Hongyang Zhao
*a,
Zhenwei Tang
a,
Shuya Cui
a,
Lirong Yang
a,
Xinjie Xiang
a,
Jianni Bai
b,
Jingying Luo
b,
Zhuojian Li
b,
Xin Li
c,
Guoqi Xiang
c,
Wuyang Ren
*b and
Xin Tong
*bcd
aSchool of Chemistry and Materials Engineering, Mianyang Normal University, Mianyang 621000, P. R. China. E-mail: hongyang.z@mtc.edu.cn
bInstitute of Fundamental and Frontier Sciences, University of Electronic Science and Technology of China, Chengdu 611731, P. R. China. E-mail: wren@uestc.edu.cn; xin.tong@uestc.edu.cn
cSolar Energy Integration Technology Popularization and Application Key Laboratory of Sichuan Province, Panzhihua University, Panzhihua 617000, P. R. China
dYunnan Key Laboratory of Electromagnetic Materials and Devices, Yunnan University, Kunming 650091, P. R. China
First published on 9th December 2025
Colloidal indium phosphide (InP) quantum dots (QDs) have emerged as a compelling class of heavy metal-free nanomaterials due to their low toxicity and size-tunable optoelectronic properties, showing great potential in solar-driven energy conversion applications. Here, a variety of synthetic techniques for preparing high-quality InP QDs, including hot-injection, heat-up, cluster-mediated growth, and cation exchange, are thoroughly reviewed. To realize enhanced photocatalytic (PC) and photoelectrochemical (PEC) performance, diverse strategies such as core/shell engineering, hybrid ligand modification and elemental doping of InP QDs are discussed in detail, which are beneficial to build various efficient QDs-based systems for hydrogen evolution, CO2 reduction, ammonia synthesis, and H2O2 production. Moreover, the main challenges and future research directions of InP QDs are briefly proposed, providing guidelines to achieve future low-cost, eco-friendly, scalable and high-efficiency QDs-based solar energy conversion technologies.
Among them, indium phosphide (InP) QDs have emerged as a leading candidate for solar energy conversion owing to their low toxicity, favorable bulk bandgap (∼1.35 eV), large exciton Bohr radius (∼10 nm), and tunable absorption across the visible to near-infrared range.10–13 As shown in Fig. 1, numerous scientific studies associated with InP QDs have been increasingly published since 1994,14 reflecting their growing significance in this field. Their covalent crystal structure also contributes to enhanced chemical stability compared to ionic II–VI QDs.15,16 However, early synthesis of high-quality InP QDs was hindered by the high In–P bond dissociation energy and challenges in controlling nucleation and surface defects.17–19 Recent progress in synthetic methodologies has alleviated these issues. Specifically, The hot-injection technique enabled rapid separation of nucleation and growth phases to improve the size uniformity;20–24 cation exchange provides a pathway to form In–P bonds under milder conditions to reduce defect formation;25–28 and continuous-flow synthesis ensures highly reproducible reaction environments, facilitating scalable production of high-quality InP QDs with enhanced optical properties.29–31 Furthermore, sophisticated core/shell designs such as InP/ZnSe/ZnS have achieved photoluminescence quantum yields (PLQY, the ratio of the emitted and absorbed photons) exceeding 90%, rivaling those of Cd-based QDs while maintaining RoHS compliance.18
![]() | ||
| Fig. 1 Timeline and number of publications on the synthesis and solar energy conversion applications of InP QDs. | ||
Advances in the material perspective of InP QDs have facilitated their applications in a variety of solar-driven processes, including PC hydrogen evolution, CO2 reduction, ammonia synthesis and H2O2 generation. Through tailored surface chemistry and hybrid structures, InP QDs have demonstrated remarkable stability and activity in solar-driven energy conversion systems, such as continuous CO2-to-CO conversion over 420 hours.32 Nevertheless, issues including charge recombination, limited large-scale production, and long-term operational stability remain obstacles to practical deployment.
Although several reviews have summarized the synthesis and optical properties of InP QDs, a comprehensive overview focusing on their solar energy conversion applications, spanning materials design, mechanistic insights, and device integration, is still lacking.33–35 This review summarizes recent developments in the synthesis, property modulation and PC/PEC applications of InP QDs, offering a forward-looking perspective on overcoming key barriers to commercialization.15,36–41 By linking materials innovation with energy-oriented functionality, this review aims to guide the development of efficient, stable, and scalable heavy metal-free QDs technologies for sustainable solar energy conversion.
As illustrated in Fig. 2(a), the hot-injection synthesis of InP QDs typically proceeds through three stages: (i) an indium precursor solution was first introduced into a hot coordinating solvent to precondition the system; (ii) the subsequent rapid injection of the phosphorus precursor induced a sharp increase in monomer concentration to trigger homogeneous nucleation; (iii) a gradual decrease in monomer concentration facilitated the diffusion-controlled Ostwald ripening, thus yielding monodisperse InP QDs.35,56 This schematic highlights how subtle variations in injection timing and sequencing profoundly influence the nucleation kinetics and final particle characteristics of QDs.
![]() | ||
| Fig. 2 Synthesis and optical properties of InP QDs prepared by hot-injection and heat-up methods. (a) Comparison of the hot-injection and heat-up synthesis processes for colloidal QDs. (b) Optical images of InP QDs prepared via different [In] precursor (InCl, InBr, and InI) and corresponding evolution of PL and ultraviolet-visible (UV-vis) absorption spectra. Reproduced with permission.20 Copyright 2023, American Chemical Society. Absorption (solid) and PL (dashed) spectra of (c) heat-up synthesized and (d) hot-injection synthesized InP QDs before (black) and after (green) InF3 treatment. (e) PLQY and (f) FWHM values of four types of InP QDs synthesized using different approaches. Reproduced with permission.70 Copyright 2024, American Chemical Society. | ||
The development of more sophisticated precursor design and reaction engineering has yielded hot-injection-synthesized InP QDs with substantially improved quality. While tris(trimethylsilyl)phosphine (P(TMS)3) remains widely used, less hazardous alternatives such as tris(diethylamino)phosphine (P(NEt2)3) were increasingly employed to improve the size distribution and batch reproducibility.20,23,35,57 Ligands, including fatty acids and long-chain amines, serve not only as colloidal stabilizers but also as reaction rate moderators by influencing precursor decomposition kinetics.58 The introduction of zinc halides as co-precursors has proven effectiveness in enhancing the crystallinity and PLQY by reducing the surface trap states of InP QDs.59,60 By effectively optimizing the reaction parameters, such as injection temperatures between 230–260 °C and growth durations within two minutes, spectral tunability across the visible range and PLQY exceeding 80% could be realized.56,61,62
Fig. 2(b) illustrated an innovative hot-injection synthesis route for InP QDs utilizing monovalent indium halides (InX, X = Cl, Br, I).20 This approach employed In(I) species as dual-functional precursors, which serve simultaneously as indium sources and reducing agents for aminophosphines. It represented a notable advance over conventional In(III)-based methods because it could eliminate the requirement of zinc additives during synthesis, a common source of structural defects and emission line broadening.12 Such synthesis technique produced well-defined tetrahedral InP QDs with edge lengths exceeding 10 nm and narrow size distribution, reflecting improved control over nanocrystal morphology and uniformity. Generally, halide identity could critically modulate the optical characteristics of InP QDs, enabling the tuning of the first excitonic absorption peaks across the visible spectrum (from 450 nm to 700 nm), with correlated shifts in the PL emission peaks. The resulting InP QDs displayed exceptionally narrow PL linewidths, as low as 112 meV at 728 nm, which indicated their highly effective surface passivation and strongly suppressed defect-associated recombination.20 The different spectral evolution behaviors observed in QD synthesis using diverse halide systems further implied the unique reaction kinetics and surface stabilization pathways, providing essential insights for precursor choice in the hot-injection synthesis of high-quality III–V QDs.63,64
Recent studies by Reiss and co-workers have shown that moderate nucleation kinetics under heat-up conditions yielded improved size uniformity of InP QDs, as evidenced by a sharper excitonic absorption feature compared to the hot-injection synthesized InP QDs (Fig. 2(c) and (d), black curves).70 Their work identified heating rate as a vital factor to influence the nanocrystal quality, with optimal values between 5–10 °C min−1 markedly reducing the overlap between nucleation and growth. This controlled synthetic method contributed to a higher PLQY, reaching up to 93% after InF3 treatment in heat-up synthesized QDs (Fig. 2(c), green curve), surpassing the maximum PLQY of 71% obtained via hot-injection under comparable conditions (Fig. 2(d), green curve).70
Building upon the superior size uniformity achieved through moderate nucleation kinetics, the heat-up method further demonstrates its advantage in radiative efficiency. As corroborated by the statistical data in Fig. 2(e), InP QDs prepared via the heat-up method achieve far higher PLQY than those from the hot-injection route across a range of post-synthetic treatments. The insight from this comparison is that the superior PLQY of the heat-up QDs does not result from improved size distribution, as evidenced by their comparable full width at half maximum (FWHM, a measure of the width of a peak or distribution at half of its maximum amplitude) values to the hot-injection samples (Fig. 2(f)). Instead, the higher radiative efficiency is originated from an intrinsic superiority in nanocrystal quality that fewer structural defects and non-radiative trap states afforded by the more controlled, moderate nucleation kinetics of the heat-up process. This finding solidified the argument that the optimized heating profile was a pivotal synthetic parameter, which concurrently enhanced the size uniformity and suppressed the defect formation to yield highly luminescent InP QDs.
![]() | ||
| Fig. 3 Cluster-mediated synthesis and transformation mechanism of InP-based QDs: (a) nucleation and growth mechanism illustrating MSCs destabilization during In(Zn)P QD synthesis. Adapted with permission.76 Copyright 2023 American Chemical Society. (b), (c) Schematic of MSC-derived InP QD synthesis pathway along with corresponding TEM images and size distribution histogram. Adapted with permission.77 Copyright 2022 Elsevier. (d) The proposed pathways for InP QDs formation from different MSCs, (e) time-dependent UV-vis absorption spectra showing the conversion of In37P20 clusters into InP QDs after the introduction of P(SiMe3)3, (f) kinetic traces of absorbance at 500 nm following the addition of 10 equivalents of P(SiMe3)3 to each cluster solution. (d)–(f) adapted with permission.78 Copyright 2024 American Chemical Society. Synthesis and structural evolution of InP QDs via cation exchange: (g) schematic for preparing water-soluble w-InP QDs from Cu3−xP templates by cation exchange, (h) TEM images of w-InP QDs with different sizes (scale bar: 20 nm). (g), (h) adapted with permission.26 Copyright 2024 American Chemical Society. (i) Reaction pathway from Cu0 to Cu3−xP, followed by the In/Zn-halide cation exchange and NOBF4 treatment. (j) high-angle annular dark-field and scanning-transmission electron microscopy (HAADF-STEM) images and energy-dispersive X-ray spectroscopy (EDS) mapping showing the phase evolution from Cu0 to w-InP QDs (scale bar: 5 nm). (i), (j) adapted with permission.27 Copyright 2023 American Chemical Society. | ||
Structural studies of MSC intermediates have provided critical mechanistic insights into the formation processes of InP QDs. The In37P20(O2CR)51 cluster has been identified as a key isolable intermediate whose reactivity dictates the subsequent growth kinetics (Fig. 3(d)).78 Ligand engineering, particularly through substituted phenylacetate derivatives, has demonstrated how steric effects precisely modulated the cluster stability. For instance, meta-substituents could enhance the phosphine accessibility to cluster core, while para-substituents introduced kinetic barriers that inhibited the core dissolution.79 This ligand-mediated control over cluster stability not only regulated the nucleation timeline but also enabled the isolation and crystallographic determination of an earlier In26P13(O2CR)39 intermediate, offering atomic-level insight into III–V cluster evolution. These findings established a clear structure and reactivity relationship that directly links ligand engineering to cluster-based synthesis pathways, thereby providing a rational framework for controlling the formation of high-quality InP QDs through intermediate stabilization and reaction kinetics modulation.
Kinetic analyses of the cluster-to-QDs transformation can provide a quantitative perspective on the growth mechanism of InP QDs. As tracked via the absorbance changes at 500 nm (Fig. 3(e) and (f)), the real-time monitoring of the reaction between In37P20 MSCs and P(SiMe3)3 in toluene reveals distinct conversion stages,78 showing the rapid emergence of excitonic features following cluster dissolution that confirmed the MSC-derived nucleation. In parallel, the temperature-dependent studies highlighted the influence of thermal barriers on reaction kinetics. Specifically, the elevated temperatures lowered the activation energy for cluster decomposition, accelerating the release of monomeric species as seeds for QDs growth. This thermal modulation directly governed the rate of cluster consumption and the subsequent nucleation burst, thereby enabling precise control over the size distribution and final quality of InP QDs. These kinetic insights established a direct link between cluster-based precursor design and the rational synthesis of monodisperse QDs.
Banin and coworkers illustrated the complete transformation from Cu3−xP nanocrystals to luminescent w-InP QDs,26 wherein the reaction began with Cu-to-In cation exchange to yield initially non-emissive w-InP QDs with broad absorption profiles (Fig. 3(g)). Subsequent treatment with nitrosyl tetrafluoroborate (NOBF4) was essential to activate the PL by removing residual copper species and tuning the surface stoichiometry. The transmission electron microscopy (TEM) images of QDs in Fig. 3(h) demonstrated the size control through precursor selection and etching duration, in which the copper tartrate templates produced larger QDs (5.9–6.9 nm), while those derived from copper acetate yielded smaller sized QDs (4.5–5.0 nm), illustrating the role of different Cu-based templates in achieving controllable sizes of InP QDs.
Further mechanistic insight into the multistep synthesis was provided by the same group through systematic analysis.27 Fig. 3(i) illustrated the initial formation of Cu0 nanocrystals with phosphorus precursors, with the ensuing reaction pathway dictated by the copper precursor's thermal stability. Specifically, stable copper halides directly reacted with phosphorus sources to form Cu3−xP above 260 °C, whereas less stable copper oxyanions first generated intermediate Cu0 nanocrystals, which were subsequently converted into Cu3−xP at elevated temperatures. This precursor-dependent divergence underscored the important role of precursor decomposition kinetics in directing the phase evolution pathway. Additionally, Fig. 3(j) tracked the transformation processes from Cu0 nanocrystals with surface phosphorus through mixed-phase intermediates to the final Cu3−xP products. When integrated with subsequent Cu-to-In cation exchange and NOBF4 treatment, this well-defined reaction process has provided a reproducible and controllable route to synthesize high-quality InP QDs with tunable optoelectronic properties.
A notable study by Sang Man Koo et al. demonstrated the benefits of incorporating a ZnSeS gradient interlayer in InP-based core/shell QD systems.91 Fig. 4(a) schematically depicted the graded shell architecture of InP/ZnSeS/ZnS QDs, which is designed to mitigate the lattice mismatch between the core and the outer shell. Corresponding spectroscopic data in Fig. 4(b) confirmed a high monodispersity of the initial InP cores, as evidenced by a sharp excitonic absorption peak. Following the shell growth, the absorption and emission spectra of the InP/ZnSeS/ZnS QDs are redshifted. More importantly, the marked narrowing of the emission peak and the enhancement in PL intensity provided direct evidences for effective surface passivation and a reduction in defect states. A narrow emission profile with a FWHM of 48 nm further reflected the gradient shell's effectiveness in suppressing inhomogeneous broadening of QDs.90
![]() | ||
| Fig. 4 Structural and optoelectronic properties of gradient InP-based core/shell QDs. (a) Schematic of the graded shell architecture in InP/ZnSeS/ZnS QDs and (b) corresponding absorption and UV-vis absorption spectra highlighting the narrow emission. (a), (b) Reproduced with permission.91 Copyright 2025 Wiley-VCH. (c) Synthesis pathway for WZ-phase ZnSe shells on InP core QDs. (d) Calculated exciton energy shift as a function of QD volume for InP QDs with ZB (gray) and WZ (red) ZnSe shell. (e) Band alignment diagrams comparing the ZB and WZ InP/ZnSe QDs and (f) UV-vis absorption and PL spectra during shell growth. (c)–(f) Reproduced with permission.97 Copyright 2025 American Chemical Society. | ||
In another work, Li et al. investigated the influence of shell crystal phase on the optoelectronic properties of InP core–shell QDs.97 Through a phase-controlled synthesis strategy detailed in Fig. 4(c), this group achieved the growth of a wurtzite (WZ) ZnSe shell on InP core QDs. It is indicated that this shell with WZ phase confers a stronger quantum confinement compared to its zinc blende (ZB) counterpart. The proposed mechanism attributes this effect to the distinct electronic structure inherent to the WZ lattice, which likely modifies the confinement potential landscape by creating a deeper potential well for charge carriers at the core–shell heterojunction. As shown in Fig. 4(d), the calculated exciton energy results revealed a redshift in exciton energy for InP/ZnSe-WZ core–shell QDs compared to their ZB counterparts. This was primarily attributed to the larger conduction band offset in the WZ structured core–shell QDs, which led to a higher degree of carrier localization and a more pronounced redshift in emission energy. The energy level alignment diagrams in Fig. 4(e) further illustrated a smaller conduction band offset in the ZB structured QDs, which facilitated the delocalization of the electron wave function into the shell. UV-vis absorption and PL spectra of these InP/ZnSe-WZ core/shell QDs revealed spectral shifts at different stages of the shell growth process, as shown in Fig. 4(f). As the shell thickness increased, both absorption and PL spectra exhibited a progressive redshift and ultimately achieved a high PLQY of 92%. This observation was consistent with the enhanced carrier localization effect due to the coating of a ZnSe shell with WZ phase, further confirming the successful growth of a high-quality crystalline shell under the optimized low-temperature conditions implemented to preserve the desired crystal structure.
Based on the fundamental advances of gradient core/shell architectures, the next critical step is to further optimize the optical and structural properties of InP QDs through precise control of shell composition.24,98 The introduction of alloyed shells (e.g., ZnSexS1−x) is able to enhance the confinement of photogenerated charge carriers, reduce the lattice strain, and suppress the defect-mediated non-radiative recombination.88,89 Compositional grading within the shell further promotes the exciton localization and enables spectral tunability across the visible range, underscoring the potential of InP QDs for high-performance optoelectronic devices.99
For example, Zhao et al. investigated the influence of alloyed inner shells on the optical properties of InP QDs.93 Fig. 5(a) illustrated the ZnSexS1−x interfacial layer in a InP/ZnS core–shell QD system, which is a key structural design employed to alleviate the lattice mismatch between the InP core and the outer ZnS shell for reduced interfacial defects. As exhibited in Fig. 5(b), the X-ray diffraction (XRD) was used to investigate the formation of phase-pure InP/ZnSexS1−x QDs across different compositions formed by tuning the Se/S precursor ratio during the inner shell growth. A key observation was that the diffraction peaks gradually shifted with increased Se/S ratio. For the InP/ZnSe0.7S0.3/ZnS structure, the peaks reside between those of the core and the ultimate ZnS shell, demonstrating the role of this intermediate layer in mitigating lattice mismatch and promoting coherent shell growth. Increasing the Se contents induced a redshift in both the first excitonic absorption peak (from 475 nm to 505 nm) and the PL peak (from 508 nm to 535 nm), as shown in Fig. 5(c) and (d). This red-shift is derived from a reduced conduction band offset at the InP/ZnSexS1−x interface, thus narrowing the effective band gap and enhancing the electron delocalization. A maximum PLQY of 97% and a FWHM of 35 nm were achieved at x = 0.7 (Fig. 5(d) and (e)), with the enhanced brightness visually confirmed by the photographs in Fig. 5(f).
![]() | ||
| Fig. 5 Composition-dependent optical and structural properties of InP-based core–shell QDs with engineered inner shells. (a) Schematic of the InP/ZnSexS1−x/ZnS structured QDs. (b) XRD patterns of composition-dependent QDs showing lattice parameter evolution. (c), (d) Absorption and PL spectra of InP/ZnSexS1−x/ZnS QDs with varying Se/S ratios, (e) Correlation among FWHM, PL peak position and PLQY and (f) digital photographs of QD solutions under UV illumination. (a)–(f) Reproduced with permission.93 Copyright 2022 Springer Nature. (g) Shell growth pathways using single S and dual Se/S precursors for InP/ZnSe/ZnSeS QDs. (h) Size distribution histograms and (i) normalized PL and absorption spectra of the resulting QDs. (g)–(i) Reproduced with permission.39 Copyright 2025 Wiley-VCH. (j) Linear correlation between the injected and incorporated sulfur in InP/ZnSe/ZnS QDs and (k) their tunable PL emission covering the visible spectrum. (j), (k) Reproduced with permission.100 Copyright 2022 American Chemical Society. | ||
Wu et al. compared the single sulfur precursor and dual selenium/sulfur precursor shell growth routes for InP/ZnSe/ZnSthick and InP/ZnSe/ZnSeS/ZnSthick QDs.39 Fig. 5(g) illustrated the two synthesis pathways, showing that the extended ZnS shell growth for 12 hours could yield QDs exceeding 20 nm in diameter. Incorporating the selenium near the core region reduced the lattice strain, enabling a thicker shell deposition even at identical Se/S ratios (25/75) and growth durations (Fig. 5(h)). Although both methods produced QDs emitting at 630 nm (Fig. 5(i)), the dual-precursor approach enhanced the optical absorption at the higher energy range, indicating the improved shell quality and a lower defect density. These findings demonstrated that the precursor selection and spatial distribution strongly influenced the shell growth kinetics and QDs morphology. Similarly, Zeger Hens and colleagues demonstrated that the tailored Zn(Se, S) inner shells on InP core QDs enabled broad spectral coverage while reducing inhomogeneous broadening induced by lattice mismatch.100 Fig. 5(j) showed the elemental analysis from EDS, which revealed that the resulting QDs contained sulfur in proportion to the amount of added tri-n-octylphosphine sulfur (TOP-S). As a result, a tunable PL emission of these QDs across the visible spectrum was realized, as shown in Fig. 5(k).
Tang et al. demonstrated the effectiveness of cysteamine (CTA) as a short-chain ligand on InP/ZnSe/ZnS QDs.106 As shown in Fig. 6(a), CTA facilitated a strong surface coordination to cause a substantial increase in PL intensity (Fig. 6(b)) and a longer radiative recombination lifetime (Fig. 6(c)). Furthermore, introducing a small amount of CTA into the QD solution increased the proportion of radiative recombination (Fig. 6(d)). The observed improvements in PL intensity and radiative recombination efficiency confirmed that CTA effectively passivated the surface defects of QDs and enhanced the radiative recombination rate of intrinsic excitons. In a parallel strategy, Himchan Cho et al. employed molecular metal chalcogenide complexes (MCCs) as multifunctional ligands that improved both the exciton dissociation and the colloidal stability in aqueous environments (Fig. 6(e)).107 Retention of the optical properties after ligand exchange (Fig. 6(f)) and a pronounced negative shift in zeta potential (Fig. 6(g)) confirmed the effective surface modification and increased dispersibility of as-prepared MCC-functionalized InP/ZnSe/ZnS QDs (MCC-InP QDs).
![]() | ||
| Fig. 6 Ligand engineering of InP-based QDs. (a) Schematic of InP/ZnSe/ZnS QDs with CTA ligand and relevant defect passivation mechanism. (b) Steady-state PL spectra and (c) time-resolved PL decay curves of QD and QD-CTA films. (d) Time-resolved PL decay curves of various QD solutions. (a)–(d) Reproduced with permission.106 Copyright 2024 American Chemical Society. (e) Schematic of MCC-InP QDs. (f) UV-vis absorption and PL spectra of InP/ZnSe/ZnS QDs. (g) Zeta potential of pristine and MCC-InP QDs. (e)–(g) Reproduced with permission.107 Copyright 2025 Wiley-VCH. | ||
Hybrid ligand engineering has also emerged as a promising strategy for enhancing the colloidal stability and optoelectronic properties of InP QDs.108 By combining the ligands with complementary functionalities, this approach could effectively passivate the surface defects and suppress the fluorescence blinking.24,109,110 For instance, the photobleaching and emission intermittency of InP QDs are still primary challenges for their applications in bioimaging and highly sensitive single-molecule detection. Regarding this issue, Li et al. compared the optical properties of three types of water-soluble InP/ZnSe/ZnS QDs at the single-particle level, which were modified with either mercaptopropionic acid (MPA), mercaptoundecanoic acid (MUA), or a mixture of both ligands (Fig. 7(a)).111 Fig. 7(b)–(d) showed the typical PL intensity trajectories of the three types of QDs. QDs with only MPA or MUA ligands showing frequent intensity fluctuations, indicating a pronounced emission intermittency. In contrast, QD-MPA-MUA with mixed ligands exhibited suppressed blinking and maintained a stable bright on state (Fig. 7(d) and (g)). The on-state probability refers to the percentage of time a QD remains bright under continuous excitation. Statistical analysis of 100 individual QDs revealed that the mixed ligand system achieved an on-state fraction of 95.2%, which was notably higher than the 63.7% for QD-MPA and 70.1% for QD-MUA (Fig. 7(e) and (f)). The combination of short-chain MPA and long-chain MUA ligands enabled a stable surface coverage, effectively passivating the defect sites responsible for Auger recombination and blinking. Such synergistic passivation not only enhanced the photostability but also improved the charge balance, which led to the observed high on-state probability.
![]() | ||
| Fig. 7 Hybrid ligand passivation strategies for InP QDs. (a) Schematic of QDs functionalized with MPA, MUA, and MPA–MUA hybrid ligands (left) and their aqueous dispersion stability (right). (b)–(d) Single-QD PL trajectories and (e)–(g) corresponding on-state probability statistics demonstrating blinking suppression in MPA-MUA QD systems. (a)–(g) Reproduced with permission.111 Copyright 2025 American Chemical Society. (h) A proposed mechanism of dilution-induced ligand desorption and (i)–(k) concentration-dependent PL properties highlighting the ligand stabilization. (h)–(k) Reproduced with permission.112 Copyright 2025 American Chemical Society. (l) Synthesis route and (m) PL enhancement of HF-pyridine treated QDs. (l)–(m) Reproduced with permission.113 Copyright 2025 American Chemical Society. (n) Schematic of CLIP patterning using photocleavable BNA ligands and (o) retained optical properties after ligand exchange. (n)–(o) Reproduced with permission.115 Copyright 2025 American Chemical Society. | ||
To unravel the stability mechanisms of hybrid ligand in InP QD systems, Hou's group explored ligand adsorption–desorption thermodynamics through controlled dilution experiments (Fig. 7(h)–(k)).112 Fig. 7(h) illustrated that InP QDs diluted in pure water underwent substantial ligand desorption, whereas the pre-addition of extra ligands preserved the surface coverage. In Fig. 7(i), 100-fold dilution introduced a fast decay component (∼5 ns) and reduced the average PL lifetime from 28 ns to 16 ns, indicating the enhanced non-radiative recombination. Fig. 7(j)–(k) showed that pre-adding MPA (0–20 mM) suppressed the fast decay component and restored the PLQY from 42% to 92% under dilution. These findings verified the importance of ligand pre-equilibrium for maintaining optical performance under dilute conditions.
Beyond the scope of conventional ligand strategies (Fig. 7(h)–(k)), Yang's group reported a dual strategy of HF etching and pyridine ligand exchange for synergistic passivation of InP-based core–shell QDs (Fig. 7(l)),113 wherein the HF purified the surface and the pyridine ensured strong capping to synergistically produce a dramatic PL enhancement and a pronounced blueshift (Fig. 7(m)). It is demonstrated that the combined chemical and ligand modifications could rival advanced ligand engineering in boosting the optical properties of InP-based QDs.
The cleavable ligand-induced photolithography (CLIP) strategy is fundamentally connected to the performance of optoelectronic devices including those used for solar energy conversion.114 It addresses a central challenge in fabricating high-performing devices by enabling precision processing while preserving essential optoelectronic properties. Unlike conventional hybrid ligand passivation methods that focus mainly on stability and charge transport, the CLIP approach introduced by Himchan Cho and colleagues represented a pioneering expansion of functionality.115 Its core innovation involved the dynamic transformation and controlled removal of photosensitive ligands to achieve high-resolution patterning, constituting an advanced form of ligand interface engineering. As illustrated in Fig. 7(n), ultraviolet light cleaved the ligands bonded to the InP QD surface, modifying their wettability and allowing patterning through differential solubility. A key achievement of this method was its ability to combine fine processing of InP QDs with effective passivation of their optical characteristics. The patterning process preserved 70% of the initial PLQY and maintained optical stability, as shown in Fig. 7(o). Throughout this process, the absorption and PL spectral features of the QDs remained consistent, confirming that the dynamic ligand design sustained surface passivation quality and safeguards fundamental optoelectronic properties. This approach thereby created new possibilities for scalable fabrication of high-performance InP QDs-based optoelectronic devices requiring precise patterning, including components for solar optoelectronic devices.
For instance, Jiang's group reported the Ga-doped InP QDs synthesized via Ga3+-for-In3+ cation exchange, which exhibited significantly enhanced PL intensity and a gradual decrease in UV-vis absorption with doping time (Fig. 8(a) and (b)).121 At 320 °C, the PL peak initially redshifted (609–615 nm, 10 min) due to Ostwald ripening, which was then blue-shifted to 589 nm over 50 min with further cation exchange and core etching (Fig. 8(c)). Concurrently, the PL broadened (FWHM >70 nm after 30 min), yet the PLQY increased and stabilized near 25%, demonstrating the effective optical tuning of InP QDs via Ga3+ doping. In another study, Tian et al. also employed cationic doping to modulate the optoelectronic properties of InP-based QDs,36 wherein the gradient In3+-doped InP/ZnSe QDs were constructed and the tuning of In3+ content established an internal electric field (Fig. 8(d)), which induced a transition from type-I to quasi-type-II band alignment for facilitated separation and transfer of photogenerated carriers. Consequently, as evidenced by the UV-vis absorption and PL spectra (Fig. 8(e) and (f)), both the absorption excitonic peaks and the PL emission peaks exhibited a gradual redshift as a result of the In3+ doping that effectively narrowed the overall band gap of the QDs, representing a promising pathway for tailoring their optical behavior through cationic incorporation. Liu et al. also investigated the modulation of band alignments in InP QDs through Cu and Mn doping (Fig. 8(g)).123 It is found that both dopants could reduce the optical band gap of InP QDs, in which the Cu doping led to dual emission originated from both dopant states and band-edge recombination, while the Mn doping facilitated the charge transport by introducing shallow trap states, manifesting the importance of dopant selection in tailoring the charge dynamics of InP-based QDs.
![]() | ||
| Fig. 8 Cationic doping strategy in InP-based QDs. (a) The schematic illustration for the synthesis of Ga-doped InP cores, (b) UV-vis absorption and (c) the normalized PL spectra of aliquots taken during the Ga3+ doping process for InP QDs. (a)–(c) Reproduced with permission.121 Copyright 2024 Springer Nature. (d) Band structure modulation through gradient In3+ doping in ZnSe shells, (e) UV-vis absorption and (f) PL emission spectra with increasing doping concentration. (d)–(f) Reproduced with permission.36 Copyright 2024 American Chemical Society. (g) Band alignments of Cu- and Mn-doped InP QDs derived from experiment analysis. (g) Reproduced with permission.123 Copyright 2024 Elsevier. | ||
![]() | ||
| Fig. 9 Surface treatment strategy for InP QDs. (a) Schematic of UV degradation mechanism involving ZnO formation and In diffusion and (b) comparison of PL lifetime evolution in air versus argon atmosphere for InP/ZnSe/ZnS QDs. (a), (b) Reproduced with permission.88 Copyright 2024 Springer Nature. (c) ZnF2-mediated synthesis of InP QDs with high PLQY and narrowed emission and (d) corresponding monoexponential PL decay profile. (c), (d) Reproduced with permission.130 Copyright 2022 American Chemical Society. (e) Various metal fluoride treatments highlighting the optimized InF3 treatment for enhanced surface passivation, (f) HF-free InF3 treatment of InP QDs to achieve a high PLQY over 90% and (g) associated PL lifetime extension. (e)–(g) Reproduced with permission.70 Copyright 2024 American Chemical Society. (h) In situ HF generation approach for surface reconstruction, (i) temperature-dependent PLQY improvement and (j) PL lifetime evolution with ZnCl2 co-treatment for InP QDs. (h)–(j) Reproduced with permission.131 Copyright 2022 American Chemical Society. | ||
The strategic application of fluoride-based treatments alongside sophisticated shell engineering has also emerged as an effective approach for mitigating the nonradiative recombination, which in turn improves both the optical properties and environmental stability.23,128,129 Li et al. found that the generated HF through the reaction of ZnF2 and oleic acid under heating could be used to alleviate the surface defects of InP QDs (Fig. 9(c)),130 resulting in a PLQY of 90% along with a narrower FWHM (Fig. 9(d)). It is demonstrated that while such fluoride-mediated surface passivation can enhance the PLQY of InP QDs, the involvement of direct HF treatment still poses safety risks and leads to the loss of QD material. Therefore, Arjan J. Houtepen et al. proposed a simple post synthesis HF-free treatment method based on treating the surface of InP QDs with different fluorides (InF3, ZnF2, AlF3 and MgF2) (Fig. 9(e)).70 The results showed that the InF3-treated InP QDs exhibited the highest PLQY value of 70%. On this basis, the processing temperature of InF3 was further optimized to reach a PLQY of up to 93% (Fig. 9(f)). Meanwhile, by comparing the PL decay lifetime of InP QDs before and after InF3 treatment, they found that the PL decay was almost single exponential and the lifetime prolonged considerably after treatment (Fig. 9(g)), suggesting the suppressed non-radiative recombination in QDs. The lifetime extension was originated from the InF3 treatment that effectively neutralized the non-radiative recombination centers, which is essential for developing high-performance QDs-optoelectronic devices with superior efficiency.70 Furthermore, this group proposed an alternative method involving a water-free and safe post-synthetic treatment that generated HF in situ from benzoyl fluoride to substantially enhance the PLQY of InP QDs, as shown in Fig. 9(h).131 Specifically, the PLQY maintained below 5% at both room temperature and 90 °C regardless of the treatment duration (Fig. 9(i)), while the PLQY exceeded 70% when the temperature reached 200 °C and stabilized within 5 minutes, indicating a non-negligible activation energy barrier for surface reconstruction. Subsequently, the ZnCl2 was introduced as a cooperative ligand to further enhance the PLQY to 85% and prolong the PL decay lifetime (Fig. 9(j)). Both studies successfully circumvented the safety hazards associated with direct HF usage for effective surface passivation, offering diverse pathways for synthesizing high-quality InP QDs with outstanding optical properties.
| QDs | App | Reaction process | Product evolution rate | Stability | Ref. |
|---|---|---|---|---|---|
| w-InP | PC | H2 evolution | 0.6 µmol h−1 cm−1 | 1 h | 26 |
| InP/InPS/ZnS | PC | H2 evolution | 102 µmol mg−1 h−1 | 16 h | 132 |
| InP/ZnSeS:Cu/ZnS | PC | H2 evolution | 337.5 µmol g−1 h−1 | 8 h | 123 |
| InP/ZnSe–G–In | PEC | H2 evolution | — | 50 min | 36 |
| InP/ZnSeS:Al | PEC | H2 evolution | 73.7 µmol cm−2 h−1 | 2 h | 133 |
| InP/ZnSe | PEC | H2 evolution | — | 1 h | 134 |
| InP/ZnSeS-VZn | PEC | H2 evolution | 107.1 µmol cm−2 h−1 | 13 h | 135 |
| InP/ZnS-Re | PC | CO2 reduction | TON of 52 | 6 h | 136 |
| InP | PC | CO2 reduction | TON of ∼51 000 |
420 h | 32 |
| Bacteria sporomusa ovata/InP | PC | CO2 reduction | 0.89 mmol L−1 h−1 | 168 h | 137 |
| InP | PC | N2 fixation | 430 µmol g−1 | 2 h | 138 |
| InP/ZnSe | PC | N2 fixation | ∼2.7 × 107 mol mol−1 CFUinitia−1 | 10 h | 139 |
| InP QDs-S2− | PC | N2 fixation | 0.58 µmol cm−2 h−1 | 12 h | 140 |
| MCC-InP | PC | H2O2 generation | TON of 173 000 |
12 h | 107 |
| InP/GaP/ZnSe | PEC | H2O2 generation | 1.32 µmol min−1 cm−2 | 5 h | 16 |
As illustrated in Fig. 10(a) and (b), David et al. developed a “rainbow” PC system to efficiently utilize the solar spectrum using size-graded InP QDs.26 In this case, smaller QDs (2.7 nm) absorbed the high-energy photons while larger ones (4.5–6.9 nm) captured the lower-energy photons, thus increasing the absorbed power density by 23 mW cm−2 (Fig. 10(a)). Based on this design, smaller InP QDs have demonstrated superior hydrogen generation rates and other InP QDs with different sizes yielded a collective generation rate approaching the sum of their individual performance. This synergistic effect led to an enhanced hydrogen production by merging the broad spectral absorption of larger QDs with the high efficiency of smaller ones (Fig. 10(b)). In another work, Gao's group illustrated the important influence of shell thickness in InP-based core/buffer/shell (CBS) QDs (inserting Sx–In–P1−x buffer monolayer between the core and thin ZnS shell), which resulted in an improved PC hydrogen evolution rate.132 Fig. 10(c) demonstrated that ZnS surface engineering enhanced the PC activity of QDs, with CBS-100 QDs achieving a 77-fold improvement over bare InP QDs at the optimal 1.87 monolayer thickness for H2 evolution. Fig. 10(d) revealed that the CBS-100 QDs exhibited a sevenfold higher PC H2 production performance than the optimized InP/ZnS QDs without the Sx–In–P1−x buffer layer. It is revealed that such interfacial structure minimized the lattice mismatch and facilitated the charge separation for efficient exciton tunneling, while the PC activity of QDs declined beyond the optimal thickness, since the thicker shells further impeded exciton transfer to the QDs surface.
![]() | ||
| Fig. 10 Solar-driven hydrogen evolution using InP-based QDs. (a) Absorption spectra and (b) PC H2 production rates of size-controlled InP QDs and their hybrid systems. (a), (b) Reproduced with permission.26 Copyright 2024 American Chemical Society. (c) Comparison of H2 evolution performance between bare InP and core-buffer-shell structured QDs and (d) the influence of sulfur precursors on the PC H2 generation activity of InP/ZnS QDs. (c), (d) Reproduced with permission.132 Copyright 2023 Royal Society of Chemistry. (e) PC H2 production performance of undoped, Cu-doped, and Mn-doped InP QDs and (f) transient photocurrent spectra. (e), (f) Reproduced with permission.123 Copyright 2024 Elsevier. (g) Schematic of InP core–shell QDs-based PEC cell and (h) corresponding linear sweep voltammetry curves showing photocurrent response. (g), (h) Reproduced with permission.36 Copyright 2024 American Chemical Society. (i) Schematic of the Al-doped InP/ZnSeS-based photoelectrodes for PEC H2 generation and (j) relevant H2 evolution rates. (i), (j) Reproduced with permission.133 Copyright 2024 Wiley-VCH. (k) The ηsep curves of different surface-treated InP/ZnSeS QDs-based photoanodes. (k) Reproduced with permission.135 Copyright 2025 Wiley-VCH. | ||
The charge carrier dynamics of InP-based QDs in PC H2 generation could also be effectively tuned through elemental doping. Specifically, Cu+ doping enables the introduction of hole-trapping states for suppressed charge recombination, which leads to an efficient cocatalyst-free operation to reach a H2 yield rate up to 2700 µmol g−1 (Fig. 10(e)).123 In contrast, Mn2+ doping predominantly enhanced the charge consumption and this positive impact of optimal doping was further verified by transient photocurrent responses (Fig. 10(f)), wherein the Cu-doped QDs showed a photocurrent of about 80 µA, which is substantially higher than that of the Mn-doped (≈35 µA) and undoped QDs (≈14 µA).
The optimization of dopant concentration and the selection of dopant species represent two important research directions for enhancing the solar-driven H2 evolution performance of InP-based QDs. For doping concentration control, Tian et al. employed the gradient In3+ doping in InP/ZnSe QDs to establish a built-in electric field, thereby facilitating the interfacial charge transfer in the QDs-based PEC device (Fig. 10(g)).36 This approach enabled the identification of an optimal In3+ doping level and yielded a maximum saturated photocurrent density of 8.7 mA cm−2 for PEC H2 generation (Fig. 10(h)). Similarly, our group also explored the incorporation of aluminum (Al) as a dopant in InP/ZnSeS QDs for PEC H2 evolution (Fig. 10(i)).133 The underlying mechanism involves the oxidation of Al species during PEC operation to form a passivation layer that suppressed the surface defects and trap states on the QDs. Concurrently, Al occupies zinc lattice sites as shallow donors, which effectively mitigated the non-radiative recombination and enhanced the extraction efficiency of photogenerated electrons from InP/ZnSeS QDs. Owing to this dual functional mechanism, the Al-doped InP/ZnSeS QDs achieved a PEC hydrogen evolution rate of 73.7 µmol cm−2 h−1 with a faradaic efficiency of 68.9% (Fig. 10(j)).
Following the exploration of elemental doping strategies, further advancements in surface engineering have demonstrated remarkable efficacy in regulating charge carrier dynamics of InP-based QDs for enhanced H2 evolution. Our group recently developed a two-step surface engineering approach to precisely manipulate the photoinduced charge carrier kinetics in classical InP/ZnSeS QDs,135 which was achieved by constructing Zn vacancy (VZn)-associated under-coordinated Se/S sites, followed by Cl− ligand exchange. The introduced VZn-under-coordinated sites function as effective hole-trapping centers to promote charge separation, while the subsequent Cl− ligands induced a localized internal electric field that further enhanced the carrier extraction. As compared to the maximum charge separation efficiency (ηsep) of 39% obtained for the untreated InP/ZnSeS QDs, the ηsep of the InP/ZnSeS-VZn QDs-based photoanode increased to 51% (Fig. 10(k),), underscoring the beneficial role of VZn-related sites in improving the charge separation. After Cl− treatment, the ηsep of the InP/ZnSeS-VZn-Cl-based PEC device was further elevated to 76%, confirming that the ligand-induced electric field provided an additional driving force for efficient carrier transport. As a result, the optimized QDs delivered a notable photocurrent density in a PEC hydrogen evolution system, highlighting the practical potential of synergistic vacancy and ligand engineering in solar fuel generation.
000 over 420 h without QD degradation. Additionally, as shown in Fig. 11(f), the PC CO2 reduction performance of this InP QD-TiO2-ReP hybrid PC system was further enhanced with Brønsted acid additives [(H2O), 2,2,2-trifluoroethanol (TFE), or TEOA], which is due to the facilitated CO2-to-CO conversion through proton-coupled electron transfer that resulted in efficient and durable PC CO2 reduction.
![]() | ||
| Fig. 11 InP-based QDs for solar-driven CO2 reduction. (a) The proposed reaction mechanism for size-dependent InP/ZnS QDs coupled with a Re complex and the measured PC (b) CO and (c) CH4 production rates with corresponding AQYs. (a)–(c) Reproduced with permission.136 Copyright 2024 Wiley-VCH. (d) Charge transfer mechanism in the InP QD-TiO2-ReP hybrid PC system, (e) Influence of photosensitizer concentration on CO generation and (f) CO production rates using different proton donors. (d)–(f) Reproduced with permission.32 Copyright 2022 American Chemical Society. (g) Schematic of the S. ovata/InP QDs biohybrid system for artificial photosynthesis. (h) Total acetate yield and (i) cumulative acetate production over seven days under light and dark conditions. (g)–(i) Reproduced with permission.137 Copyright 2022 Elsevier. | ||
In another work, Liu et al. demonstrated that InP/ZnSe/ZnS QDs could be engineered into anaerobic bacteria Sporomusa ovata (S. ovata) as a non-photosynthetic bacterial systems for CO2 photoreduction, enabling efficient light-driven CO2-to-acetate conversion (Fig. 11(g)).137 Under light illumination, the hybrid system showed largely enhanced acetate production compared to the dark condition and the addition of K3Fe(CN)6 as an electron mediator further improved the chemical yield (Fig. 11(h)). The temporal evolution of acetate production (Fig. 11(i)) revealed that the electron mediator accelerated the process by facilitating extracellular electron transfer. This work highlights the unique advantage of intracellular InP-based QDs in directly channeling photogenerated electrons for microbial CO2 reduction.
![]() | ||
| Fig. 12 InP-based QDs for solar-driven ammonia synthesis and H2O2 generation: (a) schematic of electrostatic interaction between surface-modified InP QDs and nitrate ions. (b) Corresponding ammonium production and nitrate conversion efficiency as a function of catalyst loading. (a), (b) Reproduced with permission.138 Copyright 2024 American Chemical Society. (c) The proposed electron transfer pathway in the InP QDs-A. vinelandii biohybrid system. (d) Comparison of ammonia production rate under light and dark conditions. (c), (d) Reproduced with permission.139 Copyright 2022 American Chemical Society. (e) Schematic of charge transfer mechanism in the sulfurized InP QDs/MIL-100(Fe) heterostructure and (f) NH3 yield and (g) faradaic efficiency at different applied potentials. (e)–(g) Reproduced with permission.140 Copyright 2024 Wiley-VCH. (h) Band alignment of MCC-InP QDs with two-electron O2 reduction and H2O oxidation potentials. (i) PC H2O2 production via 2e− ORR using InP QDs with different ligands and core/shell structures. (h), (i) Reproduced with permission.107 Copyright 2025 Wiley-VCH. (j) Charge separation and H2O2 formation mechanism in ZIS-Se photoanode modified with InP-based QDs. (k) Faradaic efficiency for H2O2 generation as a function of applied bias and (l) relationship between theoretical charge consumption and the measured H2 and H2O2. (j)–(l) Reproduced with permission.16 Copyright 2024 Elsevier. | ||
InP-based QDs have been demonstrated as potential building blocks to be integrated with the nitrogen-fixing bacterium Azotobacter vinelandii for light-driven nitrogen fixation.139 It is revealed that culturing the bacteria with InP/ZnSe core–shell QDs during exponential growth enhanced the cellular QDs uptake and substantially improved the ammonia yield when compared to the physically mixed systems. As depicted in Fig. 12(c), the InP/ZnSe QDs absorbed into the interior of bacterial cells could directly transfer the photoexcited electrons to the MoFe protein component of nitrogenase and initialized the nitrogen reduction into NH3. The photogenerated holes are efficiently scavenged by cytoplasmic glutathione (GSH) and regenerated via GSH reductase, maintaining charge balance throughout the PC cycle. Under standardized experimental conditions with equivalent QDs loading per cell, the QDs-bacteria hybrid demonstrated superior ammonia production over control groups under LED irradiation, as quantitatively shown in Fig. 12(d). This biohybrid system exemplifies an efficient strategy for solar-powered nitrogen fixation by directly coupling semiconductor nanomaterials with microbial metabolic pathways.
PEC ammonia synthesis is another promising nitrogen fixation route, yet its practical efficiency is limited by the six-electron nitrogen reduction process.155,156 Considering this limitation, Zhou and colleagues designed a hybrid catalyst by integrating sulfur ligand-modified InP QDs with MIL-100(Fe) to enhance the PEC nitrogen fixation performance (Fig. 12(e)),140 showing that the hybrid composite prepared with 5 mL of InP QDs solution could deliver the highest activity. Fig. 12(f) compared the ammonia production rates of MIL-100(Fe) and the hybrid catalyst under various applied biases, presenting significant performance improvement upon InP QDs incorporation. Specifically, at −0.2 V versus RHE, the hybrid catalyst achieved a maximum ammonia yield of 0.58 µmol cm−2 h−1, representing a 3.09-fold enhancement over the pristine MIL-100(Fe) (Fig. 12(g)). Consistently, the highest faradaic efficiency was also achieved at this potential for both catalysts, while more negative potentials were inclined to induce the competing hydrogen evolution reaction. Control experiments further demonstrated the essential role of Fe–S bonds in promoting the interfacial electron transfer, with no detectable hydrazine byproducts that confirmed the high selectivity toward ammonia formation. These findings recognized InP QDs as effective photocatalysts or photoelectrocatalysts for nitrogen fixation via nitrate and nitrite reduction pathways.
For PEC H2O2 generation system, our group designed a photoanode composed of InP/GaP/ZnSe QDs decorated on selenium-doped ZnIn2S4 (ZIS-Se) to enhance the H2O2 generation efficiency.16 The heterojunction formed between the InP-based QDs and ZIS-Se induced favorable band bending that promoted the two-electron WOR pathway toward H2O2 instead of the competing four-electron oxygen evolution (Fig. 12(j)). Faraday efficiency (FE) measurements further confirmed this behavior across a potential range of 0.65 to 1.85 V versus RHE (Fig. 12(k)), in which the ZIS-Se/QDs photoanode achieved a remarkable average H2O2 Faraday efficiency of 86% between 0.85 and 1.85 V, outperforming both the ZIS-Se and pristine ZIS electrodes. At higher applied biases, the Faraday efficiency decreased moderately to 80.6%, which is attributed to the increasing contribution of alternative water oxidation pathways such as oxygen evolution. As illustrated in Fig. 12(l), the simultaneous production of H2O2 at the photoanode and hydrogen at the cathode was quantified, showing close agreement between experimental yields and theoretical electron counts. The PEC H2O2 generation rate of as-fabricated ZIS-Se/QDs photoanode reached up to 1.32 µmol min−1 cm−2, which is competitive among various other state-of-the-art PEC H2O2 production systems.
Comparative analysis with established QD systems further highlights the competitive advantages of InP QDs. As compared to conventional toxic CdSe QDs, their heavy-metal-free nature complies with RoHS environmental regulations and the larger exciton Bohr radius of InP provides broader tunability of optical properties.161 In terms of prevailing perovskite QDs, InP-based QDs have demonstrated significantly superior environmental stability due to their robust inorganic shell structures. For example, InP/ZnSe/ZnS QDs have maintained stable operation for two years in outdoor luminescent solar concentrators (LSC) whereas perovskite QDs rapidly degrade under similar conditions.162
Beyond these comparisons with other QD systems, InP QDs also present distinct advantages over conventional bulk photovoltaic technologies. Although silicon-based photovoltaics have achieved high efficiency and low cost in standard solar panels, InP QDs offer unique benefits in specialized applications.163 Their solution processability enables low-temperature fabrication on flexible substrates, making them suitable for building-integrated photovoltaics (BIPVs) where traditional rigid silicon panels are impractical.56,60,89,164 Beyond these advantages, InP QDs enable direct solar-to-chemical conversion pathways that bypass the multi-stage energy losses inherent in conventional photovoltaic–electrolysis systems. For instance, in PC CO2 reduction, InP QDs have achieved a TON exceeding 51
000 with stable operation maintained for 420 hours, demonstrating a new route toward higher system-level energy utilization efficiency.32 These characteristics, which combine solution processability for flexible devices, tunable bandgaps for spectral optimization and robust stability for long-term operation, establish clear conditions where InP QDs can complement or compete with conventional photovoltaic routes, particularly in applications requiring durability and direct chemical production.
Nevertheless, the large-scale application of InP QDs faces critical challenges. Global indium production is limited to 800–900 tons per year with over 70% consumed by the indium tin oxide (ITO) industry, leading to supply constraints and price fluctuations between $150–400 per kilogram.165,166 This resource bottleneck necessitates both the development of more efficient synthesis methods to reduce indium consumption and the establishment of comprehensive recycling systems.56,167 Through continued innovation in these directions, InP QDs are expected to play a unique role in specific solar energy conversion fields. Beyond these fundamental constraints, additional core challenges require resolution:
Future research should prioritize the development of safer precursor chemistries, such as aminophosphines and phosphine–carbon monoxide adducts, which offer improved handling and tunable reactivity. A transformative pathway lies in adopting non-coordinating solvent systems or molten salt matrices (e.g., LiCl/KCl) for high-temperature (>400 °C) synthesis (as demonstrated for GaAs and InAs QDs),171 which facilitates atomic rearrangement and defect annihilation to realize superior optoelectronic properties.172,173 Coupling these advanced reactors with machine learning-guided optimization of kinetic parameters (heating rates, precursor stoichiometry) and in situ optical spectroscopy are also crucial for synthesizing InP QDs with precise control over size distribution, composition, and ultimately, batch-to-batch uniformity at commercially relevant scales.
Further investigations of InP QDs-based devices require a multi-faceted approach combining advanced spectroscopy, simulation, and new passivation paradigms. Operando X-ray photoelectron spectroscopy (XPS) and surface-enhanced Raman spectroscopy (SERS) can probe the chemical evolution of the QDs surface during solar-driven PC or PEC processes.174,175 Besides, the cryo-electron microscopy (cryo-EM) could be employed to resolve the atomic structure of the QD-ligand interface and core/shell heterostructures with minimal beam damage, providing direct visualization of surface reconstruction and defect sites that govern charge trapping and passivation efficiency.176,177 Density functional theory (DFT) calculations are essential to map the trap state distributions and identify the optimal ligand binding configurations.178,179 Beyond small molecules, future work may explore the atomic-layer deposition (ALD) of ultrathin, conformal oxide layers (e.g., TiO2, Al2O3) and the use of cross-linkable ligands to create a robust, inorganic–organic hybrid shell that simultaneously passivates traps, enhances stability, and facilitates charge extraction to reactants or electrodes.180,181
The strategic construction of type-II heterostructures (e.g., InP/ZnTe, InP/CdS) or gradient alloyed shells (e.g., InP/ZnSeS) is a powerful tool for spatially separating electrons and holes to prolong carrier lifetime.183–185 Moreover, integrating earth-abundant molecular cocatalysts (e.g., Ni-based complexes for proton reduction, Co-based macrocycles for CO2 reduction) requires precise control over the QDs–cocatalyst interface,186,187 while tuning the anchor group and the molecular structure of the catalyst can optimize the electronic coupling and proton relay pathways. To guide this rational design, future studies could employ transient absorption spectroscopy (TAS) and TRPL to quantify charge transfer rates and efficiencies, directly linking the interfacial structure to function relationship.134 To gain a more comprehensive picture of charge dynamics, future studies may utilize TAS together with operando X-ray absorption spectroscopy (XAS) to probe the oxidation state and local coordination environment of QDs’ active centers during the actual PC or PEC reactions, thus directly correlating the electronic structure of QDs with reaction activity and stability.174,188
Substantial improvements in operational lifetime demand innovative encapsulation strategies, such as embedding QDs within a protective matrix (e.g., a metal–organic framework, a conductive polymer or a mesoporous oxide) to physically isolate them from the electrolyte while permitting mass transport of reactants.190–192 Additionally, the design of multi-component sacrificial reagent systems or integrated redox mediators that can efficiently scavenge the holes is a promising yet underexplored avenue.193 Establishing standardized protocols for accelerated aging tests of InP QDs under realistic sunlight intensities (e.g., AM 1.5G) and operating conditions will be vital for fairly comparing QD material systems and guiding development toward commercially viable durability.
Future work should focus on designing integrated systems with tailored band structures that favor specific reaction pathways. As demonstrated in InP-based QD heterostructure system, the rational interface engineering can effectively steer reaction selectivity toward desired products like hydrogen peroxide.16 System-level optimization, including matched electrode design and efficient product separation, is also essential for achieving high solar-to-chemical conversion efficiency and economic viability.
| This journal is © The Royal Society of Chemistry 2025 |