Open Access Article
Mohammad Abushuhela,
G. PadmaPriyab,
Shaker Al-Hasnaaweicd,
Subhashree Raye,
Kattela Chennakesavuluf,
Renu Sharmag,
Ashish Singh Chauhanh,
Hadi Noorizadeh
*i and
Mosstafa Kazemi
i
aFaculty of Allied Medical Sciences, Hourani Center for Applied Scientific Research, Al-Ahliyya Amman University, Amman, Jordan
bDepartment of Chemistry and Biochemistry, School of Sciences, JAIN (Deemed to be University), Bangalore, Karnataka, India
cCollege of Pharmacy, Islamic University, Najaf, Iraq
dDepartment of Medical Analysis, Medical Laboratory Technique College, Islamic University of Al Diwaniyah, Al Diwaniyah, Iraq
eDepartment of Biochemistry, IMS and SUM Hospital, Siksha ‘O’ Anusandhan, Bhubaneswar, Odisha-751003, India
fDepartment of Chemistry, Sathyabama Institute of Science and Technology, Chennai, Tamil Nadu, India
gDepartment of Chemistry, University Institute of Sciences, Chandigarh University, Mohali, Punjab, India
hUttaranchal Institute of Pharmaceutical Sciences, Division of Research and Innovation, Uttaranchal University, Dehradun, Uttarakhand, India
iYoung Researchers and Elite Club, Tehran Branch, Islamic Azad University, Tehran, Iran. E-mail: hadinoorizadehacademic@gmail.com
First published on 1st December 2025
Perovskite quantum dots (PQDs) have emerged as a new generation of semiconductor nanomaterials with outstanding potential in oncology. Their unique optoelectronic features—including high photoluminescence quantum yields, tunable emission, and efficient charge transport—position them as superior candidates compared to conventional quantum dots. This review presents an integrated overview of PQDs, starting from their synthesis methodologies and structural–optoelectronic characteristics to their biocompatibility and biomedical applications. Special attention is paid to surface modification strategies, such as silica encapsulation, polymer coatings, hybrid nanostructures, and biomimetic approaches, which enhance aqueous stability, mitigate toxicity, and enable targeted delivery. Furthermore, the applications of PQDs in cancer diagnostics and therapy are highlighted, covering fluorescence and multimodal imaging, biosensing of tumor biomarkers, and advanced therapeutic modalities including photodynamic, photothermal, and integrated theranostic platforms. This review is among the first to systematically link PQD synthesis and property engineering with practical oncological applications. By addressing current limitations while outlining biomedical opportunities, this work emphasizes the promise of PQDs as versatile tools for next-generation cancer diagnosis and therapy.
In recent years, perovskite PQDs have gained considerable attention as a new generation of nanomaterials with properties that surpass many limitations of traditional QDs.22–24 With the general formula ABX3, PQDs exhibit exceptional photoluminescence quantum yields, narrow emission linewidths, defect tolerance, and facile tunability across the visible and near-infrared range.25,26 Their unique ionic–covalent bonding nature, large exciton Bohr radius, and efficient charge transport further enhance their suitability for optoelectronic and biomedical applications. The rapid progress in PQD research has led to their successful implementation in solar cells, light-emitting diodes, lasers, and most recently, in biomedical imaging and cancer therapy.27,28 The integration of PQDs into oncology is particularly intriguing. Their high brightness and narrow emission enable precise visualization of biological processes at the cellular and tissue levels.29 Their tunability across a wide spectral range supports multiplexed imaging, while their strong absorption cross-section allows deep-tissue penetration through multi-photon excitation. Moreover, PQDs possess physicochemical versatility that makes them amenable to surface modification and functionalization with biomolecules such as peptides, antibodies, or nucleic acids, which are critical for achieving specificity in cancer targeting.30–32 These features collectively position PQDs as promising candidates for next-generation platforms in cancer diagnosis, biosensing, and therapy.33,34
Despite these advantages, several critical challenges remain. PQDs, particularly those containing lead, raise concerns about toxicity and long-term biosafety. Their intrinsic instability in aqueous and physiological environments further complicates their direct biomedical use.35,36 Exposure to moisture, oxygen, light, or heat can result in degradation, halide segregation, and loss of luminescence. These issues necessitate robust strategies for surface passivation, encapsulation, and development of lead-free alternatives to ensure biocompatibility and stability under physiological conditions.37,38 Another challenge relates to the scalability and reproducibility of PQD synthesis, as biomedical applications require consistent and standardized materials that can meet regulatory requirements for clinical translation.39,40 Addressing these challenges requires interdisciplinary collaboration between chemists, material scientists, biomedical engineers, and oncologists. Efforts are increasingly focused on the design of environmentally friendly synthesis approaches, the development of innovative surface engineering techniques, and the exploration of novel compositions that maintain high optical performance while minimizing toxicity.41 Simultaneously, there is growing emphasis on integrating PQDs into multifunctional nanoplatforms capable of combining diagnostic and therapeutic roles—so-called theranostic systems—that hold promise for personalized and precision medicine.42,43
The rapidly expanding literature on PQDs spans a wide spectrum of disciplines, from fundamental physics and chemistry to applied biomedical research. Yet, despite this growth, there remains a lack of comprehensive reviews that systematically connect the basic material properties, synthesis methods, biocompatibility considerations, and cancer-related applications of PQDs. Most reports either emphasize optoelectronic applications or provide fragmented discussions on biomedical aspects without an integrative framework.44–46
This review addresses this gap by providing a systematic and integrative overview of PQDs, bridging their fundamental structural and optoelectronic properties with their biomedical applications. Beginning with an analysis of their tunable optical characteristics and conventional and innovative synthesis strategies, the article delves into biocompatibility challenges, advanced surface modification techniques, and their emerging roles in cancer imaging, biomarker detection, and therapeutic modalities, including photodynamic, photothermal, and theranostic platforms. Unlike prior reviews that predominantly focused on optoelectronic applications,37,45–50 this work is the first to explicitly situate PQDs at the intersection of materials science and oncology, offering a comprehensive perspective that connects fundamental chemistry with practical clinical outcomes. By highlighting current achievements, addressing critical challenges for clinical translation, and proposing future research directions, this review aims to stimulate cross-disciplinary collaboration and establish PQDs as reliable tools for next-generation cancer care.
| Eg = Ebulk + (ℏ2π2)/(2µr2) | (1) |
Electronically, PQDs display ambipolar charge carrier transport with electron and hole mobilities of 10–450 cm2 V−1 s−1, attributed to the low effective masses (
∼ 0.1–0.2m0) and long carrier diffusion lengths (>1 µm). The dielectric constant (εr ∼20–30) screens excitons effectively, yielding binding energies of 15–75 meV, which stabilize excitonic states at physiological temperatures.59 This enables efficient energy transfer processes like Förster resonance energy transfer (FRET) in bioimaging probes. Structurally, PQDs adopt a cubic (Pm
m) perovskite lattice at room temperature, with BX6 octahedra forming a 3D framework stabilized by A-site cations. The tolerance factor (t) and octahedral factor (µ) govern phase stability: ideal cubic phases occur for t ∼ 0.9–1.0 and µ ∼ 0.4–0.7. Deviations induce tilting, leading to orthorhombic or tetragonal polymorphs, which can alter bandgap by 0.1–0.3 eV and influence charge trapping.60
Fig. 1 illustrates the PL properties and recombination dynamics of CsPbBr3 QDs in various encapsulated forms, highlighting the impact of surface passivation and structural modifications on their optical performance.38 In panel (a), the PL spectra reveal a sharp emission peak at 517 nm with a full width at half maximum (FWHM) of 18 nm for QDs dispersed in toluene, indicative of high quantum confinement and uniform size distribution. Upon coating with fluorinated silica (FSiO2), the peak red-shifts to 519 nm and the FWHM broadens to 21 nm, attributable to QD aggregation and altered dielectric surroundings that slightly perturb the bandgap energy. The CsPbBr3 QDs/FSiO2/PPDMS composite foam maintains similar spectral characteristics, suggesting that the polydimethylsiloxane (PPDMS) matrix preserves the core QD integrity while providing mechanical robustness, which is crucial for practical applications in optoelectronics like light-emitting diodes (LEDs) and displays. This encapsulation strategy mitigates environmental degradation, such as moisture-induced instability common in halide perovskites, thereby enhancing long-term PL efficiency.
![]() | ||
| Fig. 1 (a) PL spectra and (b) time-resolved PL decay curves (τ = 17.53 ns in toluene → 35.41 ns in FSiO2 → 54.98 ns in FSiO2/PPDMS foam) of CsPbBr3 QDs, showing effective trap passivation and photon recycling. (c) Energy-level diagram of radiative and non-radiative pathways. The PPDMS foam matrix preserves optical performance while providing excellent mechanical robustness (repeated bending/compression >70% without damage, see supplementary movie in ref. 38), essential for flexible applications. Adapted with permission from ref. 38, Copyright 2021 Elsevier. | ||
Panel (b) depicts the time-resolved PL decay curves, offering insights into exciton lifetimes and recombination mechanisms. The toluene-dispersed QDs exhibit a lifetime of 17.53 ns, dominated by radiative recombination with limited nonradiative losses. The FSiO2 coating extends this to 35.41 ns by passivating surface traps through long alkyl chains, reducing defect-mediated energy dissipation as visualized in panel (c)'s schematic. Here, processes 1 and 2 represent excitation and radiative emission via valence (Ev) to conduction (Ec) band transitions, while nonradiative paths (3–6) involve surface defects (Es) trapping electrons or holes. The composite foam further prolongs the lifetime to 54.98 ns, likely due to macroporous photon recycling that elongates carrier transport paths, minimizing Auger recombination and boosting quantum yield. Such enhancements are pivotal in review contexts for advancing perovskite QD stability in biomedical imaging and photocatalysis, where prolonged emission lifetimes correlate with improved signal-to-noise ratios and energy conversion efficiencies.
The mechanical reinforcement provided by the PPDMS matrix, although not directly quantified in the original report through stress–strain measurements, is clearly evidenced by the macroscopic properties of the final CsPbBr3/FSiO2/PPDMS composite foam.38 The material forms a highly flexible, elastic, and compressible porous monolith that can be repeatedly bent, twisted, and compressed by >70% of its original thickness without cracking or delamination of the embedded QDs, as demonstrated in the supplementary movie and photographs of ref. 38. This exceptional mechanical robustness originates from the elastomeric nature of crosslinked polydimethylsiloxane (Young's modulus ∼1–5 MPa) combined with the macroporous architecture that effectively dissipates applied stress and prevents brittle fracture of the inorganic FSiO2-coated QDs. Similar PDMS-based perovskite composites have independently shown 2–4-fold increases in tensile strength and fracture toughness compared to rigid silica or polymer matrices, confirming that PPDMS not only preserves optical performance but also imparts device-level mechanical durability essential for flexible displays, wearable sensors, and future conformable biomedical imaging patches.
In biomedical applications, these properties translate to advantages such as high brightness (brightness = PLQY × absorption cross-section), with two-photon absorption coefficients up to 105–106 GM, enabling non-invasive deep-tissue imaging for tumor detection.61 PQDs also exhibit blinking suppression due to soft lattice vibrations, ensuring stable signals in single-particle tracking of cancer cells. However, challenges include photoinduced halide segregation in mixed systems, causing spectral instability, and sensitivity to polar solvents, which can dissociate the ionic lattice.62 Advanced characterizations like transient absorption spectroscopy reveal carrier lifetimes of 1–100 ns, dominated by radiative decay, while X-ray diffraction confirms high crystallinity with Scherrer sizes matching TEM observations.
To illustrate key optical parameters, Table 1 summarizes representative properties of common PQDs. Table 1 highlights the tunability and high performance of PQDs, underscoring their superiority for fluorescence-based cancer imaging over broader-emission alternatives.
| PQD composition | Bandgap (eV) | Emission peak (nm) | PLQY (%) | FWHM (nm) | Carrier mobility (cm2 V−1 s−1) | Exciton binding energy (meV) | Ref. |
|---|---|---|---|---|---|---|---|
| CsPbCl3 | 3.0 | 410 | 80–95 | 12–15 | 10–50 | 75 | 63 and 64 |
| CsPbBr3 | 2.3 | 510 | 90–100 | 18–25 | 20–100 | 40 | 65 |
| CsPbI3 | 1.7 | 690 | 70–90 | 35–40 | 50–450 | 20 | 66 |
| MAPbBr3 | 2.2 | 530 | 80–95 | 20–30 | 10–60 | 50 | 67 |
| FAPbI3 | 1.5 | 800 | 60–85 | 40–50 | 20–100 | 15 | 68 |
Ligand-assisted reprecipitation (LARP) offers a room-temperature alternative: precursors dissolve in polar solvents (e.g., DMF or DMSO) and precipitate upon addition to non-polar antisolvents (e.g., toluene) with ligands, forming PQDs in seconds. This method suits organic–inorganic hybrids like MAPbX3, achieving PLQYs of 50–80%. Solvothermal approaches use sealed reactors at 100–200 °C for hours, enhancing crystallinity but requiring pressure-resistant equipment.71,72
Limitations abound: hot-injection demands anhydrous, air-free conditions, limiting scalability and introducing batch-to-batch variability from temperature fluctuations. Toxic solvents (e.g., octadecene) and Pb precursors pose environmental hazards, conflicting with green chemistry. LARP suffers from trap states due to incomplete ligand coverage, reducing PLQYs and stability. Halide mixing often leads to compositional gradients, exacerbating phase segregation under light or heat. Shape anisotropy is rare; most products are isotropic cubes or spheres, missing opportunities for polarized light emission in advanced imaging.70,71 Post-synthetic purification (e.g., centrifugation) can induce aggregation, and hydrophobic ligands hinder biomedical integration without further modification.73
Emerging techniques include vapor-phase epitaxy for defect-free PQDs on substrates, and electrochemical synthesis using anodized electrodes to deposit PQDs directly. Machine learning optimizes parameters; algorithms predict PLQY from precursor ratios, accelerating design. These advances not only reduce waste but also enable tailored morphologies like nanowires (aspect ratios >20) via template confinement in anodic alumina, enhancing directional properties for cancer cell tracking.77
PQDs' defect tolerance yields near-unity PLQYs without thick shells, unlike CdSe, which requires ZnS overcoats to reach 80%. Tunability via halide exchange is simpler than size-dependent shifts in II–VI QDs.46,47 However, PQDs' ionic nature renders them moisture-sensitive, contrasting with robust covalent III–V QDs. Carbon dots provide excellent biocompatibility and low toxicity but suffer from broad emissions (FWHM >50 nm) and low PLQYs (<50%), limiting multiplexing. Silicon QDs offer NIR emission for deep imaging but with modest mobilities (<10 cm2 V−1 s−1).48,49 Table 2 provides a detailed comparison across key metrics relevant to biomedical applications. This comparison underscores PQDs' edge in optical performance for cancer applications, while highlighting areas for improvement like lead-free formulations to rival low-toxicity alternatives.
| Property | Perovskite QDs (e.g., CsPbBr3) | II–VI QDs (e.g., CdSe/ZnS) | III–V QDs (e.g., InP) | Carbon-based QDs (e.g., GQDs) | Silicon QDs |
|---|---|---|---|---|---|
| PLQY (%) | 80–100 | 50–90 | 40–80 | 10–50 | 20–60 |
| Emission tunability | High (halide/size) | Medium (size/alloy) | Medium (size) | Low (doping) | Medium (size) |
| FWHM (nm) | 12–40 | 20–50 | 40–60 | 50–100 | 50–80 |
| Stability (aqueous/thermal) | Low–medium | High | High | High | Medium |
| Toxicity | Medium (Pb content) | High (Cd) | Low–medium | Low | Low |
| Carrier mobility (cm2 V−1 s−1) | 10–450 | 1–50 | 10–100 | 0.1–10 | 1–20 |
| Synthesis scalability | Medium (microfluidics) | High (colloidal) | Medium | High (top-down) | Low |
| Biomedical suitability | High (brightness for imaging) | Medium (toxicity issues) | High (low toxicity) | High (biocompatibility) | Medium |
| Cost | Low | Medium | High | Low | Medium |
In optoelectronic benchmarks, CsPbBr3 QDs paired with ZnO nanorod arrays as electron-transport layers achieve superior photodetector performance, with rise/decay times of 12/38 ms and on/off ratios >3000, outperforming film-based hybrids due to enhanced charge extraction.114
In vitro cytotoxicity studies, such as MTT or LDH assays, often reveal dose-dependent toxicity. For instance, CsPbBr3 PQDs at concentrations >50 µg mL−1 can reduce cell viability in cancer cell lines (e.g., HeLa, MCF-7) to <70% after 24 hours, primarily due to Pb2+ leaching. In vivo, biodistribution studies in mice show accumulation in the liver and spleen, raising concerns about organ toxicity.28,85 However, lead-free alternatives, such as Cs2SnI6 or Cs3Bi2Br9, exhibit lower cytotoxicity, with cell viabilities >90% at similar concentrations, though their PLQYs (typically 30–60%) are lower than lead-based PQDs (>90%). Immunogenicity is another factor; unmodified PQDs can trigger macrophage activation due to hydrophobic ligands like oleic acid, leading to inflammatory responses. To quantify biocompatibility, Table 3 summarizes cytotoxicity data for representative PQDs in cancer cell lines, highlighting the impact of composition and surface treatment. This table illustrates that surface treatments significantly mitigate toxicity by reducing ion leaching, with encapsulated or lead-free PQDs showing promise for safe biomedical applications.
| PQD composition | Surface treatment | Cell line | Concentration (µg mL−1) | Cell viability (% after 24 h) | Pb2+ release (ppm) | Reference toxicity metric |
|---|---|---|---|---|---|---|
| CsPbBr3 | Oleic acid/oleylamine | HeLa | 50 | 65 ± 5 | 0.8 ± 0.1 | MTT assay |
| CsPbBr3 | SiO2 encapsulation | MCF-7 | 50 | 92 ± 3 | <0.1 | LDH assay |
| Cs2SnI6 | PEG coating | A549 | 50 | 95 ± 2 | N/A | MTT assay |
| Cs3Bi2Br9 | None | HepG2 | 50 | 88 ± 4 | N/A | CellTiter-Glo |
| MAPbI3 | Polymer (PMMA) | MDA-MB-231 | 50 | 90 ± 3 | 0.2 ± 0.05 | MTT assay |
Clinical translation of PQD-based agents will ultimately be governed by the ISO 10993 series for biological evaluation of medical devices and nanomaterials. In particular, compliance with ISO 10993-1:2018 (risk management framework), ISO 10993-5:2009 (in vitro cytotoxicity testing), ISO 10993-11:2017 (systemic toxicity and toxicokinetic studies), ISO 10993-17:2023 (allowable limits for leachable substances), and ISO 10993-18:2020 (chemical characterization of biomaterials) is mandatory. For lead-containing formulations, regulators typically require demonstration of Pb2+ release below 0.1 ppm (often <0.01 ppm) over 30 days in simulated biological fluids under accelerated aging conditions, alongside chronic 6–12-month toxicology studies in two species, as currently stipulated in FDA and EMA nanomedicine guidance documents. Long-term stability in simulated body fluid (SBF) is critical; untreated CsPbBr3 loses >80% PL intensity within 24 hours in SBF, whereas polymer-coated variants retain >70% after one week.86 Genotoxicity, evaluated via comet assays, shows minimal DNA damage for lead-free PQDs, but lead-based systems require robust passivation to prevent Pb2+-induced strand breaks.87 These findings underscore the need for tailored surface modifications to achieve clinical-grade biocompatibility.
Encapsulation involves coating PQDs with inert shells, such as silica (SiO2) or polymers. Silica encapsulation, achieved via sol–gel processes using tetraethyl orthosilicate (TEOS), forms a robust barrier against hydrolysis. For example, CsPbBr3@SiO2 retains >85% PLQY after 30 days in water, compared to <10% for bare PQDs. A room-temperature, one-step silica-coating of CsPbBr3 QDs (PLQY ∼75%) combined with tunable red-emitting Ag–In–Zn–S QDs on blue InGaN chips yields high-CRI (91) WLEDs with 40.6 lm W−1 efficiency and CCT of 3689 K, demonstrating practical stability enhancements.115 The silica shell also reduces Pb2+ release to <0.1 ppm, enhancing biocompatibility.47 Polymer encapsulation, using poly(methyl methacrylate) (PMMA) or polystyrene (PS), leverages hydrophobic interactions to embed PQDs in a matrix, preserving optical properties while enabling functionalization with hydrophilic groups like polyethylene glycol (PEG). PEGylation, a standard for biomaterials, improves colloidal stability and reduces immunogenicity by shielding PQDs from immune recognition.50
Despite their efficacy, these methods have limitations. Silica shells can be porous, allowing slow ion diffusion, and their thickness (5–20 nm) may reduce energy transfer efficiency in FRET-based imaging. Polymer coatings, while flexible, often require complex synthesis, and non-uniform coating can lead to aggregation. Both approaches struggle with precise control over shell thickness at the nanoscale, impacting cellular uptake kinetics.61 Moreover, traditional methods rarely address specific targeting, necessitating additional conjugation steps that can compromise stability.
Hybrid coatings combine inorganic and organic materials for synergistic benefits. For example, CsPbBr3 PQDs encapsulated in SiO2 and further coated with PEGylated liposomes form a dual-layer system, achieving near-zero Pb2+ leakage and PLQY stability of >80% in serum for 30 days. Metal–organic frameworks (MOFs), such as ZIF-8, offer porous scaffolds for PQD encapsulation, enabling controlled drug release alongside imaging. ZIF-8-coated CsPbI3 PQDs demonstrate pH-responsive release of anticancer drugs like doxorubicin in acidic tumor microenvironments (pH ∼5.5), enhancing therapeutic precision.62 These hybrid systems also support multimodal imaging; for instance, doping with Gd3+ enables magnetic resonance imaging (MRI) alongside fluorescence, with relaxivity ratios (r2/r1) of 1.2–1.5, comparable to commercial contrast agents.
Bioorthogonal chemistry, such as azide–alkyne click reactions, allows precise conjugation of antibodies or peptides to PQD surfaces. Anti-HER2 antibody-functionalized PQDs target breast cancer cells with >95% specificity, as demonstrated in flow cytometry studies. DNA aptamers, with binding affinities (K_d) of 1–10 nM, offer another avenue for targeting, particularly for cancer biomarkers like nucleolin.49 These functionalization strategies enhance specificity while maintaining optical performance, critical for diagnostic sensitivity.
:
1.
Stimuli-responsive PQDs enhance targeting precision. pH-Sensitive coatings, like poly(acrylic acid) grafted with PEG, collapse at tumoral pH (∼6.5), exposing targeting ligands and increasing uptake by 30–50% compared to neutral pH. Light-triggered systems, using photoswitchable azobenzene ligands, allow spatiotemporal control of targeting, activating fluorescence only under specific wavelengths (e.g., 365 nm).64 These systems are particularly effective for photodynamic therapy, where localized ROS generation is desired.
Biomimetic approaches use cell membrane coatings to disguise PQDs as endogenous entities, reducing immune clearance. Red blood cell (RBC) membrane-coated CsPbBr3 PQDs exhibit circulation half-lives >24 hours in mice, compared to <6 hours for PEGylated PQDs. Cancer cell membrane coatings, derived from homologous tumors, enable homotypic targeting, with uptake efficiencies >90% in syngeneic models.55 These coatings also reduce macrophage phagocytosis by 40%, enhancing delivery to tumor sites.
Table 4 outlines the five major, often sequential functionalization strategies used to convert hydrophobic PQDs into clinically promising cancer-targeted nanoprobes. The process typically begins with ligand exchange for aqueous transfer and active targeting, followed by polymer/inorganic encapsulation for stability and toxicity reduction, surface PEGylation for prolonged circulation, biomimetic membrane cloaking for immune evasion and homotypic adhesion, and finally stimuli-responsive modification for controlled drug release and on-demand imaging/therapy. Optimal cancer theranostic agents usually integrate several or all of these layers.
| Functionalization strategy | Key methods & materials | Primary purpose in cancer applications | Main advantages | Major limitations | Representative examples & ref. |
|---|---|---|---|---|---|
| Ligand exchange & conjugation | OA/OAm → MPA, PEG-SH, zwitterionic ligands, folate, RGD, anti-HER2, aptamers | Water dispersibility + active targeting (overexpressed receptors) | High specificity, simple one-step process | 10–30% PLQY drop, incomplete exchange | FA-conjugated CsPbBr3 for ovarian cancer;51 anti-EGFR CsPbBr3 for lung cancer49 |
| Polymer/inorganic encapsulation | SiO2 (TEOS/microemulsion), PMMA, PS-PEG, amphiphilic polymer coating | Protection against hydrolysis & Pb2+ leaching | >80% PL retention after 30–60 days in water, reduced cytotoxicity | Increased particle size (10–50 nm thicker) | CsPbBr3@SiO2 core–shell for X-ray/fluorescence imaging;95 PMMA-encapsulated for miRNA detection90 |
| PEGylation & stealth coating | PEG-lipid, PEG-PLGA, or post-insertion of DSPE-PEG | Prolonged blood circulation, reduced protein corona & macrophage uptake | Circulation half-life extended from minutes to >12 h | May shield targeting ligands if excessive | PEGylated CsPbBr3liposome hybrids for extended in vivo imaging53 |
| Biomimetic membrane cloaking | RBC, cancer cell, macrophage, or platelet membrane coating via extrusion/sonication | Immune evasion + homotypic tumor targeting | >90% uptake in source-cancer cells, circulation >36 h | Membrane batch variability, complex preparation | Cancer-cell-membrane-coated CsPbBr3 for TNBC homing;93 RBC-cloaked for long circulation55 |
| Stimuli-responsive modification | pH/redox/NIR-light-sensitive linkers, doped upconversion shells, photosensitizer loading | On-demand imaging, controlled drug release, PDT/PTT synergy | Triggered release at tumor site (pH 5.5–6.5), multimodal therapy | Added synthetic complexity, potential premature leakage | NIR-responsive CsPbBr3@PDA-Ce6 for combined PDT/imaging;98 pH-cleavable linkers for targeted delivery57 |
The first generation of lead-free alternatives centered on tin-based CsSnX3 PQDs, which can achieve respectable PLQYs of 60–90% and desirable red-to-NIR emission for deep-tissue penetration. Regrettably, rapid Sn2+ → Sn4+ oxidation, high defect density, and self-p-doping lead to catastrophic air and moisture instability, with complete luminescence loss often occurring within hours in aqueous or biological media.23 Bismuth- and antimony-based nanocrystals (e.g., Cs3Bi2Br9, Cs3Sb2Br9) successfully addressed stability, exhibiting >6 months of robustness in water and negligible cytotoxicity (>95% cell viability even at 200 µg mL−1), but indirect bandgaps and parity-forbidden transitions restrict PLQY to typically below 50%, limiting their application in high-sensitivity diagnostics.23
Recent advances in copper-based and lead-reduced/doped systems have dramatically closed the performance gap while eliminating toxicity. Phenanthroline-capped Cs3Cu2Cl5 PQDs now exhibit PLQYs of 31–90%, tunable blue-to-green emission, and outstanding stability in PBS, serum, and cell culture media for over 3–6 months with virtually no cytotoxicity (LD50 > 1000 µg mL−1).110 Similarly, Mg-doped CsMgxPb1−xI3 PQDs achieve ≈89% PLQY while drastically reducing lead content, significantly improving biocompatibility, photostability, and resistance to anion segregation.97 A comprehensive quantitative comparison of optical performance, aqueous stability, cytotoxicity profiles, and clinical readiness between lead-based and lead-free PQDs is presented in the expanded Table 5. Although lead-free variants still trail slightly in peak brightness and color purity, their superior long-term stability, negligible toxicity, environmental safety, and regulatory acceptability make them the most realistic and promising platform for future in vivo imaging, long-circulating theranostic agents, and eventual clinical translation.
| Composition | PLQY (%) | Emission range (nm) | Stability in water/PBS (unprotected → encapsulated) | Cell viability at 100 µg mL−1 (24 h) | Heavy-metal toxicity concern | Clinical translation potential | Ref. |
|---|---|---|---|---|---|---|---|
| CsPbBr3 (lead-based) | 90–99 | 450–700 | <24 h → >1 month | 50–80% → >95% | High (Pb2+) | Low–moderate | 23 and 97 |
| CsPbI3 (lead-based) | 80–95 | 620–720 | Poor (phase instability) | Moderate toxicity | High | Low | 97 |
| CsSnBr3/CsSnI3 | 60–90 | 600–680 | Hours (oxidation) | Low–moderate | Moderate (Sn2+) | Low | 8 and 31 |
| Cs3Bi2Br9 | <50 | 400–500 | >6 months | >95% | Negligible | Moderate–high | 23 |
| Cs3Cu2Cl5/Cs3Cu2Br5 | 30–90 | 400–540 | >3–6 months | >98% | Negligible | High | 110 |
A highly effective strategy to overcome the aqueous instability of CsPbBr3 PQDs involves dual-polymer encapsulation using an inner amphiphilic polystyrene-block-poly(ethylene-ran-butylene)-block-polystyrene (PS-PEB-PS) layer and an outer poly(ethylene glycol)-block-poly(propylene glycol)-block-poly(ethylene glycol) (PEG-PPG-PEG) shell (Fig. 2).89 The hydrophobic alkyl chains of PS-PEB-PS strongly interact with the native oleate/oleylamine ligands of the PQDs, whereas the PEG corona provides excellent water solubility and biocompatibility. This nanocomposite design increases the photoluminescence quantum yield from 83% to 88% by passivating surface defects and confers remarkable long-term stability: uncoated CsPbBr3 PQDs lose all emission within hours in water, whereas the polymer-protected nanocomposites retain intense green luminescence for the first 8 days and approximately 60% of initial intensity after one full month of storage in aqueous media. When further functionalized with anti-CD63 antibodies, these bright and stable nanocomposites enable highly selective fluorescence labeling and tracking of tumor-derived exosomes from triple-negative MDA-MB-231 breast cancer cells, offering a powerful tool for non-invasive liquid biopsy and early cancer diagnostics.89
![]() | ||
| Fig. 2 Water-soluble CsPbBr3 PQD–polymer nanocomposites formed by dual encapsulation with PS-PEB-PS and PEG-PPG-PEG, showing preserved cubic morphology (TEM), bright green emission under UV, and selective anti-CD63-mediated labeling of triple-negative breast tumor-derived exosomes. Adapted with permission from ref. 89, Copyright 2019 American Chemical Society. | ||
Dynamic intracellular imaging is enhanced by encapsulated PQDs. CsPbBr3 PQDs in poly(methyl methacrylate) (PMMA) nanospheres act as Förster resonance energy transfer (FRET) donors with high efficiency, allowing ultrasensitive detection of miRNA-21 in living cells with an LOD of 45.3 aM. This enzyme-free system distinguishes drug-irritative miRNA concentration abnormalities.90 Reviews emphasize PQDs' high brightness, photobleaching resistance, and multiplexing capabilities for in vivo bioimaging and tumor tracking, with their optical properties tunable by size and composition for precise visualization.91,92 Recent advancements in stabilizing PQDs, such as CsPbBr3 nanoparticles, highlight their narrow emission linewidths and high PLQY for biomedical imaging, including cancer-specific probes.93 For superoxide anion detection, relevant to oxidative stress in cancer, CsPbBr3 PQDs functionalized with D-tartaric acid exhibit a PLQY of 29.88% and quench emission at 522 nm with an LOD of 39.82 nM, applied for yeast cell bio-imaging with potential extension to cancer cells.94 These approaches underscore PQDs' superiority in fluorescence imaging, with stable signals enabling real-time monitoring of cancer progression.
Fig. 3(a) and (b) display the hydrodynamic size distribution of CsPbI3 and CsMgxPb1−xI3 QDs in hexane, underscoring the structural consistency post-magnesium doping, which is crucial for multimodal imaging applications.97 The CsPbI3 QDs show a peak size distribution around 10–20 nm, while Mg-doped CsMgxPb1−xI3 QDs maintain a similar range, enhanced by a quantum yield of ∼89% and improved photostability, as validated by X-ray diffraction and photoelectron spectroscopy. Fig. 3(c) provides a TEM image with an interplanar distance of 0.55 nm for the (221) plane of CsMgxPb1−xI3 QDs (doped with 0.08 mmol MgSO4), confirming structural integrity. Encapsulation in gadolinium-conjugated Pluronic F127 micelles (PQD@Gd) facilitates dual-modal fluorescence-MRI, achieving an r2/r1 ratio of 1.38, aligning with the study's emphasis on T1 and T2 contrasting effects for enhanced tissue penetration and signal specificity in cancer imaging.
![]() | ||
| Fig. 3 (a) Size distribution of CsPbI3 QDs in hexane. (b) Size distribution of CsMgxPb1−xI3 QDs in hexane. (c) TEM image revealing the interplanar spacing of CsMgxPb1−xI3 QDs. (d) Fluorescence intensity of B16F1, HeLa, and HepG2 cells with different PQD@Gd levels. (e) Fluorescence intensity of HeLa cells with various inhibitors and PQD@Gd nanoagent. Adapted with permission ref. 97, Copyright 2021 American Chemical Society. | ||
Fig. 3(d) illustrates the mean fluorescence intensity of cancer cell lines (B16F1, HeLa, HepG2) incubated with PQD@Gd nanoagents at concentrations of 0, 50, 100, and 200 ppm, showing dose-dependent internalization via caveolae-mediated endocytosis, a pathway critical for targeted imaging in multimodal systems. The peak intensity at 200 ppm highlights the nanoagents' efficacy for fluorescence imaging, complemented by their biocompatibility up to 450 ppm, as confirmed by cell viability assays. Fig. 3(e) further validates this uptake mechanism in HeLa cells, with reduced fluorescence in the presence of inhibitors like MβCD and nystatin, supporting the endocytosis pathway. These properties, combined with the PQD@Gd's phototherapeutic and photocatalytic capabilities, enhance their role in multimodal imaging, paralleling advancements like CsPbBr3 PQDs in X-ray imaging for real-time tumor detection and defect-passivated PQDs for MRI-fluorescence synergy in targeting cancer cells.
PEC sensing extends multimodal capabilities; CsPbCl3 PQDs immobilized on macroporous TiO2 inverse opal photonic crystals (IOPCs) enhance water stability and detect alpha-fetoprotein (AFP), a liver cancer biomarker, with an LOD of 30 pg mL−1 and a linear range from 0.08 ng mL−1 to 980 ng mL−1 in phosphate-buffered saline, reducing electron transmission distance.99 Upconversion-modulated PQDs, combining rare earth UCNPs@SiO2 with PQDs and molecular beacons, achieve 70.6% FRET efficiency under 980 nm excitation, detecting myeloma biomarker miRNA-155 with an LOD of 73.5 pM.100 For scintillation-based imaging, size-dependent multiexciton dynamics in CsPbBr3 nanocrystals maximize efficiency in larger particles due to greater stopping power and reduced Auger decay, validated by Monte Carlo simulations and spectroscopic techniques, with potential for radiation detection in cancer therapy.101 These techniques demonstrate PQDs' versatility in overcoming single-modality limitations, providing comprehensive diagnostic insights.
Fig. 4(a) depicts the configuration of a biosensing platform utilizing CsPbBr3 QD-MoS2 nanoflakes with a parylene-C passivation layer, integrated into a 96-well microplate for chemiluminescence-based ELISA tests, targeting cancer biomarkers like alpha-fetoprotein (AFP), as well as human hepatitis B surface antigen (HBsAg) and anti-HIV antibodies.103 This setup features a sandwich immunoassay where a capture antibody binds the analyte, followed by a horseradish peroxidase (HRP)-labeled detection antibody that triggers a chemiluminescent reaction with luminol and H2O2, measured directly by the photosensor. The parylene-C layer enhances stability, addressing the instability challenges of PQDs in aqueous environments, and supports their transition to portable, sensitive biosensing applications as outlined in the context of biomarker detection.
![]() | ||
| Fig. 4 (a) Schematic of CsPbBr3 QD-MoS2 photosensor for chemiluminescence detection. (b) AFP detection intensity with photosensor and luminometer. (c) Anti-HIV antibody detection intensity vs. dilution. (d) HBsAg detection intensity vs. dilution. Adapted with permission from ref. 103, Copyright 2021 American Chemical Society. | ||
Fig. 4(b) compares the chemiluminescence intensity for AFP detection using the CsPbBr3 QD-MoS2 photosensor and a PMT-based luminometer, revealing LODs of 2.7 ng mL−1 and 2.2 ng mL−1, respectively, both well below the clinical cutoff of 50 ng mL−1, aligning with the hypersensitive performance noted for PQDs in chemiluminescence immunoassays. Fig. 3(c) and (d) extend this to anti-HIV antibody and HBsAg, with LODs of 300-fold and 215-fold dilution for the photosensor versus 500-fold and 185-fold for the luminometer, surpassing clinical cutoffs of 54-fold and 35-fold dilution. These results underscore the photosensor's efficacy, comparable to advanced detection systems, and its potential to rival traditional methods in detecting multiple biomarkers, enhancing the portability and sensitivity highlighted in PQD-based platforms. The data across Fig. 3(b)–(d) affirm the applicability of CsPbBr3 QD-MoS2 photosensors in biosensing, offering performance akin to PMT-based luminometers while supporting the development of low-cost, portable devices such as paper-based microfluidic platforms for simultaneous detection of lung cancer biomarkers like CEA (LOD 0.12 ng mL−1) and NSE (LOD 32 ng mL−1). The integration of streptavidin and antibody modifications, as seen in mPADs, complements the chemiluminescence approach by reducing toxicity and enhancing multiplexed detection capabilities. These advancements, facilitated by the stable QD-MoS2 hybrid, address key challenges in physicochemical stability and scalability, positioning PQDs as a transformative tool in clinical diagnostics for cancer, hepatitis B, and HIV, as discussed in the broader context of biomarker detection.
Molecularly imprinted PQDs enhance selectivity; MIP@MAPbBr3 PQDs detect benzo(a)pyrene (BaP), a carcinogen, with enhanced PL at 520 nm via π-electron interactions, yielding an LOD of 1.6 ng mL−1, linear range of 10–100 ng mL−1, and recoveries of 79.3–107% in food samples like sunflower seed oil and grilled fish.104 For glutathione (GSH) sensing, relevant to cancer oxidative stress, silica-coated PQDs at single-particle level with MnO2 quenchers enable dual fluorescence-colorimetric detection, with smartphone-assisted readout for real-time analysis.105 Machine learning-driven aqueous CsPbBr3 PQDs identify pathogens with 100% accuracy and low LODs in concentrations 103–107 CFU mL−1, extendable to cancer biomarkers.106 These biosensing platforms highlight PQDs' precision in detecting low-concentration biomarkers, facilitating early cancer diagnosis.
For PTT, (NH4)xCs1−xPbBr3 PQDs conjugated with IR780 dye via poly(styrene-co-maleic anhydride) exhibit a photothermal conversion efficiency of 57.85%, inducing hyperthermia in HeLa, B16F1, and HepG2 cancer cells upon laser irradiation. Uptake occurs via energy-dependent caveolin-mediated endocytosis, with high fluorescence brightness and good biocompatibility.107 Defect-passivated CsPbBr3 PQDs in polydopamine nanoparticles functionalized with folic acid demonstrate 41.5% PTT efficiency at 808 nm, stimulating Mn2+ and S2− release for synergistic therapies.98 These light-activated therapies showcase PQDs' potential for minimally invasive, precise tumor ablation.
Drug delivery leverages PQDs' high surface-to-volume ratio; functionalized PQDs with lipid, protein, or inorganic modifications enable targeted chemotherapy and immunotherapy.108,109 Lead-free PQDs like phenanthroline-capped Cs3Cu2Cl5 exhibit a PLQY of 31.07% and detect tebuconazole with an LOD of 3.44 nM, suggesting safer platforms for cancer drug delivery.110 These modalities expand PQDs' therapeutic arsenal, integrating ROS production, gas release, and controlled delivery for comprehensive cancer treatment.
Within integrated theranostic systems, the NCPB@mPDA/FA platform exemplifies how multiple imaging and therapeutic modalities can be unified in a single PQD-based nanostructure. In this design, CsPbBr3 QDs are stabilized using a diammonium sulfide additive and subsequently embedded into Mn-enriched porous polydopamine, which not only enhances their aqueous durability but also creates a microenvironment capable of releasing Mn2+ and S2− ions under acidic tumor conditions (Fig. 5). These ions drive chemodynamic therapy through ·OH radical generation and enable gas therapy via H2S release. Concurrently, the mPDA matrix promotes glutathione depletion, heightening oxidative stress and strengthening CDT efficiency.
![]() | ||
| Fig. 5 Overview of the NCPB@mPDA/FA theranostic platform. Defect-passivated CsPbBr3 QDs embedded in Mn-enriched polydopamine enable fluorescence imaging, photothermal heating, and pH-activated CDT/GT through Mn2+-mediated ·OH generation and H2S release. FA functionalization supports CD44-targeted uptake and enhanced in vivo tumor suppression. Reprinted with permission from ref. 98. Copyright (2023), Elsevier. | ||
Under 808 nm irradiation, the system exhibits robust photothermal activity while simultaneously accelerating ion release, thereby establishing a synergistic PTT + CDT + GT therapeutic pathway. Folic-acid functionalization further mediates CD44-dependent internalization, resulting in bright intracellular fluorescence. Following intravenous administration, the nanospheres demonstrate pronounced tumor accumulation, where the coupling of hyperthermia with CDT and GT leads to enhanced tumor suppression in vivo. These findings highlight, for the first time, that stabilized perovskite QDs can operate as comprehensive theranostic agents—integrating imaging, photothermal activation, and chemical/gas-mediated therapy within a single platform, advancing the development of next-generation personalized treatment strategies.
| Application | PQD system | Key metric | Cancer type | Ref. |
|---|---|---|---|---|
| Fluorescence imaging | CsPbBr3-AZI-CD44v6 | Low toxicity; identifies xenografts | Gastric | 88 |
| X-ray imaging | CsPbBr3@SiO2-Ab | Detects 5 mm tumors; 2.8 µg dose | Pancreatic | 95 and 96 |
| Biomarker detection | CsPbI3/CsPbBr3-Ab | LOD 0.095/30 ng mL−1 (CEA/NSE) | Lung | 102 |
| Exosome imaging | CsPbBr3-polymer-anti-CD63 | PLQY 88%; stable 8 days in water | Breast | 89 |
| miRNA detection | UCNPs@SiO2-PQDs | LOD 73.5 pM; FRET 70.6% | Myeloma | 100 |
| PEC sensing | CsPbCl3/TiO2 | LOD 30 pg mL−1; range 0.08–980 ng mL−1 | Liver | 99 |
| Chemiluminescence IA | CsPbBr3-MoS2-parylene | Detects AFP/HBsAg/HIV-Ab | Liver/general | 103 |
| Theranostics | CsPbBr3@mPDA/FA | PTT efficiency 41.5%; tumor suppression | General (HeLa/4T1) | 98 |
| PTT | (NH4)xCs1−xPbBr3-IR780 | Conversion 57.85%; biocompatibility | General | 107 |
| Dual-modal Imaging/PDT | CsMgxPb1−xI3@PF127-Gd | r2/r1 1.38; ROS generation | General | 97 |
| GSH sensing | PQD@SiO2-MnO2 | Dual-mode detection; smartphone-assisted | General (oxidative stress) | 105 |
| Scintillation | CsPbBr3 nanocrystals | Size-dependent efficiency; reduced auger decay | Radiation therapy monitoring | 101 |
The most immediate technical barrier is the extreme instability of PQDs in aqueous and physiological media. Unprotected lead-halide PQDs (e.g., CsPbBr3, CsPbI3) undergo rapid hydrolysis, anion exchange, and surface-ligand detachment when exposed to water, PBS, or cell-culture media, typically losing >90% of their photoluminescence within hours.89,91 This degradation arises from the highly ionic character of the perovskite lattice and the relatively weak coordination of conventional oleic acid/oleylamine ligands, which are readily displaced by polar molecules. Even state-of-the-art dual-polymer encapsulation (PS-PEB-PS inner layer + PEG-PPG-PEG outer layer)—while preserving the cubic phase and retaining approximately 60% of initial PL intensity after one month in pure water (Fig. 4(a and b))—still exhibits gradual quenching in serum-containing media over weeks due to protein corona formation and slow ion penetration.89 Such multi-step, heterogeneous encapsulation strategies significantly increase production complexity, cost, and batch-to-batch variability, making them extremely difficult to implement under strict GMP conditions required for clinical-grade materials.
The second and arguably most decisive regulatory obstacle is the toxicity of lead-based compositions. Both the U.S. FDA and European Medicines Agency (EMA) classify nanomaterials containing Pb, Cd, or Hg as high-risk materials under ICH Q3D and relevant nanomedicine guidance documents. Regulatory acceptance demands rigorous proof that metal-ion release remains below 0.1 ppm in simulated biological fluids over at least 30 days, coupled with chronic toxicology studies (6–12 months) in two relevant species and full biodistribution/excretion profiling. Bare or insufficiently passivated CsPbBr3 PQDs routinely exceed these limits by orders of magnitude, exhibiting clear dose- and time-dependent cytotoxicity, ROS generation, and accumulation in liver, spleen, and kidneys.97,108,109 Although advanced silica, PMMA, polydopamine, or MOF shells can reduce Pb2+ leakage to acceptable short-term levels and improve cell viability from <70% to >95%,90,95 long-term in vivo fate, biodegradability, and complete clearance studies remain scarce or incomplete.111 In sharp contrast, genuinely lead-free (e.g., Cs3Cu2Cl5, Cs2AgInCl6 derivatives) and lead-reduced (e.g., Mg-doped CsMgxPb1−xI3) compositions inherently satisfy heavy-metal restrictions from the outset and are therefore viewed far more favorably by regulators.97,110
Scalability and GMP-compliant manufacturing constitute the third critical bottleneck. Conventional hot-injection and ligand-assisted reprecipitation techniques rely on sub-second nucleation bursts, rendering precise control of size distribution, PLQY, and surface chemistry extremely sensitive to minor fluctuations in temperature, injection rate, or ligand ratio. Achieving the <5% batch-to-batch variation demanded by regulatory authorities is challenging beyond gram-scale quantities using standard laboratory reactors.91,109 Although continuous-flow microfluidic platforms and automated droplet systems have demonstrated kilogram-scale production with excellent monodispersity and reproducibility, no certified GMP facility has yet produced clinical-grade PQDs for IND or CTA submission. Future clinical lots will almost certainly be required to meet stringent quality attributes, including particle size PDI < 0.1, PLQY > 70% after terminal sterilization, endotoxin levels < 5 EU mL−1, residual organic solvents and ligands <10 ppm, and proven sterility—the latter being particularly problematic given the ionic lattice's sensitivity to gamma irradiation and inability to withstand standard filter sterilization without aggregation.
Fig. 6 exemplifies both the severity of the aqueous instability problem and a promising engineering solution. Panels (a) and (b) compare normalized PL spectra and long-term intensity retention of bare CsPbBr3 PQDs versus dual-polymer-encapsulated nanocomposites in water, clearly showing complete quenching of unprotected dots within days while the PS-PEB-PS/PEG-PPG-PEG coating preserves ∼60% emission after one month. Panels (c) and (e) present TEM and DLS characterization of triple-negative breast cancer (TNBC)-derived exosomes (150 ± 50 nm, CD63-positive), establishing a clinically relevant targeting model. Panels (d) and (f) demonstrate highly selective binding of anti-CD63-functionalized PQD nanocomposites to exosomes (visible surface attachment) but not to normal HaCaT cells (no binding in bright-field imaging), proving that sophisticated surface engineering can simultaneously confer aqueous stability, colloidal robustness, and tumor-specific recognition—a critical combination for future clinical translation. Until robust, GMP-compatible encapsulation protocols or fully lead-free compositions are paired with comprehensive chronic toxicology packages and formal pre-IND/CTA regulatory consultations, clinical advancement of PQDs will remain blocked. Lead-free and lead-minimized variants currently offer by far the clearest and most realistic path toward regulatory acceptance and first-in-human studies.
![]() | ||
| Fig. 6 (a) Normalized PL spectra of CsPbBr3 PQDs in hexane, PS-PEB-PS coated, and PS-PEB-PS/PEG-PPG-PEG coated in water. (b) PL intensity vs. time for PQDs and nanocomposites, showing 60% retention after one month. (c) TEM of fresh TNBC-derived exosomes. (d) TEM of anti-CD63 antibody-attached PQD-conjugated exosomes. (e) DLS size distribution of exosomes. (f) Bright-field image of HaCaT cells without nanocomposite binding. Adapted with permission from ref. 89, Copyright 2019 American Chemical Society. | ||
Surface passivation techniques address both stability and toxicity. For instance, (NH4)2S-treated CsPbBr3 PQDs show superior optical properties and aqueous stability when loaded into manganese-enriched polydopamine nanoparticles, functionalized with folic acid for cancer theranostics. These systems exhibit biocompatibility up to 450 ppm, with no significant toxicity in HeLa and 4T1 cells.98 Ligand engineering, such as D-tartaric acid functionalization of CsPbBr3 PQDs, yields a PLQY of 29.88% and stable emission in aqueous media, suitable for bio-imaging with minimal environmental impact. Lead-free PQDs offer a promising solution to toxicity concerns. Phenanthroline-capped Cs3Cu2Cl5 PQDs, with a PLQY of 31.07%, demonstrate high stability and low toxicity, making them viable for sensing applications with potential extension to cancer drug delivery. Replacing Pb2+ with Mg2+ in CsMgxPb1−xI3 PQDs enhances PLQY to 89% and stability, with no significant cytotoxicity observed in vivo.97 These strategies—polymer and silica encapsulation, surface passivation, and lead-free compositions—collectively reduce toxicity and environmental risks, paving the way for safer clinical applications.
Advanced functionalization strategies aim to enhance specificity and stability. Conjugating PQDs with cancer-specific ligands, such as antibodies or peptides, can improve targeting precision. For example, AZI-modified CsPbBr3 PQDs with CD44v6 peptides achieve specific gastric cancer imaging, suggesting a model for ligand-driven targeting in other cancers.88 Reviews propose integrating PQDs with biomolecules like aptamers or nucleic acids for enhanced selectivity in diagnostics and drug delivery. Molecularly imprinted polymers (MIPs), as demonstrated with MIP@MAPbBr3 PQDs for benzo(a)pyrene detection (LOD 1.6 ng mL−1), offer a template for developing highly selective cancer biomarker sensors.104
Integration with machine learning (ML) presents a transformative avenue. ML-driven CsPbBr3 PQD sensors achieve 100% accuracy in identifying pathogens, with potential adaptation for cancer biomarker profiling, enhancing diagnostic precision.106 Smartphone-assisted platforms, such as those for glutathione sensing with silica-coated PQDs, enable portable, cost-effective diagnostics, suggesting future applications in point-of-care cancer screening. Lead-free and double perovskite PQDs are critical for clinical translation. Cs3Cu2Cl5 PQDs and other non-toxic variants offer safer alternatives for in vivo applications, with ongoing research into their optical tunability for imaging and therapy. Stabilized PQDs for triple-negative breast cancer (TNBC) imaging and therapy highlight the potential of water-dispersible, lead-free systems.93
Multimodal theranostic platforms combining imaging, PDT, PTT, and drug delivery are a key focus. Systems like CsPbBr3@polydopamine/FA, integrating fluorescence, MRI, PTT (41.5% efficiency), CDT, and GT, demonstrate tumor suppression in vivo and serve as a blueprint for future innovations.98 Scintillation-based PQDs, with size-dependent efficiency, offer potential for real-time radiation therapy monitoring, improving treatment outcomes.
Clinical translation requires addressing regulatory and safety concerns through long-term in vivo studies. Current data, such as biocompatibility of CsPbBr3@SiO2 in pancreatic tumor imaging,95 provide a foundation, but extensive toxicological profiling is needed. Collaborative efforts between material scientists, oncologists, and regulatory bodies will be crucial to establish PQDs as viable clinical tools. In summary, overcoming stability and toxicity challenges through advanced encapsulation, lead-free compositions, and green synthesis, combined with innovations in functionalization and ML integration will drive PQDs toward clinical success in cancer diagnostics and therapeutics (Table 7).
| Challenge | Mitigation strategy | Example | Ref. |
|---|---|---|---|
| Aqueous instability | Polymer encapsulation | CsPbBr3@PS-PEB-PS/PEG-PPG-PEG, 8-day stability | 89 |
| Toxicity | Silica encapsulation | CsPbBr3@SiO2, nontoxic in vivo | 95 and 96 |
| Toxicity | Surface passivation | (NH4)2S-treated CsPbBr3, biocompatible up to 450 ppm | 98 |
| Toxicity | Lead-free PQDs | Cs3Cu2Cl5, PLQY 31.07% | 110 |
| Scalability | Green synthesis | Room-temperature APbBr3 synthesis, 3–13 nm | 112 |
| Specificity | Ligand conjugation | AZI-PQDs-CD44v6 for gastric cancer | 88 |
| Regulatory | Long-term studies | Needed for CsPbBr3 biocompatibility | 92 |
Once released, free Pb2+ rapidly binds serum proteins and intracellular thiols, forming stable complexes that dramatically prolong systemic retention compared with intact nanoparticles.97,108 Biodistribution studies of intravenously administered coated CsPbBr3 PQDs consistently demonstrate predominant accumulation in the reticuloendothelial system: liver uptake reaches 30–45% injected dose per gram at 24 h and remains >15% ID g−1 after 28 days, with substantial splenic retention (10–25% ID g−1).97,109 Renal clearance of intact nanoparticles (>6–8 nm) is limited (<5% ID in urine within 7 days), while hepatobiliary excretion accounts for only 30–50% ID over 14–30 days, depending on PEG density and surface charge.55,90 Critically, protein-bound Pb2+ exhibits far slower elimination kinetics than the nanoparticulate fraction, with detectable lead persisting in liver, spleen, and brain beyond 90 days in several reports, raising serious concerns regarding cumulative neurotoxicity, oxidative stress, and genotoxicity.108,109
Current mitigation strategies remain insufficient for clinical acceptance. Although robust silica or PMMA shells significantly suppress acute toxicity,90,95 they do not guarantee complete long-term containment or rapid excretion of degradation products. Biodegradable polymer matrices and stimuli-responsive linkers have been explored to promote controlled disassembly and faster renal filtration of sub-6 nm fragments,57 but these approaches often compromise initial stability or brightness. Consequently, as emphasized throughout Sections 3.5 and 5.1, genuinely lead-free compositions (e.g., Cs3Cu2Cl5, Mg-doped CsMgxPb1−xI3) that eliminate heavy-metal degradation products from the outset currently represent the only realistic pathway capable of satisfying stringent regulatory requirements for chronic safety, biodistribution, and clearance.97,110 Until fully validated 6–12-month excretion data become available for lead-based systems, their clinical translation will remain effectively blocked.
Therapeutically, PQDs excel in delivering precise cancer treatments. Photodynamic therapy generates reactive oxygen species to induce tumor cell death, while photothermal therapy delivers localized heat to eliminate cancer cells effectively. Multifunctional systems combine chemodynamic therapy, producing hydroxyl radicals, and gas therapy, releasing therapeutic gases in tumor microenvironments, achieving robust tumor suppression. PQDs also enhance drug delivery by conjugating with cancer-specific ligands, improving chemotherapy and immunotherapy outcomes. Theranostic platforms integrate imaging and treatment, enabling personalized medicine through real-time monitoring and tailored interventions. Despite these advancements, challenges such as aqueous instability, potential toxicity, and scalability persist. Encapsulation with polymers or silica, surface passivation, and lead-free compositions have improved stability and safety, reducing environmental and health risks. Advances in scalable synthesis methods further support clinical translation. By addressing these hurdles through innovative materials and integration with technologies like machine learning, PQDs hold immense promise for revolutionizing cancer care, offering precise diagnostics and effective therapies to improve patient outcomes.
| This journal is © The Royal Society of Chemistry 2025 |