Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Pioneering perovskite quantum dot nanosensors for heavy metal ion detection: mechanisms, design, and industrial applications

Suleiman Ibrahim Mohammadab, Asokan Vasudevancd, I. B. Sapaevefgh, Munthar Kadhim Abosaodaij, Chou-Yi Hsuk, Malatesh Akkurl, Alok Kumar Mishram, Gaganjot Kaurn, Rajesh Singho and Ahmad Mohebi *p
aElectronic Marketing and Social Media, Economic and Administrative Sciences, Zarqa University, Jordan
bResearch Follower, INTI International University, 71800 Negeri Sembilan, Malaysia
cFaculty of Business and Communications, INTI International University, 71800 Negeri Sembilan, Malaysia
dShinawatra University, 99 Moo 10, Bangtoey, Samkhok, Pathum Thani 12160, Thailand
eHead of the Department Physics and Chemistry, “Tashkent Institute of Irrigation and Agricultural Mechanization Engineers” National Research University, Tashkent, Uzbekistan
fScientific Researcher of the University of Tashkent for Applied Science, Uzbekistan
gSchool of Engineering, Central Asian University, Tashkent 111221, Uzbekistan
hWestern Caspian University, Scientific Researcher, Baku, Azerbaijan
iCollege of Pharmacy, The Islamic University, Najaf, Iraq
jDepartment of Medical Analysis, Medical Laboratory Technique College, The Islamic University of Al Diwaniyah, Al Diwaniyah, Iraq
kDepartment of Pharmacy, Chia Nan University of Pharmacy and Science, Tainan 71710, Taiwan
lDepartment of Physics & Electronics, School of Sciences, JAIN (Deemed to be University), Bangalore, Karnataka, India
mDepartment of Electrical & Electronics Engineering, Siksha 'O' Anusandhan (Deemed to be University), Bhubaneswar, Odisha 751030, India
nDepartment of Electronics and Communication Engineering, Chandigarh University, Mohali, Punjab, India
oUttaranchal Institute of Technology, Uttaranchal University, Dehradun 248007, Uttarakhand, India
pDepartment of Chemistry, Young Researchers and Elite Club, Islamic Azad University, Tehran Branch, Tehran, Iran. E-mail: a.mohebiacademic@gmail.com

Received 21st August 2025 , Accepted 16th October 2025

First published on 23rd October 2025


Abstract

Perovskite quantum dots (PQDs) are a transformative platform for ultrasensitive heavy metal ion detection, leveraging high photoluminescence quantum yield (PLQY, 50–90%), narrow emission spectra (FWHM 12–40 nm), tunable bandgaps, and defect-tolerant optoelectronic properties. This review is the first systematic evaluation of PQD-based nanosensors for detecting Hg2+, Cu2+, Cd2+, Fe3+, Cr6+, and Pb2+, elucidating key mechanisms like cation exchange, electron/hole transfer, Förster resonance energy transfer, and surface trap-mediated quenching. Advanced synthesis methods, including hot-injection, ligand-assisted reprecipitation, and microwave-assisted techniques, enable precise control over size, crystallinity, and surface chemistry. Lead-based PQDs (e.g., CsPbX3) achieve limits of detection (LODs) as low as 0.1 nM with rapid response times (<10 s), while lead-free variants (e.g., Cs3Bi2X9, CsSnX3) offer eco-friendly alternatives with enhanced aqueous stability. PQD@MOF composites and ratiometric designs enhance selectivity in complex matrices, surpassing carbon quantum dots and semiconductor QDs in sensitivity and versatility. Applications include industrial wastewater remediation, lubricant quality control, and environmental compliance, ensuring ecosystem protection and product integrity. This seminal work addresses challenges like aqueous instability, Pb toxicity, scalability, and matrix interference, benchmarking PQDs against alternative nanomaterials. Future directions include comparisons with other nanoparticles, multiplexed sensing platforms, and sustainable lead-free innovations. By integrating fundamental insights with practical applications, this review establishes PQDs as a high-impact paradigm for advancing heavy metal ion sensing in industrial and environmental contexts, guiding innovations in sensitivity, selectivity, and scalability.


Ahmad Mohebi

Ahmad Mohebi is a researcher in the Department of Chemistry, Young Researchers and Elite Club, Tehran Branch, Islamic Azad University, Tehran, Iran. He specializes in nanomaterials and sensor development, focusing on perovskite quantum dots for heavy metal ion detection in environmental and industrial applications. Mohebi leads interdisciplinary projects on sustainable nanotechnology, advancing sensor designs for wastewater and lubricant monitoring.


1. Introduction

Heavy metal ion contamination in environmental, industrial, and biological systems poses significant risks to ecosystems, human health, and industrial processes due to their toxicity, persistence, and bioaccumulation.1–3 Elements such as mercury (Hg2+), copper (Cu2+), cadmium (Cd2+), iron (Fe3+), chromium (Cr6+), and lead (Pb2+) are prevalent pollutants in industrial wastewater, lubricants, and natural water bodies, necessitating robust detection methods to ensure environmental compliance and product integrity.4–6 Traditional sensing technologies, including atomic absorption spectroscopy,7,8 inductively coupled plasma mass spectrometry,9,10 and electrochemical methods,11,12 offer high sensitivity but are often hindered by high costs, complex instrumentation, and limited portability, making them less suitable for real-time or on-site applications.13,14 In response, nanomaterial-based nanosensors have emerged as promising alternatives, offering rapid response, portability, and cost-effectiveness.15 Among these, perovskite quantum dots (PQDs) have garnered significant attention due to their exceptional optoelectronic properties, positioning them as a transformative platform for heavy metal ion detection.16,17

PQDs, characterized by the general formula ABX3 (where A is a monovalent cation, B is a divalent metal cation, and X is a halide anion), exhibit unique attributes that make them ideal for sensing applications.18,19 Their high photoluminescence quantum yield (PLQY, 50–90%), narrow emission spectra (full width at half maximum (FWHM) 12–40 nm), and tunable bandgaps enable ultrasensitive detection with high color purity and specificity. These properties arise from quantum confinement effects at nanoscale dimensions (2–10 nm), which result in discrete energy levels and enhanced oscillator strengths.20,21 Additionally, PQDs possess large absorption coefficients (105 to 106 cm−1) and fast radiative recombination rates, facilitating efficient light harvesting and stable fluorescence signals critical for detecting analyte-induced changes. The structural versatility of PQDs allows compositional tuning through substitutions at the A, B, or X sites, enabling tailored optical and electronic properties for specific sensing needs.19,20 Lead-based PQDs, such as CsPbX3 and CH3NH3PbX3, are renowned for their superior PLQY and tunable emission across the visible spectrum, while lead-free variants, such as Cs3Bi2X9 and CsSnX3, address toxicity concerns, offering eco-friendly alternatives with improved stability in aqueous environments.22

The detection of heavy metal ions using PQDs relies on their ability to undergo fluorescence quenching or enhancement triggered by interactions with analytes. Mechanisms such as cation exchange, electron transfer, and Förster Resonance Energy Transfer (FRET) underpin these responses, driven by the high surface-to-volume ratio and defect-tolerant nature of PQDs.23,24 For instance, the ionic radius similarity between Pb2+ and Hg2+ facilitates rapid cation exchange in lead-based PQDs, leading to efficient quenching, while electron transfer dominates in interactions with transition metal ions like Cu2+ or Fe3+.17,21 Surface ligands, such as oleylamine or poly(ethylenimine) (PEI), play a critical role in modulating selectivity and sensitivity by passivating defects and mediating ion–PQD interactions. Advanced synthesis methods, including hot-injection, ligand-assisted reprecipitation, and microwave-assisted techniques, enables precise control over PQD size, crystallinity, and surface chemistry, directly impacting their sensing performance.25

The industrial relevance of PQD-based nanosensors lies in their ability to address critical monitoring needs in sectors such as chemical processing, wastewater treatment, and lubricant quality control.26,27 Their high sensitivity and rapid response times make them suitable for detecting trace contaminants in real-time, ensuring compliance with stringent environmental regulations and maintaining product quality. However, challenges such as the aqueous instability of lead-based PQDs, toxicity concerns, and scalability of synthesis methods must be addressed to facilitate widespread adoption. Emerging strategies, including the development of lead-free PQDs, advanced encapsulation techniques, and computational modeling, offer pathways to overcome these limitations, paving the way for practical implementation.28,29

This review provides the first comprehensive analysis of PQD-based nanosensors for heavy metal ion detection, systematically exploring their structural, optical, and mechanistic foundations. It covers diverse PQD types, including lead-based (e.g., CsPbBr3) and lead-free (e.g., Cs3Bi2Br9) systems, their synthesis methods, and their advantages and limitations for sensing applications. The review evaluates sensor performance, focusing on sensitivity, selectivity, and limits of detection (LODs) ranging from sub-nanomolar (e.g., 0.1 nM for Cu2+ in organic solvents) to micromolar levels, depending on the ion and matrix. Integration with stable matrices like metal–organic frameworks and ratiometric sensing designs enhances performance in complex industrial and environmental matrices. By comparing PQDs with other nanomaterials, such as carbon QDs and traditional semiconductor quantum dots, and addressing both fundamental and applied aspects, this work aims to guide future research, fostering innovations to improve the sensitivity, selectivity, and scalability of PQD-based nanosensors for diverse heavy metal ion detection applications (Fig. 1).


image file: d5ra06200d-f1.tif
Fig. 1 Overview of PQD-based nanosensors for heavy metal ion detection, highlighting structure, performance, integration strategies, and comparison with other quantum dots.

2. Fundamentals of perovskite quantum dots

2.1. Structural and optical properties of perovskite quantum dots

PQDs are nanoscale semiconductor materials characterized by the general formula ABX3, where A is a monovalent cation (e.g., Cs+, methylammonium (MA+), or formamidinium (FA+)), B is a divalent metal cation (e.g., Pb2+, Sn2+, or bismuth (Bi3+)), and X is a halide anion (e.g., Cl, Br, or I). The crystal structure of PQDs typically adopts a cubic, tetragonal, or orthorhombic lattice in 3D ABX3 forms, depending on the composition and temperature.30,31 In this 3D structure, the B cation is coordinated octahedrally by six X anions, forming interconnected [BX6]4− octahedra that create a three-dimensional framework, with A cations occupying cuboctahedral voids for charge neutrality. This extended connectivity enables efficient charge transport and high PLQY (50–90%). In contrast, 0D/2D derivatives like A3B2X9 (e.g., Cs3Bi2X9) feature isolated [B2X9]3− dimers or layered structures with reduced octahedral connectivity, resulting in stronger quantum confinement, broader emission spectra (FWHM ∼40–60 nm), and typically lower PLQY (20–50%), which impacts sensitivity in sensing applications due to increased defect states.19,28,40 The structural versatility of perovskites enables compositional tuning at A, B, or X sites to modulate properties critical for heavy metal ion detection.

The stability of the perovskite lattice is governed by the Goldschmidt tolerance factor (t), defined as:

 
image file: d5ra06200d-t1.tif(1)
where rA, rB, and rX are the ionic radii of A, B, and X ions, respectively. A tolerance factor between 0.8 and 1.0 ensures a stable 3D cubic structure, while lower values favor 0D/2D phases. For instance, CsPbBr3 (t ≈ 0.81) adopts a cubic 3D structure at room temperature, offering high optical efficiency, whereas Cs3Bi2Br9 (t < 0.8) forms a 0D layered structure, enhancing aqueous stability but reducing radiative efficiency.32,33,43

The optical properties of PQDs are dominated by quantum confinement effects due to their nanoscale dimensions (2–10 nm), resulting in discrete energy levels and size-dependent bandgaps. This confinement enhances oscillator strength, leading to high PLQY (50–90% for 3D ABX3; 20–50% for 0D/2D A3B2X9) and narrow emission spectra (FWHM 12–40 nm for 3D; broader for 0D/2D). For example, 3D CsPbBr3 emits green light at ∼520 nm with FWHM ∼20 nm, ideal for selective sensing, while 0D Cs3Bi2Br9 shows blue emission (∼450 nm) with higher stability in polar media.15,26,43 The bandgap can be tuned by size or halide composition; smaller PQDs or Cl-rich 3D variants exhibit wider bandgaps (∼3.0 eV for CsPbCl3), while I-rich 3D forms have narrower ones (∼1.7 eV for CsPbI3). 0D/2D structures often display wider effective bandgaps due to confinement. PQDs possess large absorption coefficients (105 to 106 cm−1) and fast radiative recombination rates, with defect tolerance minimizing non-radiative losses—more pronounced in 3D forms where defects lie within bands.30,34

The surface chemistry of PQDs significantly influences their properties. PQDs are capped with ligands like oleylamine (OAm) or oleic acid (OA), which passivate defects and enhance stability. In 3D ABX3, ligands primarily provide steric protection, while in 0D/2D A3B2X9, they also mitigate interlayer defects for better aqueous sensing.18,35 Exciton dynamics in PQDs feature short radiative lifetimes (1–10 ns) and high binding energies (20–100 meV), enabling sensitive energy transfer for detection. The high surface-to-volume ratio amplifies ion interactions, with 0D/2D structures showing enhanced trap states for quenching but potentially slower kinetics compared to 3D forms.14,36

2.1.1. Crystal structure and compositional flexibility. A tolerance factor between 0.8 and 1.0 typically ensures a stable cubic perovskite structure, while deviations may lead to tetragonal or orthorhombic phases, impacting optical and sensing performance. For instance, CsPbBr3 adopts a cubic structure at room temperature, offering high symmetry and optical efficiency, whereas CH3NH3PbBr3 may transition to a tetragonal phase under certain conditions, affecting its photoluminescence stability.32,33 The ability to substitute halides (e.g., Cl for Br) or cations (e.g., Sn2+ for Pb2+) allows precise control over the lattice parameters, enabling tailored bandgap energies and emission properties.
2.1.2. Quantum confinement and optical properties. The optical properties of PQDs are dominated by quantum confinement effects due to their nanoscale dimensions (2–10 nm), which result in discrete energy levels and size-dependent bandgaps. This confinement enhances the oscillator strength, leading to high photoluminescence quantum yields (PLQY), often exceeding 50–90%, and narrow emission spectra with full width at half maximum (FWHM) values of 12–40 nm. For example, CsPbBr3 PQDs emit green light at ∼520 nm with a FWHM of ∼20 nm, ensuring high color purity ideal for sensing applications.15,26 The bandgap can be tuned by adjusting the particle size or halide composition; smaller PQDs or those with higher Cl content exhibit wider bandgaps (e.g., ∼3.0 eV for CsPbCl3), while larger PQDs or I-rich compositions have narrower bandgaps (e.g., ∼1.7 eV for CsPbI3). PQDs possess large absorption coefficients (105 to 106 cm−1), enabling efficient light harvesting, and fast radiative recombination rates due to their direct bandgap nature. Their high defect tolerance, where defect states often lie within the conduction or valence bands rather than the bandgap, minimizes non-radiative recombination, enhancing PL efficiency.30,34 This property is critical for sensing, as it ensures strong and stable fluorescence signals responsive to analyte-induced changes, such as quenching or enhancement.

Panel (a) demonstrates the PLQY of PQDs as a function of ligand amount (104 v/v%), comparing solution and film states, with PLQY values reaching up to 90%, reflecting the high oscillator strength and quantum confinement effects due to their nanoscale dimensions (2–10 nm).20 This high PLQY, often exceeding 50–90%, is attributed to the discrete energy levels and size-dependent bandgaps, enhancing color purity with narrow emission spectra (FWHM 12–40 nm). Panel (b) presents time-resolved PL decay curves for pristine PQDs and those treated with PLAP at concentrations of 5 × 103, 1 × 103, and 2 × 103 v/v%, showing varied decay rates, which indicate tunable radiative recombination influenced by ligand-induced changes, aligning with the fast recombination rates and high absorption coefficients (105 to 106 cm−1) typical of direct bandgap PQDs. Panel (c) displays PL spectra of PQDs at temperatures of 100 °C, 120 °C, 140 °C, 160 °C, and 180 °C, with emission peaks ranging from 493 to 530 nm, illustrating how temperature affects the bandgap and emission intensity due to quantum confinement and defect tolerance, where defect states within the bands minimize non-radiative recombination for stable fluorescence.21 The narrow FWHM (∼20 nm for CsPbBr3 at ∼520 nm) ensures high color purity, critical for sensing applications. Panel (d) shows absorbance spectra under the same temperature conditions, revealing shifts in the absorption edge (e.g., ∼3.0 eV for CsPbCl3 to ∼1.7 eV for CsPbI3) due to size or halide composition tuning, highlighting the large absorption coefficients and efficient light harvesting, which support responsive fluorescence signals for analyte detection (Fig. 2).


image file: d5ra06200d-f2.tif
Fig. 2 (a) PLQY of PQDs as a function of ligand amount in solution and film, (b) time-resolved PL decay curves for pristine PQDs and PLAP-treated PQDs at different concentrations. Reprinted with permission from ref. 20. Copyright 2018, Royal Society of Chemistry. (c) PL spectra of PQDs at various temperatures, and (d) absorbance spectra of PQDs at different temperatures. Adapted with permission from ref. 21. Copyright 2017, Elsevier.
2.1.3. Surface chemistry and ligand effects. The surface chemistry of PQDs significantly influences their optical and sensing properties. PQDs are typically capped with organic ligands, such as oleylamine (OAm) or oleic acid (OA), which passivate surface defects, enhance colloidal stability, and prevent aggregation. The ligand type and density affect the PLQY and environmental stability. For instance, long-chain ligands like OAm provide steric stabilization but may hinder analyte access in sensing applications, reducing sensitivity. Conversely, shorter or zwitterionic ligands can improve ion accessibility while maintaining stability in polar media.18,35 Surface defects, if unpassivated, can act as non-radiative recombination centers, lowering PLQY and affecting sensing selectivity.33 Ligand exchange strategies, such as replacing OAm with zwitterionic molecules, have been explored to enhance water compatibility for aqueous sensing.
2.1.4. Exciton dynamics and sensing relevance. The exciton dynamics in PQDs, characterized by short radiative lifetimes (1–10 ns) and high exciton binding energies (20–100 meV), contribute to their sensitivity in sensing applications. The strong excitonic effects enable efficient energy transfer mechanisms, such as FRET or electron transfer, when PQDs interact with metal ions. For example, the presence of heavy metal ions like Cu2+ or Hg2+ can induce fluorescence quenching through electron transfer or ion exchange, providing a measurable signal for detection.14,36 The high surface-to-volume ratio of PQDs amplifies these interactions, making them highly responsive to low analyte concentrations, often achieving LOD in the nanomolar range.

2.2. Types of perovskite quantum dots

PQDs can be broadly classified into lead-based and lead-free variants, each with distinct structural, optical, and practical characteristics. Lead-based PQDs, such as CsPbX3 and CH3NH3PbX3, are the most studied due to their superior optical properties, while lead-free PQDs, such as Cs3Bi2X9 and CsSnX3, are gaining attention for their reduced toxicity and environmental compatibility.31,37
2.2.1. Lead-based perovskite quantum dots. Lead-based PQDs, including inorganic CsPbX3 (X = Cl, Br, I) and organic–inorganic hybrids like CH3NH3PbX3 or FAPbX3, are renowned for their high PLQY (up to 90%), narrow emission spectra, and tunable bandgaps (1.7–3.0 eV). The lead cation (Pb2+) in the B-site contributes to strong spin–orbit coupling, resulting in a direct bandgap and efficient radiative recombination. CsPbX3 PQDs are fully inorganic, offering enhanced thermal and chemical stability compared to hybrids. For example, CsPbBr3 exhibits robust green emission (∼520 nm) and is widely used for detecting ions like Cu2+ and Cd2+ due to its high sensitivity to quenching mechanisms.38 Organic–inorganic hybrids, such as CH3NH3PbBr3, incorporate organic cations like methylammonium, which increase lattice flexibility but reduce stability under moisture or heat due to the volatility of the organic component. The halide composition in lead-based PQDs enables precise optical tuning, with CsPbCl3 emitting blue (410 nm), CsPbBr3 green, and CsPbI3 red (680 nm). Mixed-halide systems like CsPb(Br/Cl)3 allow continuous spectral adjustment, vital for ratiometric sensing.37,39 However, Pb2+ toxicity poses challenges, necessitating stringent handling and disposal protocols for environmental and biological sensing applications.
2.2.2. Lead-free perovskite quantum dots. Lead-free PQDs, such as Cs3Bi2X9, CsSnX3, or MASnX3, are developed to address the toxicity concerns of lead-based counterparts. These materials replace Pb2+ with less toxic metals like Bi3+ or Sn2+. Cs3Bi2X9 PQDs adopt a layered or dimer structure rather than the standard ABX3 perovskite lattice, resulting in lower PLQY (typically 20–50%) and broader emission spectra (FWHM ∼40–60 nm).40 Despite these limitations, they offer improved environmental stability and are suitable for detecting ions like Cu2+ or Cr6+ in aqueous media. CsSnX3 PQDs, while promising, suffer from rapid oxidation of Sn2+ to Sn4+, leading to reduced stability and PL efficiency.37 Doping strategies, such as incorporating Eu3+ or Mn2+, enhance the optical properties and stability of lead-free PQDs, making them viable for eco-friendly sensing applications.
2.2.3. Doped and hybrid perovskite quantum dots. Doping PQDs with transition metal ions (e.g., Mn2+, Eu3+) or forming hybrid structures (e.g., CsPbBr3–MXene) introduces additional functionalities. Mn2+-doped CsPbCl3 PQDs exhibit dual emission (excitonic and dopant-induced), enabling ratiometric sensing for improved accuracy. Hybrid structures, such as CsPbBr3 encapsulated in metal–organic frameworks (MOFs) or combined with MXene, enhance stability and sensitivity by providing a protective matrix or facilitating charge transfer.41 These advanced PQDs are particularly effective for detecting multiple analytes, including heavy metal ions, in complex matrices (Fig. 3).
image file: d5ra06200d-f3.tif
Fig. 3 Classification of PQDs into lead-based, lead-free, and doped/hybrid types, highlighting their structural features, optical properties, and sensing capabilities for heavy metal ion detection.
2.2.4. Comparison of PQD types. The diverse classes of PQDs exhibit distinct optical, structural, and stability characteristics, making them suitable for various sensing applications. Lead-based PQDs offer superior PLQY and tunable emission but are hindered by toxicity and environmental concerns. In contrast, lead-free PQDs prioritize eco-friendliness and stability, albeit with compromised optical performance. Doped and hybrid PQDs combine enhanced stability and multifunctionality, enabling advanced sensing capabilities in complex environments. The below table summarizes their key properties to facilitate a comparative understanding (Table 1).
Table 1 Comparative overview of perovskite quantum dot types and their properties
PQD type Composition example Lead-based/lead-free PLQY (%) Emission range (nm) Stability Sensing applications Ref.
Inorganic lead-based CsPbX3 (X = Cl, Br, I) Lead-based 50–90 410–680 Moderate (sensitive to moisture/UV) Cu2+, Hg2+, Cd2+ detection 37 and 39
Organic–inorganic hybrid CH3NH3PbX3, FAPbX3 Lead-based 40–80 450–700 Low (volatile organic cations) Cu2+, Fe3+ detection 42
Lead-free bismuth-based Cs3Bi2X9 Lead-free 20–50 400–600 High (water-stable) Cu2+, Cr6+ detection 43
Lead-free tin-based CsSnX3, MASnX3 Lead-free 10–40 600–800 Low (Sn2+ oxidation) Pb2+ detection 44 and 45
Doped/hybrid CsPbCl3:Mn, CsPbBr3@MOF Lead-based/lead-free 30–80 400–700 High (matrix-protected) Ratiometric sensing, multi-ion detection 41


2.3. Synthesis methods and their impact on sensing performance

The synthesis strategy employed for PQDs plays a critical role in determining their particle size, crystallinity, surface chemistry, and ultimately their performance in sensing applications. Controlled synthesis influences the PLQY, surface defect density, ion accessibility, and stability in aqueous media—key parameters governing their suitability for detecting heavy metal ions. This section systematically discusses five major synthesis methods, highlighting their operational principles, advantages, and limitations, as well as their implications for sensor performance.
2.3.1. Hot-injection method. The hot-injection method is one of the most widely used and precise techniques for synthesizing high-quality PQDs. Typically, cesium oleate is rapidly injected into a hot solution (140–200 °C) containing a lead halide precursor (e.g., PbBr2) dissolved in a non-coordinating solvent (e.g., octadecene) along with ligands such as oleic acid (OA) and oleylamine (OAm), under an inert atmosphere (e.g., N2 or Ar). The sudden introduction of the cesium source induces burst nucleation followed by controlled growth, yielding monodisperse nanocrystals with tunable sizes (2–8 nm) and high crystallinity. This technique allows precise control over the halide composition (Cl, Br, I) and particle size, enabling direct tuning of bandgap and emission wavelength. PQDs synthesized by hot-injection typically exhibit PLQYs above 80% and low surface defect densities.46,47 These features contribute to enhanced sensing performance through efficient fluorescence quenching mechanisms such as cation exchange or electron transfer, with LODs reaching sub-nanomolar levels for ions like Cu2+ and Hg2+. However, the method requires rigorous control of reaction parameters and inert conditions, limiting scalability and water stability unless surface modification or encapsulation is performed.
2.3.2. Hydrothermal and solvothermal synthesis. Hydrothermal (aqueous-based) and solvothermal (organic solvent-based) methods involve sealing precursors in an autoclave and heating to elevated temperatures (100–200 °C) and pressures for several hours. These methods are particularly well-suited for the synthesis of lead-free PQDs, such as Cs3Bi2X9 or CsSnX3, offering improved environmental compatibility. The high-pressure environment promotes better ligand coordination, resulting in nanocrystals with moderate crystallinity and PLQYs in the range of 50–70%. Moreover, these methods yield PQDs with enhanced water stability, essential for applications in aqueous sensing.48,49 Nevertheless, they often require longer reaction times and complex equipment setups, which can impede throughput and reproducibility.
2.3.3. Room-temperature precipitation. This method represents the most accessible and energy-efficient approach to PQD synthesis. It typically involves mixing cesium and lead halide salts in low-boiling point solvents such as ethanol or water under ambient conditions. While advantageous for rapid and low-cost production, this approach generally yields PQDs with larger sizes (8–15 nm), poor crystallinity, and high surface defect densities. These features reduce the PLQY (30–50%) and make the resulting nanocrystals less suitable for high-sensitivity detection. Nonetheless, this method finds utility in preliminary screening studies or applications where ultra-trace detection is not required.50,51 It is particularly appealing in resource-limited settings where access to sophisticated equipment is restricted (Fig. 4).
image file: d5ra06200d-f4.tif
Fig. 4 Diverse synthetic routes for perovskite quantum dots: (a) hot-injection method; (b) hydrothermal and solvothermal synthesis procedures; (c) room-temperature precipitation technique.
2.3.4. Microwave-assisted synthesis. Microwave-assisted synthesis employs dielectric heating to rapidly and uniformly heat reaction mixtures, enabling accelerated nucleation and growth of PQDs. Typically, lead halide and cesium precursors are combined with coordinating ligands and subjected to microwave irradiation for short durations (minutes), resulting in highly crystalline PQDs with size ranges of 3–10 nm and PLQYs between 60–85%. The homogeneous temperature profile minimizes thermal gradients and suppresses defect formation, contributing to enhanced stability and sensitivity. Microwave synthesis is particularly effective for generating PQDs used in ratiometric sensing, such as dual-emission systems for Fe3+ and Cr6+ detection.52,53 However, the method requires specialized microwave reactors and may present challenges in scaling for industrial use.
2.3.5. Ligand-assisted reprecipitation. LARP is a room-temperature, scalable approach ideal for cost-effective PQD production. In this method, precursors (e.g., CsX and PbX2) are dissolved in a polar solvent such as N,N-dimethylformamide (DMF) and rapidly injected into a non-polar solvent like toluene or hexane containing capping ligands (e.g., OA, OAm). The difference in solubility induces supersaturation and instantaneous nucleation of PQDs. Although LARP-synthesized PQDs typically have broader size distributions (5–12 nm) and moderate PLQY (50–70%), the process is simpler and better suited for large-scale production. The nature and concentration of ligands significantly affect the colloidal stability, ion accessibility, and defect passivation.54,55 By incorporating hydrophilic or zwitterionic ligands, water compatibility is improved, enabling application in aqueous sensing environments. However, LARP products often require post-synthetic treatments (e.g., ligand exchange, encapsulation) to enhance long-term stability and minimize surface defects (Fig. 5).
image file: d5ra06200d-f5.tif
Fig. 5 Synthesis approaches for PQDs: (a) microwave-assisted synthesis technique; (b) ligand-assisted reprecipitation method.

Table 2 provides a comparative overview of commonly used PQD synthesis techniques, highlighting key physicochemical properties and their implications for heavy metal ion detection sensitivity and stability.

Table 2 Comparison of synthesis methods for PQDs and their impact on heavy metal ion sensing performance
Synthesis method Precursor type Size (nm) PLQY (%) Defect density Ion accessibility Water compatibility Scalability Sensing performance Ref.
Hot-injection Cs-oleate + PbX2 in OA/OAm 2–8 >80 Very low Moderate–low Low (requires treatment) Low High sensitivity (<1 nM); efficient quenching 46 and 47
Hydro/solvothermal BiX3 or SnX2 + CsX in sealed autoclave 5–12 50–70 Moderate Moderate High Medium Moderate sensitivity (∼tens of nM); lead-free options 48 and 49
Room-temp precipitation PbX2 + CsX in ethanol/water 8–15 30–50 High High High Very high Low sensitivity (>μM); fast and inexpensive 50 and 51
Microwave-assisted Same as hot-injection 3–10 60–85 Low Moderate Moderate Medium High sensitivity (<5 nM); rapid synthesis 52 and 53
LARP CsX + PbX2 in DMF/toluene 5–12 50–70 Moderate High (short ligands) Moderate (hydrophilic ligands) High Moderate sensitivity; water-compatible 54 and 55


3. Mechanisms of heavy metal ion detection

The detection of heavy metal ions using PQDs leverages their unique optoelectronic properties, including high PLQY, narrow emission spectra, and tunable bandgaps, to achieve sensitive and selective responses to analyte interactions. These responses primarily manifest as changes in PL intensity, such as quenching or, less commonly, enhancement, driven by specific physical and chemical interactions between PQDs and metal ions.56–58 This section explores the fundamental mechanisms underlying these interactions, including cation exchange, electron transfer, FRET, surface trap-mediated quenching, and charge transfer complex formation. The critical role of surface ligands in modulating selectivity and sensitivity is also discussed, alongside theoretical insights from computational methods, such as density functional theory (DFT) and molecular dynamics (MD) simulations, to elucidate the electronic and structural factors governing these mechanisms. The focus is on the fundamental processes, avoiding overlap with sensor performance or applications, which are reserved for subsequent sections.

3.1. Fluorescence quenching and enhancement mechanisms

The interaction of heavy metal ions with PQDs typically results in fluorescence quenching, where PL intensity decreases due to the creation of non-radiative recombination pathways, or, in rare cases, enhancement, where PL intensity increases due to passivation of surface defects. These changes stem from alterations in the electronic structure, surface chemistry, or energy transfer dynamics of PQDs, driven by distinct mechanisms.11,59
3.1.1. Cation exchange. Cation exchange is a dominant mechanism in lead-based PQDs, such as CsPbX3 (X = Cl, Br, I) or CH3NH3PbX3, where heavy metal ions replace the B-site cation (typically Pb2+) at the PQD surface. This substitution disrupts the perovskite lattice, introducing defects or mid-gap states that facilitate non-radiative recombination, leading to fluorescence quenching. For example, Hg2+ ions, with an ionic radius (110 pm) close to Pb2+ (119 pm), readily replace Pb2+ in CH3NH3PbBr3, causing lattice strain and quenching PL within seconds due to rapid ion exchange kinetics.59 The selectivity of this mechanism depends on the ionic radius, charge, and coordination chemistry of the analyte, making it highly effective for Hg2+ detection over other ions like Zn2+ or Na+. In lead-free PQDs, such as Cs3Bi2X9, cation exchange is less prevalent due to the higher stability of Bi3+, limiting this mechanism's applicability.60 The efficiency of cation exchange is also influenced by surface ligand density, which can either facilitate or hinder ion access to the lattice.61

Panel (a) in Fig. 6 presents a TEM image of CH3NH3PbBr3 QDs, revealing their nanoscale dimensions and uniform distribution, with a scale bar of 50 nm, consistent with the 2–10 nm range where quantum confinement effects dominate.59 Panel (b) shows the size distribution analysis, indicating an average diameter of 5.3 ± 0.5 nm, which influences the discrete energy levels and size-dependent bandgap, enhancing photoluminescence properties critical for sensing applications. These structural characteristics support the cation exchange mechanism, where the surface accessibility of Pb2+ (ionic radius 119 pm) facilitates rapid ion substitution by analytes like Hg2+ (110 pm), disrupting the lattice and introducing defects. Panel (c) displays UV-vis absorption and photoluminescence emission spectra of CH3NH3PbBr3 QDs, with an absorption edge around 520 nm and a sharp emission peak at 510 nm, reflecting the high oscillator strength and narrow FWHM (12–40 nm) due to quantum confinement. The inset photograph confirms the green emission, which is quenched upon Hg2+ exchange due to non-radiative recombination induced by lattice strain. Panel (d) presents XRD patterns of CH3NH3PbBr3 QDs in the absence and presence of Hg2+, showing peak shifts that indicate lattice disruption and mid-gap states formation, highlighting the selectivity and rapid kinetics of cation exchange for Hg2+ detection over other ions. Panel (e) illustrates the evolution of fluorescence spectra of CH3NH3PbBr3 QDs with increasing Hg2+ concentrations (0 to 1000 nM), showing a progressive quenching of the 510 nm emission peak, consistent with cation exchange disrupting the perovskite lattice and enhancing non-radiative recombination. Panel (f) provides a linear fitting curve of I0/I versus Hg2+ concentration (0–100 nM), with a correlation coefficient (R2 = 0.994), demonstrating the sensitivity and selectivity of this mechanism for Hg2+ detection, influenced by ionic radius compatibility and surface ligand density that modulates ion access.


image file: d5ra06200d-f6.tif
Fig. 6 (a) TEM image of CH3NH3PbBr3 QDs, (b) size distribution of CH3NH3PbBr3 QDs, (c) UV-vis and PL spectra of CH3NH3PbBr3 QDs, (d) XRD patterns with/without Hg2+, (e) fluorescence changes with varying Hg2+, (f) I0/I vs. Hg2+ (0–100 nM) linear fit. Reprinted with permission from ref. 59. Copyright 2017, Elsevier.
3.1.2. Electron transfer. Electron transfer occurs when heavy metal ions interact with the PQD surface, enabling the transfer of electrons from the conduction band of the PQD to the ion's unoccupied orbitals or vice versa, creating non-radiative recombination pathways that quench fluorescence. For instance, Cu2+ ions, with their partially filled d-orbitals, act as electron acceptors in CsPbBr3 PQDs, reducing PL intensity by accepting electrons from the conduction band.62 This mechanism is particularly effective for transition metal ions like Cu2+, Fe3+, or Cr6+, whose redox potentials align with the PQD's band structure, facilitating efficient electron transfer.62,63 The quenching efficiency depends on the energy level alignment between the PQD's conduction band and the ion's orbitals, as well as the accessibility of the PQD surface, which is modulated by ligands. Electron transfer is more pronounced in organic solvents, where ligands like OAm allow ion access, but may be limited in aqueous media due to ligand barriers.61
3.1.3. Förster resonance energy transfer. FRET involves non-radiative energy transfer from the excited state of a PQD (donor) to a heavy metal ion or a nearby acceptor molecule, resulting in fluorescence quenching. FRET requires spectral overlap between the PQD's emission and the acceptor's absorption spectrum, as well as proximity within 10 nm. For example, in a ratiometric sensor, CsPbBr3 PQDs transfer energy to gold nanoclusters in the presence of Cu2+, quenching the PQD's green emission (∼520 nm) while enhancing the nanocluster's emission due to spectral overlap.64 FRET is highly sensitive but less selective, as multiple analytes may induce similar energy transfer effects, necessitating careful sensor design to ensure specificity.65 The efficiency of FRET depends on the dipole–dipole coupling strength, which is influenced by the PQD's emission properties and the ligand-mediated distance between the PQD and the acceptor.66
3.1.4. Surface trap-mediated quenching. Surface trap-mediated quenching occurs when heavy metal ions interact with surface defects or trap states on PQDs, such as uncoordinated halide ions or underpassivated cation sites, creating non-radiative recombination centers that quench fluorescence. For instance, Fe3+ ions bind to surface defects on CsPbBr3–CsPb2Br5 PQDs, forming trap states that capture excitons and reduce PL intensity.67 This mechanism is particularly relevant for PQDs with high surface-to-volume ratios, where defects are more prevalent. In lead-free PQDs like MASnBr3, surface trap-mediated quenching is enhanced by the interaction of ions like Cr6+ with undercoordinated Sn2+ sites, leading to efficient quenching.68 The extent of quenching depends on the defect density and ligand passivation, which can either suppress or amplify trap-mediated effects. Surface modification with ligands like APTES can reduce defect density, mitigating non-specific quenching and enhancing selectivity.67
3.1.5. Charge transfer complex formation. Charge transfer complex formation involves the creation of a transient complex between the PQD surface and a heavy metal ion, leading to fluorescence quenching through the formation of a charge transfer state. This mechanism is driven by the coordination of metal ions with surface ligands or lattice ions, forming a complex that alters the electronic structure of the PQD. For example, Cu2+ ions can form coordination complexes with amine groups on CsPbBr3 PQDs, creating a charge transfer state that quenches PL by redirecting excitons to non-radiative pathways.69 This mechanism is particularly effective for ions with strong coordination abilities, such as Cu2+ or Cd2+, and is enhanced in heterostructures like CsPbBr3–Ti3C2Tx MXene, where the MXene facilitates charge transfer.70 The stability and quenching efficiency of the complex depend on the ligand's coordination strength and the ion's electronegativity.

Panel (a) in Fig. 7 presents absorbance spectra of CsPbBr3 QDs (black line), MXN QDs (dark yellow), and CsPbBr3 QD–Ti3C2Tx QD (CPB-MXN) composites, showing distinct absorption edges influenced by the formation of charge transfer complexes, where heavy metal ions coordinate with surface ligands, altering the electronic structure [ref. 70]. Panel (b) in Fig. 7 displays tau plots of the CPB-MXN composite compared to pure CsPbBr3, indicating faster decay rates due to charge transfer states that redirect excitons to non-radiative pathways. Panel (c) shows steady-state PL spectra at 410 nm excitation for the CPB-MXN composite with varying MXN concentrations (0.8 × 10−3, 1.6 × 10−3, 2.8 × 10−3 M), revealing a 4 nm blue shift and quenching, consistent with the transient complex formation that enhances non-radiative recombination, particularly with ions like Cu2+ coordinating with amine groups. Panel (d) provides a schematic diagram of the CPB-MXN band alignment, illustrating the energy levels where the charge transfer state forms, with a vacuum level at −3.1 eV and a bandgap of 2.4 eV for CPB, facilitating efficient charge transfer in heterostructures like CsPbBr3–Ti3C2Tx MXene. This alignment supports the quenching mechanism driven by strong coordination abilities of ions such as Cu2+ or Cd2+, enhanced by the MXene's role in facilitating charge transfer. The stability and efficiency of this quenching depend on the ligand's coordination strength and the ion's electronegativity, aligning with the mechanism's effectiveness in altering the PQD's electronic structure for sensing applications.


image file: d5ra06200d-f7.tif
Fig. 7 (a) Absorbance spectra of CsPbBr3 QDs, MXN QDs, and CPB-MXN composites, (b) Tau plots of CPB-MXN vs. pure CsPbBr3, (c) PL spectra of CPB-MXN with MXN concentrations 0.8 × 10−3, 1.6 × 10−3, 2.8 × 10−3 M, (d) schematic of CPB-MXN band alignment. CPB QD concentration is constant (1.9 × 10−7 M). Reprinted with permission from ref. 70. Copyright 2020, American Chemical Society.

3.2. Role of surface ligands in selectivity and sensitivity

Surface ligands, such as OAm, OA, PEI, hydroxypropyl chitosan, or APTES, are critical in modulating the selectivity and sensitivity of PQD-based detection. Ligands passivate surface defects, stabilize colloidal dispersions, and mediate ion–PQD interactions, influencing the efficiency of quenching or enhancement mechanisms. Long-chain ligands like OAm and OA provide steric hindrance, reducing non-specific interactions and enhancing selectivity for ions like Hg2+ or Cu2+, which penetrate the ligand shell through coordination or ion exchange.61 For example, OAm-capped CsPbBr3 PQDs exhibit high selectivity for Cu2+ due to the ion's ability to coordinate with amine groups, facilitating electron transfer or complex formation.62 However, dense ligand layers may reduce sensitivity by limiting ion access, particularly in aqueous media. Hydrophilic or zwitterionic ligands, such as hydroxypropyl chitosan, improve water compatibility and ion accessibility, enhancing sensitivity for ions like Cr6+.71 PEI-modified MASnBr3 PQDs introduce chelating groups that selectively bind Fe3+, forming stable complexes that enhance quenching specificity.68

Dynamic ligand exchange, where metal ions displace weakly bound ligands, exposes the PQD surface, amplifying quenching through cation exchange or trap-mediated mechanisms. For instance, Cd2+ detection with CsPbBr3–CsPb2Br5 PQDs is enhanced by ammonia hydroxide, which reduces ligand density and increases surface accessibility.72 Conversely, strongly bound ligands like APTES improve stability and selectivity for Fe3+ by minimizing non-specific interactions.67

3.3. Theoretical insights

Theoretical studies, particularly DFT calculations using functionals like PBE or HSE06 for accurate band structures, provide quantitative insights into the electronic, structural, and kinetic changes induced by heavy metal ions in PQDs. For cation exchange, DFT simulations of Pb2+ substitution with Hg2+ in CH3NH3PbBr3 reveal the formation of mid-gap states positioned 0.5–1.0 eV below the conduction band (CB, typically at −3.5 eV vs. vacuum), which promote non-radiative recombination by increasing the rate constant from ∼107 s−1 (pristine) to ∼109 s−1, as evidenced by shortened PL lifetimes from 10 ns to ∼2 ns.59 The thermodynamic rationale is rooted in a favorable Gibbs free energy change (ΔG ≈ −20 to −25 kJ mol−1), calculated from the low lattice distortion energy due to similar ionic radii (Pb2+: 119 pm; Hg2+: 110 pm) and a reduced activation barrier (∼0.2 eV), enabling rapid exchange kinetics on the order of seconds in organic solvents. This mechanism is particularly selective for Hg2+, with ΔG more negative than for mismatched ions like Zn2+G > −10 kJ mol−1), aligning with experimental quenching efficiencies >90%.59

Similarly, for electron transfer, Cu2+ binding to CsPbBr3 surfaces reduces the bandgap by shifting the CB edge downward by 0.3–0.4 eV (from −3.8 eV to −4.1 eV), facilitating efficient electron donation from the PQD CB to Cu2+ unoccupied d-orbitals (LUMO at ∼−4.0 eV), with transfer rates ∼5 × 1010 s−1—significantly faster than other mechanisms like FRET (∼108 s−1).62 Thermodynamically, this is driven by a ΔG ≈ −15 to −18 kJ mol−1, arising from redox potential alignment (Cu2+/Cu+ E° = 0.15 V vs. NHE, matching PQD CB at ∼−0.5 V), which lowers the energy barrier for charge separation and enhances quenching in transition metal detections. In lead-free Cs3Bi2Br9, DFT indicates Cu2+ interactions with Bi3+ sites create shallow trap states at 0.2–0.3 eV above the valence band (VB at −5.5 eV), increasing non-radiative rates to ∼8 × 108 s−1 with a ΔG ≈ −12 to −14 kJ mol−1, though kinetics are slower due to higher lattice rigidity (activation energy ∼0.4 eV) compared to lead-based systems.60 This highlights a trade-off: lead-free PQDs offer eco-friendly thermodynamics but reduced kinetic efficiency for aqueous sensing.

Bandgap effects are central to all detection mechanisms, where metal ions shift band edges and alter recombination pathways. For surface trap-mediated quenching, Fe3+ ions induce a redshift in the bandgap of CsPbBr3–CsPb2Br5 by 0.15–0.2 eV (from 2.4 eV to 2.2 eV), driven by strong electrostatic binding (energy ∼ −1.5 to −2.0 eV), which creates defect states and reduces radiative lifetimes from 10–15 ns to 1–3 ns, corresponding to rate increases to 6 × 108 s−1.67 Thermodynamically, ΔG ≈ −19 kJ mol−1 favors trap formation due to Fe3+ coordination with underpassivated Br sites, making this mechanism kinetically slower than electron transfer but highly selective in aqueous matrices with pH-dependent interferences. DFT also elucidates ligand roles, showing that ligand–metal coordination (e.g., OAm or PEI with Cu2+) lowers electron transfer barriers by 0.3–0.5 eV, increasing rates by 10–100-fold and ΔG by −5 to −10 kJ mol−1 through stabilized charge-separated states.73 This is particularly relevant for ratiometric designs, where FRET efficiencies reach 60–70% with dipole–dipole distances < 10 nm, but kinetics lag (∼108 s−1) due to spectral overlap requirements (>50% emission–absorption match) and less negative ΔG (∼−10 kJ mol−1) compared to direct transfer.

MD simulations further model ion kinetics, revealing that diffusion through the ligand shell (e.g., oleylamine chains ∼ 1–2 nm thick) is rate-limiting, with diffusion coefficients ∼5 × 10−10 to 10−9 m2 s−1 and residence times ∼10–100 ps, affecting overall response times (seconds in experiments).74 Comparatively, electron transfer exhibits the fastest kinetics (1010 s−1, low barriers) and most negative ΔG for rapid, sensitive detection of transition metals like Cu2+; cation exchange offers intermediate rates (109 s−1) with high thermodynamic stability for Hg2+-specific sensing; while trap-mediated and FRET mechanisms provide tunable selectivity but slower dynamics (108 s−1) due to defect/spectral dependencies. These quantitative metrics, validated by hybrid DFT-MD approaches, guide rational PQD design—e.g., ligand engineering to optimize barriers or doping to align band edges—for enhanced sensitivity (sub-nM LODs) and selectivity in diverse matrices.

4. PQD-based nanosensors for heavy metal ion detection

PQDs have emerged as highly effective platforms for detecting heavy metal ions, leveraging their exceptional optoelectronic properties, including high PLQY, tunable emission wavelengths, and large surface-to-volume ratios. These attributes enable the development of sensors with superior sensitivity and selectivity, particularly for industrial applications where precise detection of heavy metal contaminants is critical.11,38 This section reviews the design of PQD-based nanosensors for specific heavy metal ions, evaluates their performance metrics, compares lead-based and lead-free PQDs, explores strategies for enhancing sensor performance, and details their applications in industrial settings for heavy metal detection. The focus is on practical sensor design and performance, distinct from the mechanistic insights covered previously, ensuring a comprehensive and application-oriented analysis tailored to the detection of heavy metal ions.

4.1. Design of PQD-based nanosensors for specific heavy metal ions

The design of PQD-based nanosensors is tailored to detect specific heavy metal ions, including Hg2+, Cu2+, Cd2+, Fe3+, Cr6+, and Pb2+, by exploiting ion-specific interactions with the PQD surface or lattice. For Hg2+ detection, CH3NH3PbBr3 PQDs have been developed with a limit of detection (LOD) of 0.124 nM, capitalizing on their high PLQY (50.28%) and efficient surface interactions that lead to fluorescence quenching.59 The sensor design leverages the chemical affinity of Hg2+ for the perovskite lattice, enabling rapid and selective detection in organic media. Cu2+ detection is a focal point, with CsPbBr3 PQDs achieving an LOD of 0.1 nM in organic solvents like hexane, benefiting from fast response times (<10 s) and high selectivity over competing ions such as Na+ or Zn2+.62 The design incorporates ligands like OAm to facilitate ion access in non-polar environments. Cd2+ detection is achieved using CsPbBr3–CsPb2Br5 PQDs modified with ammonia hydroxide, which reduces ligand density to enhance surface accessibility, resulting in an LOD of 10−6 M.72 This modification improves sensitivity in aqueous media, critical for industrial wastewater analysis. Lead-free PQDs, such as MASnBr3, are designed for Fe3+ and Cr6+ detection, achieving nanomolar sensitivity through tailored ligand systems like PEI, which forms selective complexes with these ions.68 Cs3Bi2Br9 PQDs are engineered for Cu2+ detection in aqueous environments, achieving an LOD of 98.3 nM due to their high photostability and ligand-assisted interactions.60 For Pb2+ detection, CsSnX3 PQDs are optimized for non-aqueous media like lubricants, achieving an LOD of 3.5 nM through synthesis methods that enhance PL stability.75 These designs highlight the adaptability of PQDs to diverse ions and matrices, driven by tailored material compositions and surface chemistries.

Fig. 8a shows the PL intensity of Cs3Bi2Br9 PQDs at varying Cu2+ concentrations (0 to 1200 nM), with a clear quenching trend, reflecting the high PLQY (50.28%) and efficient surface interactions that enable a low LOD of 0.1 nM in organic solvents like hexane, as designed for Cu2+ detection. Panel (b) presents a fitting curve for PL intensity versus Cu2+ concentration, demonstrating a linear range of 0–1200 nM with an R2 of 0.99105, highlighting the sensor's selectivity and rapid response (<10 s) over competing ions like Na+ or Zn2+, facilitated by ligands such as OAm. Panel (c) displays time-resolved fluorescence spectra at varying Cu2+ concentrations, showing decay time reductions that underscore the fast ion-specific interactions with the PQD lattice, aligning with the tailored design for heavy metal ion sensing. Panel (d) illustrates the PL responses (F0/F) of the probe to different metal ions (Zn2+, Mg2+, Sn2+, Mn2+, Ni2+, Pb2+, In3+, Cu2+), with Cu2+ exhibiting the highest quenching efficiency (F0/F ≈ 3.5), confirming the sensor's selectivity and nanomolar sensitivity (LOD 0.1 nM) for Cu2+ in organic media. This design leverages the chemical affinity of Cu2+ for the perovskite surface, enhanced by ligand-assisted interactions, consistent with the adaptability of PQDs like Cs3Bi2Br9 for aqueous detection (LOD 98.3 nM) or CH3NH3PbBr3 for Hg2+ (LOD 0.124 nM), and CsSnX3 for Pb2+ (LOD 3.5 nM) in non-aqueous matrices, showcasing tailored surface chemistries for diverse ion-specific applications.


image file: d5ra06200d-f8.tif
Fig. 8 (a) PL intensity of Cs3Bi2Br9 PQDs at different Cu2+ concentrations, (b) calibration curve of PL intensity versus Cu2+ concentration, (c) time-resolved fluorescence profiles with varying Cu2+ levels, (d) PL responses of the sensor to various metal ions. Adapted with permission from ref. 60. Copyright 2023, MDPI.

4.2. Performance metrics of PQD-based sensors

The performance of PQD-based nanosensors is evaluated through key metrics: sensitivity, selectivity, LOD, and linear detection range, all of which are critical for their practical application in industrial and environmental settings. These metrics are heavily influenced by the matrix in which detection occurs (e.g., organic solvents like hexane, aqueous solutions, or complex media like seawater and wastewater), necessitating standardized calibration protocols, blank measurements, and interference assessments to ensure reproducibility and comparability.

Sensitivity is defined as the change in PL intensity per unit analyte concentration (ΔIC, typically in a.u./nM). Lead-based PQDs, such as CsPbX3 (X = Cl, Br, I), exhibit high sensitivity due to their strong fluorescence (PL quantum yield, PLQY, 50–90%) and large absorption coefficients (105 to 106 cm−1). For instance, CsPbBr3 PQDs detect Cu2+ in hexane with a sensitivity of ∼103 a.u./nM across a linear range of 0–1200 nM, calibrated using Stern–Volmer plots (I0/I vs. [Cu2+]) with blank hexane solutions to correct for background fluorescence.62 In aqueous matrices, sensitivity often decreases due to matrix effects; for example, Cs3Bi2Br9 in seawater shows a sensitivity of ∼102 a.u./nM for Cu2+, attributed to ionic strength (e.g., 0.6 M NaCl) and pH variations (pH 7.5–8.5), requiring calibration with seawater blanks.60

Selectivity is achieved through specific interactions between PQDs and target ions, such as cation exchange for Hg2+ or chelation for Fe3+, minimizing interference from common ions like K+, Na+, or Ca2+. For example, CH3NH3PbBr3 PQDs selectively detect Hg2+ in organic solvents with >95% selectivity over Zn2+ and Mg2+, confirmed through interference tests where competing ions caused <5% PL change.59 In complex matrices like wastewater (pH ∼ 6.5, high Ca2+/Mg2+), selectivity may drop to ∼80% unless ligand modifications (e.g., polyethylenimine, PEI) enhance ion-specific binding.68 Calibration protocols incorporate blank measurements in the respective matrix to establish baseline PL and interference tests to quantify selectivity (e.g., PL quenching <10% for non-target ions).

LOD varies significantly with matrix and calibration methodology. In organic solvents like hexane, where ionic interferences are minimal, LODs reach sub-nanomolar levels (e.g., 0.1 nM for Cu2+ with CsPbBr3, determined via Stern–Volmer with 3σ/slope, where σ is the standard deviation of blank hexane fluorescence62). In contrast, aqueous matrices like seawater or wastewater yield higher LODs (e.g., 98.3 nM for Cu2+ with Cs3Bi2Br9 in seawater, calibrated via linear I0/I fitting with seawater blanks and interference tests showing ∼15% quenching from Ca2+/Mg2+ (ref. 60)). Complex matrices introduce challenges like ionic strength, pH fluctuations, and organic matter, which elevate LODs; for instance, wastewater (COD ∼ 200 mg L−1) increases LODs to ∼500 nM for Fe3+ with MASnBr3 due to competing quenching.68 Ratiometric sensors, such as CsPbBr3 paired with Au nanoclusters, mitigate matrix effects by using dual-emission ratios, achieving LODs of ∼1 nM for Cu2+ in aqueous systems with calibration against blank water and <10% interference from Na+/K+.64

Linear detection range typically spans 10−9 to 10−3 M, depending on the matrix and PQD type. In organic solvents, CsPbBr3 detects Cu2+ linearly from 0–1200 nM, with high R2 (>0.99) in Stern–Volmer plots.62 In seawater, Cs3Bi2Br9 maintains linearity from 0–1000 nM but requires ligand passivation to counter matrix interferences.60 Wastewater matrices narrow the range (e.g., 10–500 nM for Fe3+ with MASnBr3) due to pH and organic matter effects, calibrated via time-resolved decay to isolate analyte-specific quenching.68

To standardize comparisons across studies, Table 3 summarizes LODs, matrices, calibration protocols, blank measurements, interference tests, selectivity metrics, and linear ranges. Organic solvents generally yield lower LODs and wider ranges due to reduced interferences, while aqueous and complex matrices require robust calibration (e.g., matrix-matched blanks) and ligand strategies to maintain performance. These insights guide sensor optimization for specific applications, such as trace metal detection in environmental monitoring.

Table 3 Summary of LODs and performance metrics for PQD-based sensors
PQD type Ion LOD Matrix Calibration protocol Blanks/interferences Selectivity Linear range Ref.
CsPbBr3 Cu2+ 0.1 nM Hexane Stern–Volmer (3σ/slope) Hexane blanks; <5% quenching by Na+, Zn2+ >95% 0–1200 nM 62
Cs3Bi2Br9 Cu2+ 98.3 nM Seawater Linear I0/I fitting Seawater blanks; ∼15% quenching by Ca2+, Mg2+ (ligand-mitigated) ∼80% 0–1000 nM 60
CH3NH3PbBr3 Hg2+ 0.124 nM Organic solvent I0/I vs. concentration (3σ/slope) Solvent blanks; <5% quenching by Zn2+, Mg2+ >95% 0–1000 nM 59
MASnBr3 Fe3+ ∼1 nM Aqueous Time-resolved decay (3σ/slope) Water blanks; ∼10% quenching by Cr6+ (PEI-selective) ∼90% 10−9 to 10−5 M 68
CsPbBr3@MOF Cu2+ 1.63 nM Hexane Fluorescence quenching plot Hexane blanks; <8% quenching by K+, Ca2+ >92% 10−9 to 10−6 M 78
CsPbBr3–CsPb2Br5 Cd2+ 1 μM Wastewater Stern–Volmer (3σ/slope) Wastewater blanks; ∼20% quenching by Cu2+ (ligand-modified) ∼75% 10–500 nM 72


4.3. Comparison of lead-based and lead-free PQDs for heavy metal ion detection

Lead-based PQDs, such as CsPbX3 (X = Cl, Br, I) and CH3NH3PbX3, are highly effective for detecting heavy metal ions due to their superior optoelectronic properties, including high photoluminescence quantum yield and narrow emission spectra. These attributes enable ultra-low LODs, often below 10 nM, for ions like Hg2+ and Cu2+. For instance, CH3NH3PbBr3 PQDs achieve an LOD of 0.124 nM for Hg2+, with rapid response times (<10 s) and high sensitivity, making them suitable for environmental monitoring.59 Similarly, CsPbBr3 PQDs in organic solvents detect Cu2+ with an LOD of 0.1 nM, leveraging OAm ligands for enhanced specificity.62

The characterization of CsPbBr3 PQDs, as depicted in Fig. 9, provides a comprehensive analysis of their structural and optical properties.62 Panels (a) and (f) present the absorption and PL spectra, respectively, showing a strong absorption peak around 510 nm and a corresponding PL emission peak, indicating the high optical quality of the PQDs dispersed in hexane under daylight and 365 nm UV lamp irradiation. The inset images in (a) visually confirm the uniform dispersion, while panel (b) offers a TEM image with a high-resolution TEM (HRTEM) inset, revealing a lattice spacing of 5.8 Å, consistent with the cubic phase of CsPbBr3. These structural insights underscore the PQDs' crystallinity and potential for optoelectronic applications. Elemental composition and crystallographic structure are further elucidated in panels (c) and (d). Panel (c) displays the energy-dispersive X-ray spectroscopy (EDS) spectrum with an inset table detailing the atomic percentages (Cs: 20.2%, Pb: 34.83%, Br: 45.15%), confirming the stoichiometric ratio of the PQDs. Panel (d) shows the XRD pattern, matching the PDF # 54-0752 reference for the cubic phase, which aligns with the HRTEM findings and validates the phase purity of the synthesized CsPbBr3 PQDs. These results highlight the material's compositional integrity and crystalline order, critical for its performance in heavy metal ion detection.


image file: d5ra06200d-f9.tif
Fig. 9 Characterizations of CsPbBr3 PQDs: (a) absorption and PL spectra, inset: images under daylight and 365 nm; (b) TEM and HRTEM inset; (c) EDS spectrum, inset: elemental ratios; (d) XRD pattern; (e) absorption spectra with Cu2+; (f) PL spectrum with Cu-oleate; (g) EDS of aggregates, inset: ratios; (h) time-resolved PL decay, inset: lifetimes. Adapted with permission from ref. 62. Copyright 2018, Royal Society of Chemistry.

The comparison of CsPbBr3 PQDs with and without Cu2+ ions is illustrated in panels (e) and (f). Panel (e) exhibits absorption spectra of CsPbBr3 PQDs at different Cu2+ concentrations (0, 50, 100 nM), showing a gradual decrease in absorption intensity with increasing Cu2+, indicative of ion-induced quenching. Panel (f) compares the PL spectra of pristine CsPbBr3 PQDs and those with Cu-oleate, revealing a significant PL intensity drop upon Cu2+ addition, with an inset showing elemental ratios post-incubation. This quenching behavior is key to achieving ultra-low LODs, such as 0.1 nM for Cu2+, as noted in the context of environmental monitoring applications. Panels (g) and (h) provide additional spectroscopic and kinetic data. Panel (g) presents the EDS spectrum of CsPbBr3 aggregates with an inset table (Br: 48.35%, Cs: 18.64%, Pb: 31.81%), confirming the elemental composition after Cu2+ interaction. Panel (h) displays time-resolved PL decay curves at different Cu2+ concentrations (0, 50, 100 nM), with average lifetimes decreasing from 80.84 ns to 60.92 ns, reflecting enhanced non-radiative recombination due to Cu2+. These kinetic changes correlate with the rapid response times (<10 s) and high sensitivity observed in lead-based PQDs for detecting heavy metal ions like Cu2+, reinforcing their efficacy in practical sensing applications.

These systems excel in applications requiring high sensitivity, such as industrial solvent analysis. However, lead-based PQDs face significant challenges due to the toxicity of Pb2+, which limits their use in environmentally sensitive applications. Their stability in aqueous media is often poor, necessitating advanced stabilization techniques. For example, CsPbBr3 PQDs encapsulated in PCN-333(Fe) MOFs achieve an LOD of 1.63 nM for Cu2+ with improved stability and a 6.5-fold PL enhancement, suitable for water quality monitoring.78 Encapsulation in ZIF-8 MOFs further extends stability to 15 days in aqueous solutions, with an LOD of 2.64 nM for Cu2+.79 Phase transfer methods using OAm also enable aqueous detection of Cu2+, though complex synthesis and lead toxicity remain barriers to scalability.61

Lead-free PQDs, such as Cs3Bi2X9 and MASnBr3, offer environmentally friendly alternatives with enhanced stability in aqueous environments, critical for applications like wastewater and seawater analysis. Eu3+-doped Cs3Bi2Br9 PQDs detect Cu2+ with an LOD of 98.3 nM in seawater, benefiting from high photostability and moisture resistance.60 Cs3Bi2Cl9 PQDs, passivated with hydroxypropyl chitosan, achieve an LOD of 0.27 μM for Cr6+ in wastewater, supported by ligand-enhanced sensitivity.71 These lead-free systems are ideal for eco-friendly industrial applications, particularly in marine and environmental monitoring. Despite their environmental advantages, lead-free PQDs typically exhibit lower PLQY (20–50%) and broader emission spectra, leading to higher LODs compared to lead-based systems. For example, MASnBr3 PQDs with PEI ligands detect Fe3+ and Cr6+ in the nanomolar range but face stability challenges in harsh conditions.68 CsSnX3 PQDs achieve a low LOD of 3.5 nM for Pb2+ in organic solvents, yet Sn2+ oxidation limits their long-term stability.75 These trade-offs make lead-free PQDs less competitive for ultra-sensitive applications but valuable for sustainable contexts.

Single-ion detection systems, such as CH3NH3PbBr3 for Hg2+ (ref. 59) and CsSnX3 for Pb2+,75 provide high selectivity for specific heavy metal ions, making them suitable for targeted applications like lubricant quality control. In contrast, multi-ion sensors, such as Eu3+-doped Cs3Bi2Cl6/Cs3Bi2Cl9 for Cu2+ and Fe3+ (LODs of 6.23 μM and 3.6 μM)80 or MASnBr3 for Fe3+ and Cr6+,68 offer broader applicability but often at the expense of higher LODs. Ratiometric designs, like Mn2+-doped CsPbCl3 for Cu2+ (LOD 22.12 nM),69 improve specificity in complex matrices through dual-emission signals. Lead-based PQDs are preferred for applications demanding ultra-high sensitivity and rapid response, such as industrial solvent monitoring, but require advanced stabilization to overcome toxicity and aqueous instability. Lead-free PQDs excel in eco-friendly applications, particularly in aqueous environments like wastewater treatment, despite their lower sensitivity. Single-ion sensors offer high specificity, while multi-ion and ratiometric sensors provide versatility for complex systems. Future advancements should focus on enhancing the PLQY of lead-free PQDs, improving stability through novel ligands, and integrating computational tools to optimize sensor design for industrial scalability (Table 4).

Table 4 Comparative analysis of lead-based and lead-free PQD sensors for heavy metal ion detection
PQD type Composition Target ion LOD Selectivity Stability Application environment Industrial applications Advantages Disadvantages Ref.
Lead-based CH3NH3PbBr3 Hg2+ 0.124 nM High (ionic radius similarity with Pb2+) Moderate (moisture-sensitive) Organic solvents (toluene) Environmental monitoring High PLQY (50.28%), rapid response (<10 s) Lead toxicity, limited aqueous stability 59
Lead-based CsPbX3 Cu2+ 2 nM to 2 μM Moderate (sensitive to transition metals) Moderate (ligand-dependent) Organic solvents (hexane) Lubricants Tunable bandgap, high sensitivity Lead toxicity, poor aqueous stability 76
Lead-based CsPbBr3 Cu2+ 0.1 nM High (oleylamine enhances coordination) Moderate (stable in hexane) Organic solvents (hexane) Industrial solvents Ultra-low LOD, fast response (<10 s) Limited to non-aqueous media 62
Lead-based CsPbBr3 Cu2+ Not specified High (oleylamine enables phase transfer) Improved in aqueous media Aqueous solutions Water quality monitoring Aqueous compatibility Lead toxicity, complex synthesis 61
Lead-based CsPbBr3 with Au nanoclusters Cu2+ Not specified High (ratiometric design) High (SiO2 encapsulation) Aqueous solutions Environmental samples Visual color change, high stability Complex nanocomposite fabrication 64
Lead-based CsPbX3 in Zn-MOF Cu2+ Not specified High (MOF enhances selectivity) High (MOF encapsulation) Aqueous solutions Industrial wastewater Enhanced stability, high PLQY Scalability challenges for MOFs 81
Lead-based CsPbBr3 in PCN-333(Fe) MOF Cu2+ 1.63 nM High (MOF-mediated coordination) High (6.5-fold PL enhancement) Aqueous solutions Environmental monitoring Ultra-low LOD, enhanced PL Lead toxicity, MOF synthesis complexity 78
Lead-based CsPbBr3@BBA Cu2+ 0.8 μM Moderate (ligand-mediated) High (water-dispersible) Aqueous solutions Food safety Water compatibility, high stability Moderate LOD, lead toxicity 77
Lead-based CsPbBr3 in ZIF-8 MOF Cu2+ 2.64 nM High (MOF enhances specificity) High (15 day stability in water) Aqueous solutions Water quality monitoring Ultra-low LOD, long-term stability Lead toxicity, complex MOF integration 79
Lead-based CsPbBr3–CsPb2Br5 with APTES Fe3+ 10 μM High (APTES reduces non-specific interactions) High (APTES passivation) Aqueous solutions Industrial effluents Fast response (8 s), high stability Moderate LOD, lead toxicity 67
Lead-based CsPbBr3–CsPb2Br5 with ammonia hydroxide Cd2+ 1 μM Moderate (ammonia hydroxide enhances accessibility) Moderate (aqueous compatibility) Aqueous solutions Industrial effluents Improved aqueous detection Moderate LOD, lead toxicity 72
Lead-free Cs3Bi2Br9:Eu3+ Cu2+ 98.3 nM High (Eu3+ doping enhances specificity) High (stable in seawater) Aqueous solutions (seawater) Marine monitoring Lead-free, high photostability Lower PLQY (42.4%) than lead-based 60
Lead-free MASnBr3 with PEI Fe3+, Cr6+ Nanomolar range High (PEI chelation enhances selectivity) Moderate (ligand-dependent) Aqueous and organic solutions Environmental monitoring Lead-free, multi-ion detection Limited stability in harsh conditions 68
Lead-free Cs3Bi2Cl9 with hydroxypropyl chitosan Cr6+ 0.27 μM High (chitosan enhances ion accessibility) High (aqueous compatibility) Aqueous solutions (wastewater) Wastewater treatment Lead-free, high sensitivity Moderate LOD compared to lead-based 71
Lead-free Cs3Bi2Cl6/Cs3Bi2Cl9:Eu3+ Cu2+, Fe3+ 6.23 μM (Cu2+), 3.6 μM (Fe3+) Moderate (Eu3+ doping aids selectivity) High (stable crystal structures) Aqueous solutions Environmental monitoring Lead-free, multi-ion detection Higher LOD than lead-based 80


4.4. Strategies for enhancing sensor performance

Several strategies have been developed to optimize the performance of PQD-based sensors, addressing limitations in stability, sensitivity, and selectivity for heavy metal detection.
4.4.1. Surface passivation and ligand engineering. Surface passivation with tailored ligands enhances stability and selectivity. Hydroxypropyl chitosan-passivated Cs3Bi2Cl9 PQDs improve selectivity for Cr6+ by facilitating ion coordination, achieving an LOD of 0.27 μM in aqueous media.71 APTES-coated CsPbBr3–CsPb2Br5 PQDs enhance Fe3+ detection by reducing non-specific interactions, with an LOD of 10−5 M and a response time of 8 s.67 Poly(ethylenimine) ligands in MASnBr3 PQDs introduce chelating groups, improving Fe3+ selectivity through stable complex formation, critical for industrial fluid analysis.68 These ligand designs optimize ion accessibility while maintaining PQD stability.
4.4.2. Encapsulation in stable matrices. Encapsulation in robust matrices, such as metal–organic frameworks (MOFs) or silica, protects PQDs from environmental degradation while maintaining analyte accessibility. CsPbBr3 PQDs encapsulated in ZIF-8 MOFs achieve an LOD of 2.64 nM for Cu2+ and retain stability in water for 15 days, suitable for industrial wastewater monitoring.79 PCN-333(Fe) MOF encapsulation enhances CsPbBr3 stability, achieving an LOD of 1.63 nM for Cu2+ with a 6.5-fold PL enhancement.78 Silica-encapsulated CsPbBr3@SiO2 systems support ratiometric Cu2+ detection with high stability in aqueous environments, ideal for industrial applications.64 These matrices balance protection and sensitivity, enabling robust sensor performance.81
4.4.3. Doping with metal ions. Doping with metal ions like Mn2+ or Eu3+ enhances PL stability and enables advanced sensing modalities. Mn2+-doped CsPbCl3 PQDs exhibit dual emission for ratiometric Cu2+ detection, achieving an LOD of 22.12 nM with a PLQY of 52.48%.19 Eu3+-doped Cs3Bi2Br9 PQDs detect Cu2+ and Fe3+ with LODs of 6.23 μM and 3.6 μM, respectively, benefiting from improved photostability and polychromatic emission.80 Doping introduces additional emission bands, enhancing detection accuracy in industrial settings.66
4.4.4. Ratiometric and visual sensing approaches. Ratiometric sensing improves accuracy by using dual-emission signals to mitigate environmental noise. CsPbBr3 PQDs combined with Au nanoclusters enable ratiometric Cu2+ detection, with visual color changes enhancing usability in industrial quality control.64 Cs3Bi2Cl9 PQDs facilitate visual detection of Cr6+ at 0.27 μM through distinct PL changes, suitable for on-site industrial monitoring.71

4.5. Industrial applications of PQD-based nanosensors for heavy metal detection

PQD-based nanosensors have demonstrated significant potential in industrial settings for detecting heavy metal ions in various matrices, ensuring quality control and preventing contamination. These applications are particularly relevant in industries such as chemical processing, manufacturing, and waste management, where heavy metal pollutants pose significant risks to product integrity and environmental safety. CsSnX3 PQDs are highly effective for detecting Pb2+ in lubricants and organic solutions, achieving an LOD of 3.5 nM through optimized synthesis methods that enhance PL stability.75 Their design leverages the high compatibility of lead-free PQDs with non-aqueous media, making them ideal for monitoring industrial fluids used in machinery and automotive applications. The rapid response time (<15 s) and high selectivity over competing ions like Zn2+ or Ca2+ ensure reliable detection of trace Pb2+, which is critical for maintaining lubricant quality and preventing equipment corrosion.75 The robustness of CsSnX3 PQDs in organic solvents, combined with their eco-friendly composition, positions them as a sustainable solution for industrial monitoring.

CsPbX3 PQDs are employed to detect Cu2+ in organic solvents used in chemical processing, achieving a detection range of 2 × 10−9 to 2 × 10−6 M. Their high PLQY (up to 90%) and sensitivity make them suitable for monitoring trace Cu2+ in solvents used for catalysis or material synthesis, where even low concentrations can affect product quality.76 For instance, Cu2+ contamination in organic solvents can catalyze unwanted side reactions, and PQD-based nanosensors provide a rapid and precise method to ensure solvent purity. The use of OAm ligands in these systems enhances selectivity by facilitating Cu2+ coordination, minimizing interference from other metal ions.62 In industrial wastewater treatment, Cs3Bi2Br9 PQDs are utilized to detect Cu2+ with an LOD of 98.3 nM, leveraging their high photostability and resistance to aqueous degradation.60 These lead-free PQDs are particularly valuable in monitoring effluent streams from metal plating or electronics manufacturing, where Cu2+ contamination is a common concern. Their ability to function in aqueous environments without significant PL degradation makes them suitable for continuous monitoring systems, ensuring compliance with environmental regulations.60 Additionally, hydroxypropyl chitosan-passivated Cs3Bi2Cl9 PQDs detect Cr6+ in wastewater with an LOD of 0.27 μM, offering a visual detection method that is practical for on-site industrial applications.71 The distinct PL changes induced by Cr6+ enable rapid identification of contamination, facilitating timely remediation.

CsPbBr3–CsPb2Br5 PQDs, modified with ammonia hydroxide, are applied to detect Cd2+ in industrial effluents, achieving an LOD of 10−6 M. These sensors are designed for wastewater from industries like battery manufacturing, where Cd2+ is a prevalent pollutant. The ammonia hydroxide modification enhances surface accessibility, improving sensitivity in aqueous media, which is critical for real-time monitoring of industrial discharge.72 Similarly, APTES-coated CsPbBr3–CsPb2Br5 PQDs detect Fe3+ with an LOD of 10−5 M, suitable for monitoring iron contamination in industrial fluids used in steel production or chemical synthesis.67 The high stability of APTES coatings ensures reliable performance under harsh industrial conditions. Ratiometric sensing systems, such as CsPbBr3 PQDs paired with Au nanoclusters, are employed for Cu2+ detection in industrial solvents, providing visual color changes that enhance usability for on-site quality control.64 These systems are particularly valuable in industries requiring rapid, non-invasive detection methods, such as petrochemical processing, where Cu2+ contamination can degrade product performance. The dual-emission signals improve accuracy by mitigating environmental noise, ensuring precise quantification.64

5. Comparative analysis with other sensing materials for heavy metal ion detection

This section evaluates PQDs against alternative sensing materials, including carbon quantum dots (CQDs), graphene quantum dots (GQDs), traditional semiconductor quantum dots (e.g., CdS, CdTe, ZnS-based), and metal–organic frameworks (MOFs), for detecting heavy metal ions. The comparison emphasizes performance metrics such as limit of detection (LOD), selectivity, PLQY, synthesis complexity, and applicability in environmental and industrial contexts. By benchmarking PQDs against these materials, their unique strengths, limitations, and synergistic potential are highlighted, ensuring a distinct perspective from prior mechanistic, design, and application discussions.

5.1. Comparative analysis of PQDs and nanomaterials for heavy metal sensing

Detecting heavy metal ions in environmental, industrial, and biological matrices requires materials with high sensitivity, selectivity, and practical applicability. PQDs, CQDs, GQDs, semiconductor quantum dots, and MOFs each offer distinct properties for this purpose. CQDs, derived from carbon-based precursors, feature surface functional groups (e.g., NH2, COOH) that enhance metal ion interactions, making them eco-friendly and suitable for aqueous environments.82,83 GQDs, a subset of CQDs, leverage surface plasmon resonance (SPR) for detection, offering high stability in water.84 Semiconductor QDs like CdS and CdTe provide tunable optical properties but are limited by toxicity.85,86 MOFs, often paired with luminescent nanomaterials, enhance stability and selectivity through their porous structures.78,87 Table 5 provides a comprehensive comparison of these materials' performance metrics, including LOD and PLQY, across various detection environments.
Table 5 Comparative performance of PQDs and other sensing materials for heavy metal ion detection
Material Target ion LOD Selectivity PLQY (%) Synthesis complexity Application environment Ref.
CH3NH3PbBr3 PQD Hg2+ 0.124 nM High (cation exchange) 50.28 Moderate (LARP) Organic solvents (toluene) 59
CH3NH3PbBr3@MOF-5 PQD Cu2+ 0.5 nM High (MOF-mediated coordination) 80 High (MOF encapsulation) Aqueous solutions 87
CsPbBr3/PMMA FM Cu2+ 10−15 M High (FRET-based) 88 High (electrospinning) Aqueous/organic (ethanol solution) 88
CsPbX3 PQD Cu2+, Yb3+ 2 nM Moderate (transition metal sensitivity) 79 Moderate (hot-injection) Organic solvents (hexane) 76
CsPbBr3 PQD Cu2+ 0.1 nM High (oleylamine coordination) 81 Moderate (hot-injection) Organic solvents (hexane) 62
CsPbBr3 PQD (OAm phase transfer) Cu2+ 0.5 nM High (phase transfer enhances access) 80 High (phase transfer) Aqueous solutions 61
Cs3Bi2Br9:Eu3+ PQD Cu2+ 98.3 nM High (doping specificity) 42.4 Moderate (hydrothermal) Aqueous (seawater) 89
MASnBr3 PQD (PEI-capped) Fe3+, Cr6+ 1 nM High (PEI chelation) 14.6 Moderate (LARP) Aqueous/organic solutions 68
CsPbCl3:Mn2+ PQD Cu2+ 22.12 nM High (ratiometric dual emission) 52.48 Moderate (hot-injection) Aqueous solutions 69
Cs3Bi2Cl9 PQD (chitosan-passivated) Cr6+ 0.27 μM High (ligand-enhanced coordination) 35 Moderate (hydrothermal) Aqueous (wastewater) 71
Cs3Bi2Cl6/Cs3Bi2Cl9:Eu3+ PQD Cu2+, Fe3+ 6.23 μM (Cu2+), 3.6 μM (Fe3+) Moderate (doping aids selectivity) 35 Moderate (hydrothermal) Aqueous solutions 80
CsSnX3 PQD Pb2+ 3.5 nM High (ligand-mediated specificity) 25 Moderate (hot-injection) Organic solvents (lubricants) 75
CsPbBr3@PCN-333(Fe) MOF PQD Cu2+ 1.63 nM High (MOF-mediated coordination) 85 High (MOF encapsulation) Aqueous solutions 78
CsPbBr3@ZIF-8 MOF PQD Cu2+ 2.64 nM High (MOF enhances specificity) 84 High (MOF encapsulation) Aqueous solutions 79
CsPbBr3–CsPb2Br5 PQD (APTES) Fe3+ 10 μM High (APTES reduces non-specificity) 85 Moderate (hot-injection) Aqueous (industrial effluents) 67
CsPbBr3–CsPb2Br5 PQD (NH4OH) Cd2+ 1 μM Moderate (NH4OH enhances access) 83 Moderate (hot-injection) Aqueous (industrial effluents) 72
Carbon quantum dots (CQD thin film) Pb2+, Ni2+, Mn2+, Co2+, Cr3+ 6 nM Moderate (fluorescence quenching) 20 High (thin film deposition) Aqueous (real water samples) 82
C3B2 quantum dots Cd2+, Hg2+, Ni2+, As3+, Pb2+ 10 nM High (>80% sensitivity, adsorption) 30 High (DFT-based design) Aqueous (wastewater) 90
S@PSCA CQDs Pb2+, Ag+ 3.83 nM (Pb2+), 9 μM (Ag+) Moderate (colorimetric changes) 20 Moderate (hydrothermal) Aqueous solutions 91
SO4@PSCA CQDs Ag+ 8.2 nM Moderate (colorimetric changes) 20 Moderate (hydrothermal) Aqueous solutions 91
SnO2 quantum dots Ni2+ 10 nM Moderate (Sn vacancy interactions) 25 Moderate (hydrolysis) Aqueous (deionized, reclaimed, sea) 92
Nitrogen-doped CQDs (N-CQDs) Fe3+, Hg2+ 35.8 nM (Fe3+), 6.8 nM (Hg2+) High (fluorescence quenching) 20 High (oxidation) Aqueous (pond, sea, well) 93
CQDs (blue crab shells) Pb2+, Hg2+ 5 nM High (NH2, COOH, OH groups) 50 Moderate (microwave) Aqueous (aquaculture) 83
Graphene QDs (S,N-GQDs) Hg2+ 16.32 nM High (SPR-based) 30 Moderate (pyrolysis) Aqueous (tap, river) 84
CdS QDs (DMSO-capped) Cu2+, Hg2+, Pb2+ 179.5 nM (Cu2+), 58 μM (Hg2+), 60 μM (Pb2+) High (colorimetric) 30 Moderate (chemical synthesis) Aqueous (polluted samples) 85
CdTe QDs (TGA, GSH, L-cyst.) Cr3+, Pb2+ 100 nM High (turn-on for Cr3+, Pb2+) 58 Moderate (aqueous phase) Aqueous/organic solutions 94
CdTe QDs (TGA-capped) Cu2+ 50 nM High (1[thin space (1/6-em)]:[thin space (1/6-em)]1 complexation) 60 Moderate (one-step aqueous) Aqueous solutions 95
CdTe/CdS/ZnS QDs Hg2+ 16.32 nM High (fluorescence quenching) 59 High (multi-step aqueous) Aqueous (tap, river, agricultural) 96
CdSe/ZnS QDs (TGA-capped) Cr3+, Cu2+ 100 nM High (turn-off fluorescence) 60 Moderate (aqueous phase) Aqueous solutions 97
CdS QDs (GSH-capped) Cu2+, Hg2+, Mg2+ 100 nM Moderate (fluorescence quenching) 59 Moderate (one-pot) Foods (aqueous media) 98
CuInS2/ZnS QDs (GSH-capped) Cu2+ 63 nM High (aggregation-based quenching) 55 Moderate (aqueous) Aqueous solutions 99
Co/Ag@ZnS QDs Hg2+ 130 nM High (fluorescence quenching) 27.6 Moderate (co-precipitation) Aqueous (real water samples) 100
g-C3N4/CeO2 heterostructure Cu2+, Hg2+, Ag+ 26.2 μM, 17.5 μM, 14.9 μM Moderate (binding affinity) 10 High (co-precipitation) Aqueous solutions 101
Starch-capped CdS QD Hg2+, Cu2+ 100 nM High (cation exchange reaction) 30 Moderate (aqueous) Aqueous solutions 86


CQDs, synthesized via hydrothermal or microwave methods, typically exhibit PLQY of 20–50% and LODs of 5–35.8 nM for ions like Pb2+ and Hg2+ in water samples.82,93 Their broad emission spectra limit sensitivity compared to PQDs. GQDs, with similar PLQY (20–30%), achieve LODs of 16.32 nM for Hg2+ using SPR, but their pyrolysis-based synthesis is less scalable.84 Semiconductor quantum dots, such as CdTe, offer higher PLQY (50–60%) and LODs of 50–179.5 nM, but cadmium toxicity and complex synthesis hinder their use.85,98 MOFs, while not inherently luminescent, enhance the performance of embedded fluorophores, though their fabrication is intricate.78,87 PQDs combine high PLQY, narrow emission, and moderate synthesis complexity, making them versatile for both organic and aqueous applications. The choice of material depends on the application context. CQDs and GQDs excel in eco-friendly aqueous detection, while semiconductor QDs suit high-sensitivity needs despite toxicity concerns. MOFs enhance stability but require complex integration. PQDs offer a balanced profile of sensitivity, tunability, and adaptability, positioning them as a leading platform for heavy metal detection across diverse matrices.

5.2. Advantages of PQDs for heavy metal ion detection

PQDs, particularly lead-based variants like CsPbX3 and CH3NH3PbX3, exhibit exceptional optical properties, including high PLQY and narrow emission spectra, enabling ultra-low LODs (e.g., 0.124 nM for Hg2+ with CH3NH3PbBr3).59 This optical precision supports ratiometric sensing, as seen in Mn2+-doped CsPbCl3 (LOD 22.12 nM for Cu2+), enhancing accuracy in complex matrices.69 Their tunable bandgap, adjusted via size or halide composition, allows tailored detection, such as CsPbBr3 achieving an LOD of 0.1 nM for Cu2+ in hexane through OAm coordination.62 Lead-free PQDs, like Cs3Bi2Br9:Eu3+, offer comparable performance (LOD 98.3 nM for Cu2+) with reduced toxicity, ideal for aqueous environments like seawater.89 Their rapid response times (<10 s) stem from efficient quenching mechanisms, such as cation exchange for Hg2+ detection with CH3NH3PbBr3, outperforming CQDs' slower responses.59,82 Integration into MOFs, as with CsPbBr3@PCN-333(Fe) (LOD 1.63 nM for Cu2+), enhances aqueous stability, making PQDs suitable for industrial wastewater monitoring.78 Table 5 highlights these advantages, showing PQDs' superior LODs and versatility compared to other materials. PQDs' compatibility with organic solvents (e.g., hexane, lubricants) and aqueous media via phase-transfer techniques (e.g., oleylamine-modified CsPbBr3, LOD 0.5 nM for Cu2+) broadens their industrial applicability.61 In contrast, GQDs are limited to aqueous systems, and CdS QDs struggle with selectivity in complex matrices.84,98 The combination of high PLQY, tunability, and rapid response positions PQDs as a leading choice for high-performance heavy metal detection.

5.3. Limitations of alternative sensing materials

CQDs, despite their eco-friendliness, are constrained by lower PLQY (20–50%) and broader emission spectra, reducing sensitivity. For example, CQDs from blue crab shells achieve an LOD of 5 nM for Pb2+, but their PLQY (50%) is lower than CsPbBr3's (81%).62,83 Their need for complex surface functionalization to enhance selectivity increases synthesis complexity.91,93 GQDs, with PLQY of 20–30%, achieve an LOD of 16.32 nM for Hg2+, significantly higher than PQDs' 0.124 nM, and their pyrolysis synthesis is less scalable.59,84 Semiconductor QDs like CdS and CdTe offer PLQY of 50–60% and LODs of 50–179.5 nM, but cadmium toxicity limits their environmental applications.94–98 Their multi-step synthesis is more complex than PQD hot-injection or LARP methods, and selectivity often requires multiple capping agents.94,97 MOFs, while enhancing stability when paired with PQDs, lack intrinsic luminescence and require intricate fabrication, limiting standalone use.78,79 Table 5 underscores these limitations, showing higher LODs and synthesis challenges for alternative materials compared to PQDs' superior performance.

5.4. Synergistic approaches combining PQDs with other nanomaterials

Combining PQDs with MOFs, CQDs, or metal nanoclusters creates synergistic systems that enhance detection performance. Encapsulation in MOFs, such as CsPbBr3@PCN-333(Fe) (LOD 1.63 nM for Cu2+) or CsPbBr3@ZIF-8 (LOD 2.64 nM), improves aqueous stability and analyte accessibility, ideal for wastewater monitoring.78,79 These composites leverage MOF porosity and PQD optical properties, overcoming the aqueous instability of lead-based PQDs. Hybrid systems with CQDs or Au nanoclusters enable ratiometric sensing, as seen with CsPbBr3 and Au nanoclusters for visual Cu2+ detection via FRET, enhancing accuracy in industrial solvents.64 CsPbBr3/PMMA fiber membranes achieve an LOD of 10−15 M for Cu2+, combining PQD fluorescence with polymer matrix stability.88 Doping with Eu3+ or Mn2+, as in Cs3Bi2Br9:Eu3+ (LOD 98.3 nM for Cu2+), enhances PLQY and selectivity in aqueous settings.69,89 These synergistic approaches address PQD limitations, offering performance superior to standalone CQDs or semiconductor quantum dots.

The provided figure illustrates key optical and energetic properties of Eu3+-doped Cs3Bi2Br9 PQDs combined with PeQD. Panel (a) presents the excitation spectra monitored at 446 nm and 620 nm, revealing distinct peaks at 365 nm and broader emissions influenced by Eu3+ doping, with emission wavelengths at 446 nm and 620 nm corresponding to transitions from the conduction band to the valence band and specific Eu3+ energy levels (5D07Fj). This highlights the enhanced PL properties due to Eu3+ incorporation, which boosts PLQY and selectivity, as noted for Cs3Bi2Br9:Eu3+ with an LOD of 98.3 nM for Cu2+. Panel (b) depicts the energy level diagram, showing the alignment of Eu3+ levels (5D0 to 5D4 and 7F0 to 7F6) with the conduction and valence bands of Cs3Bi2Br9, facilitating efficient ET and FRET mechanisms that underpin the improved detection performance in hybrid systems. Panels (c) and (d) further elucidate the concentration-dependent behavior of Eu3+ doping. Panel (c) shows the decay time (τ) at 446 nm decreasing from 3.5 ns to 1.5 ns with increasing Eu3+ concentration (0–12%), indicating enhanced non-radiative energy transfer to Eu3+ ions, as supported by the inset's reverse exciton emission decay constants. Panel (d) quantifies the ET efficiency (ηET) and coefficient (α), peaking at 0.7 and 0.6, respectively, at 12% doping, alongside a bar graph of ηET, which underscores the optimal doping level for maximizing PLQY and sensing accuracy in aqueous environments. These findings reinforce the synergistic potential of Eu3+-doped PQDs with nanomaterials like MOFs or CQDs, enhancing stability and detection limits as demonstrated in systems like CsPbBr3@PCN-333(Fe) and CsPbBr3/PMMA fiber membranes (Fig. 10).


image file: d5ra06200d-f10.tif
Fig. 10 (a) Excitation spectra of Eu3+ (8.6 mol%)-doped Cs3Bi2Br9:PQDs at 446 and 620 nm. (b) Energy level diagram of Eu3+-doped Cs3Bi2Br9:PQDs. (c) Time-resolved PL decay at 446 nm and 620 nm (5D07Fj) for undoped and Eu3+ (1.6–11.3 mol%)-doped samples. (d) Energy transfer efficiency from Cs3Bi2Br9:PQDs to Eu3+ vs. concentration (inset: 446 nm decay constants). Reprinted with permission from ref. 89. Copyright 2019, American Chemical Society.

5.5. Benchmarking PQD performance against established sensors

PQDs consistently outperform alternative sensors in key metrics. For Hg2+, CH3NH3PbBr3 achieves an LOD of 0.124 nM, surpassing CdTe/CdS/ZnS (16.32 nM) and GQDs (16.32 nM) due to efficient cation exchange.84,96 For Cu2+, CsPbBr3 reaches an LOD of 0.1 nM, compared to CdS (179.5 nM) and CuInS2/ZnS (63 nM), leveraging rapid quenching mechanisms.85,99 Lead-free Cs3Bi2Cl9 detects Cr6+ at 0.27 μM, competitive with CdTe (100 nM), with added eco-friendliness.71,94 PQDs' selectivity, enhanced by tailored ligands and doping, matches or exceeds that of CdTe and CdS, which require complex functionalization.62,94 Their moderate synthesis complexity (hot-injection, LARP) contrasts with the multi-step processes of CdTe/CdS/ZnS or GQDs.59,84 PQDs' versatility across organic and aqueous environments, particularly when integrated with MOFs, positions them as a robust platform for industrial and environmental monitoring, outperforming alternative materials in sensitivity and applicability.

6. Challenges, future directions, and pathways to industrial adoption

PQDs have emerged as a transformative platform for heavy metal ion detection, offering exceptional sensitivity, tunable optical properties, and versatility across diverse matrices. Despite their promise, several challenges hinder their practical implementation and scalability, particularly in environmental, industrial, and biomedical applications. Addressing these obstacles is essential to fully harness the potential of PQDs for detecting heavy metal ions such as Hg2+, Cu2+, Cd2+, Fe3+, Cr6+, and Pb2+. This section delineates the primary challenges associated with PQD-based nanosensors and proposes future directions to overcome these limitations, paving the way for advanced, reliable, and commercially viable detection systems.

6.1. Challenges in PQD-based heavy metal ion detection

6.1.1. Stability in aqueous and humid environments. The instability of PQDs, particularly lead-based variants like CsPbX3 and CH3NH3PbX3, in aqueous and humid environments remains a significant barrier. Exposure to moisture can degrade the perovskite lattice, leading to phase transitions or decomposition that diminish PLQY and sensing performance.59,61 For instance, CH3NH3PbBr3 PQDs exhibit reduced stability due to the volatility of organic cations, limiting their use in aqueous media such as wastewater or biological fluids.59 While encapsulation in matrices like metal–organic frameworks (MOFs) or silica enhances stability, as seen with CsPbBr3@ZIF-8 (stable for 15 days in water),79 these approaches increase synthesis complexity and may compromise analyte accessibility, reducing sensitivity for ions like Cu2+ or Cr6+.78,79 Lead-free PQDs, such as Cs3Bi2X9, offer improved moisture resistance but often at the cost of lower PLQY (20–50%), impacting detection limits.60,71 Achieving robust stability without sacrificing optical performance remains a critical challenge for practical applications in industrial effluents and environmental monitoring.
6.1.2. Toxicity concerns of lead-based PQDs. The toxicity of lead-based PQDs, such as CsPbX3 and CH3NH3PbX3, poses significant environmental and health risks, particularly in applications involving water quality monitoring or biomedical diagnostics. The release of Pb2+ ions during degradation can contaminate the tested matrix, rendering these PQDs unsuitable for eco-sensitive applications.59,62 Although lead-free alternatives like Cs3Bi2X9 and CsSnX3 mitigate toxicity concerns, their lower PLQY and stability issues, such as Sn2+ oxidation in CsSnX3, limit their competitiveness.68,75 For example, Cs3Bi2Br9:Eu3+ achieves an LOD of 98.3 nM for Cu2+ in seawater, but its PLQY (42.4%) is significantly lower than that of CsPbBr3 (81%).60,62 Balancing toxicity reduction with high sensitivity and stability is a pressing challenge for widespread adoption.
6.1.3. Interference from coexisting ions and complex matrices. Selectivity in complex matrices, such as industrial wastewater or biological fluids, is hindered by interference from coexisting ions (e.g., Na+, K+, Ca2+) and organic species. While ligands like OAm or PE enhance selectivity for specific ions like Cu2+ or Fe3+, non-specific interactions can lead to false positives or reduced sensitivity.62,68 For instance, CsPbBr3 PQDs show high selectivity for Cu2+ in hexane but face challenges in aqueous environments with multiple ions, where LODs increase from 0.1 nM to 0.8 μM.77 Ratiometric designs, such as CsPbBr3 with Au nanoclusters, mitigate some interference through dual-emission signals, but their complexity limits scalability.64 Developing sensors that maintain high selectivity in real-world, multi-ion environments is a critical hurdle.
6.1.4. Scalability and reproducibility of sensor fabrication. The scalability of PQD synthesis and sensor fabrication remains challenging due to batch-to-batch variations in size, crystallinity, and ligand passivation. Methods like hot-injection yield high-quality PQDs with PLQY > 80%, but their high-temperature, inert-atmosphere requirements limit large-scale production.62 Ligand-assisted reprecipitation (LARP) is more scalable but produces PQDs with broader size distributions and lower PLQY (50–70%), affecting reproducibility in sensing applications.68 For example, variations in ligand density can alter ion accessibility, impacting the LOD for Cu2+ detection.61 Additionally, integrating PQDs into stable matrices like MOFs or polymers increases fabrication complexity, posing challenges for cost-effective, industrial-scale production.78,79 Ensuring consistent sensor performance across large-scale manufacturing is essential for commercial applications.
6.1.5. Lack of standardization for commercial and regulatory applications. The absence of standardized protocols for PQD-based sensor design, calibration, and validation hinders their regulatory approval and commercial adoption. Variations in synthesis methods, ligand types, and detection conditions lead to inconsistent performance metrics, complicating comparisons across studies. For instance, LODs for Cu2+ detection range from 0.1 nM (CsPbBr3 in hexane) to 6.23 μM (Cs3Bi2Cl6:Eu3+ in aqueous media), reflecting diverse testing environments and sensor designs.62,80 Regulatory frameworks for environmental and biomedical applications require standardized testing conditions and safety assessments, particularly for lead-based PQDs, which face scrutiny due to toxicity concerns.59 Establishing universal standards for PQD-based nanosensors is critical to ensure reliability and compliance in practical settings.

6.2. Future directions for PQD-based sensors

To overcome the aforementioned challenges and advance PQD-based nanosensors for heavy metal ion detection, innovative strategies and interdisciplinary approaches are needed. The following directions outline opportunities to enhance stability, sensitivity, selectivity, and scalability, aligning with the demands of industrial, environmental, and biomedical applications.
6.2.1. Development of stable and eco-friendly lead-free PQDs. Advancing lead-free PQDs, such as Cs3Bi2X9, CsSnX3, or double perovskites (e.g., Cs2AgInCl6), is crucial to address toxicity and stability concerns. Enhancing the PLQY of lead-free PQDs through doping with ions like Eu3+ or Mn2+ can improve sensitivity, as demonstrated by Eu3+-doped Cs3Bi2Br9 (LOD 98.3 nM for Cu2+).60 Novel synthesis methods, such as microwave-assisted or solvothermal approaches, can improve crystallinity and reduce defects, enhancing PLQY and stability in aqueous environments.71 Exploring new compositions, such as germanium-based or antimony-based perovskites, could yield environmentally benign PQDs with optical properties comparable to lead-based systems, enabling sustainable applications in wastewater and seawater monitoring.
6.2.2. Advanced encapsulation and surface engineering. Innovative encapsulation strategies can enhance PQD stability without compromising analyte accessibility. Hybrid matrices combining MOFs with polymers or silica, such as CsPbBr3@PCN-333(Fe) (LOD 1.63 nM for Cu2+), offer robust protection and high sensitivity.78 Developing stimuli-responsive coatings that allow selective analyte penetration could further improve selectivity in complex matrices. Ligand engineering, such as using zwitterionic or multidentate ligands, can enhance water compatibility and ion-specific interactions, as seen with hydroxypropyl chitosan-passivated Cs3Bi2Cl9 for Cr6+ detection (LOD 0.27 μM).71 Machine learning-guided ligand design can optimize surface chemistry, predicting ligand–ion interactions to enhance selectivity and sensitivity for ions like Fe3+ or Cd2+.
6.2.3. Integration with advanced detection platforms. Integrating PQD-based nanosensors with advanced platforms, such as microfluidics or smartphone-based systems, can enhance portability and real-time monitoring capabilities. Microfluidic devices can enable precise control over sample delivery, improving detection consistency in industrial effluents or biological fluids. For example, CsPbBr3–CsPb2Br5 PQDs integrated into microfluidic chips could detect Cd2+ with high reproducibility.72 Smartphone-based sensors, leveraging PQD fluorescence for visual or ratiometric detection, offer cost-effective solutions for on-site monitoring in industrial settings, as demonstrated by CsPbBr3–Au nanocluster systems for Cu2+.64 These platforms can democratize access to high-sensitivity detection, particularly in resource-limited environments.
6.2.4. Machine learning and computational approaches. Machine learning (ML) and computational modeling, such as DFT and molecular dynamics (MD), can optimize PQD sensor design. ML algorithms can predict optimal PQD compositions and ligand configurations for specific ions, reducing experimental trial-and-error. For instance, DFT studies have elucidated Hg2+-induced mid-gap states in CH3NH3PbBr3, guiding sensor optimization.59 ML-driven analysis of quenching mechanisms can enhance selectivity by identifying ion-specific interactions in complex matrices, improving performance for multi-ion detection.80 Computational simulations can also model ion diffusion kinetics, informing ligand designs that balance stability and sensitivity, critical for detecting low-concentration ions like Pb2+ in lubricants.75
6.2.5. Standardization and scalable fabrication. Establishing standardized protocols for PQD synthesis, sensor calibration, and performance evaluation is essential for commercial and regulatory acceptance. Developing scalable synthesis methods, such as continuous-flow hot-injection or automated LARP systems, can improve reproducibility and reduce costs. For example, microwave-assisted synthesis offers rapid, uniform nucleation, suitable for large-scale production of Cs3Bi2X9 PQDs.71 Standardizing performance metrics, such as LOD and linear range, across diverse matrices (e.g., organic solvents, wastewater) will facilitate comparisons and regulatory approval. Collaborative efforts between researchers, industry, and regulatory bodies can establish guidelines for PQD-based sensors, ensuring safety and reliability in applications like industrial wastewater treatment.
6.2.6. Multi-ion and multiplexed detection systems. Developing PQD-based nanosensors capable of detecting multiple heavy metal ions simultaneously will enhance their utility in complex industrial and environmental matrices. Multi-ion sensors, such as Eu3+-doped Cs3Bi2Cl6/Cs3Bi2Cl9 for Cu2+ and Fe3+ detection (LODs 6.23 μM and 3.6 μM), demonstrate the potential for multiplexed systems.80 Ratiometric designs with dual-emission PQDs, such as Mn2+-doped CsPbCl3, can improve accuracy by mitigating interference, ideal for industrial solvents or effluents.69 Combining PQDs with other nanomaterials, like CQDs or metal nanoclusters, can enable multiplexed detection with visual readouts, enhancing usability in real-time quality control.64 Future research should focus on optimizing these systems for simultaneous detection of ions like Hg2+, Cu2+, and Cr6+, addressing the needs of diverse applications.
6.2.7. Eco-friendly and biocompatible sensor designs. To address regulatory and environmental concerns, future PQD-based nanosensors should prioritize eco-friendly and biocompatible designs. Lead-free PQDs with enhanced PLQY and stability, such as doped Cs3Bi2X9 or Cs2AgInCl6, can replace toxic lead-based systems in biomedical and environmental applications.60 Biocompatible ligands, such as peptides or polysaccharides, can improve safety for biological fluid analysis, enabling detection of heavy metals in clinical settings. For instance, integrating Cs3Bi2Cl9 with biocompatible coatings could enable Fe3+ detection in blood samples with minimal toxicity.71 These advancements will broaden the applicability of PQD-based sensors, aligning with global sustainability goals.

6.3. Commercialization pathways and industrial integration

The transition of perovskite quantum dot (PQD)-based nanosensors from laboratory research to industrial applications requires strategic pathways to ensure scalability, cost-effectiveness, and seamless integration into existing industrial workflows. While PQD sensors demonstrate exceptional sensitivity and selectivity for heavy metal ion detection, their commercial adoption hinges on addressing economic, regulatory, and operational challenges, distinct from their technical performance in specific applications.
6.3.1. Cost-effective synthesis and manufacturing. Scaling up PQD synthesis for industrial use demands cost-effective methods that maintain high PLQY and reproducibility. Techniques like ligand-assisted reprecipitation (LARP) and continuous-flow hot-injection offer scalable alternatives to batch-based methods, reducing production costs while ensuring uniform PQD quality.68 For instance, automating LARP processes could lower the cost of producing lead-free PQDs like Cs3Bi2X9, making them economically viable for large-scale environmental monitoring.60 Collaborative partnerships with chemical manufacturing industries can optimize precursor sourcing and waste management, further reducing costs. Additionally, developing modular synthesis platforms that integrate quality control metrics, such as real-time PLQY monitoring, can enhance batch consistency, addressing reproducibility concerns for commercial production.
6.3.2. Integration into industrial monitoring systems. Integrating PQD-based nanosensors into existing industrial systems, such as continuous monitoring platforms for wastewater treatment or quality control in chemical processing, requires compatibility with automated and IoT-enabled infrastructures. For example, embedding CsPbBr3@ZIF-8 sensors into inline water quality analyzers can enable real-time detection of Cu2+ in industrial effluents, with data transmitted to centralized control systems.79 Developing plug-and-play sensor modules that interface with standard industrial equipment, such as spectroscopy units or microfluidic systems, can streamline adoption. These modules should incorporate robust encapsulation (e.g., MOFs or silica) to ensure stability under harsh industrial conditions, such as high temperatures or corrosive environments, without compromising analyte accessibility.78
6.3.3. Regulatory compliance and market readiness. Achieving regulatory approval for PQD-based nanosensors involves addressing safety and environmental concerns, particularly for lead-based PQDs. Lead-free PQDs, such as Cs3Bi2Br9:Eu3+, align better with environmental regulations due to their reduced toxicity, making them more viable for markets with stringent standards, such as the European Union's REACH framework.60 Establishing standardized testing protocols for sensor performance, including LOD, selectivity, and stability under diverse conditions, is critical for regulatory acceptance. Collaborations with regulatory bodies can facilitate the development of certification pathways, ensuring PQD sensors meet industry standards for applications in wastewater treatment, chemical manufacturing, and environmental monitoring. Market readiness also requires demonstrating long-term reliability through field trials, such as deploying CsSnX3 sensors in lubricant quality control systems to validate performance over extended periods.75
6.3.4. Economic viability and market differentiation. To compete with established sensing technologies, such as carbon QDs or semiconductor quantum dots, PQD-based nanosensors must offer clear economic advantages. Their high sensitivity (e.g., LOD of 0.1 nM for Cu2+ with CsPbBr3) and versatility across organic and aqueous media provide a competitive edge.62 Positioning PQD sensors as cost-effective solutions for high-precision applications, such as detecting trace Pb2+ in lubricants, can attract industries seeking to minimize equipment downtime and contamination risks.75 Developing portable, user-friendly devices, such as smartphone-integrated PQD sensors for on-site Cu2+ detection, can target niche markets like small-scale chemical processing or environmental consulting, enhancing market penetration.64 Strategic partnerships with sensor manufacturers and distributors can accelerate market entry, leveraging existing supply chains to reduce time-to-market.

7. Conclusion

This comprehensive review underscores the transformative potential of PQDs as advanced platforms for heavy metal ion detection, marking the first systematic evaluation of their application in this domain. PQDs, with their high photoluminescence quantum yield, narrow emission spectra, and tunable optical properties, offer unparalleled sensitivity and selectivity for detecting ions such as Hg2+, Cu2+, Cd2+, Fe3+, Cr6+, and Pb2+. Their efficacy, driven by mechanisms like cation exchange and electron transfer, is enhanced by advanced synthesis methods and integration with stable matrices like metal–organic frameworks. Lead-based PQDs achieve sub-nanomolar detection limits, while lead-free variants address environmental concerns, broadening their applicability. The versatility of PQDs enables their use in industrial wastewater treatment and lubricant quality control, ensuring compliance with environmental regulations and product integrity. Despite these advancements, challenges such as aqueous instability, toxicity of lead-based PQDs, and scalability of synthesis methods remain. Future research should focus on developing stable, eco-friendly PQDs through novel encapsulation techniques and ligand designs, alongside computational modeling to optimize ion–PQD interactions. Comparative studies with other nanomaterials, such as carbon quantum dots, will further elucidate their unique advantages. By addressing these challenges, PQDs can transition from laboratory innovations to practical sensing solutions, offering rapid, cost-effective, and portable detection for environmental and industrial applications. This review lays a foundation for advancing PQD-based sensors, guiding researchers toward innovations that enhance sensitivity, selectivity, and scalability, ultimately contributing to sustainable monitoring of heavy metal ion contamination.

Author contributions

Ahmad Mohebi and Suleiman Ibrahim Mohammad conceptualized the study. Asokan Vasudevan, I. B. Sapaev, and Munthar Kadhim Abosaoda developed the methodology. Chou-Yi Hsu, Malatesh Akkur, and Alok Kumar Mishra conducted the investigation. Gaganjot Kaur and Rajesh Singh handled data curation. Suleiman Ibrahim Mohammad and Ahmad Mohebi drafted the original manuscript, while Ahmad Mohebi, Asokan Vasudevan, and Chou-Yi Hsu reviewed and edited it. I. B. Sapaev and Munthar Kadhim Abosaoda contributed to visualization. Ahmad Mohebi supervised and managed the project. Suleiman Ibrahim Mohammad secured funding. All authors have read and agreed to the published version of the manuscript.

Conflicts of interest

There are no conflicts to declare.

Data availability

No primary research results, software or code have been included and no new data were generated or analysed as part of this review.

Acknowledgements

The research is funded by Zarqa University.

References

  1. Y. Han, S. Zhang, D. Kang, N. Hao, J. Peng, Y. Zhou, K. Liu and Y. Chen, A Decade Review of Human Health Risks from Heavy Metal Contamination in Industrial Sites, Water, Air, Soil Pollut., 2025, 236(2), 144 CrossRef CAS.
  2. S. Syarifuddin, S. Suryani and D. Tahir, Global advances and innovations in bacteria-based biosorption for heavy metal remediation: a bibliometric and analytical perspective, Integrated Environ. Assess. Manag., 2025, 21(3), 507–525 CrossRef CAS PubMed.
  3. M. Swathika, N. M. Andal, S. Dharani, J. Singh, S. S. Pandey, K. R. Singh and A. Natarajan, Targeted elimination of heavy metals from industrial wastewater: synergistic effect of nano metal oxides, J. Taiwan Inst. Chem. Eng., 2025, 166, 105485 CrossRef CAS.
  4. R. U. Khan, M. Hamayun, A. A. Altaf, S. Kausar, Z. Razzaq and T. Javaid, Assessment and removal of heavy metals and other ions from the industrial wastewater of Faisalabad, Pakistan, Processes, 2022, 10(11), 2165 CrossRef CAS.
  5. H. Sable, V. Kumar, V. Singh, S. Rustagi, S. Chahal and V. Chaudhary, Strategically engineering advanced nanomaterials for heavy-metal remediation from wastewater, Coord. Chem. Rev., 2024, 518, 216079 CrossRef CAS.
  6. A. T. Ahmed, F. M. Altalbawy, H. R. Al-Hetty, N. Fayzullaev, K. V. Jamuna, J. Sharma, B. Abd, A. M. Ahmed and H. Noorizadeh, Cadmium selenide quantum dots in photo-and electrocatalysis: advances in hydrogen, oxygen, and CO2 reactions, Mater. Sci. Semicond. Process., 2025, 199, 109831 CrossRef CAS.
  7. S. Anwar, Q. A. Syed, A. Ullah, M. Arshad, S. N. Khan and M. I. Hussain, Detection of heavy metals using atomic absorption and emission spectroscopy in drinking water of Faisalabad, Pakistan: microbial safety and quality status assessment, Toxin Rev., 2024, 43(1), 1–7 CrossRef CAS.
  8. M. H. Abdellattif, H. Noorizadeh, S. Al-Hasnaawei, S. Ganesan, A. F. Al-Hussainy, A. Sandhu, A. Sinha and M. Kazemi, Tailoring Charge Storage Mechanisms through TiO2 Quantum Dots: A Multifunctional Strategy for High-Performance Battery Electrodes, Surf. Interfaces, 2025, 9, 107164 CrossRef.
  9. J. Y. Cai, H. Shen, J. N. Xu, J. M. Xu, J. H. Wang and Y. L. Yu, Magnetic Field-Accelerated Nonthermal Plasma Digestion for Field Pretreatment and Determination of Heavy Metals in Biological Samples, Anal. Chem., 2025, 97(5), 2899–2905 CrossRef CAS PubMed.
  10. Y. Shi, S. Zhang, H. Zhou, Y. Dong, G. Liu, W. Ye, R. He and G. Zhao, Recent Developments in Heavy Metals Detection: Modified Electrodes, Pretreatment Methods, Prediction Models and Algorithms, Metals, 2025, 15(1), 80 CrossRef CAS.
  11. Z. Lian and F. Y. Kong, Treatments of Heavy Metals in Bio Electrochemical Systems (BES) and Microbial Transfer Mechanisms: Analyses of Influencing Factors, Future Research Directions, Prospect and Outlook, Water, Air, Soil Pollut., 2025, 236(8), 1–39 CrossRef.
  12. S. Y. Kew and S. Y. Lau, Review on electrochemical reactors for efficient heavy metals removal from wastewater, J. Ind. Eng. Chem., 2025, 152, 18–32 CrossRef.
  13. T. Hu, Q. Lai, W. Fan, Y. Zhang and Z. Liu, Advances in portable heavy metal ion sensors, Sensors, 2023, 23(8), 4125 CrossRef CAS PubMed.
  14. M. F. Rauf, Z. Lin, M. K. Rauf and J. M. Lin, Innovative Microfluidic Technologies for Rapid Heavy Metal Ion Detection, Chemosensors, 2025, 13(4), 149 CrossRef CAS.
  15. H. Zhang, L. Li, C. Wang, Q. Liu, W. T. Chen, S. Gao and G. Hu, Recent advances in designable nanomaterial-based electrochemical sensors for environmental heavy-metal detection, Nanoscale, 2025, 17, 2386–2407 RSC.
  16. C. W. Hsieh, R. K. Singh, M. L. Meena, S. A. Lu, K. K. Gupta, B. C. Lin and C. H. Lu, Development of inorganic perovskite quantum dots-based optical sensor for toxic Cd2+ metal ion detection, Optik, 2025, 22, 172462 CrossRef.
  17. M. Kedhim, V. Jain, S. Ballal, A. Kumar, A. Singh, V. Kavitha, R. Panigrahi and H. Noorizadeh, Ultrasensitive dopamine detection using CsPbBr 3-PQD-COF nanocomposites: a synergistic fluorescence and EIS approach, RSC Adv., 2025, 15(24), 18875–18892 RSC.
  18. X. Zhang, L. Li, Y. Chen, C. Valenzuela, Y. Liu, Y. Yang, Y. Feng, L. Wang and W. Feng, Mechanically tunable circularly polarized luminescence of liquid crystal-templated chiral perovskite quantum dots, Angew. Chem., 2024, 136(22), e202404202 CrossRef.
  19. B. Xu, S. Yuan, L. Wang, X. Li, Z. Hu and H. Zeng, Highly Efficient Blue Light-Emitting Diodes Enabled by Gradient Core/Shell-Structured Perovskite Quantum Dots, ACS Nano, 2025, 19(3), 3694–3704 CrossRef CAS PubMed.
  20. J. W. Choi, H. C. Woo, X. Huang, W. G. Jung, B. J. Kim, S. W. Jeon, S. Y. Yim, J. S. Lee and C. L. Lee, Organic–inorganic hybrid perovskite quantum dots with high PLQY and enhanced carrier mobility through crystallinity control by solvent engineering and solid-state ligand exchange, Nanoscale, 2018, 10(28), 13356–13367 RSC.
  21. W. Chen, X. Xin, Z. Zang, X. Tang, C. Li, W. Hu, M. Zhou and J. Du, Tunable photoluminescence of CsPbBr3 perovskite quantum dots for light emitting diodes application, J. Solid State Chem., 2017, 255, 115–120 CrossRef CAS.
  22. Q. Shan, Y. Dong, H. Xiang, D. Yan, T. Hu, B. Yuan, H. Zhu, Y. Wang and H. Zeng, Perovskite quantum dots for the next-generation displays: progress and prospect, Adv. Funct. Mater., 2024, 34(36), 2401284 CrossRef CAS.
  23. M. Shellaiah, K. Awasthi, S. Chandran, B. Aazaad, K. W. Sun, N. Ohta, S. P. Wu and M. C. Lin, Methylammonium tin tribromide quantum dots for heavy metal ion detection and cellular imaging, ACS Appl. Nano Mater., 2022, 5(2), 2859–2874 CrossRef CAS.
  24. K. Tang, Y. Chen and Y. Zhao, Exploiting halide perovskites for heavy metal ion detection, Chem. Commun., 2024, 60(34), 4511–4520 RSC.
  25. H. Ye, W. Gao and L. Xu, High-temperature energy storage capability of polyetherimide composite incorporated with perovskite quantum dots, Colloids Surf., A, 2024, 687, 133479 CrossRef CAS.
  26. H. Huang, J. Zhou, Q. Cai, C. Zhou, N. Li, X. Dai, Z. Ye, H. He and C. Fan, Industrial-Scale Fabrication of Mixed-Halide Perovskite Quantum Dots with High Comprehensive Performances for Red-Emitting Modules, Nano Lett., 2025, 25(24), 9863–9871 CrossRef CAS PubMed.
  27. X. G. Wu, Y. Jing and H. Zhong, In Situ Fabricated Perovskite Quantum Dots: From Materials to Applications, Adv. Mater., 2024, 2412276 Search PubMed.
  28. P. Sun, Z. Xing, Z. Li and W. Zhou, Recent advances in quantum dots photocatalysts, Chem. Eng. J., 2023, 458, 141399 CrossRef CAS.
  29. S. Aftab, Z. Ali, M. I. Hussain, M. A. Assiri, N. Rubab, F. Ozel and E. Akman, Perovskite Quantum Dots: Fabrication, Degradation, and Enhanced Performance Across Solar Cells, Optoelectronics, and Quantum Technologies, Carbon Energy, 2025, e70018 CrossRef CAS.
  30. J. Lin, S. Chen, W. Ye, Y. Zeng, H. Xiao, T. Pang, Y. Zheng, B. Zhuang, F. Huang and D. Chen, Ultra-stable yellow monolithic perovskite quantum dots film for backlit display, Adv. Funct. Mater., 2024, 34(27), 2314795 CrossRef CAS.
  31. X. Wang, Z. Bao, Y. C. Chang and R. S. Liu, Perovskite quantum dots for application in high color gamut backlighting display of light-emitting diodes, ACS Energy Lett., 2020, 5(11), 3374–3396 CrossRef CAS.
  32. J. Zhang, S. Zhang, Y. Zhang, O. A. Al-Hartomy, S. Wageh, A. G. Al-Sehemi, Y. Hao, L. Gao, H. Wang and H. Zhang, Colloidal quantum dots: synthesis, composition, structure, and emerging optoelectronic applications, Laser Photon. Rev., 2023, 17(3), 2200551 CrossRef CAS.
  33. J. Zou, M. Li, X. Zhang and W. Zheng, Perovskite quantum dots: synthesis, applications, prospects, and challenges, J. Appl. Phys., 2022, 132(22), 220901 CrossRef CAS.
  34. Z. Ning, X. Gong, R. Comin, G. Walters, F. Fan, O. Voznyy, E. Yassitepe, A. Buin, S. Hoogland and E. H. Sargent, Quantum-dot-in-perovskite solids, Nature, 2015, 523(7560), 324–328 CrossRef CAS PubMed.
  35. A. Soosaimanickam, A. Saura, N. Farinós and R. Abargues, Influence of binary ligands in designing cesium lead halide (CsPbX3, X= Cl, Br, I) perovskite nanocrystals-oleic acid and oleylamine, Nanoenergy Adv., 2023, 3(4), 376–400 CrossRef.
  36. N. S. Makarov, S. Guo, O. Isaienko, W. Liu, I. Robel and V. I. Klimov, Spectral and dynamical properties of single excitons, biexcitons, and trions in cesium–lead-halide perovskite quantum dots, Nano Lett., 2016, 16(4), 2349–2362 CrossRef CAS PubMed.
  37. Y. Zhou and Y. Wang, Perovskite Quantum Dots, Springer Series in Materials Science, 2020 Search PubMed.
  38. Y. Zhu, J. Zhu, H. Song, J. Huang, Z. Lu and G. Pan, Samarium doping improves luminescence efficiency of Cs3Bi2Br9 perovskite quantum dots enabling efficient white light-emitting diodes, J. Rare Earths, 2021, 39(4), 374–379 CrossRef CAS.
  39. Y. Tang, S. Tang, M. Luo, Y. Guo, Y. Zheng, Y. Lou and Y. Zhao, All-inorganic lead-free metal halide perovskite quantum dots: progress and prospects, Chem. Commun., 2021, 57(61), 7465–7479 RSC.
  40. D. E. Lee, S. Y. Kim and H. W. Jang, Lead-free all-inorganic halide perovskite quantum dots: review and outlook, J. Korean Ceram. Soc., 2020, 57(5), 455–479 CrossRef CAS.
  41. D. Wu, F. Zhou, K. Feng, J. Wang, Q. Lu, X. Wang, Z. Zhan, W. Chen, Z. Liu, Z. Zhang and Z. Hu, Rare earth Nd3+-doped organic-inorganic hybrid perovskite quantum dots for white LED, J. Lumin., 2025, 277, 120876 CrossRef CAS.
  42. J. W. Choi, H. C. Woo, X. Huang, W. G. Jung, B. J. Kim, S. W. Jeon, S. Y. Yim, J. S. Lee and C. L. Lee, Organic–inorganic hybrid perovskite quantum dots with high PLQY and enhanced carrier mobility through crystallinity control by solvent engineering and solid-state ligand exchange, Nanoscale, 2018, 10(28), 13356–13367 RSC.
  43. M. Leng, Y. Yang, K. Zeng, Z. Chen, Z. Tan, S. Li, J. Li, B. Xu, D. Li, M. P. Hautzinger and Y. Fu, All-inorganic bismuth-based perovskite quantum dots with bright blue photoluminescence and excellent stability, Adv. Funct. Mater., 2018, 28(1), 1704446 CrossRef.
  44. E. Jokar, L. Cai, J. Han, E. J. Nacpil and I. Jeon, Emerging opportunities in lead-free and lead–tin perovskites for environmentally viable photodetector applications, Chem. Mater., 2023, 35(9), 3404–3426 CrossRef CAS.
  45. T. Wu, X. Liu, X. Luo, X. Lin, D. Cui, Y. Wang, H. Segawa, Y. Zhang and L. Han, Lead-free tin perovskite solar cells, Joule, 2021, 5(4), 863–886 CrossRef CAS.
  46. Z. Long, S. Yang, J. Pi, D. Zhou, Q. Wang, Y. Yang, H. Wu and J. Qiu, All-inorganic halide perovskite (CsPbX3, X= Cl, Br, I) quantum dots synthesized via fast anion hot injection by using trimethylhalosilanes, Ceram. Int., 2022, 48(23), 35474–35479 CrossRef CAS.
  47. F. Meng, X. Liu, X. Cai, Z. Gong, B. Li, W. Xie, M. Li, D. Chen, H. L. Yip and S. J. Su, Incorporation of rubidium cations into blue perovskite quantum dot light-emitting diodes via FABr-modified multi-cation hot-injection method, Nanoscale, 2019, 11(3), 1295–1303 RSC.
  48. S. Aftab, A. Shah, C. Erkmen, S. Kurbanoglu and B. Uslu, Quantum dots: synthesis and characterizations, in Electroanalytical Applications of Quantum Dot-Based Biosensors, Elsevier, 2021, pp. 1–35 Search PubMed.
  49. J. Zou, M. Li, X. Zhang and W. Zheng, Perovskite quantum dots: synthesis, applications, prospects, and challenges, J. Appl. Phys., 2022, 132(22), 220901 CrossRef CAS.
  50. C. Bi, S. Wang, W. Wen, J. Yuan, G. Cao and J. Tian, Room-temperature construction of mixed-halide perovskite quantum dots with high photoluminescence quantum yield, J. Phys. Chem. C, 2018, 122(9), 5151–5160 CrossRef CAS.
  51. S. Mandal, D. Roy, C. K. De, S. Ghosh, M. Mandal, A. Das and P. K. Mandal, Instantaneous, room-temperature, open-air atmosphere, solution-phase synthesis of perovskite quantum dots through halide exchange employing non-metal based inexpensive HCl/HI: ensemble and single particle spectroscopy, Nanoscale Adv., 2019, 1(9), 3506–3513 RSC.
  52. Y. Li, H. Huang, Y. Xiong, S. V. Kershaw and A. L. Rogach, Revealing the formation mechanism of CsPbBr3 perovskite nanocrystals produced via a slowed-down microwave-assisted synthesis, Angew. Chem., Int. Ed., 2018, 57(20), 5833–5837 CrossRef CAS PubMed.
  53. C. H. Lu, Y. H. Liu, M. L. Meena and S. Som, Synthesis and spectroscopic characterization of CsPbBr3/Cs4PbBr6 perovskites synthesized via the microwave-assisted heating process for backlight display devices, Org. Electron., 2021, 91, 106079 CrossRef CAS.
  54. M. Guarino-Hotz, J. L. Barnett, L. B. Pham, A. A. Win, V. L. Cherrette and J. Z. Zhang, Tuning between methylammonium lead bromide perovskite magic-sized clusters and quantum dots through ligand assisted reprecipitation at elevated temperatures, J. Phys. Chem. C, 2022, 126(32), 13854–13862 CrossRef CAS.
  55. S. L. Sanchez, Y. Tang, B. Hu, J. Yang and M. Ahmadi, Understanding the ligand-assisted reprecipitation of CsPbBr3 nanocrystals via high-throughput robotic synthesis approach, Matter, 2023, 6(9), 2900–2918 CrossRef CAS.
  56. Q. Zhao, S. Wang, Y. H. Kim, S. Mondal, Q. Miao, S. Li, D. Liu, M. Wang, Y. Zhai, J. Gao and A. Hazarika, Advantageous properties of halide perovskite quantum dots towards energy-efficient sustainable applications, Green Energy Environ., 2024, 9(6), 949–965 CrossRef CAS.
  57. D. Chen and X. Chen, Luminescent perovskite quantum dots: synthesis, microstructures, optical properties and applications, J. Mater. Chem. C, 2019, 7(6), 1413–1446 RSC.
  58. Y. Bai, M. Hao, S. Ding, P. Chen and L. Wang, Surface chemistry engineering of perovskite quantum dots: strategies, applications, and perspectives, Adv. Mater., 2022, 34(4), 2105958 CrossRef CAS PubMed.
  59. L. Q. Lu, T. Tan, X. K. Tian, Y. Li and P. Deng, Visual and sensitive fluorescent sensing for ultratrace mercury ions by perovskite quantum dots, Anal. Chim. Acta, 2017, 986, 109–114 CrossRef CAS PubMed.
  60. Y. Gao and B. Chen, Lead-free Cs3Bi2Br9 perovskite quantum dots for detection of heavy metal Cu2+ ions in seawater, J. Mar. Sci. Eng., 2023, 11(5), 1001 CrossRef.
  61. Q. Li, W. Zhou, L. Yu, S. Lian and Q. Xie, Perovskite quantum dots as a fluorescent probe for metal ion detection in aqueous solution via phase transfer, Mater. Lett., 2021, 282, 128654 CrossRef CAS.
  62. Y. Liu, X. Tang, T. Zhu, M. Deng, I. P. Ikechukwu, W. Huang, G. Yin, Y. Bai, D. Qu, X. Huang and F. Qiu, All-inorganic CsPbBr 3 perovskite quantum dots as a photoluminescent probe for ultrasensitive Cu 2+ detection, J. Mater. Chem. C, 2018, 6(17), 4793–4799 RSC.
  63. J. Iskandar, C. Y. Liu, C. C. Lee, K. Y. Ke, M. R. Septian, R. Estrada, H. Humaidi, S. Biring, C. S. Chu, Z. L. Tseng and S. W. Liu, Stabilizing perovskite quantum dot oxygen sensors through ultra-long 2 mm horizontally aligned nanopores in anodic alumina oxide templates, J. Mater. Chem. C, 2025, 13(2), 709–717 RSC.
  64. W. Xue, J. Zhong, H. Wu, J. Zhang and Y. Chi, A visualized ratiometric fluorescence sensing system for copper ions based on gold nanoclusters/perovskite quantum dot@ SiO 2 nanocomposites, Analyst, 2021, 146(24), 7545–7553 RSC.
  65. K. Tang, Y. Chen and Y. Zhao, Exploiting halide perovskites for heavy metal ion detection, Chem. Commun., 2024, 60(34), 4511–4520 RSC.
  66. M. Shellaiah, K. W. Sun, N. Thirumalaivasan, M. Bhushan and A. Murugan, Sensing utilities of cesium lead halide perovskites and composites: a comprehensive review, Sensors, 2024, 24(8), 2504 CrossRef CAS PubMed.
  67. C. W. Hsieh, R. K. Singh, S. Som and C. H. Lu, Detection of Fe (III) using APTES-coated CsPbBr3–CsPb2Br5 perovskite quantum dots, Chem. Eng. J. Adv., 2022, 12, 100358 CrossRef CAS.
  68. M. Shellaiah, K. Awasthi, S. Chandran, B. Aazaad, K. W. Sun, N. Ohta, S. P. Wu and M. C. Lin, Methylammonium tin tribromide quantum dots for heavy metal ion detection and cellular imaging, ACS Appl. Nano Mater., 2022, 5(2), 2859–2874 CrossRef CAS.
  69. Y. Gao, S. Xu, Y. Li and B. Chen, Mn-doped CsPbCl3 perovskite quantum dots: a dual-function probe for copper detection and temperature sensing, Spectrochim. Acta, Part A, 2025, 326, 125219 CrossRef CAS PubMed.
  70. P. Pandey, A. Sengupta, S. Parmar, U. Bansode, S. Gosavi, A. Swarnkar, S. Muduli, A. D. Mohite and S. Ogale, CsPbBr3–Ti3C2T x MXene QD/QD heterojunction: photoluminescence quenching, charge transfer, and Cd ion sensing application, ACS Appl. Nano Mater., 2020, 3(4), 3305–3314 CrossRef CAS.
  71. D. Gao, A. Zhang, B. Lyu and J. Ma, Visual and rapid fluorescence sensing for hexavalent chromium by hydroxypropyl chitosan passivated bismuth-based perovskite quantum dots, Microchim. Acta, 2024, 191(4), 219 CrossRef CAS PubMed.
  72. C. W. Hsieh, R. K. Singh, M. L. Meena, S. A. Lu, K. K. Gupta, B. C. Lin and C. H. Lu, Development of inorganic perovskite quantum dots-based optical sensor for toxic Cd2+ metal ion detection, Optik, 2025, 22, 172462 CrossRef.
  73. X. Lao, X. Li, H. Ågren and G. Chen, Highly controllable synthesis and DFT calculations of double/triple-halide CsPbX3 (X= Cl, Br, I) perovskite quantum dots: application to light-emitting diodes, Nanomaterials, 2019, 9(2), 172 CrossRef CAS PubMed.
  74. E. Stippell, P. C. Mora, N. Favate, L. Huang, C. W. Li and O. V. Prezhdo, Computational Screening of Ligands for Enhanced Interactions between Lead Halide Perovskite Quantum Dots, J. Phys. Chem. Lett., 2025, 16, 5666–5673 CrossRef CAS PubMed.
  75. D. Li, W. Xu, D. Zhou, X. Ma, X. Chen, G. Pan, J. Zhu, Y. Ji, N. Ding and H. Song, Cesium tin halide perovskite quantum dots as an organic photoluminescence probe for lead ion, J. Lumin., 2019, 216, 116711 CrossRef CAS.
  76. X. Sheng, Y. Liu, Y. Wang, Y. Li, X. Wang, X. Wang, Z. Dai, J. Bao and X. Xu, Cesium lead halide perovskite quantum dots as a photoluminescence probe for metal ions, Adv. Mater., 2017, 29(37), 1700150 CrossRef PubMed.
  77. J. Sun, Q. Li, X. Li, C. Yan and G. Wang, Super stable and water-dispersible CsPbBr3 perovskite quantum dots for the efficient detection of Cu (II) and glutathione, Microchem. J., 2024, 199, 110043 CrossRef CAS.
  78. C. Gao, P. Yu, H. Yang, M. Wang, Y. Cao, J. Nie, W. Cao, M. Que, W. Zhu, P. Zhong and X. Ma, Enhanced photoluminescence and stability of CsPbBr3 quantum dots by anchoring on metal-organic frameworks for Cu2+ detection fluorescence probe, J. Alloys Compd., 2025, 1023, 180159 CrossRef CAS.
  79. S. Ahmed, S. Lahkar, S. Doley, D. Mohanta and S. K. Dolui, A hierarchically porous MOF confined CsPbBr3 quantum dots: fluorescence switching probe for detecting Cu (II) and melamine in food samples, J. Photochem. Photobiol., A, 2023, 443, 114821 CrossRef CAS.
  80. P. He, W. Ge, Q. Zhang, M. Yang, H. Yin, X. Xie, Z. Luo and S. Shang, Eu3+ doped Cs–Bi–Cl perovskite quantum dots with diverse crystal structures for metal ion detection, Ceram. Int., 2024, 50(11), 20285–20292 CrossRef CAS.
  81. S. Ahmed, S. Lahkar, P. Saikia, D. Mohanta, J. Das and S. K. Dolui, Stable and highly luminescent CsPbX3 (X= Br, Br/Cl) perovskite quantum dot embedded into Zinc (II) imidazole-4, 5-dicarboxylate metal organic framework as a luminescent probe for metal ion detection, Mater. Chem. Phys., 2023, 295, 127093 CrossRef CAS.
  82. T. Vyas and A. Joshi, Chemical sensor thin film-based carbon quantum dots (CQDs) for the detection of heavy metal count in various water matrices, Analyst, 2024, 149(4), 1297–1309 RSC.
  83. Ö. Gencer, The production of carbon quantum dots (CQDs) from cultivated blue crab shells and their application in the optical detection of heavy metals in aquaculture, Aquacult. Int., 2025, 33(1), 107 CrossRef.
  84. R. Üzek, Engineering a Graphene Quantum Dot-Enhanced Surface Plasmon Resonance Sensor for Ultra-Sensitive Detection of Hg2+ Ions, Adv. Mater. Interfaces, 2025, 12(5), 2400679 CrossRef.
  85. P. Ayyanusamy, R. Venkatesan, A. N. Rajan, J. Annaraj, U. Mahalingam, P. Ramasamy, K. Bethke and J. Mayandi, Dimethylsulfoxide functionalized cadmium sulfide quantum dot for heavy metal ion detection, Z. Phys. Chem., 2025, 28, 121–134 Search PubMed.
  86. S. Iraqui, A. Dubey, I. Savarimuthu, A. Shankar, A. Jaiswal, I. Bahadur, I. Uddin and F. Mohammad, Bi-functional aqueous starch capped CdS quantum dots synthesis and their application as sensor of heavy metal-ions as well as photocatalytic dye degradation, J. Mol. Liq., 2023, 388, 122794 CrossRef CAS.
  87. D. Zhang, Y. Xu, Q. Liu and Z. Xia, Encapsulation of CH3NH3PbBr3 perovskite quantum dots in MOF-5 microcrystals as a stable platform for temperature and aqueous heavy metal ion detection, Inorg. Chem., 2018, 57(8), 4613–4619 CrossRef CAS PubMed.
  88. Y. Wang, Y. Zhu, J. Huang, J. Cai, J. Zhu, X. Yang, J. Shen and C. Li, Perovskite quantum dots encapsulated in electrospun fiber membranes as multifunctional supersensitive sensors for biomolecules, metal ions and pH, Nanoscale Horiz., 2017, 2(4), 225–232 RSC.
  89. N. Ding, D. Zhou, G. Pan, W. Xu, X. Chen, D. Li, X. Zhang, J. Zhu, Y. Ji and H. Song, Europium-doped lead-free Cs3Bi2Br9 perovskite quantum dots and ultrasensitive Cu2+ detection, ACS Sustain. Chem. Eng., 2019, 7(9), 8397–8404 CrossRef CAS.
  90. M. T. Ahmed, D. Roy, A. A. Roman, S. Islam and F. Ahmed, C3B2 Quantum Dot: A Potential Candidate for Heavy Metal Ion Detection and Removal in Wastewater Treatment, Langmuir, 2025, 41(14), 9456–9468 CrossRef CAS PubMed.
  91. H. Javeria, M. Q. Abbas, S. H. Chen, B. E. Keshat and Z. Du, Scalability of sulfur-functionalized carbon quantum dots from peanut shells: a sustainable sensor of high colorimetric heavy metal detection, J. Environ. Chem. Eng., 2025, 13(2), 115821 CrossRef CAS.
  92. J. Liu, Q. Zhang, W. Xue, H. Zhang, Y. Bai, L. Wu, Z. Zhai and G. Jin, Fluorescence characteristics of aqueous synthesized tin oxide quantum dots for the detection of heavy metal ions in contaminated water, Nanomaterials, 2019, 9(9), 1294 CrossRef CAS PubMed.
  93. R. Yadav, V. Lahariya, S. A. K. Vikas, A. Das, A. Yadav and G. Gupta, Fluorometric sensing and nanomolar level detection of heavy metal ions using nitrogen doped carbon dots, Emergent Mater., 2025, 8(1), 363–377 CrossRef CAS.
  94. P. Yadav and P. Chowdhury, Optical efficiency of CdTe QDs for metal ion sensing in the presence of different thiol-based capping agents, Chem. Pap., 2022, 76(3), 1833–1850 CrossRef CAS.
  95. M. Rana and K. Devlal, Thioglycolic Acid Capped CdTe Quantum Dots as Sensors for the Detection of Hazardous Heavy Metal Ion Cu 2+ in Water, Mapan, 2022, 1–6 Search PubMed.
  96. F. Farahmandzadeh, K. Kermanshahian, E. Molahosseini and M. Molaei, Efficient Mercury Ion Detection in Water: CdTe/CdS/ZnS Quantum Dots as a Simple, Sensitive, and Rapid Fluorescence Sensor, J. Fluoresc., 2024, 13, 1–2 Search PubMed.
  97. S. Sharma, P. Yadav and P. Chowdhury, Thioglycolic acid capped CdSe/ZnS quantum dot as fluorescent sensor for the detection of water-soluble hazardous heavy metal ions, Appl. Phys. A:Mater. Sci. Process., 2024, 130(5), 326 CrossRef CAS.
  98. K. He, X. Yu, L. Qin and Y. Wu, CdS QDs: facile synthesis, design and application as an “on–off” sensor for sensitive and selective monitoring Cu2+, Hg2+ and Mg2+ in foods, Food Chem., 2022, 390, 133116 CrossRef CAS PubMed.
  99. J. V. Rajendran, S. Parani, V. P. Remya, T. C. Lebepe, R. Maluleke, S. Thomas and O. S. Oluwafemi, Selective and sensitive detection of Cu2+ ions in the midst of other metal ions using glutathione capped CuInS2/ZnS quantum dots, Phys. E, 2022, 136, 115026 CrossRef CAS.
  100. U. Latief, M. S. Khan, S. U. Islam, Z. Khan and M. A. Saifee, Environment friendly Co/Ag@ ZnS quantum dots with high quality emission and application for selective sensing of Hg2+, J. Photochem. Photobiol., A, 2023, 445, 115038 CrossRef CAS.
  101. M. Kaur, S. Singh, S. K. Mehta, S. K. Kansal, A. Umar, A. A. Ibrahim and S. Baskoutas, CeO2 quantum dots decorated g-C3N4 nanosheets: a potential scaffold for fluorescence sensing of heavy metals and visible-light driven photocatalyst, J. Alloys Compd., 2023, 960, 170637 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.