Rima Heider
Al Omari
a,
Anjan
Kumar
b,
Ali Fawzi
Al-Hussainy
c,
Shaker
Mohammed
de,
Aashna
Sinha
f,
Subhashree
Ray
g and
Hadi
Noorizadeh
*h
aFaculty of Allied Medical Sciences, Hourani Center for Applied Scientific Research, Al-Ahliyya Amman University, Amman, Jordan
bDepartment of Electronics and Communication Engineering, GLA University, Mathura-281406, India
cCollege of Pharmacy, Ahl Al Bayt University, Kerbala, Iraq
dCollege of Pharmacy, The Islamic University, Najaf, Iraq
eDepartment of Medical Analysis, Medical Laboratory Technique College, The Islamic University of Al Diwaniyah, Al Diwaniyah, Iraq
fSchool of Applied and Life Sciences, Division of Research and Innovation, Uttaranchal University, Dehradun, Uttarakhand, India
gDepartment of Biochemistry, IMS and SUM Hospital, Siksha ‘O’ Anusandhan (Deemed to Be University), Bhubaneswar, Odisha-751003, India
hYoung Researchers and Elite Club, Tehran Branch, Islamic Azad University, Tehran, Iran. E-mail: hadinoorizadeh@yahoo.com
First published on 15th August 2025
Indium phosphide (InP) quantum dots (QDs) offer a sustainable, low-toxicity alternative to heavy-metal-based nanomaterials for environmental detoxification. This critical review evaluates their potential as green photocatalysts, focusing on their ability to degrade organic pollutants, including dyes, pesticides, and polycyclic aromatic hydrocarbons (PAHs), under visible-light irradiation. Innovations in ligand functionalization, core/shell architectures, and eco-friendly synthesis enhance colloidal stability, photostability, and charge separation, surpassing traditional photocatalysts such as CdSe/ZnS and TiO2 in efficiency and safety. By elucidating structure–property relationships, this work provides a novel framework for designing scalable, biocompatible nanomaterials, paving the way for advanced nanoremediation technologies to address global pollution challenges.
Environmental significanceThis review highlights the transformative potential of indium phosphide (InP) quantum dots as eco-friendly photocatalysts for environmental detoxification, addressing the urgent need for sustainable solutions to combat organic pollutants such as dyes, PAHs, and pesticides. By synthesizing advancements in surface engineering, synthesis methods, and photocatalytic mechanisms, it underscores InP QDs' advantages over toxic cadmium-based alternatives, including low toxicity, visible-light activity, and robust aqueous stability. The critical evaluation of their performance, environmental fate, and safety offers a roadmap for designing scalable, green nanotechnology platforms. These insights are pivotal for developing efficient, regulatory-compliant remediation technologies, contributing to cleaner water systems and reduced ecological risks in complex environmental matrices. This is the first comprehensive review specifically addressing InP QDs in environmental detoxification. |
Consequently, nanotechnology-based approaches have gained increasing attention as promising alternatives for next-generation environmental detoxification.22–25 Among various nanomaterials, semiconductor quantum dots (QDs) have emerged as highly effective photocatalysts due to their size-tunable bandgaps, strong light absorption, high surface area, and photostability.26–35 Traditionally, QDs composed of cadmium-based materials such as CdSe and CdTe have dominated the field, demonstrating exceptional photocatalytic and optoelectronic performance. However, the inherent toxicity and regulatory limitations associated with cadmium have led to growing concerns regarding their environmental and biomedical applications.36–42
This shift in regulatory and societal priorities has driven a surge of interest in developing heavy-metal-free alternatives that retain the functional advantages of QDs while minimizing ecological risks.43–47
To highlight the increasing research interest in this area, Fig. 1 presents the annual publication trends from 2014 to 2024 related to QDs in environmental applications. To underscore the growing research interest in this field, Fig. 1 illustrates the annual publication trends and citation counts for QDs in environmental applications from 2014 to 2024, drawing on data from papers.48–62 These references highlight advancements in nanomaterial photocatalysis, including 2D materials, MXenes, S-scheme heterojunctions, and others, which provide a broader context for QD development. To enhance readability and provide historical context, Table 1 presents a timeline of key academic research developments in InP QD synthesis, surface engineering, and environmental applications over the past decade. This timeline discusses general trends in nanomaterial synthesis and photocatalysis and highlights progress in green synthesis, core/shell engineering, and photocatalytic applications, offering a clear visual summary of QD research evolution.57–64
Year | Milestone | Description |
---|---|---|
2014 | Early exploration of InP QDs | Initial studies on InP QD synthesis and optoelectronic properties, establishing them as cadmium-free alternatives, inspired by advances in 2D material synthesis techniques (e.g., CVD growth) |
2016 | Advances in surface passivation | Development of ZnS and ZnSe shells to enhance photostability and reduce surface defects, improving aqueous compatibility, paralleling surface modification strategies for non-oxide photocatalysts |
2018 | Green synthesis methods | Adoption of safer precursors (e.g., tris(dimethylamino)phosphine) and ligand-assisted approaches for eco-friendly InP QD production, aligning with sustainable nanomaterial synthesis trends |
2020 | Core/shell engineering | Optimization of multi-shell (e.g., InP/ZnSe/ZnS) and graded-alloy structures to improve charge separation and photocatalytic efficiency, drawing on heterostructure design principles from MXenes |
2022 | Photocatalytic applications | Demonstration of InP QDs for efficient degradation of organic dyes and PAHs under visible-light irradiation in complex aqueous environments, leveraging insights from S-scheme heterojunction photocatalysis |
2023 | Scalable synthesis techniques | Adoption of microfluidic and continuous-flow synthesis for reproducible, high-throughput InP QD production, reflecting scalable fabrication strategies for 3D porous carbon-based materials and MXenes |
2024 | Environmental stability and toxicity | Studies on InP QD stability under high-salinity and variable pH conditions, with reduced ecotoxicity through advanced surface engineering, informed by environmental safety assessments of MXene/CNT hybrids and electro-Fenton processes |
Addressing InP QD challenges requires a deep understanding of their surface chemistry, structural dynamics, and interfacial behavior. Advances in core/shell engineering, such as ZnS, ZnSe, and ZnSeS alloyed shells, have improved colloidal stability, oxidation resistance, and charge carrier lifetimes. Hydrophilic and zwitterionic ligands have enhanced aqueous dispersion across a wide pH and ionic strength range, enabling applications in wastewater and surface water systems. Multi-shell and graded-alloy structures, such as InP/ZnSe/ZnS or InP/Zn0.25Cd0.75S, have further boosted performance in environmental settings.65,66 Synthesis techniques, including hot-injection, heat-up, microwave-assisted, and microfluidic methods, allow precise control over particle size, crystallinity, and surface ligand binding. Green and continuous-flow synthesis protocols align with sustainability goals by reducing hazardous reagents, energy use, and waste, promoting scalability for practical remediation.67,68 The photocatalytic activity of InP QDs hinges on charge carrier dynamics—generation, separation, and transport of photoinduced electrons and holes. Band structure tailoring through alloying or doping, surface state engineering, and integration with conductive matrices (e.g., TiO2, g-C3N4, or MXenes) enhance reactive oxygen species (ROS) formation, pollutant adsorption, and degradation rates. InP-based systems have shown impressive performance in degrading organic dyes, PAHs, and pesticides under visible-light irradiation, even under saline or complex wastewater conditions.69,70Fig. 2 illustrates InP QDs for photocatalytic pollutant degradation, highlighting challenges and engineering strategies.
![]() | ||
Fig. 2 InP QDs synthesized via green methods, surface-engineered for visible-light-driven degradation of pollutants, with lower toxicity and higher efficiency than conventional photocatalysts. |
Despite advances in photocatalytic nanomaterials, significant gaps remain in the literature regarding InP QDs for environmental detoxification. Existing studies on cadmium-based QDs and traditional photocatalysts such as TiO2 lack a unified approach to address eco-friendly synthesis, surface engineering, and scalable applications. Reviews on related materials, such as nitrogen-doped magnetic biochar and graphdiyne, highlight the need for integrated strategies to enhance efficiency and sustainability. This review fills this gap by being the first comprehensive synthesis of InP QD research, benchmarking their performance against CdSe/ZnS and TiO2, and proposing a predictive model for optimization. These contributions provide a novel framework for designing safe, scalable nanotechnology solutions, addressing current limitations in water remediation research. We explore how structural and interfacial modifications impact photocatalytic behavior, assess the environmental fate and stability of these nanostructures, and benchmark their performance against conventional and emerging nanomaterials. By integrating insights from materials chemistry, environmental science, and surface/interface engineering, this work aims to guide the rational design of InP QDs for sustainable remediation technologies. The implications extend beyond laboratory-scale demonstrations, offering pathways toward safe, scalable, and high-efficiency nanotechnologies for addressing global pollution challenges.
The relative band edge positions of common II–VI and III–V semiconductor materials (e.g., ZnS, ZnSe, InP, CdS, and CdSe) can be considered on an absolute energy scale referenced to the vacuum level. This energy alignment is critical for rational heterostructure design, as it determines the nature of charge confinement and separation. Materials such as ZnS and ZnSe, with deeper valence band maxima and higher conduction band minima, often serve as effective shell materials for Type I structures due to their wider bandgaps. Meanwhile, staggered alignments with materials such as CdSe or InP enable the realization of Type II or Quasi-Type II architectures, where asymmetry in carrier localization is exploited for specific electronic and photophysical properties. Normalized PL spectra for core/shell QDs with increasing shell thickness (in monolayers, ML) of ZnS and ZnSe demonstrate key optical properties.87–89 A pronounced red shift in the PL peak is observed with increasing shell thickness, indicative of modified quantum confinement and dielectric screening effects. For the ZnS system, the emission red-shift is accompanied by spectral narrowing, consistent with reduced nonradiative recombination due to surface passivation. Similarly, ZnSe shell growth leads to tunable emission, underlining the impact of shell composition and thickness on excitonic recombination dynamics. These optical tunabilities are pivotal when optimizing QDs for optoelectronic and energy-harvesting applications.90,91
From a crystallographic standpoint, high-resolution transmission electron microscopy (HRTEM) typically reveals well-defined lattice fringes consistent with a face-centered cubic (FCC) phase. However, during the rapid burst of nucleation, particularly in hot-injection methods, stacking faults and twin planes may develop, disrupting the periodic potential and serving as trap sites.70,92 Such structural irregularities introduce shallow or deep defect states within the bandgap, decreasing PLQY and reducing carrier mobility—detrimental effects that compromise photocatalytic efficiency. The suppression of these defects demands meticulous control over precursor conversion kinetics, ligand–shell interactions, and thermal gradients within the reaction medium.93–95 Furthermore, anisotropy in growth rates along different crystallographic axes can lead to non-spherical morphologies such as rods or tetrahedra, especially in ligand-rich environments where facet-selective passivation occurs. Such morphological deviations affect the spatial localization of electronic wavefunctions, the dipole moment distribution, and the surface-area-to-volume ratio—all of which critically determine reactivity and interaction with environmental species. These morphological and structural considerations form the foundational framework upon which surface and interface chemistry must be engineered.96–98 Furthermore, the incorporation of alloyed Zn0.25Cd0.75S shells has been demonstrated to confer exceptional colloidal stability and high photocatalytic performance of InP-based QDs under real wastewater treatment conditions. This structural adaptation effectively mitigates surface oxidation and ensures long-term functionality in harsh aquatic environments.99,100
Fig. 3a presents a representative transmission electron microscopy (TEM) image of the as-synthesized InP QDs, clearly illustrating their spherical morphology and uniform dispersion. The associated size distribution histogram (inset) confirms a narrow diameter range centered around 2.7 ± 0.5 nm, which is critical for controlling surface-to-volume ratios that influence both surface reactivity and defect density.105 This size uniformity is particularly relevant in the context of defect engineering, as smaller QDs inherently possess a greater proportion of surface atoms, thus increasing susceptibility to the formation of undercoordinated sites and native oxides. Given the high surface energy of InP and its sensitivity to oxidative conditions during post-synthetic purification, these structural observations underscore the importance of surface stabilization strategies to mitigate uncontrolled formation of In2O3 or InPO4 phases. Fig. 3b displays the X-ray diffraction (XRD) pattern of the InP QDs, featuring broad reflections characteristic of nanocrystalline domains and indexing to the zinc blende crystal structure. The (111), (220), and (311) diffraction planes are evident, although peak broadening is pronounced due to the nanoscale size, which aligns with the Scherrer effect. Notably, these diffraction features confirm the presence of a well-formed crystalline core, but do not preclude the existence of an amorphous or partially oxidized surface layer. The absence of distinct signals from oxide phases suggests that native oxide formation, while pervasive at the surface (as corroborated by XPS and solid-state NMR in the full study), remains below the detection threshold of XRD. This disparity highlights the necessity of complementary surface-sensitive spectroscopies to detect ultrathin oxide layers or substoichiometric surface species that adversely affect charge carrier dynamics. Fig. 3c illustrates the optical absorption and PL spectra of the InP QDs, revealing two key emissive features: a sharp excitonic peak near 590 nm and a broader, red-shifted band centered around 710 nm. The narrow excitonic emission corresponds to band-edge recombination within the crystalline core, whereas the lower-energy emission is attributed to radiative transitions involving surface trap states, often associated with undercoordinated indium atoms or oxidized phosphorus moieties. These mid-gap states act as recombination centers that quench PL efficiency, consistent with the proposed model of defect-mediated non-radiative Auger recombination. The co-existence of these emissions underscores the duality of the InP QDs' surface: a crystalline interior capable of efficient quantum confinement and a chemically reactive periphery susceptible to oxidation, which together dictate the photostability and catalytic viability of the nanocrystals in real-world applications.
![]() | ||
Fig. 3 (a) TEM image and size distribution of InP QDs (∼2.7 nm); (b) XRD pattern showing the zinc blende crystal structure; (c) UV-Vis absorption and PL spectra highlighting band-edge emission (∼590 nm) and surface trap emission (∼710 nm). Reproduced with permission from ref. 105, copyright 2023 American Chemical Society. |
Importantly, toxicological assessments of InP/ZnS core/shell structures have also confirmed their stability and reduced adverse effects in aqueous systems, supporting their environmental compatibility and functional deployment in detoxification applications. These findings further validate the robustness of surface-engineered InP-based QDs under diverse aquatic conditions.106–108 Additionally, surface reconstruction phenomena can alter the local coordination environment. Density functional theory (DFT) calculations indicate that indium-rich facets tend to exhibit metallic surface states, whereas phosphorus-terminated surfaces favor insulating character. These reconstructions are dynamic and can change upon ligand desorption, pH fluctuations, or interaction with redox species in environmental matrices. Furthermore, the presence of acidic or basic species (e.g., H+, OH−, or carbonate ions) can catalyze surface hydrolysis or promote the formation of metastable hydroxide intermediates, accelerating degradation under realistic environmental conditions.109 Suppressing surface oxidation and defect formation has thus become a key target in InP QD engineering. Strategies include the use of tris(trimethylsilyl)phosphine [(TMS)3P] or less pyrophoric alternatives such as tris(dimethylamino)phosphine (TDMP) under rigorously air-free conditions. Chelating ligands such as carboxylates, thiols, or phosphonates can passivate dangling bonds by forming stable coordination complexes with surface atoms. However, such ligands must simultaneously fulfill criteria of solubility, colloidal stability, and environmental compatibility—a trifecta rarely achieved without trade-offs.110,111
Advanced shell growth protocols such as successive ion layer adsorption and reaction (SILAR) or continuous injection methods have enabled precise control over shell thickness, composition, and uniformity. These protocols offer tunability over the energy offset at the core–shell interface, critical for aligning conduction or valence bands with redox potentials in photocatalysis.120 In certain configurations, Type-II band alignment can be achieved, wherein electrons and holes are spatially separated into different regions of the QD, thereby extending carrier lifetimes and enhancing charge utilization.110 Ligand engineering is an equally critical element of surface passivation. Native ligands such as oleic acid or trioctylphosphine are often replaced with short-chain or functionalized moieties (e.g., dihydrolipoic acid, polyethylene glycol, or zwitterionic ligands) to improve solubility, facilitate interfacial interactions, or enable bioconjugation. In photocatalytic systems, electron-donating ligands can enhance carrier extraction, while hydrophilic moieties improve aqueous stability and dispersion.121–123 Additionally, bifunctional ligands that serve both as passivators and anchoring agents for cocatalysts (e.g., TiO2 and carbon dots) enable the construction of hybrid nanocomposites with synergistic functionalities.101 Notably, shell growth must be optimized to avoid electron/hole tunneling barriers that can restrict interfacial reactivity. Excessively thick shells, while improving stability, can insulate the core from the environment, rendering the QDs inactive in photocatalytic processes. Therefore, rational design must balance core protection with functional accessibility, especially under harsh chemical or photonic conditions encountered in real-world remediation scenarios.124–128 Tailored ligand functionalization and core/shell designs significantly enhance InP QD stability and photocatalytic efficiency, offering a promising avenue for practical water remediation applications.
Comparative characterization underscores the limitations of bare InP cores, where unpassivated surfaces lead to defect-mediated recombination and chemical instability.131,135 Shell engineering, employing materials such as ZnS, ZnSeS, or multi-layered designs, transforms these properties by encapsulating the core and mitigating surface traps.138 For instance, ZnS shells enhance chemical durability, while alloyed shells reduce interfacial discontinuities, optimizing charge carrier dynamics.137 Multishell or doped configurations introduce additional control over electronic properties, achieving superior luminescent efficiency through tailored band offsets.140 These advancements are critical for applications requiring high radiative efficiency or prolonged stability, such as in vivo imaging or photocatalytic systems.
Table 2 presents a comprehensive comparison of experimentally characterized InP QD architectures, highlighting the impact of core/shell engineering on key structural, optical, and surface-related parameters. This synthesis-independent analysis reveals how varying shell materials and surface ligands modulate bandgap energies, PLQYs, and surface oxidation levels—factors that critically determine QD performance in environmental and optoelectronic applications. Unlike earlier discussions focused on synthetic protocols or environmental stability, this table emphasizes the correlative role of surface engineering and experimental outcomes. Unpassivated InP cores consistently show low PLQY values (15–18%) and elevated surface oxidation (up to 15%), attributed to prevalent surface traps such as In2O3 and InPO4. Subsequent shelling with ZnS or ZnSe improves confinement and reduces trap density, with PLQYs rising to ∼60% and oxidation levels dropping below 7%. Multishell or alloyed structures, such as InP/GaP/ZnS and InP/Zn1−xCdxSe, further enhance these properties, reaching PLQYs above 90% due to better lattice matching, defect suppression, and optimized band alignment. The inclusion of gradient shells and intermediate buffer layers significantly mitigates interfacial strain, improving both optical output and long-term colloidal stability. Ligand selection—particularly MPA, sulfide, and zwitterionic types—plays a non-trivial role in tuning surface charge, colloidal behavior, and interaction with the aqueous environment, without introducing the synthetic overlap discussed in prior sections. Collectively, these findings underscore how rational core/shell design, validated by precise characterization metrics, enables the development of environmentally robust, high-performance InP QDs tailored for advanced functional integration.
Sample structure | Crystal phase | Morphology | Surface defects | Ligand/surface | Advantages and disadvantages |
---|---|---|---|---|---|
InP core | Zinc-blende | Spherical | Dangling bonds, In2O3, InPO4 | None | Adv: simple structure, cadmium-free. Disadv: high surface defects, low PLQY, unstable in aqueous environments |
InP core | Zinc-blende | Spherical | P-Vacancies, In2O3 | TDMP | Adv: improved stability with TDMP. Disadv: limited PLQY, persistent oxidation issues |
InP/ZnS core/shell | Zinc-blende | Spherical | Reduced oxides | MPA | Adv: enhanced PLQY, better stability. Disadv: lattice mismatch, moderate oxidation |
InP/ZnS core/shell | Zinc-blende | Spherical | Minimal oxides | Cysteine | Adv: biocompatible ligand, good PLQY. Disadv: slightly lower PLQY than alternatives |
InP/ZnSe core/shell | Zinc-blende | Spherical | Lower strain, reduced oxides | MPA | Adv: reduced strain, good stability. Disadv: moderate PLQY, complex shell deposition |
InP/Zn0.25Cd0.75S core/shell | Zinc-blende | Tetrahedral | Interfacial strain, reduced oxides | MPA | Adv: high PLQY, tunable bandgap. Disadv: Cd presence, interfacial strain issues |
InP/GaP/ZnS multishell | Zinc-blende | Spherical | Intermediate GaP layer, minimal oxides | MPA | Adv: high PLQY, excellent stability. Disadv: complex synthesis, higher cost |
InP/Zn(Mg)Se/ZnS core/gradient shell | Zinc-blende | Spherical | Mg doping, reduced defects | MPA | Adv: very high PLQY, defect suppression. Disadv: Mg doping complexity, scalability challenges |
InP/Zn1−xCdxSe core/shell | Zinc-blende | Spherical | Type-II alignment, minimal oxides | Zwitterionic | Adv: tunable bandgap, high PLQY. Disadv: Cd presence, complex synthesis |
InP/ZnSe/ZnS core/shell/shell | Zinc-blende | Spherical | Double shell, suppressed defects | Sulfide | Adv: exceptional PLQY, robust stability. Disadv: multi-step synthesis, cost-intensive |
InP/Cu:ZnS core/shell | Zinc-blende | Spherical | Cu doping, hole capture, reduced oxides | MPA | Adv: enhanced charge dynamics, high PLQY. Disadv: Cu doping may introduce toxicity |
InP/InPS/ZnS core/buffer/shell | Zinc-blende | Spherical | Buffer layer, minimal lattice mismatch | Sulfide | Adv: high PLQY, excellent stability. Disadv: complex buffer layer synthesis |
Table 2 compares the structural and optical properties of InP QDs with various core/shell configurations, highlighting the impact of shell engineering on performance. Bare InP cores exhibit low PLQY (15–18%) and high surface oxidation (9–15%), limiting their environmental applicability due to instability and defect-related recombination. Core/shell structures, particularly InP/ZnSe/ZnS and InP/InPS/ZnS, achieve high PLQY (up to 95%) and low oxidation (3–5%), demonstrating significant improvements in photostability and efficiency through defect suppression and lattice matching. However, the inclusion of cadmium in some shells (e.g., InP/Zn0.25Cd0.75S) raises toxicity concerns, undermining the cadmium-free advantage of InP QDs. Complex multi-shell or doped systems (e.g., InP/Zn(Mg)Se/ZnS) offer superior performance but face scalability and cost challenges. Future research should prioritize fully cadmium-free, scalable multi-shell designs to balance high PLQY, stability, and environmental safety, leveraging insights from heterostructure engineering in 2D materials.
Ligand-assisted and green synthesis methods prioritize sustainability and morphology control. These approaches use non-pyrophoric precursors (e.g., phosphorus trichloride or acylphosphines) and ligands such as oleic acid or alkylphosphines to stabilize indium complexes and modulate growth kinetics. Microwave-assisted heating (150–200 °C, 5–15 minutes) enhances energy efficiency, producing QDs with sizes of 3–8 nm and PLQY up to 70% with narrow full width at half maximum (FWHM < 40 nm). These methods reduce hazardous waste, aligning with eco-friendly goals for environmental applications.143,144 Heat-up synthesis, a non-injection approach, combines all precursors in a single pot, heating gradually to 180–250 °C. This simpler method yields QDs with sizes of 3–10 nm and PLQY of 30–50%, suitable for applications such as environmental sensing where moderate optical quality suffices. Its scalability and compatibility with green precursors make it industrially viable, though size distribution is broader compared to hot-injection.145,146
Microfluidic and continuous-flow synthesis enable scalable, reproducible production. Precursors and ligands are injected into microchannels or reactors at 180–220 °C, with controlled flow rates (0.1–1 mL min−1), yielding QDs with sizes of 2–8 nm and PLQY up to 70%. Aqueous-compatible ligands such as mercaptopropionic acid (MPA) can be integrated during synthesis, enhancing environmental compatibility. These methods minimize waste and optimize conditions, though reactor fouling remains a challenge.149,150 Cation exchange synthesis converts CdSe or ZnS QDs into InP by replacing Cd2+/Zn2+ with In3+. While preserving template morphology (3–8 nm), this method yields lower PLQY (20–40%) due to impurities, limiting its environmental relevance (Fig. 4).151,152
Synthesis method | Scalability | Environmental compatibility | Advantages | Disadvantages | Ref. |
---|---|---|---|---|---|
Hot-injection | Moderate | Low (hazardous precursors) | Precise size control, good crystallinity | Hazardous reagents, complex purification | 141 and 142 |
Ligand-assisted/green | Moderate | High (green precursors) | Eco-friendly, versatile morphology | Slower process, ligand optimization needed | 143 and 144 |
SILAR | Medium | Moderate (purification needed) | High PLQY, stable shells | Time-consuming, lattice mismatch risks | 147 and 148 |
Microfluidic/continuous-flow | High | High (green synthesis feasible) | Scalable, low waste | Fouling issues, equipment complexity | 149 and 150 |
Heat-up | High | Moderate (green options) | Simple, cost-effective | Lower uniformity, reduced PLQY | 145 and 146 |
Cation exchange | Low | Low (residual impurities) | Morphology retention | Low PLQY, impurity risks, complex process | 151 and 152 |
Microfluidic and ligand-assisted/green methods stand out for their high scalability and environmental compatibility, achieving PLQY up to 80% and supporting sustainable precursors, aligning with green nanotechnology goals.49,51 Hot-injection offers precise control but relies on hazardous pyrophoric precursors, limiting its environmental viability. SILAR provides high PLQY but is time-intensive and prone to lattice mismatch issues. Heat-up methods are simple and scalable but suffer from broad size distributions and lower PLQY. Cation exchange is limited by impurities and low efficiency, making it less practical. Future efforts should focus on optimizing microfluidic and green synthesis methods to combine high PLQY, uniform size distribution, and eco-friendly processes, drawing on scalable fabrication strategies from MXenes and 3D carbon-based materials.51,54
Ligand identity governs colloidal behavior under complex environmental conditions. Carboxylate-terminated ligands, such as oleic acid (OA), achieve high phase transfer efficiency (∼4.2 mg mL−1) in humic acid-rich media due to strong electrostatic interactions, outperforming amine-functionalized ligands (e.g., octadecylamine, ODA) prone to flocculation at high ionic strength.50 Dissolved organic matter (DOM), such as humic and fulvic acids, modulates stability by adsorbing onto QD surfaces, either stabilizing via steric repulsion or inducing bridging flocculation based on molecular weight and charge.129,155 In high-ionic-strength settings (e.g., seawater), charge-screening effects collapse electrical double layers, but sterically hindered or zwitterionic ligands mitigate aggregation.68
Temporal stability is challenged by ligand desorption or degradation under UV illumination or oxidative conditions, leading to aggregation, sedimentation, or release of ionic species (e.g., In3+ and PO43−).129,140 ROS, such as hydroxyl radicals, oxidize surface atoms or cleave ligands, initiating destabilization.70 Advanced core/shell architectures, such as InP/GaP/ZnS, reduce lattice mismatch and surface defects, sustaining stability across diverse conditions.133 Encapsulation in mesoporous silica or antifouling polymer coatings (e.g., pH-responsive PEGylation) further enhances durability and resists fouling in heterogeneous matrices such as wastewater or biofilms.70,156 Functional ligands (e.g., –COOH and –NH2) enable selective pollutant binding (e.g., heavy metals and pesticides), supporting simultaneous capture and photodegradation.110
Transformation pathways depend on surface chemistry, pH, and redox potential. Low pH accelerates ligand desorption, exposing InP cores to hydrolysis, while high pH triggers etching and In3+ release, mitigated by phosphonate ligands.129 Analytical techniques, including field-flow fractionation (FFF), inductively coupled plasma mass spectrometry (ICP-MS), and synchrotron-based X-ray absorption spectroscopy, trace QD dissolution, sedimentation, and bioavailability in real matrices.99,129 Emerging trends include smart ligands responsive to environmental stimuli and hybrid systems (e.g., InP QDs with TiO2) for enhanced stability and reactivity.99 These strategies optimize InP QD performance, ensuring effective and sustainable environmental remediation.
Advanced shell architectures and encapsulation strategies enhance environmental persistence. InP/ZnS QDs with gradient-alloyed ZnSexS1−x shells show only a 15% PLQY reduction during phase transfer compared to 50% for conventional InP/ZnS, due to reduced lattice strain and fewer defects.161 Mesoporous silica coatings suppress phosphorus oxidation by up to 90%, preserving photocatalytic activity in oxidative or high-ionic-strength wastewater matrices.156 Polymer coatings further stabilize QDs but must balance structural robustness with redox accessibility to maintain functionality. These advancements ensure InP QDs withstand dynamic aqueous environments, enabling applications in wastewater treatment and pollutant degradation.
Temporal stability remains a challenge. Ligand desorption under UV or oxidizing conditions leads to aggregation and sedimentation, with aged InP/ZnS QDs showing increased hydrodynamic diameters even at neutral pH.140 Phosphonic acid-capped QDs retain stability at pH 5–8 and ionic strengths up to 150 mM NaCl, suitable for natural and wastewater systems.154 InP/GaP/ZnS core/shell/shell structures with an intermediate GaP layer reduce lattice mismatch, sustaining high PLQY and stability across pH 4–10 and ionic strengths up to 200 mM NaCl.133,161 DOM, such as humic and fulvic acids, complicates stability by adsorbing onto QD surfaces, either stabilizing via steric repulsion or inducing flocculation based on DOM's molecular weight and charge.155 These findings underscore the need for robust surface engineering to ensure long-term stability in complex aquatic ecosystems.
The surface structure, including exposed facets and shell composition, influences reactivity. Anisotropic InP/ZnS QDs (e.g., nanorods and tetrapods) exhibit localized charge carrier behavior and enhanced pollutant adsorption on high-index facets, enabling pollutant-specific catalysis.154 Core/shell/shell architectures such as InP/GaP/ZnS or InP/InPS/ZnS create interfacial energy gradients, facilitating charge migration and selective pollutant activation, as confirmed by transient absorption spectroscopy and electrochemical impedance analysis.133,162 Zwitterionic ligands balance hydrophilicity and antifouling properties, minimizing matrix interference while supporting pollutant capture.163
Table 4 summarizes surface modification strategies enhancing InP QD versatility under environmental stressors. PEGylation suppresses In3+ leaching in DOM-rich media via hydration shells, while silica encapsulation provides a robust barrier, improving biocompatibility.156 Alloyed shells (e.g., ZnSexS1−x) mitigate lattice strain and passivate defects, sustaining PLQY and stability under high-ionic-strength or oxidative conditions.161,164 Strongly binding ligands such as TDMP enhance air stability without compromising optoelectronic properties.154 These strategies address ionic fluctuations, DOM interactions, and oxidative degradation, establishing InP QDs as adaptable platforms for environmental remediation.
Strategy/ligand type | Environmental challenge | Mechanism | Advantages and disadvantages | Ref. |
---|---|---|---|---|
Responsive –NH2 ligands | Biofouling in wastewater | Functional charge-based interactions | Adv: adaptive to pH changes, antifouling. Disadv: limited stability in high salinity | 99 |
MPA (mercaptopropionic acid) | pH/ionic strength variability | Electrostatic repulsion via –COOH, –SH | Adv: broad pH stability, biocompatible. Disadv: moderate oxidation protection | 154 |
InP@ZnSexS1−x shell alloying | Lattice strain, surface traps | Bandgap engineering, defect suppression | Adv: enhanced stability, high PLQY. Disadv: complex synthesis, potential Cd traces | 147 |
PEGylation (MW > 5000) | DOM-rich water | Hydration barrier, antifouling | Adv: excellent antifouling, low ion release. Disadv: high MW PEG costly, bulky | 161 |
Silica encapsulation | Oxidation, bioaccumulation | Physical barrier, biocontainment | Adv: robust protection, biocompatible. Disadv: reduced photocatalytic accessibility | 158 |
TDMP ligand coating | Air-sensitivity, degradation | Strong P-ligand bonding | Adv: strong passivation, air stability. Disadv: limited aqueous dispersibility | 165 |
PEGylation and silica encapsulation excel in reducing ion release and biofouling, ensuring stability in complex aqueous environments, similar to surface modification approaches in non-oxide photocatalysts.56 MPA and –NH2 ligands offer broad pH stability and antifouling properties but provide less robust oxidation protection. ZnSexS1−x alloying improves PLQY and stability through defect suppression but may introduce trace cadmium, compromising environmental safety. TDMP coatings enhance air stability but limit aqueous dispersibility, restricting their use in wastewater. Future research should prioritize hybrid strategies combining ligand functionalization (e.g., MPA and PEG) with alloyed shells to achieve optimal stability, selectivity, and reactivity while ensuring cadmium-free compositions, drawing on S-scheme heterojunction principles.53
Surface PEGylation is a key strategy to reduce toxicity by minimizing protein corona formation and nonspecific membrane adsorption. PEG-modified InP QDs exhibit extended circulation times, reduced macrophage uptake, and lower ROS generation in cellular assays.157,159 High-molecular-weight PEG (MW > 5000) suppresses In3+ release by up to 95% in aqueous media, as confirmed by ICP-MS.161 Mesoporous silica encapsulation further enhances safety, reducing cellular uptake by 98% in HeLa and A549 cell lines while preserving optical properties and structural integrity.158,161 Dynamic coatings, such as pH- or redox-responsive ligands, enable site-specific QD disassembly, minimizing non-target interactions.158 Thicker ZnS shells also reduce bioactivity, limiting bio-interactions by enhancing shell integrity.140 These advanced surface engineering strategies align with regulatory demands for sustainable nanomaterials, supported by lifecycle analyses and environmental exposure modeling.159,160 By integrating robust coatings such as PEGylation and silica encapsulation, InP QDs achieve minimal ecological footprints while retaining photocatalytic efficacy. Iterative design ensures biocompatibility and environmental safety, enabling their safe deployment in remediation applications.
InP QDs also benefit from band-edge engineering via controlled shell composition. Reverse Type-II band alignment is one of the most effective configurations, where both the valence and conduction bands of the shell lie at higher energy than those of the core. This structure causes photogenerated holes to migrate toward the shell while electrons remain confined within the core, resulting in spatially separated carriers. Recent studies on InP/Zn0.25Cd0.75S QDs demonstrated this architecture's remarkable capability to facilitate solar-driven hydrogen evolution with turnover numbers exceeding two million per QD.167 Ligand engineering plays a pivotal role in improving exciton dynamics. Replacing traditional long-chain insulating ligands with inorganic sulfide ions not only enhances surface passivation but also dramatically improves charge extraction. Sulfide ligands serve as efficient hole acceptors, capturing valence band holes and reducing the barrier for interfacial charge transfer. Sulfide-capped InP and InP/ZnS QDs have demonstrated turnover numbers up to 128000 in the hydrogen evolution reaction, with internal quantum yields reaching 31%.168 These outcomes suggest that ligand-induced modifications at the QD surface can be as critical as the core–shell composition in optimizing photocatalytic performance.
In Fig. 5a, control experiments underscore the indispensability of all photocatalytic system components—Ni2+ as the proton reduction catalyst, ascorbic acid (H2A) as the sacrificial electron donor, and illumination—without which hydrogen evolution is negligible.168 This finding validates the optimized environment required for exciton generation and utilization in InP and InP/ZnS QDs. Fig. 5b directly compares InP–S QDs with two different CdSe-based systems (CdSe-A and CdSe-B) under identical conditions, revealing superior hydrogen evolution efficiency of InP–S, which highlights their potential as non-toxic, heavy-metal-free alternatives for photocatalysis. Furthermore, Fig. 5c shows that QDs with excitonic absorption peaks centered around 525 nm exhibit the highest photocatalytic performance, confirming the importance of optimal light absorption for maximizing exciton utilization. As shown in Fig. 6d, QDs capped with sulfide ions exhibit a nearly tenfold increase in hydrogen evolution compared to QDs modified with organic ligands such as MPA, MUA, and OLA. The superior performance of sulfide-capped QDs is corroborated in Fig. 6e by their markedly enhanced transient photocurrent response, signifying more efficient photoinduced charge separation. Electrochemical impedance spectroscopy (Fig. 5f) further confirms this enhancement, with sulfide-capped QDs showing the lowest charge transfer resistance (Rct), indicative of facilitated interfacial charge mobility. These results directly support the assertion that ligand-induced modifications are pivotal in optimizing exciton separation and reducing recombination.
![]() | ||
Fig. 5 Photocatalytic and charge transfer behavior of InP and InP/ZnS QDs with various ligands. (a) All components are required for H2 evolution. (b) InP–S QDs outperform CdSe–S analogs. (c) Peak activity occurs at 525 nm. (d) Sulfide ligands yield higher H2 than OLA, MPA, and MUA. (e) Photocurrent confirms improved charge separation. (f) Sulfide-capped QDs exhibit the lowest impedance. (g) Inorganic S2− ligands outperform Cl− and PO43−. (h) EPR shows light-induced signals only for S2−-capped QDs. (i) SPV confirms enhanced hole extraction and charge separation with sulfide ligands. Reproduced from Yu et al., Nat. Commun., 2018, 9, 4009, under the Creative Commons Attribution 4.0 International License (CC BY 4.0).168 |
![]() | ||
Fig. 6 (a) Schematic of O2 activation via charge transfer in InP/g-C3N4 under visible light. (b) PL spectra show reduced recombination with InP loading. (c) Time-resolved PL confirms shorter carrier lifetime and improved charge separation. Reproduced from Cao et al., Chin. Chem. Lett., 2020, 31, 2689–2692, under the Creative Commons CC BY license.172 |
The influence of various inorganic ligands on photocatalytic performance is further dissected in Fig. 5g–i. As seen in Fig. 6g, InP/ZnS QDs functionalized with Cl− or PO43− exhibit significantly lower hydrogen evolution yields compared to their sulfide-functionalized counterparts, reinforcing the notion that sulfide ions uniquely act as efficient hole acceptors. This is further supported by Fig. 6h, where electron paramagnetic resonance (EPR) spectroscopy reveals a strong signal at g = 2.006 in sulfide-capped QDs under illumination—an indicator of stabilized photogenerated electrons due to effective hole transfer. In contrast, QDs capped with Cl− or PO43− show no such response. Finally, Fig. 6i displays the surface photovoltage (SPV) spectra, in which sulfide-capped QDs yield the highest photovoltage, affirming their enhanced charge separation and lower recombination probability. These photophysical and electrochemical characteristics collectively validate the role of sulfide ligands in maximizing photocatalytic performance by facilitating rapid, barrier-free hole transfer from the QD interior to the surrounding medium. These figures provide compelling evidence that ligand engineering—particularly through the incorporation of sulfide ions—is not merely a surface-level modification but a deeply influential strategy that governs core photophysical behaviors. The observed improvements in charge carrier separation, interfacial transfer kinetics, and reduced recombination losses directly align with the mechanisms described earlier for InP/ZnSe/ZnS and InP/Zn0.25Cd0.75S QDs.
Moreover, doping strategies provide additional levers for tuning charge dynamics. For example, copper-doping in InP/ZnSeS/ZnS QDs introduces shallow trap states within the shell that act as hole sinks, suppressing recombination and allowing more photogenerated electrons to participate in redox reactions. In contrast, Mn doping—although structurally beneficial—can introduce competing redox activity where Mn4+ captures electrons, thereby reducing hydrogen evolution efficiency.169 These insights reinforce the importance of careful selection of dopants based on their redox chemistry and energy level alignment relative to the InP core. Another innovative approach to enhance exciton separation involves forming heterojunctions with conductive two-dimensional materials. InP QDs integrated into Ti3C2Tx MXene sheets have shown exceptional improvements in interfacial charge separation due to built-in electric fields that promote unidirectional carrier migration. These composite systems achieve hydrogen peroxide production rates more than 100 times higher than pristine InP, while simultaneously improving COD degradation in real wastewater.99
The structural symmetry and crystallinity of InP QDs also influence charge transport pathways. Wurtzite InP QDs, synthesized via cation exchange methods, offer elongated exciton diffusion lengths due to their anisotropic crystal field effects. When synthesized in mixed-size distributions, small QDs contribute high-energy photon utilization while larger ones extend absorption into the red and near-infrared regions, leading to more comprehensive solar harvesting. This size-distribution strategy has demonstrated superior solar-to-hydrogen conversion efficiencies compared to monodisperse systems.170 To further improve carrier lifetimes, researchers have also investigated the role of host matrices. For instance, embedding InP QDs in protective scaffolds such as mesoporous silica or TiO2 matrices not only minimizes oxidation but also provides additional interfaces for charge extraction. InP-QD-decorated TiO2 composites exhibit electron injection times as short as 25 ps, which is essential for fast charge separation and efficient redox conversion.171
Charge recombination remains one of the major challenges in InP QD systems. Photogenerated electrons and holes often undergo nonradiative decay via surface trap states or interfacial defects. To combat this, strategies such as mid-gap state engineering, ligand replacement with short-chain donors, and energy-gradient shelling have been developed. Among these, gradient alloy shells such as ZnSexS1−x have proven particularly effective in smoothing the electronic potential at the core–shell interface, leading to reduced wavefunction mismatch and enhanced PLQY.167 The manipulation of charge carrier dynamics in InP QDs involves a synergistic balance of crystallographic design, band alignment, ligand engineering, doping, and interface coupling. Each of these factors contributes uniquely to the suppression of recombination and promotion of exciton dissociation, ultimately determining the photocatalytic performance. The successful deployment of InP QDs for environmental remediation depends not only on material synthesis but also on their charge separation.
An emerging area of interest is the coupling of InP QDs with advanced semiconductors to broaden pollutant scope and improve degradation efficiency. For example, InP/g-C3N4 heterostructures have been shown to activate molecular oxygen under visible light, promoting the formation of O2− radicals. This configuration significantly enhances NO removal efficiency while suppressing NO2 generation, indicating a selective oxidation pathway facilitated by interfacial charge transfer.172 This is particularly relevant for atmospheric pollutant control, where oxidative selectivity is critical. This selective oxidative mechanism is further illustrated by the interfacial interaction model shown in Fig. 6a, where InP QDs are anchored onto g-C3N4 nanosheets to establish a built-in electric field. Upon visible-light illumination, photogenerated electrons transfer from g-C3N4 to InP QDs, enabling molecular O2 adsorption and activation to O2− species. This interfacial configuration facilitates directional electron flow and lowers the activation barrier for molecular oxygen, driving the formation of reactive oxygen intermediates and enhancing NO-to-NO3− conversion while minimizing undesirable NO2 accumulation. PL measurements further confirm the improved charge separation efficiency. As depicted in Fig. 6b, the steady-state PL intensity of g-C3N4 significantly decreases with increasing InP QD loading, indicating suppressed recombination of photogenerated carriers. The time-resolved PL spectra in Fig. 7c support this observation, showing a reduction in average electron lifetime from 6.86 ns for pristine g-C3N4 to 6.16 ns for 0.5% InP/g-C3N4. These findings validate that InP QDs serve as effective electron sinks, accelerating charge transfer and sustaining active radical generation, thereby underpinning the enhanced photocatalytic performance of the heterostructure.
![]() | ||
Fig. 7 Structural and optical characterization of InP/ZnS core/shell QDs: (a) TEM image; (b) XRD pattern confirming the core/shell structure; (c) UV-Vis absorption and PL emission spectra; (d) thermal stability of fluorescence at various annealing temperatures; (e) schematic of energy level alignment. Reproduced from Kuo et al., Nanoscale Res. Lett., 2017, 12, under the Creative Commons Attribution 4.0 International License (CC BY 4.0).154 |
In hybrid configurations with TiO2 nanoparticles, InP or InP/ZnSe QDs act as photosensitizers, enhancing charge separation and enabling dye degradation even in the visible range. The sulfonate-functionalized InP/ZnSe QDs electrostatically bind to TiO2 surfaces, forming stable assemblies under neutral and acidic conditions. After conjugation, the PL intensity of the QDs decreases, indicating efficient charge transfer to TiO2 and suppressed recombination. This process enhances the degradation rate of rhodamine B and other dye molecules via singlet oxygen or hydroxyl radical-mediated pathways.173
Photocatalytic activity of InP QDs can be optimized by tailoring surface ligands and the reaction medium. Sulfide-capped InP QDs exhibit enhanced pollutant interaction due to their compact size, elevated charge density, and improved aqueous dispersion. Sulfide-capped all-inorganic InP/ZnS QDs were shown to degrade pollutants through a mechanism driven by rapid hole transfer to the sulfide ligands, promoting oxidation reactions at the QD surface, enabling selective and efficient degradation, particularly in hydrogen-evolving systems with dual photocatalytic functionality.168 Additionally, the design of multicomponent photocatalytic platforms extends the pollutant degradation landscape. In the work by Chon et al., InP QDs were immobilized on TiO2 loaded with Re-based CO2 reduction catalysts. In this ternary system (InP–TiO2–ReP), photogenerated electrons from InP are transferred to TiO2 and then to ReP, enabling selective CO2-to-CO conversion under mild conditions. Although originally designed for carbon reduction, the system's electron-transfer architecture is equally applicable to complex pollutant reduction pathways, offering a framework for multistep photochemical remediation.171
Environmental compatibility and toxicity considerations are equally important when evaluating degradation systems. Chen et al. studied the toxicity of InP/ZnS QDs on the embryos of Gobiocypris rarus, highlighting that these QDs, when accumulated in biological systems, interfere with embryo hatching and gene expression related to developmental pathways. This finding underscores that degradation intermediates and unreacted QDs may pose ecological risks, necessitating controlled deployment and post-reaction purification strategies.140 The kinetics of pollutant degradation typically follows pseudo-first-order or Langmuir–Hinshelwood models, depending on pollutant type and catalyst surface properties. InP-based systems show a clear dependency of degradation rate on light intensity, QD loading, and pollutant concentration. For example, in systems using InP/Ti3C2Tx composites, degradation of dyes such as acid orange 7 and the generation of H2O2 follow a linear correlation with irradiation time, supporting a radical-mediated degradation mechanism. The reported degradation rate of 99.9% within 60 minutes indicates efficient surface activation and high ROS turnover.99
Recent studies have also investigated the degradation of more complex molecules such as pharmaceuticals and pesticides. InP QDs have been shown to degrade deltamethrin with over 90% efficiency, following a mechanism involving ester bond hydrolysis and subsequent radical-induced fragmentation.131 These results point to the potential of InP QDs in addressing emerging contaminants, which often resist degradation via traditional photochemical or oxidative approaches. Furthermore, multi-scale separation techniques such as stepwise agglomeration have been applied to isolate InP/ZnS QDs with high PLQY and minimal byproducts. Rezvani et al. showed that fractions rich in small QDs exhibited superior ROS generation capacity, highlighting the role of size uniformity and surface quality in pollutant degradation kinetics.174 Studies involving hybrid doping and shell thickness optimization reveal that pollutant degradation efficiency can be systematically tuned. Zeng et al. demonstrated that the introduction of a ZnSe midshell in InP/ZnS QDs enhances electron transfer while retaining stability, which in turn improves photocatalytic degradation rates for persistent organic pollutants166 and other researchers confirm that doping with Cu+ ions suppresses recombination and increases the lifetime of reactive intermediates.169
The incorporation of a ZnSe intermediate shell not only mitigates lattice mismatch but also enhances exciton delocalization. InP/ZnSe/ZnS QDs with a ZnSe midshell were found to reduce exciton–phonon coupling, thereby extending carrier lifetime and improving transfer efficiency.166 This three-layered architecture sustains high PLQY and offers superior thermal and photostability, ideal for prolonged environmental applications. Doping provides an additional strategy for band structure modulation. Incorporating Cu+ ions into the ZnS shell introduces shallow trap states that enhance charge extraction. InP/Cu:ZnS QDs demonstrated reduced recombination and accelerated electron transfer to redox acceptors such as benzoquinone, resulting in improved visible-light-driven hydrogen evolution without additional cocatalysts.175 These results underscore the effectiveness of electronic structure tuning through doping for optimizing QD performance in energy and environmental applications.
In parallel, hybrid architectures that combine InP QDs with electron-conducting scaffolds have demonstrated substantial advantages. For example, the integration of InP QDs into MXene (Ti3C2Tx) nanosheets leverages the high surface area, conductivity, and chemical compatibility of the 2D host. These heterostructures exhibit excellent interfacial charge separation, enabling both sacrificial-agent-free hydrogen peroxide generation and >99% degradation of complex dyes in wastewater.99 Their performance underscores how hybrid systems can overcome the limitations of colloidal QDs alone, especially in complex media. The crystallography of QDs significantly influences their photocatalytic efficiency. Wurtzite-phase InP QDs, distinguished by anisotropic charge transport properties, demonstrate enhanced spectral harvesting, particularly in the red and near-infrared regions of the solar spectrum. Size-tuned wurtzite InP QDs were synthesized, revealing that larger dots extended absorption to longer wavelengths but exhibited reduced activity, while smaller dots showed higher catalytic efficiency. Combining both size regimes resulted in a synergistic effect, improving solar-to-fuel conversion rates through wider spectral absorption and optimized reactivity.170 Another promising strategy is the conjugation of InP QDs with visible-light-active semiconductor photocatalysts such as graphitic carbon nitride (g-C3N4). InP/g-C3N4 hybrids promote efficient molecular oxygen activation, leading to the generation of superoxide radicals that can oxidize NO directly to nitrate. The use of g-C3N4 also improves the long-term chemical stability of InP QDs by offering a protective matrix that resists photobleaching.172 These systems illustrate the broader utility of InP QDs beyond aqueous pollutant degradation, extending into gas-phase and atmospheric remediation.
Multistep agglomeration techniques provide an effective approach for purifying and stabilizing InP/ZnS QDs from a synthetic perspective. Size-selective separation was employed to isolate QDs with uniform PL characteristics and reduced shelling byproducts.174 Fractions containing smaller, more consistent InP/ZnS QDs exhibited higher quantum yields and enhanced resistance to photodegradation, demonstrating that post-synthesis refinement significantly improves photocatalytic robustness. Long-term stability under operational conditions remains a major challenge for photocatalytic QDs. Photobleaching, oxidation, and ligand detachment are commonly observed, particularly in aerobic or high-intensity illumination environments. To mitigate these effects, several protective strategies have been developed. For example, sulfide-capped all-inorganic InP/ZnS QDs exhibit enhanced oxidative stability and retain photocatalytic activity over extended cycles. These improvements stem from the compact ionic shell structure, which resists water infiltration and ligand displacement.168
In environmental applications, the stability of InP QDs must be evaluated with ecological safety in mind. InP/ZnS QDs were found to induce developmental toxicity in aquatic organisms through ROS generation and chorion disruption.140 Consequently, surface passivation strategies that minimize ROS leakage into surrounding media are crucial for safe deployment. Coating QDs with inert silica or embedding them in polymeric matrices represent promising approaches currently under exploration. InP QD performance in photocatalytic applications is deeply influenced by advanced band structure engineering, hybridization with functional substrates, and long-term structural and chemical stability. Strategies such as alloyed shelling, reverse Type-II architecture, size-controlled synthesis, and dopant incorporation provide tunable control over electronic and optical properties. Hybrid systems further elevate performance by facilitating carrier transport and stabilizing QDs under operational stress. Together, these innovations are rapidly pushing InP QDs toward real-world, scalable applications in environmental nanotechnology. The superior charge separation and visible-light response of InP/ZnS QDs underscore their potential as advanced photocatalysts for pollutant degradation.
Table 5 summarizes the photocatalytic degradation of organic pollutants by InP QDs, demonstrating high efficiencies (>95% for dyes and 90–99.6% for PAHs) driven by ROS generation and charge transfer mechanisms. Dyes degrade rapidly (60–90 min) due to favorable carboxylate ligands and Type-II alignments, while pesticides and PAHs require longer times (120–180 min) due to complex molecular structures and matrix effects. Hybrid matrices and alloyed shells enhance performance, but slower kinetics for pesticides and matrix sensitivity for PAHs remain challenges. Compared to non-oxide photocatalysts,56 InP QDs excel in visible-light activity but need optimization for faster degradation in complex wastewater. Future work should focus on hybrid InP QD systems with conductive matrices (e.g., g-C3N4 and MXenes) to enhance kinetics and broaden pollutant applicability.54
Pollutant type | Examples | Mechanism | Key features | Advantages | And disadvantages | Ref. |
---|---|---|---|---|---|---|
Dyes | Crystal violet, methylene blue | ˙OH, O2− generation, charge transfer | Carboxylate ligands, Type-II alignment | High efficiency, fast kinetics | Limited to specific dyes, ligand-dependent | 131, 161 and 157 |
Pesticides | Malathion, chlorpyrifos | Hole oxidation, ROS production | Hydrophilic ligands, alloyed shells | Effective for pesticides, stable shells | Slower kinetics, complex matrices | 99 and 161 |
PAHs | Phenanthrene, naphthalene | ˙OH-mediated ring cleavage | Hybrid matrices, anisotropic morphologies | High efficiency for PAHs, versatile | Long degradation times, matrix complexity | 99, 131 and 154 |
Additionally, Tran et al. (2023) reported that InP QDs, when surface-modified with sulfide ligands for aqueous dispersion, could degrade caffeic acid under visible light with a removal efficiency of 60%, significantly outperforming TiO2, which exhibited negligible activity under similar conditions.176 These results underline the visible-light responsivity of InP QDs, attributed to their direct bandgap (∼1.34–2.3 eV depending on size and shell composition), efficient charge separation in Type-II band structures, and surface states that facilitate redox reactions with target molecules. InP QDs outperform several alternative nanomaterials in photocatalytic applications. TiO2 and ZnO, though widely applied, are restricted to UV activation and suffer from fast charge recombination and photocorrosion during long-term use.177–180 Carbon-based materials such as graphene oxide and carbon dots are known for their chemical stability but generally exhibit lower photocatalytic efficiency than InP-based QDs in degrading dyes and pesticides, mainly due to their lower absorption cross-section in the visible range.181,182 CuInS2/ZnS QDs, while cadmium-free and responsive to visible light, tend to have broader emission spectra and higher polydispersity, which can result in less predictable photocatalytic performance compared to well-engineered InP/ZnS QDs.163 Although CdSe/ZnS QDs show excellent efficiency, their cadmium content raises serious environmental and regulatory issues, making InP-based QDs a more sustainable and environmentally friendly option for advanced photocatalysis.183
Beyond simple degradation rates, reaction kinetics and byproduct analysis provide further evidence of InP QDs' superior performance. The photocatalytic degradation of PAHs by InP/ZnS follows first-order kinetics, with pathway analysis indicating a primary mechanism via phthalic acid intermediates—a marker of deep ring cleavage, which signifies complete mineralization rather than mere transformation.131 Equally important is the structural adaptability of InP QDs. Alloying strategies such as Zn0.25Cd0.75S shells and the use of intermediate layers (e.g., GaP or Mg-doped ZnSe) enable bandgap tuning and suppression of interfacial defects, leading to higher photostability and quantum yield (up to 94%). Such architectures have been shown to extend charge carrier lifetimes and reduce surface oxidation, both of which are critical for retaining high photocatalytic turnover over repeated use cycles. Importantly, the colloidal and chemical stability of InP QDs in realistic aquatic environments further distinguishes them. InP/ZnS QDs functionalized with hydrophilic ligands (e.g., MPA or PEG–thiol) resist aggregation in high-ionic-strength solutions, such as seawater, retaining consistent activity across a wide pH range (4–10) and ionic strength up to 200 mM NaCl.153 This is in contrast to many oxide-based catalysts, which exhibit loss of activity due to surface fouling or instability under similar conditions.184,185
Nevertheless, InP QDs are not without challenges. Oxidative degradation of surface ligands or the core/shell interface under prolonged light exposure can lead to performance decline. However, encapsulation strategies—such as silica coating or hybridization with antioxidant nanomaterials—are under investigation and have shown promising results in extending QD longevity in reactive environments.105 From a photocatalytic standpoint, InP-based QDs rival or outperform many conventional and alternative nanomaterials, offering a unique balance between high photonic activity, environmental compatibility, and structural tunability. Their performance in degrading persistent organic pollutants—without relying on UV light or sacrificial agents—and their demonstrated compatibility with real wastewater systems highlight their utility in sustainable nanoremediation technologies.
Aquatic toxicity is a critical factor for nanomaterials used in water remediation. The impact of InP/ZnS QDs on rare minnow (Gobiocypris rarus) embryo development was investigated across concentrations from 50 to 400 nM, revealing dose-dependent inhibition of embryo hatching. Effects included altered gene expression related to hatching enzymes, structural damage to the chorion, and evidence of oxidative stress.140 Although these findings highlight potential ecotoxicity at higher concentrations, the effects are notably less severe than those observed with Cd-based QDs, which are associated with embryo lethality and long-term reproductive dysfunction in fish models.186 Nonetheless, the findings emphasize that “low toxicity” does not equate to “no toxicity.” Continued research is needed to fully map the ecological risks of InP QDs, especially under chronic exposure and in complex environmental matrices.
Unlike the relatively inert TiO2 nanoparticles, InP QDs undergo chemical transformations in the presence of oxygen, light, and water. Surface phosphorus and indium atoms readily oxidize to form In2O3, InPO4, and PO43− species, many of which create mid-gap trap states that impair PL and increase solubility in water.135 These oxidative processes not only degrade optical performance but also influence environmental fate by potentially increasing the mobility and bioavailability of indium ions. Advanced surface passivation techniques, including gradient shells (e.g., ZnSeS), bifunctional ligands, or silica encapsulation, substantially mitigate ecotoxicity effects. Replacing HF etching with a greener metal-oleate treatment was shown to yield more stable core–shell structures, suppressing reoxidation of the InP surface during prolonged exposure.154 This approach is essential for preventing the leaching of indium and phosphorus compounds into aquatic environments. Compared to CuInS2-based QDs, which also suffer from surface dissolution and ion release, InP QDs can be rendered more chemically inert with appropriate surface modifications. In contrast, TiO2 particles, while chemically stable, can catalyze the formation of ROS under UV light, inadvertently leading to secondary pollution through oxidation of natural organic matter or even DNA damage in aquatic organisms.164,187
Fig. 7a presents a TEM image of InP/ZnS core/shell QDs synthesized through a green solvothermal approach, revealing monodisperse, quasi-spherical nanocrystals with an average diameter of ∼4 nm.51 This size uniformity and surface morphology are vital for suppressing mid-gap trap states that typically arise from surface oxidation of indium and phosphorus atoms in bare InP QDs. Encapsulation within a ZnS shell provides steric and chemical protection, reducing the susceptibility of the core to oxygen and moisture—factors that would otherwise lead to the formation of In2O3 and PO43− species. This encapsulation thus directly addresses the environmental and functional limitations of unprotected InP systems. Panel (b) shows the powder XRD patterns for InP, ZnS, and InP/ZnS QDs. The diffraction peaks of the core/shell QDs are observed at intermediate positions between those of pure InP and ZnS phases, consistent with coherent heteroepitaxial growth and lattice strain accommodation at the interface. This structural shift confirms the presence of a conformal ZnS shell, which is crucial for passivating reactive InP surfaces and preventing the formation of defect states that enhance solubility and ion leaching. The importance of this heterostructure lies in mitigating oxidative degradation pathways that compromise optical stability and increase the mobility and bioavailability of indium ions in aqueous media. Panel (c) displays the UV-Vis absorption and fluorescence spectra of the InP/ZnS QDs, showing a strong excitonic absorption around 480 nm and a sharp emission centered at ∼530 nm with a FWHM of 55 nm. The high PL intensity and narrow emission profile indicate effective surface passivation by the ZnS shell, which suppresses nonradiative recombination associated with surface oxidation. The inset shows a bright green fluorescence under UV light, further demonstrating optical stability. Panel (d) examines thermal stability across various annealing temperatures (25–150 °C). Notably, the fluorescence intensity remains relatively stable up to 70 °C over several hours, with a gradual decline at higher temperatures, suggesting partial thermal reactivation or annealing of surface traps followed by degradation. This reinforces the importance of surface engineering—such as gradient shells and ligand strategies—for enhancing resistance to thermal oxidation and preserving long-term luminescence under variable environmental conditions.
Understanding the behavior of QDs in real water matrices is essential for predicting their long-term environmental impacts. A detailed study compared the phase transfer behavior of InP/ZnS, CuInS/ZnS, and CdSe/ZnS QDs under varying environmental conditions, including pH, ionic strength, and dissolved organic matter. InP/ZnS QDs showed intermediate transfer rates and lower ion release compared to Cd-based QDs, though their phase behavior was significantly affected by the ligand shell, especially under acidic or high-salt conditions.153 The maximum ion release was reported for indium-based QDs in acidic media, emphasizing the need for pH-stable ligand coatings. This finding is particularly relevant for water remediation in industrial or coastal zones where variable pH and salinity levels can rapidly destabilize nanomaterials. InP-based QDs offer a promising balance of performance and environmental safety compared to traditional cadmium-based systems. InP QDs exhibit reduced systemic toxicity, minimal ion release under standard conditions, and high adaptability via surface engineering, positioning them as promising candidates for eco-friendly nanotechnologies. However, challenges such as oxidative degradation under environmental exposure, subtle ecotoxicity at elevated doses, and surface chemistry sensitivity remain. Overcoming these issues requires robust shelling strategies, long-term ecotoxicity evaluations in complex ecosystems, and comprehensive life cycle analyses. By addressing these aspects proactively, InP QDs can be responsibly deployed in various environmental applications, including sensing, catalysis, pollutant degradation, and sustainable water treatment solutions.165,176
Surface engineering plays a pivotal role in optimizing InP/ZnS QDs for photocatalysis. Ligand choice significantly affects water dispersibility and charge separation. For example, Tran et al. showed that sulfide ligand exchange on InP QDs enhances aqueous stability, achieving a 58% degradation of caffeine under visible light within 4 hours, far surpassing TiO2, which shows negligible activity under similar conditions.140 Similarly, Abbasi et al. reported that InP/ZnS QDs functionalized with MPA achieved degradation efficiencies of 91.78% for Congo red and 99.6% for phenanthrene, attributed to reduced surface defects and improved electron–hole separation.131 In contrast, oleylamine (OAM)-coated InP/ZnS QDs exhibit aggregation in high-ionic-strength media (e.g., seawater), reducing photocatalytic efficiency due to limited pollutant access to active sites.174 Advanced shell architectures further enhance performance. Gradient ZnSeS or Zn0.25Cd0.75S shells mitigate lattice mismatch, suppress non-radiative recombination, and boost quantum yields up to 94%, ensuring sustained activity over extended irradiation cycles.161
Additionally, Chen et al. showed that bifunctional metal-oleate treatments minimize surface reoxidation, preserving photocatalytic efficiency during prolonged use.189Fig. 8a outlines the dual-function mechanism for surface treatment of InP QDs using bifunctional metal oleate precursors. The process begins with native InP QDs capped by surface oxides (InPOx), which typically act as non-radiative recombination centers. Treatment involves two steps: (1) etching and removal of InPOxvia oleic acid (OA), and (2) surface passivation using metal oleate complexes (Zn-oleate or Cd-oleate). The resulting QDs—ZnOA–InP—are stabilized by oleate ligands coordinated with zinc ions. This approach avoids the safety hazards of conventional HF etching and is better at preventing reoxidation, forming oxide-free surfaces critical for high PLQY and shell growth compatibility. Panels (b) and (c) show UV-Vis absorption spectra and corresponding PLQY values for InP QDs treated with increasing concentrations of ZnOA and CdOA, respectively. The systematic blue shifts in the first excitonic absorption peak with increasing ZnOA/CdOA ratio reflect the mild etching effect that reduces core size. Importantly, the PLQY increases significantly with ZnOA (up to ∼3.3%) and CdOA (up to ∼2.3%) treatments (insets), indicating effective surface passivation and suppression of trap states. The zinc-based treatment achieves a more consistent and narrower size distribution (sharper excitonic peak), while the cadmium variant introduces slight redshifts, possibly due to defect state formation or exciton delocalization onto surface states. Panel (d) evaluates the temporal photostability of three InP QD samples (ZnOA-, CdOA-, and HF-treated) under atmospheric aging for up to 144 hours. ZnOA–InP exhibits a steady increase in PLQY over time, possibly due to slow surface ligand rearrangement and stabilization. In contrast, HF–InP shows a sharp PLQY drop, underscoring its poor ambient durability and fast surface reoxidation. CdOA–InP demonstrates moderate stability but lower performance than ZnOA, reinforcing the latter's superior long-term passivation. These findings underscore the benefit of bifunctional ligands in providing both chemical protection and optical performance, key to prolonged photocatalytic or optoelectronic applications. Panel (e) compares PL properties—PLQY and FWHM—after overcoating the core QDs with ZnSe/ZnS shells. ZnOA–InP QDs show a marked improvement in both metrics: they achieve a PLQY of ∼34.4% and exhibit the narrowest emission linewidth (FWHM ∼ 48 nm), indicating a well-defined core/shell interface with minimal defect-mediated broadening. This contrasts with the broader spectra (FWHM > 60 nm) observed in CdOA- and HF-treated samples. Thus, ZnOA treatment not only enhances core quality but also optimally supports subsequent shell growth—critical for producing high-performance InP-based QDs suitable for sustainable photocatalysis or display technologies.
![]() | ||
Fig. 8 (a) Schematic of bifunctional metal oleate treatment for simultaneous surface oxide removal and passivation of InP QDs. (b) and (c) Absorption spectra and PLQY trends of ZnOA- and CdOA-treated InP QDs at varying metal ion ratios. (d) Aging stability of PLQY over time for ZnOA-, CdOA-, and HF-treated QDs. (e) PLQY and FWHM of ZnSe/ZnS-shelled QDs. Reproduced from Chen et al., Nanomaterials, 2022, 12, 573, under the Creative Commons Attribution 4.0 International License (CC BY 4.0).189 |
Environmental conditions profoundly influence QD behavior in aquatic systems. Marşan et al. found that InP/ZnS QDs retain colloidal stability across a pH range of 4–10, but acidic conditions (pH < 4) accelerate indium ion release, forming InPO4 or In2O3 and creating mid-gap trap states that impair performance. The presence of humic acid (HA) can stabilize QDs by forming a protective corona at low concentrations (<10 mg L−1), enhancing dispersibility; however, high HA levels (>10 mg L−1) shield pollutants, reducing degradation rates.174 Ionic strength also plays a critical role. In high-salinity environments (e.g., 200 mM NaCl), hydrophilic ligands such as MPA or PEG–thiol prevent aggregation, ensuring consistent photocatalytic activity, unlike CuInS2/ZnS QDs, which destabilize above 50 mM NaCl.173,174 Prolonged light exposure poses another challenge. Oxidative degradation of surface ligands under visible light can diminish efficiency, but encapsulation strategies, such as silica coating or hybridization with antioxidant nanomaterials, mitigate this issue, extending QD longevity.135
Compared to alternative nanomaterials, InP/ZnS QDs offer distinct advantages. TiO2 and ZnO, while widely used, require UV activation and suffer from photocorrosion, limiting their efficacy in visible-light-driven systems.140 CuInS2/ZnS QDs, though cadmium-free, exhibit broader size distributions and inconsistent photocatalytic output due to polydispersity.85 CdSe/ZnS QDs, despite high efficiency, pose environmental risks from cadmium leaching, making InP/ZnS a safer, sustainable alternative.183
Colloidal stability across variable environmental conditions is critical for real-world deployment. InP/ZnS QDs functionalized with hydrophilic ligands such as MPA or PEG–thiol exhibit robust dispersion in high-salinity environments (up to 200 mM NaCl) and a broad pH range (4–10).153 This contrasts sharply with CuInS2/ZnS QDs, which tend to aggregate at ionic strengths above 50 mM, and bare InP cores, which rapidly oxidize and precipitate. Although carbon dots and g-C3N4 maintain moderate dispersion in freshwater, their stability in complex matrices often deteriorates unless chemically modified. The matrix confirms that surface passivation and ligand engineering are indispensable for maintaining functional colloidal behavior. Environmental transformation is another critical metric captured in the matrix. While InP/ZnS QDs are free from heavy metal leaching, they are not chemically inert. Under oxidative conditions, surface degradation can lead to the formation of InPO4 and In2O3 species, which may affect both photocatalytic efficiency and ecological fate.135,165 Nevertheless, encapsulation strategies—such as silica coating or gradient shell engineering—have been shown to dramatically reduce ion release, preserve surface integrity, and extend photostability.65,189 Compared to CdSe/ZnS QDs, which exhibit significant cadmium leaching under environmental stress, InP-based QDs offer a demonstrably safer alternative.
Another advantage of InP-based QDs lies in their tunable band alignment and selective reactivity. As indicated in the matrix, InP/ZnS structures support first-order reaction kinetics in pollutant degradation and facilitate intermediate formation indicative of complete mineralization pathways.131 Hybrid architectures and core/shell/shell designs enhance spatial charge separation and electron–hole lifetime, attributes rarely found in single-phase oxides or carbon nanostructures. This makes InP QDs particularly suitable for applications demanding high selectivity and advanced degradation mechanisms, such as multi-step oxidation of aromatic pollutants or degradation of recalcitrant pesticides. From a toxicity and regulatory standpoint, InP QDs offer a relatively low-risk profile. Studies indicate that their systemic and aquatic toxicity is considerably lower than cadmium-based analogs, especially when encapsulated or surface-functionalized.140,157 Although some adverse effects have been observed at high concentrations, such as embryo hatching inhibition in fish models, these are largely mitigated through surface engineering. This makes InP/ZnS QDs more viable for long-term environmental deployment than CdSe/ZnS or CuInS2-based systems, which face growing restrictions due to potential ion leaching and bioaccumulation risks.153,187
Interpreting the matrix from an application perspective reveals alignment between specific nanomaterials and deployment contexts. InP/ZnS QDs are especially well-suited for wastewater treatment systems with high salinity or fluctuating pH, owing to their colloidal robustness and visible-light activation.157 In contrast, TiO2 may still be favored in UV-rich environments where cost constraints outweigh efficiency. For biological sensing or imaging applications, InP's low systemic toxicity and high PLQY under environmentally relevant conditions further reinforce its utility. The matrix thus enables rational decision-making by mapping structure–function relationships to environmental scenarios. Despite the advantages of InP-based QDs, the matrix also highlights persistent gaps—particularly regarding sustainability metrics such as recyclability, synthetic energy cost, and life-cycle analysis. These dimensions remain underexplored but are increasingly critical for evaluating the true environmental footprint of nanomaterials.190–194 Moving forward, comparative frameworks should integrate functional performance with ecological and economic parameters to provide a holistic view of nanomaterial suitability. In this context, InP QDs show great potential, but comprehensive assessments across their entire life cycle are necessary to validate their green credentials.
The comparative matrix substantiates the superior performance, tunability, and environmental compatibility of InP/ZnS QDs relative to conventional photocatalysts. Their ability to combine structural resilience with high catalytic turnover—without relying on hazardous elements such as cadmium—positions them as front-runners for next-generation nanoremediation technologies. By offering a structured decision-making framework, the matrix approach bridges the gap between materials science and environmental application, promoting the strategic deployment of InP-based nanomaterials across diverse remediation contexts.195,196 Overall, the 10–20% efficiency edge of InP/ZnS over TiO2 and CdSe/ZnS (e.g., 91.8% Congo red degradation) highlights its viability as a safer, more effective alternative.
Table 6 compares InP-based QDs with competing nanomaterials, highlighting their superior visible-light activity (e.g., 91.8–99.6% degradation for InP/ZnS) and improved safety profiles compared to CdSe/ZnS, which suffers from high toxicity and cadmium leaching. InP/ZnS QDs offer excellent colloidal stability and low ion release, making them suitable for environmental applications, though their complex synthesis is a drawback. TiO2 and carbon dots are limited by UV dependence and low visible-light efficiency, respectively, while g-C3N4 and MoS2 show promise but lag behind InP in degradation efficiency. InP-loaded g-C3N4 demonstrates synergistic potential, doubling NO removal rates.53 Future research should focus on simplifying InP QD synthesis and integrating them with eco-friendly matrices such as g-C3N4 or MXenes54 to enhance performance while maintaining sustainability and regulatory compliance.
Property/material | InP QDs (bare core) | CdSe/ZnS QDs | CuInS2/ZnS QDs | Carbon dots/GO | g-C3N4 | MoS2 QDs |
---|---|---|---|---|---|---|
Light activation | Visible (520–600 nm) | Visible, UV-enhanced | Visible | Visible-NIR | Visible | Visible |
Photocatalytic degradation | CR: 88.1%, CV: 74.5% | >90%, varies | 60–85%, unstable | 20–40% | InP-loaded: ×2 NO removal | ∼75% dyes |
Stability in water | Unstable without ligands | Moderate | Poor > 50 mM NaCl | Good in freshwater | Moderate | High with PEG |
Oxidation/photocorrosion | Oxidation to InPO4 | Cd2+ leach | Sulfide release | Low oxidation | Light-sensitive | Low recombination |
Aquatic toxicity | Moderate (embryos) | High (lethal) | Moderate | Low | Mild | Low |
Reaction kinetics | First-order | Variable | Broad | Mixed | Enhanced with InP | 1st order |
Colloidal stability | Low ionic tolerance | Moderate | Poor | Moderate | HA-stable | PEG-stabilized |
Ligand/interface effects | High; S2− enhances activity | Ox-sensitive | Ligand-dependent | Tunable | Enhanced via InP | Passivated |
Environmental transformation | In3+/PO43− release | High Cd leach | S2−/Cu+ ion release | DOM corona | Low persistence | Inert forms |
Sustainability/regulation | Cd-free; tunable | Highly restricted | Cd-free but inconsistent | Eco-friendly | Metal-free | Green material |
Advantages | Cd-free, tunable bandgap | High efficiency | Cd-free, visible-light active | Eco-friendly, low toxicity | Metal-free, enhanced with InP | Green, low toxicity |
Disadvantages | Poor stability, high ion release | Toxic Cd leaching, restricted use | Unstable, inconsistent performance | Low efficiency in visible light | Light-sensitive, moderate stability | Lower efficiency than InP |
First-principles DFT calculations complement these experimental approaches by providing molecular-level insights into InP QD properties. DFT studies have elucidated surface reconstruction, band alignment, and defect-mediated recombination in InP QDs, revealing that indium-rich facets exhibit metallic states, while phosphorus-terminated surfaces favor insulating behavior.109 These calculations guide the design of core/shell heterostructures, such as InP/ZnSeS or InP/Zn0.25Cd0.75S, to optimize charge separation and redox potentials for photocatalysis.147,189 Recent reviews on S-scheme heterojunctions, carbon-based QDs, MXene hybrids, and single-atom catalysts highlight the synergy of in situ techniques and DFT in understanding charge transfer mechanisms and enhancing photocatalytic efficiency.52–54 For example, in situ XPS and DFT have been used to study S-scheme heterojunctions, revealing enhanced charge separation in InP-based hybrids with Ti3C2Tx MXenes for H2O2 production.99 These methods also inform strategies to mitigate surface oxidation and ligand desorption, critical challenges in aqueous environments. By integrating in situ characterization with computational modeling, researchers can rationally design InP QDs with tailored surface chemistry and band structures, paving the way for high-efficiency, sustainable photocatalytic systems for environmental remediation.
The table below provides a comparative overview of InP QDs, carbon-based QDs, g-C3N4, and Ti3C2Tx MXenes for photocatalytic environmental detoxification, evaluating their photocatalytic efficiency, aqueous stability, scalability, and toxicity. InP QDs, while offering a cadmium-free platform with tunable bandgap properties, are limited by moderate photoluminescence quantum yield (40–70%), aqueous instability due to ligand aggregation, costly synthesis, and potential In3+ toxicity, which can be mitigated through green synthesis (e.g., microfluidic reactors) and hybrid integration with conductive supports such as MXenes, drawing inspiration from advancements in carbon-based QDs and non-oxide photocatalysts (Table 7).49,54,181,207
Material | Photocatalytic efficiency | Aqueous stability | Scalability | Toxicity | Key advantages | Key limitations | References |
---|---|---|---|---|---|---|---|
InP QDs | Moderate (PLQY ∼ 40–70%) | Low (ligand aggregation) | Limited (costly precursors) | Moderate (In3+ release) | Cd-free, tunable bandgap | Surface defects, instability | 153 and 173 |
Carbon-based QDs | High (PLQY > 80%) | High (hydrophilic) | High (green synthesis) | Low (biocompatible) | Cost-effective, stable | Limited selectivity | 183 and 212 |
g-C3N4 | High (visible-light active) | Moderate (aggregation) | Moderate (exfoliation issues) | Low | Metal-free, stable | Defect-limited efficiency | 53 and 56 |
Ti3C2Tx MXenes | High (with co-catalysts) | Moderate (oxidation) | Moderate (etching complexity) | Low | High conductivity | Surface termination issues | 54, 207 and 208 |
Fig. 9 presents a radar chart comparing the current challenges and future directions for InP QDs in environmental detoxification, highlighting six critical aspects: surface defects, ligand stability, photochemical stability, toxicity, scalability, and catalytic selectivity. The red dataset (current challenges) quantifies the severity of limitations (scored 4–8), including high surface defect density due to native oxide formation, ligand instability in aqueous environments, photochemical degradation under prolonged irradiation, potential ecotoxicity from In3+ release, scalability constraints of hot-injection synthesis, and limited selectivity in complex wastewater matrices. The blue dataset (future directions) illustrates the potential for improvement (scored 2–3) through advanced strategies such as compositionally graded shells, stimuli-responsive ligands, hybrid integration with 2D materials (e.g., MXenes and g-C3N4), biodegradable coatings, green synthesis via microfluidic reactors, as informed by recent advancements in nanomaterials.49,54,181,207–213 This visual roadmap underscores the transformative potential of interdisciplinary approaches to enhance the efficacy and sustainability of InP QDs for environmental applications.
This journal is © The Royal Society of Chemistry 2025 |