of different absorber layers can increase the overall PCE through the absorption of multiple regions of the solar spectrum.158 Among the different combinations, CuInS2 and CuInSe2 have been ideal candidates to support perovskite-based tandem devices as bottom absorbing layers. Examples of these architectures include the work of Jang et al., in which a top layer of perovskite and a bottom layer of CuInSe2 achieved a PCE of 10.3%, obtained through the optimization of parameters such as the thickness of the AZO layer, which played an important role in reducing interfacial resistance. However, device performance was close to standalone CuInSe2, which reached 10.64%. The authors claimed that the reduction in FF in the tandem device could be associated with interfacial resistance and parasitic absorption from the transparent window layers.159 As a strategy to reduce the parasitic absorption of films like IZO in tandem devices, Jian et al. developed a hydrogen-doped In2O3 layer through high-vacuum sputtering. The addition of this layer, instead of traditional transparent conductive oxides, helped increase the carrier mobility up to 129 cm2 V−1 s−1 and reduced the parasitic absorption above 600 nm caused by the ITO layer. When tested as a perovskite top cell arrangement, the device achieved a PCE of 17.3% while maintaining a transmittance of 82% in the range of 820–1300 nm. The standalone CuInSe2 device reached a PCE of 18.1%, and as a 4-terminal device, the overall PCE increased to 24.6%.160
Two-terminal devices have also been improved by combining perovskites and CuInSe2, as shown by Ruiz-Preciado et al. In their work, it was discussed that 2-terminal devices with subcells interconnected in series must match their current generation to avoid current limitation by the lowest-producing cell. For that, band gap engineering across the two cells can be done to match the generated photocurrent. Through computational modeling, they determined that the optimal band gap range must lie between 0.95–1.15 eV for the bottom cell and 1.54–1.72 eV for the top cell to achieve PCEs of up to 25%. The bottom cell band gap range can be achieved by using CuInSe2 with low gallium content, which has a band gap of approximately 1.03 eV, while the top perovskite layer should have a band gap around 1.59 eV to reach the desired 25% PCE. Using these parameters, the group produced a 2-terminal device that reached 24.9% PCE measured in-house and a certified efficiency of 23.5%.158
Another strategy to increase the PCE is the utilization of reflected light by surfaces interacting with the back of the device. For this, a transparent contact must be added to allow light to enter through the back side. An et al. replaced common non-transparent layers like Mo with a transparent conductive oxide such as ITO to fabricate a bifacial transparent CuInSe2 layer. The device performance was further tuned by modifying parameters like the CuInSe2 layer thickness (300–800 nm) and by using NaF for grain boundary passivation and suppression of donor defects like InCu and VSe. Final evaluation showed that the bifacial CuInSe2 devices achieved PCEs of 6.32% (300 nm) and 10.6% (800 nm). To test albedo conditions, the devices were illuminated from both sides under 2-sun illumination, leading to increased PCEs of 9.41% and 13.9%, respectively. Furthermore, a 4-terminal device was tested by adding a perovskite top cell, achieving PCEs of 18.8% (300 nm) and 21.1% (800 nm) without albedo; under 2-sun conditions, the PCEs rose to 23.4% and 24.4%, respectively.134
Other devices have included the use of conductive polymers or organic structures to improve charge carrier transport, such as liquid crystals, whose molecular arrangement can benefit charge transfer. However, it has been shown by Singh et al. that the presence of QDs like CuInS2/ZnS can disturb the structural arrangement of ferroelectric liquid crystals. Despite this, the presence of the liquid crystal in conjunction with the QDs enhances the capacitance of the devices compared to pure ferroelectric liquid crystals.161 The use of polymeric materials has also helped avoid the need for long capping ligands, as CuInS2 nanocrystals can be fabricated within their structure by thermal decomposition of molecules containing the metal and non-metal precursors. This was demonstrated by Rath et al., who elaborated devices through the decomposition of metal xanthates dissolved into the polymer, which was previously dispersed on the substrate by spin coating, followed by thermal treatment around 195 °C to form the nanocrystals. Further modification with 1,3-benzenedithiol aided surface passivation of the CuInS2 nanocrystals, improving the PCE from 1.87% to 2.56% when 1,3-benzenedithiol was used as the passivating ligand.162
The combination of CuInS2 and CuInSe2 with the heavier chalcogenide tellurium (Te) has also gained interest for solar cell applications. Although bulk CuInTe2 has a band gap of 1.02 eV it can be tuned at the quantum dot level through quantum confinement, where the band gap increases as particle size decreases.18,163 Another strategy to optimize its properties is alloying CuInTe2 with other chalcogenides such as selenium. For example, Kim et al. investigated mixed chalcogenide compositions like CuInTe2−xSex. In their work, quantum dots with a stoichiometry of Cu0.23In0.36Te0.19Se0.22 showed greater stability than the Te-only composition (CuIn1.5Te2.5). XPS analysis revealed that the Te-only QDs were more susceptible to air oxidation, forming TeO2 species, while the Se/Te alloyed QDs showed no clear signs of oxidation. When applied in solar cells, the alloyed QDs achieved a PCE (PCE) of 3.1%, with a Jsc of 17.4 mA cm−2, a Voc of 0.40 V, and a FF of 44.1% under AM 1.5G illumination at 100 mW cm−2.164 In another study by Buatong et al., several compositions of CuInTe2−xSex quantum dots were tested in dye-sensitized solar cells. The authors observed that the optical absorption properties of the QDs varied depending on the ratio between the chalcogen elements. Specifically, a higher selenium content was associated with an increase in band gap energy. Among the tested stoichiometries, the best photovoltaic performance was achieved using CuInTe1.2Se0.8, which yielded the highest PCE (3.751%) under standard 1 sun illumination.163 Even though experimental studies have addressed the impact of chalcogen mixtures with tellurium, CuInTe2 holds promise as an efficient solar absorber. Computational studies indicate that a PCE of 24.21% can be achieved with a n-CdS/p-CuInTe2 single-junction solar cell with proper optimization. Furthermore, the efficiency can be elevated up to 34.32% with the introduction of a MoS2 back surface field layer.165
Apart from surface oxidation, cation mobility is a concern in CuInS2 and CuInSe2 solar cells, as cations can segregate within the structure. In memristor devices, it has been observed that under an external electric potential, Cu atoms can oxidize to Cu2+ and migrate into cathodes such as ITO, causing copper accumulation at the interface between the two layers. This mobility increases with the strength of the electric field.167 To our knowledge, no migration of Cu has been reported toward interfaces like TiO2 in sensitized solar cells, likely because the electric fields involved are much weaker than those in memristors. However, copper diffusion has been observed in CuInS2/In2S3 solar cells when CuInS2 is deposited through spray pyrolysis, negatively affecting the device performance. This issue has been addressed by depositing a double layer of CuInS2 with different stoichiometries, copper-rich near the electrode and copper-poor at the junction region, effectively reducing Cu diffusion and improving device stability.168
Another critical issue is the durability of fluorophores; multiple studies have shown the efficiency of a variety of fluorophores albeit they often rely on short-term exposures. In a 3-year study, Terricabres-Polo et al.90 tested the performance of multiple device compositions, including CuInS2/ZnS, InP/ZnSe/ZnS, CdSe/CdS/ZnS core–shell QDs (QDs), and Lumogen, a widely studied organic luminophore (Fig. 7a).90 Among all the devices, CuInS2/ZnS outperformed the others in terms of optical properties, device performance, and stability, exhibiting the broadest absorption spectrum, an absorption of 28% of the incoming light flux, and the highest median PCE over three years (Fig. 7a).90 Over the duration of the study, devices based on CuInS2 and InP showed improved performance and durability. This was attributed to the polymer, which effectively protected both CuInS2 and InP devices, maintaining their stability for up to two years.90 However, CuInS2 demonstrated better resistance to outdoor conditions than InP. This was linked to differences in their charge recombination dynamics. CuInS2 benefits from a localized hole and a delocalized electron, making it less prone to PL quenching caused by hole traps. In contrast, InP, with localized electrons and holes, was more susceptible to such quenching.90
![]() | ||
Fig. 7 (a) Multiple LSC devices fabricated with InP/ZnSe/ZnS QDs (1), CuInS2/ZnS QDs (2), CdSe/CdS/ZnS QDs and (3), Lumogen Red F305 (4) and blank LSCs (5), with their respective median power conversion efficiency over 3 years, adapted with permission from ref. 88. Copyright 2024, Wiley-VCH. (b) CuInS2/ZnS luminescent solar concentrator fabricated with thiol-ene polymers and its I–V curves with black and white backgrounds, adapted with permission from ref. 48. Copyright 2024 The Royal Society of Chemistry. (c) LSCs based on carbon dots and CuInS2/ZnS QDs, evaluation of I–V curves as single and tandem devices, adapted with permission from ref. 31. Copyright 2024 The Royal Society of Chemistry. (d) NREL certified I–V for a CuInS2/ZnS LSCs with black and mirrored background, along with averaged quantum efficiency from four locations with black and mirrored background, adapted with permission from ref. 32. Copyright 2018 American Chemical Society. |
CuInX2 materials show a large enough Stokes shift to attract interest in this field.169 However, their typically low QY values limit direct applications. A key challenge, which has seen progress over the years, is improving the QY, as strong emission is needed for high PCE. Doping, alloying, or coreshelling with materials like ZnS is often used to enhance the PL properties of CuInX2. Although these strategies increase QY, they can cause a blue Stokes shift, increasing overlap between emission and absorption spectra and reducing device efficiency.30,36 To address this, surface engineering and coupling with a polymer matrix have been explored to enhance the Stokes shift.51 In the work of Wu et al., thiol-containing monomers were used to create off-stoichiometric thiol-ene polymers.51 Possessing a high affinity for cations like Cu or Zn, the interactions coming from the thiol groups caused an increment in the Stokes shift of 151 meV, which was attributed to the modification of Cu+ defect energy states due to the electron-donating ability of thiol groups.51 This interaction also prevented the reduction in QY, as observed in non-thiol-containing polymers like PLMA and PMMA.51 The enlargement of the Stokes shift contributed to an increase in the PCE of the device, achieving a certified efficiency of 1.36% for medium-area (9 × 9 cm2) LSCs (Fig. 7b). In comparison, large-area LSCs (29 × 29 cm2) achieved 4.25%.51
Apart from designing fluorophores with near-unity QY and sufficiently large Stokes shifts to avoid reabsorption losses, the dispersibility of QDs in the polymer matrix plays a substantial role in defining the optical properties related to light transmission.89 Multiple studies have demonstrated the role of surface ligands and structure matching with the polymer matrix in determining transmittance, waveguiding, and refractive index matching.89 Common issues related to haze, which results in non-transparent polymer nanocomposites, can be solved by a better selection of the polymer matrix associated with the structure of the ligand on the surface of the QDs. However, some strategies can involve ligand exchange to more suitable structures compatible with the most common polymers, such as polymethyl methacrylate (PMMA), polycarbonate (PC), epoxy, polystyrene (PS), polyethylene (PE), polypropylene (PP), and polyethylene terephthalate (PET). The incompatibility of surface ligands further limits the application of CuInX2 in advanced polymers, such as UV-curable celloxides. For example, CuInS2/ZnS covered by hydrophilic ligands like MPA resulted in lower transmittance and more significant aggregation than cyclohexyl 3-mercaptopropionate.89 The latter was carefully selected due to its structural similarity to celloxides, which led to higher transmittance, reduced haze, and improved QY in the polymer, as opposed to the MPA-coated QDs.89
Although haze is generally avoided in the luminophore layer of an LSC, it can be beneficial when incorporated into other layers. Advanced technologies, such as thermally responsive materials, can alter the transmittance of the LSC by incorporating optical scattering to redirect light back into the luminophore layer, thereby increasing device efficiency. Chen et al.91 tested such an architecture, utilizing a thermally responsive layer composed of polydimethylsiloxane (PDMS) and an ethylene glycol solution. This layer demonstrated the ability to change the device's transmittance across a temperature range of 20–50 °C. At ambient temperatures (25 °C), the visible light transmittance (VLT) was 83.5%, which decreased to 56.5% at 50 °C. The enhanced light scattering at higher temperatures had a positive impact on the PCE, increasing from 1.21% at 25 °C to 1.36% at 50 °C.91
Although previous research has demonstrated that CuInS2 materials exhibit a high degree of stability, as shown by Wang et al., they still suffer from PLQY reductions due to environmental factors such as UV exposure,30 where a single layer device of CuInS2/ZnS had a 21% decrease of PL intensity after 8 hours of irradiation under a 395 nm, 50 W lamp.30 As shown by Wang et al., other protective and emissive materials can be added to serve as LSC devices to address issues such as photobleaching,30 where a tandem device architecture consisting of a layer of CuIn(Se,S)2/ZnS QDs and a layer of carbon dots (CDs) (Fig. 7c) was used.30 Although the PCE in the tandem devices was not greater than that of the single-layer device, the optical stability was improved, maintaining a 93% PL intensity compared to a 21% loss by the single-layer device.30
In tandem devices, each QD layer with a different band gap is designed to absorb a specific part of the solar spectrum, allowing for its further emission and absorption by the PV device. If the QD layers share the same type of PV device, energy losses can occur as a consequence of a mismatch between the QD emission and the PV band gap. This was noted by Wu et al.169 who used computational modelling to show that PV band gap matching in a QD tandem device can lead to an improvement from 3.1 to 3.8% when an appropriate PV device is selected for each layer.169
In another study, Bergren et al.31 synthesized QDs of CuInS2/ZnS with a Stokes shift >550 meV and a QY of 91% in non-polar solvents. However, the QY decreased to 66% when the dots were encapsulated in a polymer matrix. These QDs were used to create a nanocomposite sandwiched between two pieces of low-iron-content glass, which acted as a waveguide and reduced the polymer's optical losses and optical requirements.31 The device demonstrated a visible light transmittance of 43.7% and solar absorption of 35.5%. Performance testing, as shown in Fig. 7d, revealed a PCE of 2.18% for a non-reflective substrate and 2.94% when using a reflective counterpart, indicating that the device was capable of producing between 22 and 30 W m−2 under certain specifications.31 Additionally, an averaged quantum efficiency was measured from four locations, comparing reflective and non-reflective substrates. Bergren et al.31 discussed strategies to enhance the PCE further to achieve a benchmark of 50 W m−2, including optimizing the form factor of the embedded solar cell, increasing its intrinsic efficiency, reducing haze, and improving sunlight absorption.31 Their model suggested that surpassing the energy generation benchmark of 50 W m−2 would render these devices suitable for tall buildings. For an 80-story building, such performance could yield approximately 1 GWh of electricity annually, translating to an estimated economic benefit of around $200000.31 Table 3 presents other works related to LSC and their achieved performance metrics.
Material | Architecture | Deposition | PCE | Voc | Jsc | FF | Ref. |
---|---|---|---|---|---|---|---|
CuInSe2 doped with Mg | FTO/TiO2/CISe or MnCISe/polysulfide electrolyte/Cu2S/FTO | Adsorption on TiO2 | 6.28 | 687 mV | 18.11 | 50.5 | 148 |
CuInS2 | ITO/PEDOT:PSS/PCDTBT:CuInS2/Ag | Spin coating | 0.21 | 0.716 mV | 1.09 | 0.27 | 106 |
CuInS2 | Mo-SLG/CuIn(S, Se)2/CdS/i-ZnO/ITO/Ni/Al/MgF2 | Precursor annealing on Mo-coated soda lime glass | 9.46 | 0.45 V | 35.04 | 59.93 | 149 |
Zn:CuInSe2 | FTO/TiO2/CuInSe2/ZnS/polysulfide electrolyte/brass or mesoporous carbon | Deep coating | 9.4 | 603 | 26.5 | 59.3 | 72 |
Al/Zn CuInSe2 | FTO/TiO2 + QDs/ZnS/polysulfide electrolyte/Cu2S on brass | Adsorption of TiO2 films | 10.18 | 0.622 V | 27.13 | 60.3 | 150 |
CuInS2 | Mo/CIS/CdS/i-ZnO/IZO/Ni/Al + MgF2 | Co-evaporation | 18.1 | 0.590 V | 41.8 | 73.2 | 132 |
CuInSe2 | ITO/CISe QD/perovskite/PCBM, BCP/Ag | Spin coating | 8.59 | 0.80 | 16.3 | 59.3 | 25 |
CuInS2 | Au/CIS QDs/perovskite + TiO2/FTO | Spin coating | 12.33 | 864 mV | 21.69 | 66 | 135 |
CuInSe2 | ITO/SnO2/MAPbI3 + CuInSe2 QDs/spiro-OMeTAD/Ag | Spin coating | 18.041 | 1.093 V | 21.086 | 78.35 | 133 |
CuInSe2 | ITO/SnO2/CsPbI1.2Br1.8 perovskite with CISe QDs/spiro-OMeTAD/Au | Spin coating | 10.26 | 1.28 V | 11.54 | 69.40 | 151 |
CuInSe2 | ITO/CuInSe2/CdS/i-ZnO/IZO | Co-evaporation | 10.6 | 0.454 V | 33.5 | 69.8 | 134 |
CuInSe2 | SLG/Mo/CISe/CdS/i-ZnO/n-ZnO/Al | Spin coating | 12.83 | 0.491 V | 38.59 | 67.74 | 144 |
CuInSe2 | Si3N4/Mo/CISSe/CdS/i-ZnO/Al:ZnO/Ni:Al | Spin coating | 9.6 | 464.5 mV | 33.4 | 62 | 152 |
CuInS2 | Glass/ITO/PEDOT:PSS/PCDTBT:CuInS2/Ag | Spin coating | 0.23 | 0.73 | 1.16 | 27.2 | 106 |
Fluorophore | AR (nm) | Em (nm) | SS (meV) | QY (%) | Size | PCE (%) | OE | SA | Encapsulation | Ref. |
---|---|---|---|---|---|---|---|---|---|---|
a With reflective substrate.b With no reflective substrate. Polymethyl methacrylate (PMMA), poly(lauryl methacrylate-co-ethylene glycol) (PLMA), polyvinylpyrrolidone (PVP), poly(lauryl methacrylate) (PLMA), ethylene glycol dimethacrylate (EGDM), polydimethylsiloxane (PDMS). | ||||||||||
CuInS2/ZnS | 300–600 | 550 | — | 81% | — | 8.71 | — | — | PMMA | 32 |
CuInS2/CdS | 400–700 | ∼740 | — | 75% in toluene | — | 5.7 | — | — | PLMA | 171 |
CuInS2/ZnS | 300–800 | 862 | >550 | 91% in nonpolar solvent | 10 × 10 cm | 2.94a | 8.1 | 35.5 | PLMA | 31 |
2.18b | ||||||||||
CuInS2/ZnSexS1−x | 300–600 | ∼600 | — | 13–70 depending on the shell Se:S composition in the solution | 10 × 10 cm | 0.53 | — | — | PVP | 82 |
9.3–65% in solid | ||||||||||
CuInS2/ZnS | 300–800 | 830 | — | — | 15.24 cm × 15.24 cm | 3.6a | — | 57.2a | — | 172 |
Tandem LSC (CDs + CuInSe2−xSx/ZnS | 300–800 | 878 (CuInSeS/ZnS) | 580 (CDs), 530 (CuInSeS/ZnS) | 61%) | 10 × 10 × 0.3 cm3 | 0.46 (tandem), 0.50 (single-layer CuInSeS/ZnS) | — | — | PVP and PLMA-co-EGDM | 30 |
CuIn(S, Se)2/ZnS | 300–1000 | 833–1088 | ∼519 | 45–83% | 5 × 5 × 0.9 cm3 | 1.19–1.36 | — | — | PDMS | 91 |
CuInS2/ZnS | 400–700 | ∼500 nm in OSTE | 585 in OSTE | 85 in OSTE | — | 1.36 | — | — | Off stoichiometric thiolene | 51 |
Recent studies suggest that the photoluminescence mechanism involves a free-to-bound donor–acceptor pair, often linked to Cu+ or Cu2+ states, whose distribution depends on the sample's stoichiometry. Fully understanding the location and nature of these traps is essential for controlling PL behavior, improving QY without the need for surface passivation, and suppressing non-radiative recombination – two factors that are known to be critical for boosting device PCE. This knowledge can also guide stoichiometry control to selectively favor defects that are either beneficial for solar cells or LSCs.
As discussed in the synthesis section, sulfur-based nanomaterials benefit from readily available and easy-to-handle precursors, such as DDT. In contrast, selenium-based compositions still rely on toxic precursors or complex ligand combinations, including DDT to produce colloidally compatible selenium sources. This often prevents the formation of pure-phase CuInSe2. Designing selenium precursors that are both safe and compatible with colloidal synthesis is crucial for enabling the fabrication of phase-pure CuInSe2 nanomaterials. This is especially important for applications that require the specific optical properties of the pure selenide phase. This could also benefit the stoichiometric and size control of the selenide compositions, as it wouldn't involve the formation of sulfur-related phases.
Regarding solar cells, interfacial layer treatments have recently proven effective in increasing device efficiency. Therefore, more attention should be given to passivating these interfaces to reduce unwanted recombination and prevent energy losses.
In the case of LSCs, copper chalcogenides have proven to be efficient due to their large Stokes shifts and potential for high QY, often improved through passivation strategies, such as ZnS core–shell structures. However, efforts should be made to avoid the blueshift in emission that occurs as a result of these treatments. Minimizing this shift is crucial to reducing reabsorption losses and enhancing overall device performance.
Progress in the aspects mentioned above will help to establish ternary copper chalcogenides as key materials for the next generation of efficient, low-cost, and sustainable solar energy conversion devices.
This journal is © The Royal Society of Chemistry 2025 |