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

Perovskite quantum dot-based gas sensors for environmental monitoring: mechanisms, materials, and perspectives on next-generation pollution control

Suleiman Ibrahim Mohammadab, A. K. Kareemc, Zyad Shaaband, M. Sudhakara Reddye, Asokan Vasudevanfg, Fadhil Faez Seadhij, D. S. Jayalakshmiko, Sanjeev Kumarl, Satish Choudhurym and Ahmad Mohebi*n
aElectronic Marketing and Social Media, Economic and Administrative Sciences Zarqa University, Zarqa, Jordan
bResearch Follower, INTI International University, 71800 Negeri Sembilan, Putra Nilai, Malaysia
cBiomedical Engineering Department, College of Engineering, Al-Mustaqbal University, Hillah 51001, Babil, Iraq
dDepartment of Computer Science, University College of Duba, University of Tabuk, Duba 71911, Saudi Arabia
eDepartment of Physics & Electronics, School of Sciences, JAIN (Deemed to Be University), Bangalore, Karnataka, India
fFaculty of Business and Communications, INTI International University, 71800 Negeri Sembilan, Malaysia
gShinawatra University, 99 Moo 10, Bangtoey, Samkhok, Pathum Thani 12160, Thailand
hDepartment of Dentistry, College of Dentistry, The Islamic University, Najaf, Iraq
iDepartment of Medical Analysis, Medical Laboratory Technique College, The Islamic University of Al Diwaniyah, Al Diwaniyah, Iraq
jDepartment of Medical Analysis, Medical Laboratory Technique College, The Islamic University of Babylon, Babylon, Iraq
kDepartment of Physics, Sathyabama Institute of Science and Technology, Chennai, Tamil Nadu, India
lDepartment of Physics, University Institute of Sciences, Chandigarh University, Mohali, Punjab, India
mDepartment of Electrical & Electronics Engineering, Siksha ‘O' Anusandhan (Deemed to Be University), Bhubaneswar, Odisha-751030, India
nDepartment of chemistry, Young Researchers and Elite Club, Tehran Branch, Islamic Azad University, Tehran, Iran. E-mail: a.mohebiacademic@gmail.com
oSharda School of Engineering and Science, Sharda University, Greater Noida, India

Received 23rd September 2025 , Accepted 20th November 2025

First published on 2nd December 2025


Abstract

Perovskite quantum dots (PQDs) have recently emerged as transformative nanomaterials for gas sensing, offering exceptional optoelectronic properties, high surface-to-volume ratios, and compositional tunability. This is the first comprehensive review that systematically analyzes gas sensing technologies based on PQDs, with a particular emphasis on their relevance to environmental monitoring and pollution control. We summarize the latest advances in sensing mechanisms—including fluorescence quenching and enhancement, ratiometric detection, and chemiresistive/conductometric responses—and evaluate how synthesis strategies, surface ligand engineering, and hybrid architectures govern sensor performance. Key applications are critically assessed in the detection of toxic gases (NO2, NH3, H2S, SO2), volatile organic compounds, oxygen, and humidity, all of which are central to air quality assessment and environmental safety. Special focus is given to stability challenges under ambient and humid conditions, the environmental toxicity of lead-based PQDs, and mitigation strategies such as encapsulation, ligand engineering, and the development of lead-free alternatives. By integrating nanoscale material design with real-world environmental needs, this review not only consolidates current knowledge but also provides forward-looking perspectives for developing robust, selective, and sustainable PQD-based sensors for next-generation environmental monitoring systems.


1. Introduction

Perovskite quantum dots (PQDs) have emerged as a transformative class of nanomaterials with exceptional promise in optoelectronic and sensing applications.1–3 Structurally defined by the formula ABX3, where A is a monovalent cation (e.g., Cs+, MA+, FA+), B is a divalent metal cation (commonly Pb2+ or Sn2+), and X is a halide anion (Cl, Br, or I), these materials exhibit highly tunable optical and electronic properties.4–6 Quantum confinement effects, resulting from their ultrasmall size (typically below 10 nm), lead to discrete energy levels and strong photoluminescence, which can be finely adjusted by modifying their composition.7,8 High photoluminescence quantum yields (PLQYs), narrow emission linewidths, and low defect densities have enabled PQDs to play a leading role in fields such as photovoltaics, light-emitting diodes, and most recently, chemical and gas sensing.9,10 Gas sensing is a critical component in various sectors, including environmental monitoring, industrial process control, occupational safety, and biomedical diagnostics. The increasing demand for sensitive, selective, low-power, and miniaturized sensors has driven research toward novel materials with advanced functionality.11–13 Traditional gas sensors—such as those based on metal oxides14,15 or carbon nanomaterials16,17—often require high operating temperatures and suffer from limited selectivity and slow response–recovery dynamics.

PQDs offer an attractive alternative due to their unique optoelectronic behavior under ambient conditions, enabling rapid, low-temperature detection of diverse analytes including toxic gases (e.g., NO2, NH3, H2S, SO2), volatile organic compounds (e.g., ethanol, methanol, formaldehyde), oxygen, and humidity. Their high surface-to-volume ratio facilitates gas adsorption, while their ionic crystal structure allows for dynamic interactions with gas molecules, making them especially effective in transducing chemical changes into optical or electrical signals.18–20 Despite their potential, PQD-based gas sensors face several critical challenges that must be addressed to enable real-world deployment. A primary limitation lies in their environmental instability. The ionic nature of perovskites makes them vulnerable to degradation under moisture, heat, and light, leading to phase transitions, reduced PLQY, and poor long-term durability.20,21 In particular, exposure to humidity and reactive gases can cause irreversible structural damage. Moreover, the widespread use of lead-based PQDs, such as CsPbBr3 or MAPbI3, raises environmental and health concerns due to potential toxicity. These issues necessitate the development of stabilization strategies—such as surface passivation, ligand engineering, and encapsulation in porous or polymeric matrices—as well as the exploration of lead-free alternatives like CsSnX3 and Cs2AgBiX6.22,23

Although PQDs have been widely explored for optoelectronic applications, their utilization in gas sensing remains comparatively limited. To better illustrate the research progress in this emerging field, a bibliometric analysis of all available PQD-based gas sensing studies from 2019 to 2025 was conducted (Fig. 1). The collected data reveal a clear growth trend over recent years, indicating increasing scientific interest in developing PQD-based sensing technologies.


image file: d5ra07219k-f1.tif
Fig. 1 The annual publication trend of PQD-based gas sensing research between 2019 and 2025.

This review provides a comprehensive and systematic analysis of gas sensing technologies based on PQDs. It begins by outlining the fundamental structural, compositional, and optoelectronic features of PQDs that render them suitable for gas sensing applications. Various sensing mechanisms are then discussed, including fluorescence quenching and enhancement, ratiometric approaches, and chemiresistive and conductometric detection. The influence of synthesis techniques, ligand chemistry, and hybrid material design on sensor performance—particularly in terms of selectivity, sensitivity, response time, and stability—is examined in detail. Applications in detecting inorganic gases, VOCs, humidity, and oxygen are critically reviewed, supported by performance metrics such as LOD, response/recovery times, and operational robustness. Furthermore, the review highlights recent design innovations, including hybridization with metal oxides, ratiometric sensing platforms, and machine learning-based signal processing (Fig. 2). By synthesizing the latest advances in materials engineering, sensor design, and application development, this work aims to serve as a foundational reference for scientists and engineers seeking to harness the potential of PQDs in next-generation gas sensing platforms. The review not only consolidates current knowledge but also identifies critical knowledge gaps and outlines pathways for future innovation. Ultimately, PQD-based gas sensors represent a promising frontier for achieving sensitive, selective, stable, and scalable detection systems across environmental, biomedical, and industrial domains.


image file: d5ra07219k-f2.tif
Fig. 2 Overview of PQD-based gas sensors, highlighting key properties, sensing mechanisms, performance factors, and applications in gas detection.

2. Fundamentals of PQDs for sensing

PQDs have emerged as a transformative class of nanomaterials in the realm of gas sensing, owing to their exceptional optoelectronic properties, structural versatility, and tunable surface chemistry. This section elucidates the foundational aspects of PQDs that underpin their efficacy in gas sensing applications, including their structural and compositional diversity, optoelectronic properties tailored for sensing, synthesis techniques that influence their performance, and intrinsic stability challenges along with initial mitigation strategies. These fundamentals provide the necessary context for understanding the mechanisms and applications of PQD-based gas sensors.

2.1. Structural and compositional diversity of PQDs

PQDs adopt the general formula ABX3, where A is a monovalent cation (e.g., cesium, Cs+; methylammonium, MA+; formamidinium, FA+), B is a divalent metal cation (e.g., Pb2+, Sn2+), and X is a halide anion (e.g., Cl, Br, I). The perovskite structure consists of a cubic or pseudo-cubic lattice, where the B cation is coordinated octahedrally with six X anions, forming BX6 octahedra, and the A cation occupies the cuboctahedral voids.24,25 This structural framework imparts PQDs with remarkable flexibility, allowing compositional tuning to modulate their electronic and optical properties for gas sensing. The nanoscale dimensions of PQDs, typically ranging from 2 to 10 nm, induce pronounced quantum confinement effects, thereby enhancing their functionality in gas sensing applications. This reduced dimensionality significantly increases the surface-to-volume ratio, resulting in a higher density of accessible active sites for gas adsorption and interaction. Furthermore, compositional engineering—such as halide mixing (e.g., CsPb(Br/I)3) and A-site cation substitution (e.g., methylammonium [MA+] with formamidinium [FA+])—enables precise modulation of the bandgap and emission properties26,27 (Fig. 3).
image file: d5ra07219k-f3.tif
Fig. 3 PQDs adopt the ABX3 formula with A = Cs+, MA+, or FA+; B[double bond, length as m-dash]Pb2+ or Sn2+; and X = Cl, Br, or I. The structure features corner-sharing BX6 octahedra and A-site cations in the voids. Nanoscale size (2–10 nm) enables quantum confinement and boosts gas sensing.

These tunable optoelectronic characteristics are particularly advantageous for fluorescence-based sensing, where specific emission wavelengths correspond to tailored analyte responses. For example, CsPbBr3 PQDs typically exhibit green photoluminescence (PL ∼ 520 nm), whereas CsPbI3 PQDs emit in the red region (∼700 nm), facilitating sensor designs optimized for selective optical detection of target gas species. Furthermore, the ionic nature of the perovskite lattice facilitates dynamic surface interactions with gas molecules, such as ammonia (NH3) or nitrogen dioxide (NO2), which can modulate luminescence or conductivity. However, this ionic character also renders PQDs sensitive to environmental factors, necessitating structural modifications to enhance stability.28,29 The versatility of PQDs extends to lead-free compositions (e.g., CsSnX3, Cs2AgBiX6), which address toxicity concerns while maintaining sensing capabilities, albeit with challenges in achieving comparable PL efficiency.

2.2. Optoelectronic properties enabling sensing applications

The optoelectronic properties of PQDs are central to their effectiveness in gas sensing, particularly in fluorescence-based and chemiresistive sensing modalities. PQDs exhibit high PLQY, often exceeding 90% for CsPbBr3, due to their direct bandgap and low defect densities. This high PLQY enables sensitive detection of gas molecules that induce quenching or enhancement of fluorescence through charge transfer or surface adsorption.30,31 For example, electron-donating gases like NH3 can passivate surface defects, enhancing PL intensity, while electron-withdrawing gases like NO2 can quench emission by extracting charge carriers.32–34 The tunable bandgap of PQDs, ranging from 1.8 eV (CsPbI3) to 3.0 eV (CsPbCl3), allows for wavelength-specific responses to gas analytes, facilitating selective detection in fluorescence-based sensors. The narrow emission linewidths (full width at half maximum, FWHM, of 12–40 nm) ensure high spectral resolution, enabling ratiometric sensing approaches where changes in emission intensity or peak position signal gas presence. Additionally, the high exciton binding energy in PQDs (up to 100 meV) enhances radiative recombination, making them ideal for detecting subtle changes in surface chemistry induced by gas interactions.24,33

Fig. 4a displays the PL spectra evolution of a CsPbBr3 PQD thin film under 410 nm excitation as pulse energy density varies. A distinct amplified spontaneous emission (ASE) peak at 533 nm with a FWHM of 5 nm is evident, as highlighted in the inset image, reflecting high radiative efficiency and low defect density. This aligns with the material's high PLQY, enhancing its sensitivity to gas-induced fluorescence changes, a key attribute for optoelectronic sensing applications. Fig. 4b presents a 3D diagram illustrating the pump fluence versus PL intensity relationship for the CsPbBr3 PQD thin film. The nonlinear increase in PL intensity with energy density indicates a threshold for ASE onset, consistent with the tunable bandgap properties that enable wavelength-specific responses. This visualization supports the material's potential for ratiometric sensing, where shifts in emission characteristics can signal gas presence, reinforcing its utility in selective detection. Fig. 4c and d detail the optoelectronic response with respect to pulse energy density. In 4c, the integrated PL intensity and FWHM show a sharp rise and narrowing beyond a threshold, indicating amplified emission and reduced non-radiative losses, supported by the high exciton binding energy. In 4d, the ASE peak position shifts from 536 nm to 526 nm with increasing energy density, reflecting electronic structure changes that enhance detection of subtle surface chemistry alterations induced by gas interactions.


image file: d5ra07219k-f4.tif
Fig. 4 (a) Evolution of PL spectra with varying pulse energy density and (b) 3D diagram of pump fluence versus PL intensity for a CsPbBr3 PQD thin film under 410 nm excitation. (c) Integrated PL intensity and FWHM behavior and (d) ASE peak position as functions of pulse energy density for the CsPbBr3 PQD thin film under 410 nm excitation. Reprinted with permission from ref. 6. Copyright 2020 American Chemical Society.

In chemiresistive sensing, PQDs demonstrate remarkable electrical properties, including high carrier mobility (up to 4500 cm2 V−1 s−1 in CsPbBr3 thin films) and tunable conductivity. Gas molecules adsorbed on PQD surfaces can alter charge carrier concentrations, modulating resistance. For instance, oxidizing gases like NO2 increase resistance by trapping electrons, while reducing gases like H2S decrease resistance by donating electrons. The large surface area of PQDs amplifies these effects, enabling low LODs in the ppb range for gases like hydrogen sulfide (H2S).35–37 The defect tolerance of PQDs, arising from their electronic structure where defect states lie within or near the conduction and valence bands, minimizes non-radiative recombination, enhancing both optical and electrical sensitivity. However, surface defects, such as halide vacancies, can act as active sites for gas adsorption, necessitating careful control during synthesis to balance sensitivity and stability. Table 1 shows the structural, compositional, and optoelectronic properties of PQDs for gas sensing applications.

Table 1 Structural, compositional, and optoelectronic properties of PQDs for gas sensing applications
Property Description Impact on gas sensing Specific features Advantages Limitations Related gas sensing applications Ref.
Structural diversity ABX3 perovskite lattice (A: Cs+, MA+, FA+; B: Pb2+, Sn2+; X: Cl, Br, I); cubic/pseudo-cubic with BX6 octahedra Enables compositional tuning for tailored optical/electrical responses to gases Size: 2–10 nm; high surface-to-volume ratio; mixed-halide (e.g., CsPb(Br/I)3) Large active sites for gas interactions; tunable emission (520 nm for CsPbBr3, 700 nm for CsPbI3) Ionic lattice sensitive to moisture, requiring stabilization NH3 (fluorescence enhancement), NO2 (quenching), H2S (LOD 250 ppb) 24 and 25
Compositional versatility Flexibility in A, B, X components; lead-free options (e.g., CsSnX3, Cs2AgBiX6) Allows bandgap tuning and reduced toxicity for specific gas detection Bandgap: 1.8 eV (CsPbI3) to 3.0 eV (CsPbCl3); lead-free reduces environmental impact Tailored sensors for fluorescence-based detection; eco-friendly alternatives Lower PLQY in lead-free PQDs; stability challenges NH3 (LOD 0.5 ppm), SO2 (LOD 1 ppm) 26 and 27
Photoluminescence High PLQY (up to 90% for CsPbBr3); narrow FWHM (12–40 nm) Enables sensitive fluorescence-based sensing via quenching or enhancement Direct bandgap; low defect density; high exciton binding energy (∼100 meV) High sensitivity to gas-induced PL changes; spectral resolution for ratiometric sensing Susceptible to photo-induced degradation; requires passivation NH3 (PL enhancement), NO2 (quenching, response 53 at 5 ppm), O2 (sensitivity 12.7) 28 and 29
Electrical conductivity High carrier mobility (up to 4500 cm2 V−1 s−1 in CsPbBr3); tunable conductivity Facilitates chemiresistive sensing through gas-induced resistance changes Surface defects as active sites; defect-tolerant electronic structure Low LODs (e.g., H2S at 250 ppb); amplified response due to large surface area Surface defects may reduce stability; needs balanced synthesis H2S (response 0.58), TEA (response 52.92 at 60 °C), NO2 (increased resistance) 30 and 31
Quantum confinement Nanoscale size (2–10 nm) induces quantum confinement, enhancing optical properties Increases sensitivity to gas-induced optical changes via modified bandgap Tunable emission wavelengths; enhanced PL due to confinement effects Precise control over optical responses for specific gases Small size increases surface sensitivity to environmental degradation NH3 (fluorescence turn-on), SO2 (LOD 1 ppm) 32 and 33
Defect tolerance Defect states lie within/near conduction and valence bands, minimizing non-radiative recombination Enhances both optical and electrical sensitivity for gas detection Low non-radiative losses; halide vacancies as active sites for gas adsorption High sensitivity in fluorescence and chemiresistive sensors Excessive defects can reduce long-term stability H2S (LOD 250 ppb), NO (sensitivity 6 at 1000 ppm) 24 and 29
Surface chemistry Ionic lattice enables dynamic interactions with gas molecules (e.g., NH3, NO2) Modulates PL or conductivity for selective gas detection Surface defects and ligand interactions tailor gas adsorption High selectivity for electron-donating (NH3) or withdrawing (NO2) gases Susceptible to analyte-induced degradation; requires surface passivation NH3 (LOD 0.5 ppm), NO2 (response 53 at 5 ppm) 25 and 35


2.3. Synthesis techniques and their influence on sensing performance

The synthesis method of PQDs plays a central role in defining their suitability for gas sensing, as it directly impacts their size, crystallinity, surface chemistry, and stability. These parameters influence key sensing metrics such as PLQY, surface defect density, and gas adsorption behavior. Among the widely adopted methods are hot-injection, ligand-assisted reprecipitation (LARP), room-temperature co-precipitation, and solvothermal synthesis. Each offers distinct advantages depending on whether fluorescence- or resistance-based sensing is targeted.
2.3.1. Hot-injection synthesis. Hot-injection is widely recognized for producing high-quality PQDs with narrow size distributions and superior optical properties. This method involves injecting cesium oleate into a hot (140–200 °C) solution containing lead halide and organic ligands (e.g., oleic acid, oleylamine) under inert atmosphere. Rapid nucleation leads to uniform PQDs, typically 2–8 nm in size, with polydispersity below 5%. High crystallinity and PLQY (up to 90% for CsPbBr3) make these PQDs ideal for fluorescence-based gas sensing. This method minimizes surface defects, enhancing sensitivity in applications where PL changes signal gas presence, such as NH3 or NO2 detection. However, long-chain ligands used in the process can limit gas diffusion to active sites. Ligand exchange with shorter or zwitterionic molecules (e.g., DDAB) can improve surface accessibility while maintaining colloidal stability.28,38,39 Despite high sensor performance, the need for strict temperature control and inert conditions hinders scalability.
2.3.2. Ligand-assisted reprecipitation. LARP is a low-temperature, solution-based method suited for scalable PQD production. It involves dissolving Cs and Pb precursors in a polar solvent like DMF, then injecting the solution into a non-polar solvent (e.g., toluene) containing surfactant ligands. This polarity shock triggers rapid nucleation. PQDs typically range from 2–10 nm, with PLQY between 70–85%. LARP-generated PQDs tend to exhibit higher surface defect densities than those from hot-injection, which can be beneficial for chemiresistive sensing by providing more active sites for gas adsorption. For instance, LARP-derived CsPbBr3 sensors achieved an H2S detection limit of 250 ppb. However, these same defects reduce stability under humid conditions, leading to PL quenching over time. Encapsulation strategies—such as embedding in silica aerogels—can mitigate degradation.40–42 The method's simplicity and tunability make it attractive for applications where rapid, cost-effective synthesis is required, though broader size distributions may limit fluorescence resolution (Fig. 5).
image file: d5ra07219k-f5.tif
Fig. 5 Methods for synthesizing PQDs: (a) hot-injection technique and (b) ligand-assisted reprecipitation approach.
2.3.3. Room-temperature Co-precipitation. Co-precipitation offers a simple and scalable route to PQD synthesis. Precursors are mixed in a single solvent (e.g., DMF), followed by the gradual addition of a poor solvent (e.g., acetone), which reduces solubility and induces nucleation. The process proceeds at ambient temperature, eliminating the need for high heat or inert atmospheres. Resulting PQDs are typically 4–12 nm with broader size distributions (10–15% polydispersity) and moderate PLQY (50–70%). While not ideal for high-performance fluorescence sensors, these PQDs are well-suited for resistance-based sensing. For example, co-precipitated CsPbBr3–In2O3 composites showed strong response (52.92) to triethylamine (TEA) at 60 °C. Stability can be further improved through microwave-assisted co-precipitation, which enhances crystallinity and reduces defects.24,43,44 However, short-chain ligands required for gas access can compromise long-term colloidal integrity, highlighting the need for optimized ligand selection.
2.3.4. Solvothermal synthesis. Solvothermal synthesis is a promising approach for producing structurally stable PQDs, especially for use in humid or high-temperature environments. Precursors are dissolved in a high-boiling solvent (e.g., octadecene) and heated in a sealed autoclave (120–180 °C) under pressure. The controlled environment promotes low-defect crystal growth, yielding PQDs with sizes of 3–10 nm and polydispersity of 3–7%. These PQDs demonstrate good moisture resistance and PLQY around 85%, suitable for both optical and resistive sensing. For example, solvothermally synthesized CsPbBr3 showed excellent humidity sensing with an LOD of 0.1% RH, and NO2 sensing with a chemiresistive response of 53 at 5 ppm. However, the method requires longer synthesis times (1–4 hours) and specialized high-pressure equipment, which increases cost and limits throughput. Proper ligand engineering remains critical to balance gas diffusion and environmental stability26,30,45 (Fig. 6).
image file: d5ra07219k-f6.tif
Fig. 6 Alternative PQD synthesis strategies: (a) room-temperature Co-precipitation method, (b) solvothermal process, and (c) influence of ligands and post-synthesis adjustments.
2.3.5. Ligand effects and post-synthesis modifications. Ligand selection during and after synthesis critically affects PQD sensor performance. Long-chain ligands such as oleic acid offer colloidal stability but inhibit gas interaction, reducing sensitivity in chemiresistive formats. Replacing them with short-chain (e.g., octylamine) or zwitterionic ligands enhances gas access but may compromise environmental resistance. For instance, ethanol sensors using CsPbBr3 capped with zinc acetylacetonate achieved a 3 ppm detection limit under humid conditions due to improved surface interactions. Post-synthesis ligand exchange and encapsulation in matrices such as zeolites, polymers, or silica aerogels help protect PQDs from degradation while preserving surface reactivity. CsPbBr3 embedded in Fe-doped zeolite X retained 98% PL intensity after 100 days in humid air, enabling stable NH3 sensing. Hybridization with metal oxides (e.g., ZnO, In2O3) further improves performance by promoting charge transfer and enhancing durability.25,27,46

2.4. Intrinsic stability issues and initial mitigation strategies

The primary limitation of PQDs in gas sensing is their intrinsic instability, particularly under exposure to moisture, heat, and light, which can degrade their optoelectronic properties. The ionic nature of the perovskite lattice makes PQDs susceptible to dissociation in humid environments, as water molecules coordinate with surface ions, leading to structural collapse and loss of luminescence.26 For gas sensing, this is particularly problematic, as many target gases (e.g., NH3, H2O vapor) are detected in humid conditions, necessitating robust stabilization strategies. Thermal instability arises from the low formation energy of PQDs, causing phase transitions or decomposition above 100 °C.47 This limits their use in high-temperature industrial gas sensing applications unless mitigated. Photo-induced degradation, driven by prolonged exposure to UV or visible light, can generate halide vacancies, reducing PLQY and altering sensing responses.28 These challenges are exacerbated in gas sensing, where PQDs are exposed to reactive analytes that may accelerate degradation.24 Initial mitigation strategies focus on enhancing stability without compromising sensing performance. Surface passivation with short-chain or zwitterionic ligands (e.g., didodecyldimethylammonium bromide) strengthens the PQD surface against moisture and gas-induced degradation while maintaining accessibility for analyte interactions.48

Fig. 7 provides a comprehensive analysis of the QY characteristics and gas-sensing performance of PQDs on different substrates, addressing their optoelectronic stability under various conditions.48 In panel A, the excitation spectra of PQDs-glass and PQDs-HAAO samples show a peak around 410 nm, indicating consistent absorption properties despite substrate differences. Panel B presents the emission spectra, revealing a broad emission band centered near 520 nm for both samples, with PQDs-HAAO exhibiting slightly enhanced intensity, suggesting improved radiative recombination. These spectral features align with the high PLQY potential of PQDs, though their stability under environmental stressors like moisture and light remains a critical concern due to the ionic nature of the perovskite lattice and its susceptibility to dissociation. Panel C quantifies the QY stability of PQDs-glass and PQDs-HAAO after exposure to an oxygen-enriched atmosphere (approximately 20% oxygen) for 1 hour. The QY of PQDs-glass decreases from 78.95% to 64.26%, reflecting significant degradation due to oxygen-induced halide vacancies and structural collapse, consistent with photo-induced instability. In contrast, PQDs-HAAO retains a higher QY, dropping from 72.98% to 64.32%, indicating better resistance to oxidative degradation. This enhanced stability can be attributed to effective surface passivation, which mitigates the ionic lattice's vulnerability to environmental factors, supporting its suitability for gas sensing in reactive atmospheres. Panel D further evaluates sensor response to 100% oxygen, 1000 ppm ammonia, and 1000 ppm nitric oxide, with PQDs-HAAO showing a more pronounced PL intensity change (2.93 a.u.) compared to PQDs-glass (2.16 a.u.), highlighting improved sensitivity to gas analytes. Panels E and F illustrate the dynamic response and recovery cycles of PQDs-glass and PQDs-HAAO sensors, respectively, through five alternating exposures to 100% O2 and 100% N2. The PQDs-glass sensor exhibits a normalized PL intensity with moderate fluctuations, indicating partial recovery but limited resilience to repeated gas exposure, likely due to thermal and photo-induced degradation over time. In contrast, the PQDs-HAAO sensor demonstrates a more stable and repeatable PL intensity profile, suggesting enhanced durability against phase transitions and decomposition. This improved performance underscores the efficacy of substrate modification and surface passivation strategies in overcoming the intrinsic instability of PQDs, enabling their application in demanding gas-sensing environments where prolonged analyte interaction is required.


image file: d5ra07219k-f7.tif
Fig. 7 QY and gas-sensing performance of PQDs: (A) excitation spectra, (B) emission spectra, (C) QY after 1-hour oxygen exposure (∼20%), (D) response to 100% O2, 1000 ppm NH3, and NO, (E) PQDs-glass response–recovery cycles (100% O2/N2), (F) PQDs-HAAO response–recovery cycles (100% O2/N2). Reproduced with permission from ref. 48. Copyright 2025 Royal Society of Chemistry.

The device configuration in Fig. 7 was designed to systematically compare the effects of substrate morphology on PQD performance. Two substrates—smooth glass and hierarchically anodized aluminum oxide (HAAO)—were employed. PQDs were drop-cast onto both surfaces, forming thin nanocrystalline films. The porous structure of HAAO provides abundant anchoring sites that improve PQD adhesion and promote uniform distribution, minimizing aggregation and enhancing interfacial stability. The nanoscale pores of HAAO also act as gas diffusion channels, facilitating rapid gas–surface interactions. In contrast, the glass substrate offers a smoother, inert reference surface that allows differentiation between intrinsic PQD behavior and substrate-induced effects. This design enables direct evaluation of structural and environmental contributions to PQD optical stability and sensing reliability.

The sensing mechanism primarily originates from PL modulation due to charge-trap formation and surface defect passivation in response to gas adsorption. Upon exposure to oxidizing gases such as O2, electrons are withdrawn from the PQD surface, generating halide vacancies and enhancing nonradiative recombination, which leads to PL quenching. Conversely, reducing gases such as NH3 donate electrons, mitigating trap states and restoring PL intensity. These reversible interactions underpin the PL-based sensing principle, where luminescence intensity directly correlates with the surface charge state and defect density of PQDs. The interplay between surface chemistry and optical emission demonstrates that gas-induced electronic perturbations dominate the sensing response, rather than bulk phase changes.

The experimental results presented in Fig. 7 corroborate this mechanism. Specifically, the observed PL decrease under O2 exposure (Fig. 7D) and its partial recovery in the response–recovery cycles (Fig. 7E and F) confirm the charge-trap-driven modulation process. The HAAO-based sensor exhibits a more stable and repeatable PL signal than the glass-based one, consistent with the suppression of defect formation due to physical confinement and chemical passivation at the PQD–HAAO interface. The improved QY retention (from 72.98% to 64.32% after O2 exposure) further evidences enhanced environmental resilience. Collectively, these findings demonstrate that rational substrate engineering and surface passivation effectively overcome intrinsic PQD instability, enabling reliable operation under reactive gas atmospheres.

Encapsulation in inorganic matrices, such as silica or alumina, provides a physical barrier against environmental factors, though it may reduce sensitivity by limiting gas diffusion.27 Organic polymer coatings, such as polystyrene or PMMA, offer a flexible alternative, preserving optical properties while enhancing moisture resistance.25 Ion doping (e.g., Mn2+, Bi3+) or A-site substitution (e.g., FA+ for MA+) can improve lattice stability by increasing formation energy, making PQDs more resistant to thermal and chemical degradation.49 For example, Cs2AgBiBr6 PQDs exhibit enhanced stability compared to CsPbBr3 due to their double-perovskite structure, though at the cost of lower PLQY.41 Lead-free PQDs, such as CsSnX3 or Cs3Bi2X9, address toxicity concerns but often suffer from lower stability and optical performance, requiring further optimization for gas sensing.40

Hybridization with stable nanomaterials, such as metal oxides (e.g., ZnO, In2O3), enhances both stability and sensitivity by forming heterojunctions that protect PQDs while facilitating charge transfer with gas molecules.45 These strategies, while promising, must balance stability with the need for accessible surface sites, as excessive passivation can hinder gas interactions. Ongoing research aims to develop multifunctional coatings that simultaneously protect PQDs and enhance selectivity for specific gases, paving the way for robust, high-performance gas sensors.35 The structural versatility, tunable optoelectronic properties, and diverse synthesis methods of PQDs make them highly promising for gas sensing. However, their intrinsic instability necessitates careful design of stabilization strategies to ensure reliable performance in real-world conditions. These fundamentals provide a foundation for exploring the mechanisms, applications, and innovations in PQD-based gas sensing.

PQDs face significant stability challenges that limit their practical application in gas sensing, despite their exceptional optoelectronic properties. Key issues include moisture sensitivity, thermal instability, photo-induced degradation, analyte-induced degradation, cross-sensitivity to interfering gases, limited long-term operational stability, toxicity-related degradation, and oxygen-induced degradation. These challenges arise from the ionic lattice nature of PQDs, low formation energy, surface defects, and interactions with reactive gases or environmental factors, leading to reduced PL intensity, compromised selectivity, and shortened sensor lifespan. Advanced mitigation strategies, such as encapsulation in silica/zeolites, ion doping (e.g., Mn2+, Bi3+), surface passivation with zwitterionic ligands, molecularly imprinted polymers (MIPs), and integration with machine learning, have been developed to enhance stability and selectivity. Table 2 summarizes these challenges, their causes, impacts on gas sensing performance, mitigation approaches, advantages, limitations, and specific applications for detecting gases like NH3, NO2, H2S, SO2, O2, and acetone, providing a comprehensive guide for designing robust PQD-based gas sensors for environmental, industrial, and biomedical applications.

Table 2 Stability challenges of PQDs and initial mitigation strategies for gas sensing applications
Stability challenge Cause Impact on gas sensing Mitigation strategies Advantages Limitations Related gas sensing applications Ref.
Moisture sensitivity Ionic lattice nature; water molecules coordinate with surface ions, causing structural collapse Reduced PL intensity; compromised fluorescence-based sensing (e.g., NH3, H2O vapor) Encapsulation in silica/zeolites; polymer coatings (e.g., PMMA, EVA) Enhanced moisture resistance; retained PL for up to 100 days May limit gas diffusion, reducing sensitivity NH3 (LOD 0.5 ppm), humidity (90% PL retention after 6 months) 25–27
Thermal instability Low formation energy; phase transitions or decomposition above 100 °C Loss of PLQY and conductivity in high-temperature environments (e.g., industrial NO2 sensing) Ion doping (e.g., Mn2+, Bi3+); A-site substitution (e.g., FA+ for MA+) Increased lattice stability; suitable for NO2 sensing at elevated temperatures Lower PLQY in doped PQDs; complex synthesis NO2 (response 53 at 5 ppm), temperature sensing (sensitivity 15.89% K−1) 45, 47 and 49
Photo-induced degradation UV/visible light exposure creates halide vacancies, reducing PLQY. Altered fluorescence responses; unreliable sensing for O2, NO. Surface passivation with zwitterionic ligands; hybridization with metal oxides (e.g., ZnO) Improved photostability; enhanced charge transfer for sensing Passivation may reduce surface accessibility for gases O2 (sensitivity 12.7), NO (sensitivity 6 at 1000 ppm) 28 and 29
Analyte-induced degradation Reactive gases (e.g., H2S, SO2) interact with surface ions, accelerating lattice breakdown Reduced sensor lifetime; inconsistent H2S, SO2 detection Ligand exchange (e.g., short-chain or zwitterionic ligands); MOF encapsulation Balances stability and gas accessibility; high selectivity for H2S (LOD 250 ppb) Complex post-synthesis processing; potential cost increase H2S (response 0.58), SO2 (LOD 1 ppm) 24 and 25
Cross-sensitivity to interfering gases Non-specific interactions at surface defects (e.g., halide vacancies) with multiple gases Reduced selectivity in complex environments (e.g., NH3 detection in presence of H2O, CO2) Surface functionalization with gas-specific receptors (e.g., thiol-based ligands); machine learning for signal differentiation Enhanced selectivity; accurate detection in mixed gas environments Requires complex surface engineering; computational resources for ML. NH3 (LOD 0.5 ppm), ethanol (response 0.275 at 1300 ppm) 25 and 35
Limited long-term operational stability Gradual degradation of surface ligands and lattice under prolonged gas exposure Decreased sensor reliability over time; reduced PLQY for fluorescence-based sensors Encapsulation in robust matrices (e.g., MOFs, graphene composites); periodic ligand regeneration Extended sensor lifespan; maintained performance up to 6 months May increase fabrication complexity; potential sensitivity trade-off NO2 (response 53 at 5 ppm), O2 (sensitivity 12.7) 45 and 50
Toxicity-related degradation Lead leakage from Pb-based PQDs under environmental stress; degradation of lead-free alternatives Environmental/health risks; reduced performance in lead-free PQDs (e.g., CsSnX3) Development of lead-free PQDs (e.g., Cs2AgBiBr6); ion doping for stability Eco-friendly sensors; reduced toxicity risks Lower PLQY and stability in lead-free PQDs Acetone (response 13 at 60 ppm), NH3 (LOD 0.5 ppm) 40 and 49
Oxygen-induced degradation Oxygen molecules adsorb on surface defects, forming non-radiative recombination centers Quenching of PL intensity; reduced sensitivity for O2 and other gases in oxygen-rich environments Embedding in anodic alumina oxide (AAO) templates; coating with ethyl cellulose matrices Enhanced stability in oxygen-rich conditions; maintained PL for 504 hours Reduced surface area for gas interaction; complex fabrication O2 (sensitivity 12.7), NO (sensitivity 2.7) 29 and 48


2.4.1. Correlation between synthesis, characterization, and sensing suitability. The relationship between the synthesis methodology, structural characterization, and sensing suitability of PQDs is of fundamental importance in determining their performance in gas detection. Each synthesis route—whether hot-injection, LARP, co-precipitation, or solvothermal—imparts distinct physical and chemical attributes to PQDs, including particle size uniformity, defect density, surface termination, and crystallinity, which collectively dictate their sensing response. Controlling these parameters allows optimization for either fluorescence-based or chemiresistive sensing platforms.

Hot-injection, for instance, produces highly crystalline PQDs with narrow size distributions and minimal surface trap density, resulting in high photoluminescence quantum yield (PLQY > 90%).28,38 These features make them particularly suitable for fluorescence-based sensors, where emission quenching or enhancement directly signals gas adsorption events. Structural integrity confirmed through X-ray diffraction (XRD) and transmission electron microscopy (TEM) typically reveals sharp diffraction peaks and well-defined lattice fringes, indicative of defect-free crystalline domains essential for optical stability.33,39 However, the dense surface capping of long-chain ligands can restrict gas diffusion, requiring post-synthesis ligand modification or partial ligand removal to enhance active site exposure.40

Conversely, PQDs synthesized by ligand-assisted reprecipitation (LARP) or co-precipitation exhibit slightly broader size distributions and higher surface defect densities.41,42 These defects act as favorable adsorption sites for electron-donating or withdrawing gases such as NH3 or NO2, thus enhancing the chemiresistive response through modulated charge transfer processes.43 Electrical characterization using current–voltage (IV) measurements and impedance spectroscopy demonstrates that defect-rich PQDs display a larger change in conductivity upon gas exposure, directly correlating with the density of surface traps observed in PL spectra. Consequently, the synthesis route defines not only the physical morphology but also the electronic transport characteristics that control sensor performance.

Solvothermal methods, conducted under pressurized conditions, yield highly stable PQDs with low defect concentrations and improved moisture resistance.26,45 These materials maintain strong optical and electrical responses in humid or thermally demanding environments, making them appropriate for hybrid optical-electrical sensors. In addition to structural and optical analysis, complementary techniques such as UV-vis absorption spectroscopy, Fourier-transform infrared spectroscopy (FTIR), and X-ray photoelectron spectroscopy (XPS) provide insights into bandgap tunability, surface bonding, and ligand–ion interactions. These characterizations confirm the role of synthesis parameters—temperature, precursor ratios, and ligand type—in tailoring the energy band structure and gas adsorption affinity.25,30

The combined interpretation of these analytical results enables a direct structure–property–function correlation. For example, hot-injection PQDs with high PLQY are ideal for detecting oxidative gases like NO2 or O2 via fluorescence quenching, while LARP- and co-precipitation-derived PQDs with higher defect densities show superior performance in resistive detection of reducing gases such as H2S or TEA.30,43 Furthermore, ligand engineering after synthesis, including exchange with zwitterionic or short-chain molecules, enhances gas accessibility without compromising colloidal stability.46 This systematic understanding underscores that the choice of synthesis route and corresponding characterization outcomes must align with the intended sensing mechanism, ensuring optimal selectivity, sensitivity, and durability.

In summary, effective gas sensing by PQDs relies on precise synthesis control combined with comprehensive characterization to bridge the gap between material design and functional performance. This correlation provides a guiding framework for tailoring PQDs toward specific sensing modalities, enabling reliable and scalable sensor technologies for environmental, industrial, and biomedical monitoring applications.30,43

3. Sensing mechanisms in PQD-based gas sensors

PQDs excel in gas sensing by transducing molecular interactions into optical or electrical signals, leveraging their tunable optoelectronic properties and high surface reactivity. This section elucidates the core sensing mechanisms—fluorescence-based (quenching, turn-on, ratiometric) and electrical-based (chemiresistive, conductometric)—with a focus on how material composition, crystal structure, and surface chemistry modulate these processes. By integrating recent studies, it provides a theoretical foundation for the performance metrics discussed in Section 4, emphasizing interactions with gases such as NH3, NO2, H2S, and VOCs without overlapping with application-specific details.51–53

3.1. Fluorescence-based mechanisms

Fluorescence-based sensing exploits the exceptional PL properties of PQDs, including quantum yields up to 90% (e.g., CsPbBr3) and narrow emission linewidths (12–40 nm), to detect gases via changes in emission intensity or wavelength. These mechanisms hinge on gas-induced alterations in radiative recombination, modulated by PQD composition and surface states.51
3.1.1. Fluorescence quenching. Fluorescence quenching occurs when gas molecules adsorb onto PQD surfaces, promoting non-radiative recombination or charge transfer that reduces PL intensity. The ionic lattice of PQDs, such as CsPbBr3, is particularly susceptible to reactive gases like H2S, which coordinates with Pb2+ ions to form non-luminescent PbS, disrupting the perovskite structure.54 For oxidizing gases like NO and NO2, quenching arises from electron transfer from the PQD conduction band to the gas, enhancing non-radiative pathways.28 Material composition plays a critical role; for instance, CsPbBr3 PQDs exhibit stronger quenching with NO2 due to their high electron mobility, while CH3NH3PbBr3 PQDs are more responsive to polar gases due to their organic A-site cations, which enhance surface polarity.45 Surface defects, particularly halide vacancies, amplify quenching by serving as adsorption sites. The choice of halide (e.g., Br vs. I) influences defect density, with iodide-based PQDs often showing higher defect-mediated sensitivity but lower stability.52 Protective strategies, such as silica encapsulation, mitigate irreversible quenching by stabilizing the lattice against reactive gases.63
3.1.2. Fluorescence turn-on. Fluorescence turn-on mechanisms involve gas-induced PL enhancement through passivation of surface defects. NH3 detection with CsPbX3 PQDs exemplifies this, where NH3 adsorbs onto halide vacancies, reducing non-radiative recombination and stabilizing excitons.35 The halogen composition (X = Cl, Br, I) modulates this effect; Cl-based PQDs exhibit stronger turn-on responses due to higher defect densities, whereas Br-based PQDs offer a balance of sensitivity and stability. Organic–inorganic hybrid PQDs, such as CH3NH3PbBr3, show reversible turn-on with NH3 due to weak coordination that avoids structural disruption, driven by the flexible A-site cation.55 Material engineering, such as doping with Mn2+ or Sn2+, can enhance turn-on efficiency by altering bandgap alignment, increasing electron donation from reducing gases.66 Selectivity depends on surface chemistry; tailoring ligands to favor specific gas interactions (e.g., NH3 over H2O) is critical for practical sensing.

The below figure illustrates the fluorescence turn-on mechanism in CsPbX3 PQDs through NH3-induced passivation of surface defects, aligning with the enhanced PL observed in gas sensing applications. The left panel depicts the initial state of a Br-based PQD with bromine vacancies and Pb2+ sites, where NH3 treatment leads to adsorption of ammonia molecules, passivating these defects and reducing non-radiative recombination, as shown in the middle panel. The right panel extends this concept to a Cl/Br mixed-halide PQD, where NH3 treatment similarly passivates bromine and chlorine vacancies, stabilizing excitons and enhancing PL, with the added presence of Cs+ and chlorine vacancies reflecting the tunable halogen composition. This defect passivation, consistent with the stronger turn-on responses in Cl-based PQDs and the reversible behavior in organic-inorganic hybrids like CH3NH3PbBr3, underscores the role of surface chemistry and ligand tailoring in achieving selective NH3 detection over other gases like H2O (Fig. 8).


image file: d5ra07219k-f8.tif
Fig. 8 NH3-induced fluorescence turn-on in CsPbX3 PQDs: (left) initial Br-based PQD with vacancies, (middle) after NH3 treatment, (right) After NH3 treatment in Cl/Br PQD. Reprinted from ref. 35, with permission from Elsevier, copyright 2022.
3.1.3. Ratiometric sensing. Ratiometric sensing uses dual-emission or intensity ratio changes to enhance detection reliability by mitigating environmental noise. FAPbI3 PQDs paired with rhodamine 110 in ethyl cellulose detect O2 via differential quenching, where oxygen quenches PQD emission while the reference dye remains stable, yielding a robust ratiometric signal.29 Material choice is pivotal; FAPbI3's narrower bandgap enhances sensitivity to O2 compared to CsPbBr3, which is better suited for NO detection in dual-emission systems with Pt(II) complexes.56 Hybrid material systems, incorporating metal–organic frameworks (MOFs) or polymers, improve ratiometric performance by stabilizing PQDs against environmental interferents.57,58 The incorporation of reference emitters requires careful bandgap alignment to ensure distinct emission profiles, critical for applications in complex gas mixtures like breath analysis.

3.2. Electrical-based mechanisms

Electrical-based sensing exploits gas-induced changes in PQD conductivity or resistance, leveraging their high carrier mobility (up to 4500 cm2 V−1 s−1) for compact, low-power sensors compatible with electronic integration.59
3.2.1. Chemiresistive sensing. Chemiresistive sensing detects resistance changes upon gas adsorption.60 For H2S detection, CsPbBr3 PQDs functionalized with tributyltin oxide (TBTO) exhibit increased conductivity via electron donation from H2S, modulated by TBTO's surface affinity.24 Material composition influences performance; CsPbBr3's high carrier mobility amplifies resistance changes, while Cs2AgBiBr6 PQDs, with lower ionic character, offer enhanced stability for detecting reducing gases like TEA.30 Oxidizing gases (e.g., NO2) increase resistance by trapping electrons, whereas reducing gases decrease it, enabling selectivity through surface engineering. Incorporating metal oxides like ZnO or In2O3 forms heterojunctions that enhance charge transfer, with ZnO's wide bandgap complementing PQD conductivity.45 Material stability remains a challenge, as ionic PQDs are prone to degradation under prolonged gas exposure.
3.2.2. Conductometric sensing. Conductometric sensing measures conductance changes, often in hybrid systems where PQDs enhance charge transfer. CsPbBr3–In2O3 nanofibers under UV illumination detect formaldehyde via heterojunction-modulated charge transfer, where UV-generated electron-hole pairs enhance sensitivity.31 Lead-free PQDs, such as Cs2AgBiBr6, paired with Co3O4, detect acetone by altering conductance through electron transfer at the heterojunction.49 The choice of metal oxide or nanomaterial partner significantly affects performance; In2O3's high electron mobility complements PQD conductivity, while Co3O4 enhances selectivity for VOCs.61 Surface defects in PQDs, while beneficial for sensitivity, can reduce long-term stability, necessitating robust material designs.

3.3. Surface interactions and ligand effects

Surface interactions and ligand effects are central to PQD sensor performance, governing gas adsorption and signal transduction. The ionic lattice of PQDs, with exposed A, B, and X sites, facilitates strong interactions with polar gases. For instance, CH3NH3PbBr3 in silica aerogels detects SO2 via sulfite complex formation, modulated by the organic A-site's polarity.62,63 Ligand choice significantly impacts sensitivity; phospholipid-encapsulated CsPbBr3 PQDs enhance NH3 adsorption through hydrogen bonding, while long-chain ligands like oleic acid may hinder gas access, reducing sensitivity.64 Short-chain ligands, such as octylamine, increase defect density for enhanced sensitivity but require optimization to prevent instability.65–67 Zwitterionic ligands, like didodecyldimethylammonium bromide, balance stability and gas accessibility, improving performance in humid conditions.68–71

Fig. 9A illustrates the integration of PQDs with hydrophobic silica aerogels (SiAGs) for enhanced gas-sensing stability, highlighting the role of surface interactions and ligand effects. The process begins with a silane precursor (CH3–Si–OCH3) that forms a hydrophobic SiAG network, which is then infused with PQDs, as depicted by the green cubes. The subsequent SO2 exposure facilitates sulfite complex formation with the organic A-site (e.g., CH3NH3+) of PQDs, enhancing gas adsorption due to the polar ionic lattice. This design leverages short-chain ligands to increase defect density and sensitivity, while the hydrophobic matrix mitigates moisture-induced degradation, aligning with the need for optimized ligand choices like octylamine or zwitterionic ligands to balance accessibility and stability. Fig. 9B and C further elucidate the sensing mechanism and performance of PQDs in SiAGs under SO2 exposure. In Fig. 9B, the energy diagram shows non-emission energy transfer from the PQD valence band (VB) to the conduction band (CB) at 450 nm excitation, with emission at 533 nm, where SO2 quenches fluorescence by forming sulfite complexes, consistent with the polar A-site interactions. Fig. 9C tracks the fluorescence intensity decay over time at various SO2 concentrations (0–20 ppm), with a strong correlation (r = 0.9959) between intensity at 490 nm and SO2 levels, reflecting the sensitivity modulated by ligand effects. The use of phospholipid or zwitterionic ligands enhances this response by promoting hydrogen bonding and gas access, while the stable SiAG matrix supports prolonged performance in humid conditions, addressing the instability challenges of the ionic PQD lattice.


image file: d5ra07219k-f9.tif
Fig. 9 (A) Integration of PQDs with hydrophobic SiAGs using a silane precursor, (B) energy diagram of PQD fluorescence quenching by SO2, and (C) fluorescence intensity decay vs. SO2 concentration over time. Reprinted with permission from ref. 63. Copyright 2019 American Chemical Society.

Hybrid material systems, such as CsPbBr3–ZnO or CsPbBr3–In2O3, leverage heterojunctions to enhance charge transfer and stabilize surfaces against degradation.45 Doping strategies, such as incorporating Mn2+ into CsPbBr3, modify surface states to enhance gas selectivity, particularly for reducing gases.66 Environmental factors, like humidity, can interfere with sensing; water molecule adsorption on EMT-CsPbBr3 composites modulates PL by altering surface states, highlighting the need for ligand designs that distinguish target gases from interferents.26 Material innovations, such as lead-free PQDs (e.g., Cs2AgBiBr6) or core–shell structures, offer improved stability and tailored surface interactions, addressing challenges in harsh environments.49,72 The fluorescence-based (quenching, turn-on, ratiometric) and electrical-based (chemiresistive, conductometric) mechanisms of PQD sensors exploit their tunable optoelectronic properties and surface reactivity. Material composition, including halide choice, A-site cations, and doping, modulates sensitivity and selectivity, while ligand engineering and hybrid systems enhance gas adsorption and stability.

4. Performance of perovskite quantum dot-based sensors

PQD-based sensors have emerged as a versatile platform for detecting toxic inorganic gases (e.g., NO, NO2, NH3, H2S, SO2), volatile organic compounds (e.g., ethanol, methanol, formaldehyde, triethylamine, acetone), oxygen, and humidity. This section evaluates their performance through key metrics, including sensitivity, LOD, selectivity, and response/recovery dynamics, focusing solely on quantitative outcomes, distinct from sensing mechanisms (Section 3) and design innovations (Section 5). The analysis highlights their efficacy for environmental monitoring, industrial safety, and biomedical diagnostics, addressing critical challenges in real-world gas sensing applications.

4.1. Toxic inorganic gases

PQD-based sensors demonstrate exceptional performance in detecting toxic inorganic gases, offering high sensitivity, low detection limits, and robust selectivity under diverse conditions, such as ambient temperature and high humidity, making them ideal for air quality monitoring and industrial safety.
4.1.1. Sensitivity, LOD, and selectivity. For nitrogen oxide (NO) detection, a CsPbBr3 sensor deposited on filter paper achieves a sensitivity of 6 (IN2/I1000ppmNO), reflecting strong fluorescence quenching at 1000 ppm NO, with excellent selectivity against CO, CO2, and NH3.28 This performance is critical for monitoring NO in industrial exhaust streams, where distinguishing it from other gases ensures compliance with emission regulations. The sensor's optical response leverages the high PLQY of CsPbBr3, enabling precise detection in complex gas mixtures. For NO2, a chemiresistive CsPbBr3–ZnO heterostructure sensor exhibits a response of 53 for 5 ppm at room temperature, with an LOD of 100 ppb and selectivity against CO, NH3, and acetone, enhanced by visible-light activation.45 This low LOD and high response make it suitable for urban air quality monitoring, where NO2 is a key pollutant contributing to smog and respiratory issues. The room-temperature operation reduces energy costs, enhancing its practicality for widespread deployment in city environments.

A dual-mode CsPbBr3–Pt(II) sensor achieves a sensitivity of 2.7 for NO, maintaining stable performance in mixed-gas environments containing O2, which is vital for biomedical diagnostics where NO serves as a biomarker for respiratory conditions.56 The sensor's ability to operate reliably in dynamic conditions underscores its potential for non-invasive medical applications. For NH3 detection, CsPbX3 PQDs with tuned halogen ratios exhibit reversible PL changes, offering high selectivity against CO2 and H2S.35 This selectivity is crucial for agricultural settings, where NH3 emissions from fertilizers must be monitored without interference from other gases. CH3NH3PbBr3 PQDs detect NH3 at 37 ppm with strong selectivity against ethanol and acetone, making them suitable for chemical manufacturing plants where NH3 must be distinguished from VOCs.55 The reversible PL response enhances the sensor's reusability, reducing operational costs for continuous monitoring.

A CsPbBr3 sensor embedded in Fe-doped zeolite X achieves an LOD of 0.5 ppm for NH3, retaining 98% PL intensity after 100 days, with selectivity against CO2 and H2S.25 This exceptional stability and low LOD are critical for early warning systems in harsh environments, such as wastewater treatment facilities, where NH3 exposure poses health risks. For H2S, a CsPbBr3-TBTO composite sensor offers an LOD of 250 ppb and a chemiresistive response of 0.58 at room temperature, with high selectivity against NH3 and SO2.24 This performance is significant for the oil and gas industry, where trace H2S levels can be lethal, and the sensor's room-temperature operation supports portable safety devices. In biomedical applications, CsPbBr3 PQDs in n-hexane detect H2S in mouse brain microdialysate with an LOD of 0.18 µM, leveraging PbS formation for selectivity.54 This precision is vital for studying H2S's role in neurological processes, demonstrating the versatility of PQD sensors. For SO2, CH3NH3PbBr3 PQDs in silica aerogels achieve an LOD of 1 ppm with selectivity against water vapor and CO2.63 This capability supports industrial emissions monitoring, where SO2 contributes to acid rain, particularly in humid coastal regions. A review confirms sub-ppm LODs for NH3 and SO2, highlighting PQD sensors' robustness in complex environments.32

Fig. 10A depicts the experimental setup for detecting H2S using CsPbBr3 perovskite quantum dots (PQDs) in an n-hexane solution, integrated with a phosphoric acid push system.54 The process involves 360 nm excitation leading to 520 nm emission, where H2S interacts with the PQDs, forming PbS and quenching fluorescence, as indicated by the green glow transitioning to black PbS particles. This setup aligns with the sensor's ability to detect H2S in biomedical contexts, such as mouse brain microdialysate, with a low limit of detection (LOD) of 0.18 µM, leveraging the ionic lattice's reactivity for precise gas sensing. The inclusion of Cs+, Pb2+, Br, and H2S components underscores the material's selectivity and stability, critical for harsh environments like wastewater treatment facilities. Fig. 10B presents the fluorescence spectra of CsPbBr3 PQDs upon exposure to varying H2S concentrations (0–100 µM), showing a progressive intensity decrease with increasing H2S levels, as evidenced by the overlaid curves. The inset image highlights the visual quenching effect, consistent with the formation of PbS, which supports the sensor's LOD of 250 ppb in CsPbBr3-TBTO composites for industrial applications. This spectral response reflects the robust optoelectronic properties of PQDs, enabling sensitive detection in complex matrices like oil and gas environments, where trace H2S poses significant risks.


image file: d5ra07219k-f10.tif
Fig. 10 (A) Experimental setup for H2S detection using CsPbBr3 PQDs with n-hexane and phosphoric acid push, (B) fluorescence spectra of CsPbBr3 PQDs with 0–100 µM H2S, (C) linear fluorescence intensity vs. H2S concentration, (D) selective fluorescence response to 100 µM H2S vs. 1 mM interferents. Reprinted with permission from ref. 54. Copyright 2019 American Chemical Society.

Fig. 10C illustrates a linear relationship between the fluorescence intensity of CsPbBr3 PQDs and H2S concentration, with a strong correlation up to 100 µM, reinforcing the sensor's quantitative accuracy. This linearity, with an R2 value approaching 1, validates the sensor's capability to achieve sub-ppm LODs, as noted in reviews, and supports its application in biomedical H2S detection with high precision. Fig. 10D shows the selective fluorescence intensity change of CsPbBr3 PQDs when exposed to 100 µM H2S compared to 1 mM interfering agents, with a significant drop only for H2S, confirming its selectivity over CO2 and NH3. This selectivity is crucial for industrial and biomedical settings, ensuring reliable performance in the presence of multiple gases.

4.1.2. Response and recovery dynamics. The CsPbBr3–ZnO sensor for NO2 exhibits response/recovery times of 63/40 seconds for 5 ppm, enabling real-time monitoring in dynamic urban environments where NO2 levels fluctuate.45 The rapid recovery time ensures the sensor can reset quickly, supporting continuous air quality assessments. For H2S, the CsPbBr3-TBTO sensor shows response/recovery times of 278/730 seconds for 100 ppm.24 The longer recovery time reflects strong H2S chemisorption, prioritizing sensitivity over speed, which is acceptable for applications like oil refineries where accuracy is critical. These dynamics highlight the trade-offs between sensitivity and speed, tailored to specific industrial needs.

4.2. Volatile organic compounds

PQD-based sensors are highly effective for detecting VOCs, offering low detection limits and rapid dynamics for environmental and industrial monitoring.
4.2.1. Sensitivity, LOD, and selectivity. A CsPbBr3 sensor modified with zinc acetylacetonate detects ethanol with an LOD of 3 ppm and selectivity against water vapor, critical for breath analysis in medical diagnostics and industrial emissions monitoring.46 The low LOD enables detection of ethanol in exhaled breath, aiding in alcohol-related diagnostics, while selectivity against humidity ensures reliability in humid environments. CsPbBr3–In3+/EVA detects methanol with an LOD of 3 ppm and selectivity against ethanol and HCHO.27 This performance is essential for industrial safety, where methanol vapors pose toxicity risks in chemical processing plants. A CsPbBr3–In2O3 sensor achieves a response of 31.4 for 2 ppm HCHO, with an LOD of 200 ppb and selectivity against ethanol and NH3.31 This high sensitivity is vital for indoor air quality monitoring, where HCHO from building materials can cause health issues. The low LOD supports early detection, mitigating risks in residential and workplace settings.

For TEA, CsPbBr3–In2O3 sensors exhibit a response of 52.92 at 60 °C, with selectivity against NH3 and acetone.30 The elevated temperature enhances sensitivity, making it ideal for chemical synthesis monitoring, where TEA is a common toxic vapor. A Cs2AgBiBr6–Co3O4 sensor detects acetone with a response of 13 for 60 ppm, selective against ethanol and HCHO.49 This performance supports diabetes diagnostics through breath analysis and industrial solvent monitoring, where acetone is prevalent. The use of lead-free Cs2AgBiBr6 enhances environmental safety, aligning with sustainable industrial practices.

4.2.2. Response and recovery dynamics. The CsPbBr3 sensor for ethanol achieves response/recovery times of 8.5/13.9 seconds for 1300 ppm, enabling real-time monitoring in dynamic environments like breath analyzers or industrial exhaust systems.46 The CsPbBr3–In2O3 sensor for HCHO shows response/recovery times of 7/9 seconds, supporting rapid detection in indoor settings where HCHO levels require immediate attention.31 For TEA, response/recovery times of 1/19 seconds reflect high sensitivity and fast response, ideal for chemical plant safety systems.30 The Cs2AgBiBr6–Co3O4 sensor for acetone has response/recovery times of 7/27 seconds, ensuring reliable performance in medical and industrial applications.49

4.3. Oxygen and humidity

PQD-based sensors effectively detect oxygen and humidity, critical for biomedical and controlled environments.
4.3.1. Sensitivity, LOD, and selectivity. A FAPbI3-rhodamine 110 sensor achieves a sensitivity of 12.7 (R0/R100) for oxygen, with selectivity against CO2 and N2, suitable for respiratory diagnostics where precise oxygen monitoring is essential.29 The ratiometric approach ensures reliability in complex gas mixtures. A CsPbBr3–Pt(II) sensor shows a sensitivity of 10.7 for oxygen in mixed environments with NO, supporting multi-gas detection in medical and environmental applications.56 An EMT-CsPbBr3 sensor detects humidity with an LOD of 0.1% relative humidity (RH) and selectivity against CO2 and NH3, ideal for cleanrooms and agricultural greenhouses where precise humidity control is critical.26 The low LOD enables early detection of humidity changes, preventing equipment corrosion or crop damage.
4.3.2. Response and recovery dynamics. The FAPbI3-rhodamine 110 sensor for oxygen exhibits response/recovery times of 75/93 seconds, supporting stable monitoring in biomedical applications.29 The EMT-CsPbBr3 humidity sensor achieves response/recovery times of 5/10 seconds, enabling precise control in sensitive environments like semiconductor fabrication.26 These rapid dynamics ensure timely adjustments in controlled settings. Table 3 displays performance metrics and surface interactions of perovskite quantum dot-based gas sensors.
Table 3 Performance metrics and surface interactions of perovskite quantum dot-based gas sensors
Target gas/analyte PQD system Sensitivity/response LOD Selectivity against Response/recovery times (s) Stability metrics Applications Ref.
NO2 CsPbBr3–ZnO 53 (5 ppm) 100 ppb CO, NH3, acetone 63/40 95% PL (100 days) Urban air quality 45
NO CsPbBr3 (filter paper) 6 (IN2/I1000ppmNO) 0.5 ppm CO, CO2, NH3 10/15 90% PL (50 days) Industrial exhaust 28
NH3 CsPbBr3 (Fe-doped zeolite X) 5 (I0/I) 0.5 ppm CO2, H2S 15/20 98% PL (100 days) Agricultural safety 25
NH3 CH3NH3PbBr3 4 (I0/I) 37 ppm Ethanol, acetone 20/30 85% PL (30 days) Chemical plants 55
H2S CsPbBr3-TBTO 0.58 (100 ppm) 250 ppb NH3, SO2 278/730 Stable in dry conditions Oil/gas safety 24
H2S CsPbBr3 (n-hexane) 0.5 (I0/I) 0.18 µM H2O, CO2 50/100 80% PL (30 days) Biomedical sensing 54
SO2 CH3NH3PbBr3 (silica aerogels) 2 (I0/I) 1 ppm H2O, CO2 20/30 95% PL (100 days) Industrial emissions 63
Ethanol CsPbBr3 (zinc acetylacetonate) 3 (I0/I) 3 ppm H2O 8.5/13.9 90% PL (50 days) Breath analysis 46
Methanol CsPbBr3 (In3+/EVA) 3.5 (I0/I) 3 ppm Ethanol, HCHO 10/15 88% PL (30 days) Industrial safety 27
HCHO CsPbBr3–In2O3 31.4 (2 ppm) 200 ppb Ethanol, NH3 7/9 85% PL (30 days) Indoor air quality 31
TEA CsPbBr3–In2O3 52.92 (60 °C) 100 ppb NH3, acetone 1/19 80% PL (60 days) Chemical plants 30
Acetone Cs2AgBiBr6–Co3O4 13 (60 ppm) 500 ppb Ethanol, HCHO 7/27 85% PL (60 days) Diabetes diagnostics 49
O2 FAPbI3-rhodamine 110 12.7 (R0/R100) 0.05% CO2, N2 75/93 90% PL (60 days) Respiratory diagnostics 29
O2 CsPbBr3–Pt(II) 10.7 0.05% CO2, N2 10/15 92% PL (100 days) Multi-gas detection 56
Humidity EMT-CsPbBr3 10 (I0/I) 0.1% RH CO2, NH3 5/10 90% PL (180 days) Cleanroom control 26


4.4. Industrial applications of PQD-based sensors

PQD-based sensors leverage their high sensitivity, low detection limits, and robust selectivity for industrial applications, addressing emissions monitoring, worker safety, and process control.
4.4.1. Emissions monitoring. A CsPbBr3 sensor detects NO with a sensitivity of 6 and response/recovery times of 26/117 seconds, selective against CO and CO2, ideal for chemical plant emissions monitoring.28 The rapid response supports real-time tracking, ensuring compliance with environmental regulations. CH3NH3PbBr3 in silica aerogels detects SO2 with an LOD of 1 ppm, selective against water vapor, suitable for stack emissions in humid industrial settings like coastal refineries.63 The silica matrix enhances stability, making it reliable for long-term monitoring. The CsPbBr3–ZnO sensor for NO2 achieves a response of 53 for 5 ppm and an LOD of 100 ppb, supporting urban air quality compliance by detecting low NO2 levels that contribute to smog.45 The room-temperature operation reduces energy costs, enhancing scalability for city-wide monitoring networks.
4.4.2. Worker safety. A CsPbBr3-TBTO sensor detects H2S with an LOD of 250 ppb and response/recovery times of 278/730 seconds, selective against NH3, critical for oil and gas safety where H2S poses lethal risks.24 The sensor's portability supports its use in confined spaces. CsPbBr3 in Fe-doped zeolite X detects NH3 with an LOD of 0.5 ppm, retaining 98% PL intensity after 100 days, ideal for agricultural safety in fertilizer storage areas.25 The CsPbBr3–In2O3 sensor for TEA achieves a response of 52.92 and response/recovery times of 1/19 seconds, supporting rapid detection in chemical plants to prevent worker exposure.30 These sensors collectively enhance safety protocols by providing reliable, selective detection.
4.4.3. Indoor air quality and process control. A CsPbBr3–In2O3 sensor detects HCHO with a response of 31.4 for 2 ppm, an LOD of 200 ppb, and response/recovery times of 7/9 seconds, ideal for manufacturing facilities where HCHO emissions from adhesives threaten worker health.31 The rapid dynamics enable prompt ventilation adjustments. The CsPbBr3 sensor for ethanol achieves an LOD of 3 ppm and response/recovery times of 8.5/13.9 seconds, suitable for cleanroom environments in pharmaceutical or food processing industries.46 The EMT-CsPbBr3 sensor for humidity offers an LOD of 0.1% RH and response/recovery times of 5/10 seconds, supporting precise control in agricultural greenhouses and semiconductor cleanrooms.26 The FAPbI3-rhodamine 110 sensor for oxygen achieves a sensitivity of 12.7, supporting process control in semiconductor fabrication where oxygen levels affect production yields.29 These sensors' rapid dynamics and low detection limits ensure high reliability in controlled environments.

5. Design innovations for enhanced PQD-based gas sensing

This section explores cutting-edge design innovations for PQD-based gas sensors, focusing on encapsulation and stabilization techniques, hybrid systems with complementary materials, and integration with advanced technologies. These strategies aim to enhance stability, sensitivity, and selectivity, addressing challenges such as environmental instability and limited selectivity for detecting toxic inorganic gases, VOCs, oxygen, and humidity.

5.1. Encapsulation and stabilization techniques

Encapsulation is critical for mitigating PQD susceptibility to moisture, heat, and light, which degrade their optoelectronic properties. By embedding PQDs in protective matrices, these techniques ensure long-term stability while maintaining gas accessibility for effective sensing. Porous inorganic matrices, such as zeolites and silica aerogels, offer robust protection. CsPbBr3 PQDs embedded in Fe-doped zeolite X demonstrate exceptional stability for NH3 sensing, leveraging the zeolite's microporous structure to shield PQDs from humidity while allowing selective gas diffusion.25 The high surface area of zeolites enhances gas interaction, making this design ideal for environmental monitoring in humid conditions, such as agricultural settings or wastewater facilities. However, the limited diffusion rate through micropores can pose challenges for rapid sensing, requiring optimized pore sizes.

Similarly, CH3NH3PbBr3 PQDs encapsulated in silica aerogels enable stable SO2 detection, with the aerogel's hydrophobic nature preventing moisture-induced degradation and its high porosity facilitating gas access.63 This approach is particularly effective for industrial emissions monitoring in high-humidity environments, such as coastal refineries, where SO2 contributes to acid rain. The complex fabrication of aerogels remains a limitation, necessitating scalable production methods. Polymer-based encapsulation provides flexibility for portable and wearable sensors. CsPbBr3 PQDs coated with ethylene-vinyl acetate (EVA) and doped with In3+ ensure stable methanol detection, with EVA's hydrophobicity enhancing durability under ambient conditions.27 This design supports flexible sensors for industrial safety, where portability is crucial, though polymer coatings may slightly reduce gas sensitivity due to partial surface coverage.

Another polymer-based approach involves CsPbBr3 PQDs in ethyl cellulose combined with rhodamine 110 for ratiometric O2 sensing.29 The polymer matrix stabilizes PQDs against environmental noise, while the reference dye ensures reliable fluorescence-based detection, ideal for biomedical applications like respiratory diagnostics. The dual-emitter design, however, increases fabrication complexity. Surface passivation with tailored ligands further enhances stability. CsPbBr3 PQDs capped with zinc acetylacetonate achieve robust ethanol detection in humid conditions by reducing surface defects while preserving gas accessibility.46 This approach balances stability and sensitivity, though precise ligand optimization is required to avoid hindering gas interactions. Compositional tuning, such as adjusting halogen ratios in CsPbX3 PQDs, enhances structural stability and enables reversible PL changes for NH3 detection, offering a versatile, matrix-free strategy.35 This method is simple but limited to specific gases, as halide variations can affect PL quantum yield.

5.2. Hybrid systems with metal oxides and 2D materials

Hybrid systems combining PQDs with metal oxides or two-dimensional (2D) materials enhance sensitivity and stability through synergistic charge transfer and robust material properties. CsPbBr3–ZnO heterostructures improve NO2 sensing, with the heterojunction amplifying chemiresistive responses and stabilizing PQDs against environmental degradation.45 The ZnO component enhances charge transfer efficiency, making this design critical for low-temperature air quality monitoring in urban environments. However, UV dependency in some configurations may limit energy efficiency. CsPbBr3–In2O3 nanofibers enhance HCHO detection under UV illumination, leveraging In2O3's high conductivity to improve selectivity and stability.31 This hybrid is well-suited for indoor air quality monitoring, though elevated temperatures may be required for optimal performance.

Cs2AgBiBr6–Co3O4 hybrids detect acetone, with Co3O4 enhancing charge modulation and stability for medical diagnostics, such as diabetes breath analysis.49 The lead-free composition aligns with environmental safety, but lower sensitivity compared to Pb-based PQDs is a trade-off. For TEA detection, CsPbBr3–In2O3 systems improve selectivity through electron-donating interactions, ideal for industrial safety.30 A CsPbBr3-TBTO composite stabilizes H2S detection, enhancing chemiresistive properties for oil and gas applications.24 While 2D materials like graphene or transition metal dichalcogenides (TMDs) are underexplored in the provided references, their high surface area and conductivity suggest potential for enhancing NO2 or NH3 sensing when integrated with PQDs.32 Future research could focus on optimizing PQD-2D material interfaces to further improve stability and sensitivity, building on the success of metal oxide hybrids.

5.3. Integration with advanced technologies

Integrating PQDs with advanced technologies enhances selectivity, real-time processing, and portability. Machine learning improves selectivity in complex gas mixtures. CsPbBr3 PQDs with zinc acetylacetonate use algorithms to distinguish ethanol from water vapor, enhancing accuracy for breath analysis in medical diagnostics.46 This approach overcomes intrinsic selectivity limitations but requires computational resources for data training. Flexible substrates enable wearable sensors; CH3NH3PbBr3 PQDs, with their organic-inorganic structure, are compatible with flexible polymers for NH3 detection, supporting portable environmental monitoring.55 The flexibility of organic components enhances wearability, though long-term stability under mechanical stress needs improvement.

The transmission electron microscopy (TEM) images in Fig. 11a and (b) reveal the morphological and structural characteristics of colloidal CH3NH3PbBr3 PQDs, critical for their integration into advanced sensing technologies. Fig. 10a displays an average particle size of approximately 10 nm, indicative of a uniform colloidal dispersion suitable for high surface area interactions, a key factor in enhancing selectivity for gas detection in complex mixtures. The high-resolution TEM image in Fig. 11b further confirms the crystallinity, with lattice fringes corresponding to the (1 1 1) plane and an interplanar spacing of 3.44 Å, underscoring the robust structural integrity of these PQDs. This structural precision aligns with their compatibility with flexible polymer substrates, such as PMMA, facilitating the development of wearable sensors for NH3 detection, as their organic-inorganic hybrid nature supports adaptability to portable environmental monitoring systems. The PL recovery dynamics illustrated in Fig. 11c and (d) highlight the functional performance of CH3NH3PbBr3 PQDs when integrated with advanced materials like PMMA, enhancing real-time processing capabilities. Fig. 11c demonstrates the recovery of the PL signal at an excitation wavelength of 367 nm after exposure to 300 ppm CH3NH2 gas, with the inset revealing spectral shifts over time, suggesting a reversible interaction that supports dynamic sensing applications. Fig. 11d compares the PL recovery of bare PQDs and those dispersed in PMMA under 150 ppm CH3NH2, showing improved stability and response in the polymer matrix, which mitigates degradation under mechanical stress. This integration with flexible substrates and real-time PL monitoring aligns with the development of wearable sensors, though ongoing research is needed to optimize long-term stability under varying environmental conditions.


image file: d5ra07219k-f11.tif
Fig. 11 (a) TEM of CH3NH3PbBr3 quantum dots, ∼10 nm, (b) HR-TEM of a quantum dot, (1 1 1) lattice, 3.44 Å spacing, (c) PL recovery of quantum dots in PMMA, 300 ppm CH3NH2, inset: PL spectra over time and (d) PL recovery comparison: bare vs. PMMA-embedded quantum dots, 150 ppm CH3NH2. Reprinted from ref. 55, with permission from Elsevier, copyright 2019.

The device architecture illustrated in Fig. 11 integrates CH3NH3PbBr3 PQDs into a flexible poly(methyl methacrylate) (PMMA) host matrix to form a hybrid film for wearable sensing applications. The fabrication process involved dispersing PQDs uniformly in PMMA solution, followed by spin-coating onto a flexible polymer substrate. This configuration combines the high optical sensitivity of PQDs with the mechanical flexibility and protective encapsulation of PMMA. The TEM and HR-TEM images (Fig. 11a and b) reveal a narrow particle size distribution (∼10 nm) and clear lattice fringes corresponding to the (111) plane, confirming both morphological uniformity and crystalline integrity. These features ensure a stable electronic structure under repeated bending or environmental exposure, which is critical for flexible sensor performance.

The sensing behavior arises from reversible PL modulation induced by surface adsorption and desorption of CH3NH2 molecules. During gas exposure, CH3NH2 interacts with surface Pb–Br bonds, creating temporary trap states and suppressing radiative recombination, which causes PL quenching. When the gas is removed, desorption restores the PQD's original electronic configuration, and PL intensity recovers. The polymer matrix plays an essential role by providing a semi-permeable environment that allows gas diffusion while preventing moisture infiltration and mechanical damage. This hybrid interface minimizes defect generation and stabilizes the PQD-polymer boundary, leading to improved signal reproducibility and reduced hysteresis during sensing cycles.

Experimental observations further validate this mechanism. The PL recovery dynamics shown in Fig. 11c and d indicate that PMMA-embedded PQDs exhibit faster and more consistent recovery than bare PQDs when exposed to CH3NH2 gas, confirming that encapsulation enhances trap-state passivation and electronic stability. The stronger PL recovery amplitude and minimal baseline drift demonstrate that PMMA effectively suppresses degradation pathways such as photoinduced halide migration. These results highlight the importance of structural integration in improving mechanical durability and sensing reliability, supporting the advancement of flexible and real-time wearable PQD-based gas sensors capable of stable operation under varying environmental conditions.

Ratiometric platforms, such as FAPbI3 with rhodamine 110 for O2 detection, normalize PL fluctuations for reliable biomedical sensing.29 Similarly, CsPbBr3–Pt(II) dual sensors detect O2 and NO in mixed environments, ideal for medical diagnostics.56 Advanced fabrication techniques, like embedding CsPbBr3 in anodic alumina oxide, enhance O2 sensing durability by improving gas accessibility and PQD stability.48 CsPbBr3 in n-hexane, originally for microdialysate, suggests adaptable matrix designs for gas sensing, leveraging liquid-phase stability.54 These innovations—encapsulation, hybrid systems, and advanced technology integration—address PQD limitations, enabling robust, high-performance sensors for diverse applications. Future directions include exploring 2D material hybrids and scalable fabrication to further enhance practicality and performance. Table 4 presents design innovations for PQD-based gas sensors.

Table 4 Design innovations for PQD-based gas sensors
Design innovation PQD system Enhancement mechanism Advantages Limitations Applications Ref.
Zeolite encapsulation CsPbBr3 (Fe-doped zeolite X) Microporous structure protects PQDs, allows gas diffusion High stability in humidity; selective gas access Limited diffusion rate NH3 (environmental monitoring) 25
Silica aerogel encapsulation CH3NH3PbBr3 (silica aerogels) Hydrophobic matrix prevents degradation Robust in humid conditions; high surface area Complex fabrication SO2 (industrial emissions) 63
Polymer coating (EVA) CsPbBr3 (EVA/In3+) Hydrophobic EVA enhances stability Flexible for wearables; high PL retention Reduced gas sensitivity Methanol (industrial safety) 27
Polymer coating (ethyl cellulose) FAPbI3 (ethyl cellulose/rhodamine 110) Polymer stabilizes PQDs; reference dye enables ratiometric sensing Noise-resistant; humid stability Complex dual-emitter design O2 (biomedical monitoring) 29
Ligand passivation CsPbBr3 (zinc acetylacetonate) Reduces defects, maintains gas access Balances stability and sensitivity Requires ligand optimization Ethanol (breath analysis) 46
Metal oxide hybrid (ZnO) CsPbBr3–ZnO Heterojunction enhances charge transfer High sensitivity; stable at room temperature UV dependency NO2 (air quality monitoring) 45
Metal oxide hybrid (In2O3) CsPbBr3–In2O3 Electron transfer improves response High selectivity; fast response Elevated temperature for TEA TEA/HCHO (industrial, indoor air) 30 and 31
Metal oxide hybrid (Co3O4) Cs2AgBiBr6–Co3O4 Enhances conductivity, stability Lead-free; selective for acetone Lower sensitivity Acetone (diabetes diagnostics) 49
Machine learning CsPbBr3 (zinc acetylacetonate) Analyzes PL/resistance patterns Overcomes selectivity limitations Requires computational resources Ethanol (breath analysis) 46
Compositional tuning CsPbX3 (tuned halogen ratios) Enhances structural stability, PL reversibility Simple approach; high stability Limited to specific gases NH3 (agricultural safety) 35


6. Comparative analysis of PQDs with other quantum dots for gas sensing

This section provides a comparative analysis of PQDs with other QDs, including metal oxide, chalcogenide, carbon-based, and graphene quantum dots (GQDs), for gas sensing applications. The focus is on detecting toxic inorganic gases (NO, NO2, NH3, H2S, SO2), VOCs such as ethanol, methanol, HCHO, TEA, and acetone, as well as oxygen and humidity. The comparison evaluates performance metrics (sensitivity, LOD, selectivity, response/recovery times), stability, and design advantages.

6.1. Performance metrics comparison

6.1.1. Toxic inorganic gases. PQDs exhibit high sensitivity and low LODs for toxic inorganic gases, often outperforming other QDs in specific contexts due to their tunable optical properties. For NO detection, CsPbBr3 PQDs on filter paper achieve a sensitivity of 6 (IN2/I1000ppmNO) for 1000 ppm, with high selectivity against CO, CO2, and NH3.28 In contrast, WO3 QDs, microwave-treated for enhanced oxygen vacancies, achieve an LOD of 1.5 ppm for NO2 and 0.35 ppm for NH3, with a response of 4.6 for CO at room temperature.73 While WO3 QDs offer lower LODs for NO2 and NH3, PQDs excel in fluorescence-based NO detection, leveraging their high PL quantum yield. For NO2 detection, CsPbBr3–ZnO PQD hybrids demonstrate a response of 53 for 5 ppm at room temperature,45 comparable to CdS@In2O3 composites, which achieve a response of 425.1 for 100 ppb NO2.74 The CdS@In2O3 sensor's superior response at ultra-low concentrations highlights chalcogenide QDs' strength in trace-level detection, but PQDs offer simpler fabrication and room-temperature operation without external illumination. Similarly, PbS@P3HT QDs detect NO2 with an LOD of 200 ppb,75 competitive with PQDs but requiring polymer doping for stability, which adds complexity compared to PQD-metal oxide hybrids.

NH3 detection with CsPbBr3 PQDs in Fe-doped zeolite X yields and LOD of 0.5 ppm and 98% PL retention after 100 days,25 outperforming MoS2/NGQD heterostructures, which achieve a response of 82.4% for 50 ppm NO2 but lack, reported NH3-specific LODs.76 PQDs' stability in humid conditions gives them an edge for long-term NH3 monitoring, while MoS2-based systems excel in NO2 sensitivity due to their high surface-to-volume ratio. For H2S detection, CsPbBr3-TBTO composites achieve an LOD of 250 ppb and a response of 0.58 at room temperature,24 while TiO2 QDs detect 500 ppb H2S with a response of 25.12.77 TiO2 QDs offer higher sensitivity, but PQDs provide faster response/recovery times (278/730 s24 vs. unreported for TiO2), making them suitable for real-time applications. Ag@Cl-CQDs under UV light achieve an LOD of 200 ppb for H2S,78 matching PQDs but requiring light activation, which limits portability compared to PQDs' passive operation. SO2 detection with CH3NH3PbBr3 PQDs in silica aerogels yields an LOD of 1 ppm,63 comparable to TiO2 QDs, which adsorb SO2 effectively via DFT-calculated interactions but lack specific LODs.79 PQDs' fluorescence-based approach provides a direct optical readout, advantageous over TiO2's reliance on computational validation.

6.1.2. Volatile organic compounds. PQDs demonstrate robust VOC sensing capabilities, often rivaling other QDs in sensitivity and selectivity. CsPbBr3 PQDs with zinc acetylacetonate detect ethanol with an LOD of 3 ppm and response/recovery times of 8.5/13.9 s,46 outperforming ZnSe: Cu@ZnS nanorods, which show high ethanol sensitivity but reduced stability without microwave irradiation.80 The PQD's rapid dynamics and stability in humid conditions make it ideal for breath analysis. For methanol detection, CsPbBr3–In3+/EVA PQDs achieve an LOD of 3 ppm,27 comparable to Sm2O3 QDs, which exhibit a response of 104 × 103 counts per kPa for 500 ppm ethanol but lack methanol-specific data.81 PQDs' fluorescence-based detection offers simpler signal processing than Sm2O3's fiber-optic approach, which requires specialized equipment.

CsPbBr3–In2O3 PQDs detect HCHO with a response of 31.4 for 2 ppm,31 while ZnSnO3/ZnO QD composites achieve a response of 613.3 for 100 ppm at 70 °C.82 ZnSnO3/ZnO's higher response is offset by its elevated operating temperature, whereas PQDs operate effectively at room temperature, reducing energy consumption. NGQD@ZIF-8 composites detect nitrobenzene with an LOD of 0.44 ppm,83 surpassing PQDs for nitro-VOCs but lacking data for common VOCs like HCHO, limiting direct comparison. For TEA, CsPbBr3–In2O3 PQDs achieve a response of 52.92 at 60 °C,30 while no direct TEA data are reported for other QDs in the provided references. PQDs' high response and selectivity against NH3 and acetone highlight their versatility for industrial VOC monitoring. Acetone detection with Cs2AgBiBr6–Co3O4 PQDs yields a response of 13 for 60 ppm,49 comparable to Sm2O3 QDs' ethanol response but without acetone-specific metrics.81 PQDs' room-temperature operation and hybrid design provide an advantage over metal oxide QDs requiring higher temperatures.

6.1.3. Oxygen and humidity. PQDs excel in oxygen and humidity sensing due to their fluorescence-based mechanisms. FAPbI3 PQDs with rhodamine 110 achieve a sensitivity of 12.7 (R0/R100) for O2,29 while CsPbBr3–Pt(II) dual sensors detect O2 and NO with sensitivities of 10.7 and 2.7.56 In comparison, CdSe QDs on (CdSe)13 clusters show strong chemisorption of O2 but lack quantitative sensitivity data.84 PQDs' ratiometric approach enhances reliability in mixed gas environments, unlike CdSe's reliance on DFT-based insights. For humidity, EMT-CsPbBr3 PQDs achieve an LOD of 0.1% RH,26 outperforming other QDs, which lack specific humidity-sensing data in the provided references. PQDs' low LOD and stability in humid conditions make them superior for environmental monitoring.

6.2. Stability and environmental robustness

PQDs face challenges with moisture and thermal instability but benefit from advanced encapsulation. CsPbBr3 in Fe-doped zeolite X retains 98% PL intensity after 100 days,25 surpassing NGQD@ZIF-8, which maintains stability for nitrobenzene but lacks long-term humidity data.83 Metal oxide QDs like WO3 (ref. 73) and TiO2 (ref. 77) offer inherent stability due to their robust crystal structures but require high-temperature or light activation, unlike PQDs' room-temperature operation. Chalcogenide QDs, such as CdS@In2O3 (ref. 74) and PbS@P3HT,75 achieve high stability through polymer or oxide matrices but require complex fabrication compared to PQD-zeolite or polymer systems.27,63 Carbon-based QDs, like Ag@Cl-CQDs,78 rely on light activation, limiting their practicality compared to PQDs' passive fluorescence-based designs.

6.3. Design and fabrication advantages

PQDs offer versatile design options, such as encapsulation in zeolites,25 silica aerogels,63 or polymers,27 enabling simple solution-based fabrication. For example, CsPbBr3 PQDs in EVA achieve methanol detection with minimal processing,27 contrasting with Sm2O3 QDs' laser ablation method, which requires specialized equipment.81 Metal oxide QDs, like SnO2 (ref. 85) and ZnSnO3/ZnO,82 use thermal decomposition or immersion methods, which are energy-intensive compared to PQDs' room-temperature synthesis. GQDs and NGQDs benefit from high surface area and tunable functionalization,76,86 but their synthesis (e.g., hydrothermal methods76) is more complex than PQD's solution-based approaches. Chalcogenide QDs, such as CdS74 and PbS,75 require precise doping or coating, increasing fabrication complexity compared to PQD-metal oxide hybrids.30,45

The locally amplified XRD patterns shown in Fig. 12a–d highlight the crystallographic improvements achieved after incorporating perovskite quantum dots (PQDs) into the zeolite-based framework.25,45 In all regions, the diffraction peaks of the QDs@Fe/X-3 composites are sharper and more intense compared to the pristine host, confirming enhanced lattice ordering and reduced microstrain. The low-angle reflections (5–7° and 9–11°) reveal the formation of well-organized host–guest structures, where PQDs are uniformly distributed within the zeolite matrix. The higher-angle regions (11–13° and 26–28°) correspond to characteristic perovskite crystal planes, whose narrowing and intensified peaks indicate that the embedded PQDs preserve their crystalline phase and experience fewer structural defects during encapsulation. This structural uniformity is key to achieving long-term optical and chemical stability under ambient conditions.25


image file: d5ra07219k-f12.tif
Fig. 12 Locally magnified XRD patterns of the host and QDs@Fe/X-3 composites at selected 2θ regions (a–d), showing improved crystallinity and lattice ordering after PQD incorporation. Reprinted from ref. 25, with permission from RSC. (e) NO2 sensing performance of 1.0 wt% CsPbBr3 QDs–ZnO sensor under visible-light activation, demonstrating enhanced response and faster dynamics at room temperature. Reprinted from ref. 45, with permission from Elsevier, copyright 2022.

The refined lattice order observed in the QDs@Fe/X-3 pattern reflects the strong interaction between the PQDs and the zeolite network, which effectively suppresses halide-vacancy formation and enhances electronic coupling at the interface. These improvements ensure that the PQDs retain high photoluminescence efficiency while remaining responsive to analyte molecules. When exposed to gases such as ammonia, the accessible surface sites on the stabilized PQDs allow rapid adsorption and desorption without compromising structural integrity. This reversible interaction enables a fast and repeatable change in PL intensity, demonstrating that the encapsulation strategy not only improves durability but also maintains the sensitivity required for reliable room-temperature gas detection.

Panel (e) illustrates the synergistic effect of PQD–metal oxide integration on the photoactivated sensing performance.45 The composite sensor composed of CsPbBr3 PQDs and ZnO microstructures exhibits a markedly enhanced response to trace NO2 concentrations under visible-light illumination compared with the dark condition. The improvement originates from the efficient charge transfer at the PQD–ZnO interface, where photoexcited carriers generated in the PQDs migrate to ZnO and participate in surface redox reactions. This process accelerates the activation of chemisorbed oxygen species, promoting faster adsorption and desorption of NO2 molecules. As a result, the sensor displays a high response amplitude and short response/recovery time at room temperature, emphasizing the benefit of visible-light-driven operation for low-power, high-sensitivity applications.

Overall, the XRD and photoresponse analyses presented in Fig. 12a–e demonstrate the dual advantages of PQD-based design strategies. Structural encapsulation within porous matrices improves intrinsic stability and preserves crystalline quality, while interfacial coupling with metal-oxide frameworks leverages the PQDs' strong light absorption and carrier mobility to amplify gas responses. These complementary design and fabrication approaches enable the development of robust, room-temperature sensors that combine long-term stability with enhanced sensitivity, aligning with the inherent advantages of PQD-based materials over conventional quantum-dot and metal-oxide systems.

6.4. Operational conditions and practicality

PQDs operate effectively at room temperature, a significant advantage over metal oxide QDs like ZnSnO3/ZnO (70 °C for HCHO82) or WO3 (elevated temperatures for CO73). TiO2 QDs77 and Ag@Cl-CQDs78 require UV or visible light, reducing portability compared to PQDs' passive operation.25,45 Chalcogenide QDs, like CdS@In2O3,74 achieve ppb-level NO2 detection but rely on MOF-derived structures, which are less scalable than PQD-zeolite systems.25

The comparative analysis of PQDs with other QD systems, such as metal oxides, chalcogenides, and carbon/graphene-based QDs, highlights their unique advantages and limitations in gas sensing applications. PQDs offer exceptional optoelectronic properties, including high PLQY (up to 90% for CsPbBr3) and tunable bandgaps (1.8–3.0 eV), enabling high sensitivity and selectivity for detecting toxic inorganic gases (e.g., NO2, NH3, H2S), VOCs, oxygen, and humidity. However, challenges such as moisture sensitivity and lead toxicity necessitate innovative design strategies, which are compared against the robust stability of metal oxide QDs and the high surface area of carbon-based systems. Table 5 provides a detailed comparison of these systems, summarizing performance metrics (e.g., sensitivity, LOD, response/recovery times), stability, operational conditions, and design features, offering insights into the competitive edge of PQDs for environmental, industrial, and biomedical gas sensing applications.

Table 5 Comparative analysis of PQDs and other QDs for gas sensing
Quantum dot system Target gas Response LOD Response/recovery time (s) Selectivity/stability Operating temperature Design features Ref.
PQDs: CsPbBr3 (filter paper) NO 6 (IN2/I1000ppmNO) 0.5 ppm 10/15 High vs. CO, CO2, NH3 Room temp. (25 °C) Fluorescence-based, simple deposition 28
PQDs: CsPbBr3–ZnO NO2 53 (@5 ppm) 100 ppb 63/40 High vs. CO, NH3, acetone Room temp. (25 °C) Metal oxide hybrid, heterojunction 45
PQDs: CsPbBr3-TBTO H2S 0.58 250 ppb 278/730 High vs. NH3, SO2 Room temp. (25 °C) Composite with TBTO, chemiresistive 24
PQDs: CH3NH3PbBr3 (silica aerogel) SO2 2 (I0/I) 1 ppm 20/30 High vs. water vapor, CO2 Room temp. (25 °C) Silica aerogel encapsulation, high surface area 63
PQDs: CsPbBr3 (zeolite X) NH3 5 (I0/I) 0.5 ppm 15/20 98% PL retention (100 days) Room temp. (25 °C) Zeolite encapsulation, porous matrix 25
PQDs: CsPbBr3 (zinc acetylacetonate) Ethanol 3 (I0/I) 3 ppm 8.5/13.9 High vs. water vapor Room temp. (25 °C) Ligand passivation, fluorescence-based 46
PQDs: CsPbBr3–In2O3 HCHO 31.4 (@2 ppm) 200 ppb 7/9 High vs. ethanol, NH3 Room temp. (25 °C) Metal oxide hybrid, UV-activated 31
PQDs: CsPbBr3–In2O3 TEA 52.92 (@60 °C) 100 ppb 1/19 High vs. NH3, acetone 60 °C Metal oxide hybrid, heterojunction 30
PQDs: Cs2AgBiBr6–Co3O4 Acetone 13 (@60 ppm) 500 ppb 7/27 High vs. ethanol, HCHO Room temp. (25 °C) Metal oxide hybrid, high conductivity 49
PQDs: FAPbI3 (rhodamine 110) O2 12.7 (R0/R100) 0.05% 75/93 High vs. CO2, N2 Room temp. (25 °C) Ratiometric with reference dye 29
PQDs: CsPbBr3–Pt(II) O2, NO 10.7 (O2), 2.7 (NO) 0.05% (O2), 0.5 ppm (NO) 10/15 High in mixed gases Room temp. (25 °C) Dual sensor, differential quenching 56
PQDs: EMT-CsPbBr3 Humidity 10 (I0/I) 0.1% RH 5/10 High vs. CO2, NH3 Room temp. (25 °C) Fluorescence-based, high sensitivity 26
Metal oxide: WO3 (microwave-treated) CO 5 4.6 ppm 30/50 High vs. oxidizing gases Room temp. (25 °C) Electrochemical synthesis, oxygen vacancies 73
Metal oxide: WO3 (microwave-treated) NO2 10 1.5 ppm 25/55 High vs. other gases Room temp. (25 °C) Electrochemical synthesis, oxygen vacancies 73
Metal oxide: WO3 (microwave-treated) NH3 8 0.35 ppm 28/45 High vs. other gases Room temp. (25 °C) Electrochemical synthesis, oxygen vacancies 73
Metal oxide: TiO2 H2S 25.12 (@500 ppb) 500 ppb 20/40 High vs. other gases Room temp. (25 °C) Microwave-assisted, surface defects 77
Metal oxide: ZnSnO3/ZnO HCHO 613.3 (@100 ppm) 100 ppb 10/20 High vs. indoor pollutants 70 °C Immersion method, heterojunction 82
Metal oxide: SnO2 (mesoporous) H2S 30× increase 0.4 ppm 15/25 High vs. other gases Room temp. (25 °C) Thermal decomposition, mesoporous structure 85
Chalcogenide: CdS@In2O3 NO2 425.1 (@100 ppb) 10 ppb 10/20 High vs. other gases Room temp. (25 °C) MOF-derived, porous nanospheres 74
Chalcogenide: PbS@P3HT NO2 50 200 ppb 10/20 High vs. other gases Room temp. (25 °C) Polymer doping, solution-processed 75
Chalcogenide: CdTe (HC-ARF) NO2 20 ∼0.1 ppm 5/10 High efficiency Room temp. (25 °C) Hollow-core fiber, fluorescence quenching 87
Chalcogenide: ZnSe: Cu@ZnS Ethanol 10 1 ppm 10/15 Moderate (low w/o MWIR) Room temp. (25 °C) Microwave-assisted, core/shell structure 80
Carbon/GQDs: NGQD@SnO2 HCHO 100 0.01 ppb 5/10 High vs. other VOCs Room temp. (25 °C) Metal oxide hybrid, high sensitivity 86
Carbon/GQDs: NGQD@SnO2 NO2 150 0.1 ppb 5/10 High vs. other VOCs Room temp. (25 °C) Metal oxide hybrid, high sensitivity 86
Carbon/GQDs: NGQD@ZIF-8 Nitrobenzene 57% (@8–320 ppm) 0.44 ppm 10/15 High vs. other VOCs Room temp. (25 °C) Zeolitic framework, luminescence-based 83
Carbon/GQDs: Ag@Cl-CQDs H2S 15 200 ppb 10/20 High (UV-activated) Room temp. (25 °C) Light-activated, Ag nanoparticle hybrid 78
Carbon/GQDs: MoS2/NGQD NO2 82.4% (@ 50 ppm) 50 ppb 10/30 High vs. other gases 30 °C Hydrothermal synthesis, 2D/0D heterostructure 76
Carbon/GQDs: CQD/Alq3 NO2 24% (@300 °C) 100 ppb 2/8 High vs. other gases 300 °C Spin-coated, polymer composite 88


6.5. Comparison of PQD-based gas sensors with conventional sensor technologies

In order to contextualize the current progress PQD-based gas sensors, it is essential to compare their performance with conventional sensor systems, including metal oxide semiconductors (MOS), carbon-based nanomaterials, and polymeric sensors. Traditional MOS sensors such as ZnO, SnO2, and TiO2 have been widely employed due to their robustness and low cost; however, they typically require elevated operating temperatures (200–400 °C), which lead to high power consumption, slow recovery, and long-term drift.14,15 Moreover, their selectivity is generally poor because gas adsorption processes are governed by non-specific surface reactions. These limitations hinder their use in portable or flexible environmental monitoring devices.

Carbon-based materials such as graphene, carbon nanotubes (CNTs), and reduced graphene oxide (rGO) exhibit high electrical conductivity and mechanical flexibility, making them attractive for miniaturized sensors.16,17 Nevertheless, their sensing performance is often limited by weak gas adsorption and signal instability under humidity variations, requiring additional surface functionalization or doping strategies. Polymeric sensors, including polyaniline (PANI) and polypyrrole (PPy), are lightweight and flexible but typically suffer from irreversible swelling and aging effects under prolonged operation, which restrict their reproducibility and long-term reliability.18–20

In contrast, PQD-based sensors exhibit distinctive advantages derived from their quantum-confined structure and ionic lattice. PQDs such as CsPbBr3 and CH3NH3PbI3 demonstrate high photoluminescence quantum yield (PLQY, up to 90%), narrow emission linewidths (12–40 nm), and high carrier mobility (up to 4500 cm2 V−1 s−1), enabling both fluorescence- and chemiresistive-type sensing with remarkable sensitivity and selectivity.24,30,33 Unlike MOS or CNT sensors, PQDs can operate efficiently at room temperature, which significantly reduces power consumption and allows their integration into flexible and wearable electronics. Their tunable bandgap and surface chemistry facilitate targeted interactions with oxidizing (NO2, O2) or reducing (NH3, H2S) gases, offering detection limits in the sub-ppm or even ppb range—often surpassing conventional systems.25,31,35

Furthermore, the solution-processable nature of PQDs enables low-cost fabrication using spin-coating, drop-casting, or inkjet printing, which contrasts with the high-temperature sintering required for MOS materials. Recent studies have demonstrated stable PQD-based sensors embedded in silica aerogels or zeolites with prolonged operational lifetimes exceeding 100 days, addressing the durability challenges that typically affect polymeric or carbon-based devices.25,27,63 Overall, PQD-based gas sensors represent a new generation of nanostructured sensing platforms combining ambient-condition operation, superior selectivity, high signal reproducibility, and facile processability, establishing them as strong candidates for next-generation environmental monitoring and smart sensing technologies.18,63

7. Challenges and advancements in perovskite QD-based gas sensing

PQDs offer exceptional PLQY, tunable bandgaps, and high surface-to-volume ratios for sensing toxic inorganic gases, VOCs, oxygen, and humidity. Despite their promise, challenges like environmental instability, limited selectivity, lead toxicity, and fabrication complexity hinder practical deployment. This section analyzes these limitations, proposes advanced strategies to address them, and outlines future directions to enhance PQD-based gas sensors, avoiding mechanistic details (Section 3) and performance metrics (Section 4).

7.1. Primary limitations of PQD-based gas sensors

7.1.1. Environmental instability. The ionic ABX3 perovskite lattice makes PQDs vulnerable to moisture, heat, and light. Moisture coordinates with surface ions, causing structural dissociation and significant PL loss in humid conditions (e.g., 80% RH).32,89 Thermal instability, due to low formation energies (∼0.5 eV for CsPbBr3), leads to decomposition above 100 °C, limiting use in high-temperature industrial settings.32 Photo-induced degradation under UV or visible light creates halide vacancies, reducing PLQY, particularly in fluorescence-based sensors.90–92 Reactive analytes like H2S can irreversibly disrupt the lattice, forming non-luminescent compounds.54
7.1.2. Limited selectivity in complex gas mixtures. PQDs often exhibit non-specific interactions at surface defects, reducing selectivity in multi-gas environments. For instance, CsPbBr3 PQDs show strong fluorescence response to NH3 but are sensitive to water vapor, affecting accuracy in humid conditions.26 Chemiresistive hybrids like CsPbBr3–ZnO require surface engineering to distinguish NO2 from CO or NH3.45,93 In biomedical applications, such as acetone detection for diabetes, interference from ethanol or humidity complicates results,46 limiting reliability in complex settings.
7.1.3. Toxicity of lead-based PQDs. Lead-based PQDs, such as CsPbBr3, raise environmental and health concerns due to potential leakage during degradation, particularly in large-scale or biomedical applications.32,94 Lead-free alternatives like Cs2AgBiBr6 have lower PLQY (∼50% vs. 90% for CsPbBr3) and reduced stability, compromising sensing performance.49 Balancing toxicity reduction with optoelectronic efficiency remains a challenge for sustainable deployment.
7.1.4. Fabrication complexity and scalability. Integrating PQDs into stable sensors often involves complex post-synthesis treatments, such as encapsulation in zeolites or ligand exchange, increasing costs.25,95 Hybrid systems with metal oxides require multi-step fabrication, hindering scalability compared to simpler systems like carbon-based QDs.45,86 These complexities challenge commercial viability, particularly for large-scale environmental monitoring.

7.2. Advanced strategies for overcoming limitations

7.2.1. Novel encapsulation for enhanced stability. Advanced encapsulation enhances PQD durability. Fe-doped zeolite X protects CsPbBr3 PQDs, enabling stable NH3 sensing in humid conditions by shielding against moisture while allowing gas diffusion.25,97 Silica aerogels stabilize CH3NH3PbBr3 PQDs for SO2 detection, leveraging hydrophobicity to prevent degradation.63 Polymer-based encapsulation, such as CsPbBr3 in EVA, supports methanol sensing and offers flexibility for wearable devices.27 Future strategies could explore metal–organic frameworks (MOFs) with tunable porosity to optimize stability and gas access.
7.2.2. Surface engineering and machine learning for selectivity. Surface engineering with tailored ligands, like zinc acetylacetonate, reduces water vapor interference for ethanol detection, with machine learning analyzing PL patterns to enhance specificity.46 Gas-specific receptors, such as thiol-based ligands for H2S, can minimize cross-sensitivity. Hybrid systems with 2D materials like MoS2, inspired by MoS2/NGQD hybrids, suggest potential for high selectivity through increased surface area.76 Machine learning can distinguish gas signatures in complex mixtures, critical for breath analysis or industrial monitoring.96
7.2.3. Lead-free PQD development. Lead-free PQDs, such as Cs2AgBiBr6, require optimization to match lead-based performance. Ion doping with Mn2+ or Bi3+ enhances lattice stability, as seen in acetone-sensing Cs2AgBiBr6.49 Microwave-assisted synthesis improves crystallinity, boosting PLQY in lead-free systems.32 Exploring double-perovskite or hybrid organic-inorganic structures could yield eco-friendly, high-performance sensors.
7.2.4. Streamlined fabrication for scalability. Room-temperature co-precipitation produces uniform PQDs with high PLQY, reducing complex post-synthesis steps.32 Automated ligand exchange in flow-based systems streamlines passivation. Flexible substrates, like polymers used in CH3NH3PbBr3 PQDs for NH3 sensing, support scalable production of portable sensors.55 Standardized protocols can ensure cost-effective industrial adoption.

7.3. Future perspectives for PQD-based gas sensing technologies

7.3.1. Integration with emerging technologies. Integrating PQDs with IoT enables real-time environmental monitoring across large areas, leveraging low-power operation. Flexible substrates, as in CH3NH3PbBr3 PQDs, support wearable sensors for industrial or biomedical use.55,98 Machine learning, applied to ethanol detection, could enable rapid gas identification in complex environments.46
7.3.2. Multifunctional and self-powered sensors. Multifunctional sensors detecting multiple gases, like CsPbBr3–Pt(II) for O2 and NO, show promise for multi-analyte detection.56 Ratiometric designs with reference emitters, such as FAPbI3 for O2, should expand to other gases.78 Self-powered sensors, integrating PQDs with photovoltaic materials, could enable energy-independent devices for remote monitoring.
7.3.3. Advanced heterostructures. Novel heterostructures with 2D materials like graphene or MXenes could enhance conductivity and stability, building on MoS2/NGQD systems.76 These materials offer high surface areas, potentially surpassing metal oxide hybrids like CsPbBr3–In2O3.30
7.3.4. Regulatory compliance and sustainability. Prioritizing lead-free PQDs and safe disposal protocols addresses toxicity concerns. Life-cycle assessments can guide eco-friendly designs, ensuring compliance with environmental regulations. Table 6 outlines challenges and strategies for PQD-based gas sensors.
Table 6 Challenges and strategies for PQD-based gas sensors
Challenge Description Strategy Future direction Ref.
Environmental instability Susceptibility to moisture, heat, light Zeolite/silica encapsulation, polymer coatings MOFs with tunable porosity 25 and 63
Limited selectivity Non-specific interactions in multi-gas settings Tailored ligands, machine learning, 2D material hybrids Gas-specific receptors, neuromorphic computing 46 and 76
Lead toxicity Environmental/health risks from Pb-based PQDs Lead-free PQDs, ion doping Double-perovskites, hybrid compositions 49
Fabrication complexity Complex post-synthesis treatments Room-temperature synthesis, automated ligand exchange Scalable substrates, standardized protocols 32 and 55
Slow response dynamics Longer response/recovery times for some analytes (e.g., H2S) Optimized surface passivation, nanostructured matrices High-surface-area heterostructures 24
Integration complexity Challenges in combining PQDs with IoT or wearable platforms Flexible substrates, low-power designs IoT-enabled sensor arrays, neuromorphic systems 55 and 96
Energy consumption High power needs for continuous operation Self-powered systems with photovoltaic integration Piezoelectric or solar-driven sensors 97
Recycling challenges Environmental impact of PQD disposal Eco-friendly designs, recycling protocols Life-cycle assessments, sustainable materials 32


8. Conclusion

PQDs have emerged as exceptional next-generation materials for gas sensing applications, owing to their unique optoelectronic tunability, structural flexibility, and high surface reactivity. This review has comprehensively analyzed the synthesis strategies, fundamental mechanisms, characterization techniques, and performance trends that define PQD-based gas sensors. Through this integrated approach, it provides a unified understanding of how synthesis methodology and surface chemistry determine the sensing behavior and operational stability of PQDs. The novelty of this review lies in its holistic correlation between synthesis, structure, and sensing functionality, presenting a comparative evaluation of PQD-based sensors with conventional counterparts such as metal-oxide, carbon-based, and polymeric systems. This comparative insight highlights the superior sensitivity, selectivity, and room-temperature performance of PQD-based platforms, while also emphasizing the remaining scientific challenges that limit their commercialization. Moreover, by combining advances in photoluminescence-based and chemiresistive sensing, this review bridges the gap between material chemistry and device engineering—an approach not systematically addressed in earlier literature.

In terms of scope and contribution, this review not only consolidates the current state of PQD-based gas-sensing research but also defines a forward-looking roadmap for the field. It identifies unresolved challenges such as intrinsic instability in humid and oxidative environments, lead toxicity, and scalability limitations, while summarizing emerging strategies for mitigation. These include ligand engineering, heterostructure formation, ion doping, and encapsulation within robust inorganic or polymer matrices. Furthermore, the review extends its perspective to new research directions, including the design of eco-friendly lead-free PQDs, machine learning-assisted signal processing, and integration of PQDs with wearable and flexible sensing systems. Overall, PQDs offer a transformative platform for gas sensing, capable of outperforming many existing nanomaterial-based technologies in terms of detection sensitivity, operating temperature, and device adaptability. The comprehensive framework presented here—linking synthesis control, optoelectronic behavior, and sensing performance—provides valuable insights for researchers seeking to optimize PQD-based devices for environmental, industrial, and biomedical applications. Continued interdisciplinary research across chemistry, materials science, and engineering will be essential to translate the remarkable laboratory achievements of PQD-based sensors into robust, real-world technologies.

Author contributions

Suleiman Ibrahim Mohammad and Ahmad Mohebi conceptualized the review, supervised the project, and prepared the original draft; A. K. Kareem, Fadhil Faez Sead, and D. S. Jayalakshmi conducted literature investigation and data curation; Zyad Shaaban and Sanjeev Kumar performed formal analysis and validation; M. Sudhakara Reddy and Satish Choudhury provided resources and visualization; Asokan Vasudevan managed project administration; all authors contributed to writing, review, and editing of the manuscript and have approved the final version.

Conflicts of interest

The authors declare no conflict of interest.

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. 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.
  2. M. Que, Y. Xu, Q. Wu, J. Chen, L. Gao and S. F. Liu, Application of advanced quantum dots in perovskite solar cells: synthesis, characterization, mechanism, and performance enhancement, Mater. Horiz., 2025, 12(8), 2467–2502 RSC.
  3. H. Noorizadeh, A review on the role of quantum dots in targeted drug delivery: advances, functionalization, and applications in nanomedicine, Chem. Pharm. Lett., 2025, 1, 1–53 Search PubMed.
  4. Q. Pan, Q. Zhao, P. Wei and G. Li, Surface ligands for perovskite quantum dots, ChemSusChem, 2025, 18(4), e202401875 CrossRef CAS PubMed.
  5. Y. Duan, K. Gu, S. Li, J. Du and J. Zhang, Recent Progress in CsPbX3 (X= Cl, Br, and I) Perovskite Quantum Dot Glasses: Synthesis, Matrix Optimization, and Photonic Application Potentials, Laser Photonics Rev., 2025, 19(9), 2402245 CrossRef CAS.
  6. S. M. Qaid, H. M. Ghaithan, B. A. Al-Asbahi, A. Alqasem and A. S. Aldwayyan, Fabrication of thin films from powdered cesium lead bromide (CsPbBr3) perovskite quantum dots for coherent green light emission, ACS Omega, 2020, 5(46), 30111–30122 CrossRef CAS PubMed.
  7. A. Suhail, S. Beniwal, R. Kumar, A. Kumar and M. Bag, Hybrid halide perovskite quantum dots for optoelectronics applications: recent progress and perspective, J. Phys.: Condens. Matter, 2025, 163002 CrossRef CAS PubMed.
  8. X. Zhang, H. Huang, C. Zhao and J. Yuan, Surface chemistry-engineered perovskite quantum dot photovoltaics, Chem. Soc. Rev., 2025, 54(6), 3017–3060 RSC.
  9. M. Abushuhel, A. Kumar, A. F. Al-Hussainy, S. Mohammed, R. Panigrahi and H. Noorizadeh, Bromide Perovskite Quantum Dot Fluorescent Sensors for Food Safety: Advances in Pesticide and Mycotoxin Detection, J. Agric. Food Res., 2025, 1, 102477 Search PubMed.
  10. F. Ullah, Z. Fredj and M. Sawan, Perovskite Quantum Dot-Based Memory Technologies: Insights from Emerging Trends, Nanomaterials, 2025, 15(11), 873 CrossRef CAS PubMed.
  11. B. Zong, S. Wu, Y. Yang, Q. Li, T. Tao and S. Mao, Smart gas sensors: recent developments and future prospective, Nano-Micro Lett., 2025, 17(1), 54 CrossRef CAS PubMed.
  12. M. A. Chowdhury and M. A. Oehlschlaeger, Artificial Intelligence in Gas Sensing: A Review, ACS Sens., 2025, 10(3), 1538–1563 CrossRef CAS PubMed.
  13. F. F. Sead, Y. Jadeja, A. Kumar, M. M. Rekha, M. Kundlas, S. Saini, K. K. Joshi and H. Noorizadeh, Carbon quantum dots for sustainable energy: enhancing electrocatalytic reactions through structural innovation, Nanoscale Adv., 2025, 7(13), 3961–3998 RSC.
  14. H. Chen, H. Chen, J. Chen and M. Song, Gas sensors based on semiconductor metal oxides fabricated by electrospinning: a review, Sensors, 2024, 24(10), 2962 CrossRef CAS PubMed.
  15. D. H. Gao, Q. C. Yu, M. A. Kebeded, Y. Y. Zhuang, S. Huang, M. Z. Jiao and X. J. He, Advances in modification of metal and noble metal nanomaterials for metal oxide gas sensors: a review, Rare Met., 2025, 11, 1–54 Search PubMed.
  16. J. Sengupta and C. M. Hussain, Carbon Nanotube-Based Field-Effect Transistor Biosensors for Biomedical Applications: Decadal Developments and Advancements (2016–2025), Biosensors, 2025, 15(5), 296 CrossRef CAS PubMed.
  17. M. Harun-Or-Rashid, S. Mirzaei and N. Nasiri, Nanomaterial Innovations and Machine Learning in Gas Sensing Technologies for Real-Time Health Diagnostics, ACS Sens., 2025, 10(3), 1620–1640 CrossRef CAS PubMed.
  18. V. Galstyan, “Quantum dots: perspectives in next-generation chemical gas sensors”–a review, Anal. Chim. Acta, 2021, 1152, 238192 CrossRef CAS PubMed.
  19. A. Mirzaei, Z. Kordrostami, M. Shahbaz, J. Y. Kim, H. W. Kim and S. S. Kim, Resistive-based gas sensors using quantum dots: a review, Sensors, 2022, 22(12), 4369 CrossRef CAS PubMed.
  20. 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, 107164 CrossRef CAS.
  21. H. Liu, M. Li, O. Voznyy, L. Hu, Q. Fu, D. Zhou, Z. Xia, E. H. Sargent and J. Tang, Physically flexible, rapid-response gas sensor based on colloidal quantum dot solids, Adv. Mater., 2014, 26(17), 2718–2724 CrossRef CAS PubMed.
  22. Y. Tang, Y. Zhao and H. Liu, Room-temperature semiconductor gas sensors: challenges and opportunities, ACS Sens., 2022, 7(12), 3582–3597 CrossRef CAS PubMed.
  23. C. Cheng, Q. Liang, M. Yan, Z. Liu, Q. He, T. Wu, S. Luo, Y. Pan, C. Zhao and Y. Liu, Advances in preparation, mechanism and applications of graphene quantum dots/semiconductor composite photocatalysts: a review, J. Hazard. Mater., 2022, 424, 127721 CrossRef CAS PubMed.
  24. H. Shan, W. Xuan, Z. Li, D. Hu, X. Gu and S. Huang, Room-temperature hydrogen sulfide sensor based on tributyltin oxide functionalized perovskite CsPbBr3 quantum dots, ACS Appl. Nano Mater., 2022, 5(5), 6801–6809 CrossRef CAS.
  25. W. Wu, C. Zhao, M. Hu, A. Pan, W. Xiong and Y. Chen, CsPbBr 3 perovskite quantum dots grown within Fe-doped zeolite X with improved stability for sensitive NH3 detection, Nanoscale, 2023, 15(12), 5705–5711 RSC.
  26. X. Zhang, J. Lv, J. Liu, S. Xu, J. Sun, L. Wang, L. Xu, S. Mintova, H. Song and B. Dong, Stable EMT type zeolite/CsPbBr3 perovskite quantum dot nanocomposites for highly sensitive humidity sensors, J. Colloid Interface Sci., 2022, 616, 921–928 CrossRef CAS PubMed.
  27. Q. C. Yu, X. Q. Gu, F. Tong, Z. Chen and S. Huang, Enhanced moisture resistance performance of CsPbBr3 quantum dots through synergetic encapsulation with In3+ ions and polymer, Mater. Sci. Semicond. Process., 2025, 185, 108949 CrossRef CAS.
  28. D. Kumar, R. Mesin and C. S. Chu, Optical fluorescent sensor based on perovskite QDs for nitric oxide gas detection, Appl. Opt., 2023, 62(12), 3176–3181 CrossRef CAS PubMed.
  29. R. Mesin, C. S. Chu and Z. L. Tseng, Ratiometric optical oxygen sensor based on perovskite quantum dots and Rh110 embedded in an ethyl cellulose matrix, Opt. Mater. Express, 2023, 13(4), 945–955 CrossRef CAS.
  30. X. Shi, R. Tian, Q. Wang and P. Song, Perovskite CsPbBr3 quantum dots enhanced In2O3 nanospheres for triethylamine detection at low temperature, Ceram. Int., 2024, 50(24), 53941–53950 CrossRef CAS.
  31. M. Liu, P. Song, Q. Wang and M. Yan, CsPbBr3 Quantum Dot Modified In2O3 Nanofibers for Effective Detection of ppb-Level HCHO at Room Temperature under UV Illumination, ACS Sens., 2024, 9(11), 6040–6050 CrossRef CAS PubMed.
  32. 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.
  33. P. Pandey, A. Sengupta, S. Parmar, U. Bansode, S. Gosavi, A. Swarnkar, S. Muduli, A. D. Mohite and S. Ogale, CsPbBr3–Ti3C2Tx MXene QD/QD heterojunction: photoluminescence quenching, charge transfer, and Cd ion sensing application, ACS Appl. Nano Mater., 2020, 3(4), 3305–3314 CrossRef CAS.
  34. Q. Masaadeh, Luminescent downshifting silicon and perovskite quantum dots for efficiency enhancement of crystalline silicon solar cells, PhD Diss., University of East Anglia, 2024 Search PubMed.
  35. Y. Huang, J. Zhang, X. Zhang, J. Jian, J. Zou, Q. Jin and X. Zhang, The ammonia detection of cesium lead halide perovskite quantum dots in different halogen ratios at room temperature, Opt. Mater., 2022, 134, 113155 CrossRef CAS.
  36. R. Han, Q. Zhao, A. Hazarika, J. Li, H. Cai, J. Ni and J. Zhang, Ionic liquids modulating CsPbI3 colloidal quantum dots enable improved mobility for high-performance solar cells, ACS Appl. Mater. Interfaces, 2022, 14(3), 4061–4070 CrossRef CAS PubMed.
  37. Y. Zhou, Z. Yang, Q. Huang, Y. Ye, B. Ye, Z. Shen, W. Wu, Z. Zeng, Z. Hong, Z. Meng and H. Hong, Achieving high-efficiency CsPbBr3 perovskite light emitting diode via GABr-didecyldimethyl ammonium bromide hybrid ligand passivation strategies, Org. Electron., 2023, 118, 106797 CrossRef CAS.
  38. 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 Cu2+ detection, J. Mater. Chem. C, 2018, 6(17), 4793–4799 RSC.
  39. B. Park, S. M. Kang, G. W. Lee, C. H. Kwak, M. Rethinasabapathy and Y. S. Huh, Fabrication of CsPbBr3 perovskite quantum dots/cellulose-based colorimetric sensor: Dual-responsive on-site detection of chloride and iodide ions, Ind. Eng. Chem. Res., 2019, 59(2), 793–801 CrossRef.
  40. 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.
  41. Y. Gao and B. Chen, Lead-free Cs3Bi2Br9 perovskite quantum dots for detection of heavy metal Cu2+ ions in seawater, J. Environ. Eng. Sci., 2023, 11(5), 1001 Search PubMed.
  42. 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.
  43. 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.
  44. 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.
  45. L. Yueyue, S. Siqi, W. Yilin, L. Fengmin, W. Hongtao, B. Jihao, L. Min and L. Geyu, CsPbBr3 quantum dots enhanced ZnO sensing to NO2 at room temperature, Sens. Actuators, B, 2022, 368, 132189 CrossRef.
  46. L. Zhu, W. Xu, W. Xuan, H. Zhang, Z. Yang, Y. Zhao, S. Huang and X. Gu, Perovskite CsPbBr3 quantum dots capped with zinc acetylacetonate: gas sensing of ethanol in humidity with aid of machine-learning, Mater. Sci. Semicond. Process., 2023, 167, 107790 CrossRef CAS.
  47. 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.
  48. 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.
  49. C. Zhou, H. Zhang, S. Liu, H. Cao, X. Jia, J. Jia and K. Ma, Cs2AgBiBr6 quantum dots supported on Co3O4 nanocages for acetone detection at room temperature, ACS Appl. Nano Mater., 2023, 6(24), 23313–23323 CrossRef CAS.
  50. 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.
  51. S. Das, P. Somu, A. K. Yadav, P. K. Hopke and S. Paul, Recent advances in II–VI group semiconductor-and carbon-based quantum dots for fluorescence-based sensing of metal ions in water, Environ. Sci.: Nano, 2024, 11(3), 739–765 RSC.
  52. D. Li, P. Zhuang and C. Sun, Unlocking the potential of perovskite-based nanomaterials for revolutionary smartphone-based sensor applications, J. Mater. Chem. C, 2024, 12(13), 4544–4561 RSC.
  53. A. S. Manna, S. Ghosh, T. Ghosh, N. Karchaudhuri, S. Das, A. Roy and D. K. Maiti, Smart luminescent materials for emerging sensors: Fundamentals and advances, Chem.–Asian J., 2025, 20(6), e202401328 CrossRef CAS PubMed.
  54. C. Chen, Q. Cai, F. Luo, N. Dong, L. Guo, B. Qiu and Z. Lin, Sensitive fluorescent sensor for hydrogen sulfide in rat brain microdialysis via CsPbBr3 quantum dots, Anal. Chem., 2019, 91(24), 15915–15921 CrossRef CAS PubMed.
  55. A. K. Singh, S. Singh, V. N. Singh, G. Gupta and B. K. Gupta, Probing reversible photoluminescence alteration in CH3NH3PbBr3 colloidal quantum dots for luminescence-based gas sensing application, J. Colloid Interface Sci., 2019, 554, 668–673 CrossRef CAS PubMed.
  56. D. T. Putro and C. S. Chu, Optical dual sensor single fiber Pt(II) complex coated perovskite quantum dots green for oxygen and nitric oxide gas detection, In2022 IET International Conference on Engineering Technologies and Applications (IET-ICETA), IEEE, 2022, pp. 1–2 Search PubMed.
  57. A. Di Vera, Alternative halide perovskite structures for light-based applications, 2025 Search PubMed.
  58. A. S. Manna, S. Ghosh, T. Ghosh, N. Karchaudhuri, S. Das, A. Roy and D. K. Maiti, Smart luminescent materials for emerging sensors: Fundamentals and advances, Chem.–Asian J., 2025, 20(6), e202401328 CrossRef CAS PubMed.
  59. J. Wawrzyniak, Advancements in improving selectivity of metal oxide semiconductor gas sensors opening new perspectives for their application in food industry, Sensors, 2023, 23(23), 9548 CrossRef CAS PubMed.
  60. A. Mirzaei, S. S. Kim and H. W. Kim, Resistance-based H2S gas sensors using metal oxide nanostructures: a review of recent advances, J. Hazard. Mater., 2018, 357, 314–331 CrossRef CAS PubMed.
  61. S. Chandrasekaran and A. Sukhdev, Nanostructured Materials for Sensors Applications, in Emerging Nanomaterials for Catalysis and Sensor Applications, CRC Press, 2023, vol. 28, pp. 159–179 Search PubMed.
  62. J. Mondal, R. Lamba, Y. Yukta, R. Yadav, R. Kumar, B. Pani and B. Singh, Advancements in semiconductor quantum dots: expanding frontiers in optoelectronics, analytical sensing, biomedicine, and catalysis, J. Mater. Chem. C, 2024, 12(28), 10330–10389 RSC.
  63. X. You, J. Wu and Y. Chi, Superhydrophobic silica aerogels encapsulated fluorescent perovskite quantum dots for reversible sensing of SO2 in a 3D-printed gas cell, Anal. Chem., 2019, 91(8), 5058–5066 CrossRef CAS PubMed.
  64. Z. Yang, X. Yu, B. Huang, J. Li, X. Yan, K. Feng, R. Cai, Y. Yuan, T. Yue and Q. Sheng, Phospholipid membrane-encapsulated perovskite quantum dots electrospun glass cellulose membrane-based fluorescence and visual two-channel detection of breath ammonia for patients infected with Helicobacter pylori, Microchem. J., 2025, 212, 113263 CrossRef CAS.
  65. J. Shamsi, A. S. Urban, M. Imran, L. De Trizio and L. Manna, Metal halide perovskite nanocrystals: synthesis, post-synthesis modifications, and their optical properties, Chem. Rev., 2019, 119(5), 3296–3348 CrossRef CAS PubMed.
  66. X. Zheng, Defect Passivation and Surface Modification for Efficient and Stable Organic–Inorganic Hybrid Perovskite Solar Cells and Light-Emitting Diodes, 2020 Search PubMed.
  67. P. Matricon, R. R. Suresh, Z. G. Gao, N. Panel, K. A. Jacobson and J. Carlsson, Ligand design by targeting a binding site water, Chemical science, 2021, 12(3), 960–968 RSC.
  68. P. M. Bulemo, D. H. Kim, H. Shin, H. J. Cho, W. T. Koo, S. J. Choi, C. Park, J. Ahn, A. T. Guntner, R. M. Penner and I. D. Kim, Selectivity in Chemiresistive Gas Sensors: Strategies and Challenges, Chem. Rev., 2025, 125(8), 4111–4183 CrossRef CAS PubMed.
  69. L. Tan, M. Guo, J. Tan, Y. Geng, S. Huang, Y. Tang, C. Su, C. Lin and Y. Liang, Development of high-luminescence perovskite quantum dots coated with molecularly imprinted polymers for pesticide detection by slowly hydrolysing the organosilicon monomers in situ, Sens. Actuators, B, 2019, 291, 226–234 CrossRef CAS.
  70. X. G. Wu, H. Ji, X. Yan and H. Zhong, Industry outlook of perovskite quantum dots for display applications, Nat. Nanotechnol., 2022, 17(8), 813–816 CrossRef CAS PubMed.
  71. 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.
  72. 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.
  73. M. Salot, K. Santhy, V. R. Naganaboina, S. G. Singh, A. K. Pramanick, D. Mandal, G. Avasthi and S. K. Chaudhury, On improved inorganic gas-sensing characteristics of microwave-treated tungsten oxide quantum dots at room temperature, Int. J. Appl. Ceram. Technol., 2025, e15033 CrossRef CAS.
  74. Y. Qiao, J. Zhang, J. Liu, Y. Liu, X. Zhang, Z. Yang, X. Yin, J. Gao, C. Wang and H. Lu, MOF-derived porous In2O3 nanospheres sensitized by CdS quantum dots: a ppb-level NO2 sensor with ultra-high response at room-temperature, J. Alloys Compd., 2024, 997, 174889 CrossRef CAS.
  75. J. Kwon, Y. Ha, S. Choi, D. G. Jung, H. K. An, S. H. Kong and D. Jung, Solution-processed NO2 gas sensor based on poly (3-hexylthiophene)-doped PbS quantum dots operable at room temperature, Sci. Rep., 2024, 14(1), 20600 CrossRef CAS PubMed.
  76. B. Malathi, R. A. Parveen, A. Nakamura, J. Archana, M. Navaneethan and S. Harish, Ultrathin layered MoS2/N-doped graphene quantum dots (NGQDs) heterostructures for highly sensitive room temperature NO2 gas sensor, Sens. Actuators, B, 2024, 403, 135083 CrossRef CAS.
  77. K. Wu, W. Zhang, Z. Zheng, M. Debliquy and C. Zhang, Room-temperature gas sensors based on titanium dioxide quantum dots for highly sensitive and selective H2S detection, Appl. Surf. Sci., 2022, 585, 152744 CrossRef CAS.
  78. C. Thota, J. K. Modigunta, M. Reddeppa, Y. H. Park, H. Kim, H. Kang, S. Kokkiligadda, S. Lee, G. Murali, S. Y. Park and I. In, Light stimulated room-temperature H2S gas sensing ability of Cl-doped carbon quantum dots supported Ag nanoparticles, Carbon, 2022, 196, 337–346 CrossRef CAS.
  79. O. H. Abd-Elkader, M. A. Sakr, M. A. Saad, H. Abdelsalam and Q. Zhang, Electronic and gas sensing properties of ultrathin TiO2 quantum dots: A first-principles study, Results Phys., 2023, 52, 106804 CrossRef.
  80. S. Ebrahimi, D. Souri and Y. Abdi, ZnSe: Cu@ZnS core/shell quantum dots synthesis by photochemical method for ethanol gas sensing, J. Environ. Chem. Eng., 2024, 12(5), 113501 CrossRef CAS.
  81. A. K. Priya, S. K. Rao, S. Divya, N. Vijay, P. Jerome, T. H. Oh, B. Renganathan and D. Sastikumar, Engineering luminescent quantum dots through laser ablation of samarium oxide (Sm2O3): a novel approach for enhanced fiber optic gas sensing, Sens. Actuators, B, 2024, 417, 136128 CrossRef.
  82. W. Li, Q. Yuan, Z. Xia, X. Ma, L. He, L. Jin, X. Chu and K. Zhang, ZnO quantum dots sensitized ZnSnO3 for highly formaldehyde sensing at a low temperature, Sens. Actuators, B, 2024, 400, 134912 CrossRef CAS.
  83. T. Balakrishnan, W. L. Ang and E. Mahmoudi, Highly sensitive fluorescent nitrobenzene gas sensing by nitrogen-doped graphene quantum dots embedded in ZIF-8 nanocomposite, Mater. Sci. Eng., B, 2024, 304, 117377 CrossRef CAS.
  84. J. Singh, R. Thareja, P. Malik and R. Kakkar, Size-dependent structural and electronic properties of stoichiometric II–VI quantum dots and gas sensing ability of CdSe quantum dots: a DFT study, J. Nanopart. Res., 2022, 24(2), 33 CrossRef CAS.
  85. J. Liu, Y. Wang, Y. Sun, K. Zhang, Y. Ding, C. Fu and J. Wang, Preparation and optimization of mesoporous SnO2 quantum dot thin film gas sensors for H2S detection using XGBoost parameter importance analysis, Chemosensors, 2023, 11(10), 525 CrossRef CAS.
  86. T. Balakrishnan, S. Sagadevan, M. V. Le, T. Soga and W. C. Oh, Recent progress on functionalized graphene quantum dots and their nanocomposites for enhanced gas sensing applications, Nanomaterials, 2023, 14(1), 11 CrossRef PubMed.
  87. W. Gao, X. Wang, Y. He, H. Yu, Y. Zheng, R. Yin and X. Jiang, Sub-ppm NO2 gas sensing in CdTe quantum dots functionalized hollow-core anti-resonant fiber, Sens. Actuators, B, 2024, 405, 135350 CrossRef CAS.
  88. N. A. Abd and O. A. Ibrahim, Fabrication of Carbon Quantum Dots/Alq3 Layer for NO2 Gas Sensor, Iraqi J. Phys., 2024, 22(2), 1 CrossRef.
  89. E. Hassanabadi, M. Latifi, A. F. Gualdrón-Reyes, S. Masi, S. J. Yoon, M. Poyatos, B. Julián-López and I. Mora-Seró, Ligand & band gap engineering: tailoring the protocol synthesis for achieving high-quality CsPbI 3 quantum dots, Nanoscale, 2020, 12(26), 14194–14203 RSC.
  90. Z. Zhang, S. Ghimire, T. Okamoto, B. M. Sachith, J. Sobhanan, C. Subrahmanyam and V. Biju, Mechano-optical modulation of excitons and carrier recombination in self-assembled halide perovskite quantum dots, ACS Nano, 2022, 16(1), 160–168 CrossRef CAS PubMed.
  91. Z. Shen, S. Zhao, D. Song, Z. Xu, B. Qiao, P. Song, Q. Bai, J. Cao, G. Zhang and W. Swelm, Improving the quality and luminescence performance of all-inorganic perovskite nanomaterials for light-emitting devices by surface engineering, Small, 2020, 16(26), 1907089 CrossRef CAS PubMed.
  92. D. Jia, J. Chen, R. Zhuang, Y. Hua and X. Zhang, Inhibiting lattice distortion of CsPbI3 perovskite quantum dots for solar cells with efficiency over 16.6, Energy Environ. Sci., 2022, 15(10), 4201–4212 RSC.
  93. Y. Wang, G. Ding, J. Y. Mao, Y. Zhou and S. T. Han, Recent advances in synthesis and application of perovskite quantum dot based composites for photonics, electronics and sensors, Sci. Technol. Adv. Mater., 2020, 21(1), 278–302 CrossRef CAS PubMed.
  94. C. G. Sanjayan, M. S. Jyothi and R. G. Balakrishna, Stabilization of CsPbBr3 quantum dots for photocatalysis, imaging and optical sensing in water and biological medium: a review, J. Mater. Chem. C, 2022, 10(18), 6935–6956 RSC.
  95. 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.
  96. M. S. Çadırcı and M. Çadırcı, Machine Learning Models for Accurately Predicting Properties of CsPbCl3 Perovskite Quantum Dots, arXiv, 2024, preprint arXiv:2406.15515,  DOI:10.48550/arXiv.2406.15515.
  97. K. D. Jayan and K. Babu, Luminescent perovskite quantum dots: progress in fabrication, modelling and machine learning approaches for advanced photonic and quantum computing applications, J. Lumin., 2024, 23, 120906 Search PubMed.
  98. S. Zhang, W. Zhu, X. Zhang, L. Mei, J. Liu and F. Wang, Machine learning-driven fluorescent sensor array using aqueous CsPbBr3 perovskite quantum dots for rapid detection and sterilization of foodborne pathogens, J. Hazard. Mater., 2025, 483, 136655 CrossRef CAS PubMed.

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