Abstract
CsPbBr3 perovskite quantum dots (QDs) are renowned for their exceptional optical properties, including high quantum efficiency, strong exciton binding energy, and tunable emission wavelengths. However, their practical application is hindered by their inherent susceptibility to environmental degradation. In this study, we introduce a CsPbBr3@UiO-66 composite material, where CsPbBr3 QDs self-assemble within the microporous framework of UiO-66, a robust metal-organic framework (MOF). This encapsulation strategy significantly enhances the environmental stability of CsPbBr3 QDs, maintaining luminescence for over 30 months under ambient conditions and several hours underwater. Temperature-dependent and time resolved photoluminescence (TRPL) measurements further revealed the exciton–phonon interaction within the CsPbBr3@UiO-66 material. We distributed CsPbBr3@UiO-66 into a hybrid microcavity (MC) and observed strong exciton–polariton coupling, showcasing the remarkable light–matter interaction capabilities of the composite. These findings highlight the potential of CsPbBr3@UiO-66 as a robust platform for advanced polaritonic applications, paving the way for next-generation optoelectronic devices and quantum technologies.
Keywords: microcavity, polariton, inorganic perovskite, QD, metal-organic framework
1. Introduction
Quantum dots (QDs) are pivotal in modern optoelectronics due to their quantum confinement effects, offering tunable emission, high photoluminescence quantum yields, and excellent color purity. These properties have enabled their applications in displays, lasers, sensors, and quantum technologies [1], [2], [3], [4], [5], [6], [7], [8], [9], [10]. Among QDs, lead halide perovskites stand out for their high power conversion efficiency (PCE), superior charge transport, and cost-effective fabrication [11], [12], [13], [14], [15]. Moreover, the versatility of perovskites enables their integration into various optoelectronic components, such as LEDs, lasers, and photodetectors [16], [17], [18], [19], [20], [21], [22], [23], [24], [25], [26], [27], [28]. However, instability against moisture, heat, and light limits their performance and use [29], [30], [31], [32], [33], [34]. This instability degrades perovskite-based devices, limiting their performance and viability. Moisture has been identified as a major factor contributing to perovskite degradation. To address these challenges, template-assisted methods have been explored [35], [36], [37], [38], [39]. Porous materials like alumina, graphene, and MOFs provide spatial confinement, improving stability [40], [41], [42], [43], [44], [45], [46]. MOFs feature tunable internal pores, providing nanoscale spaces to isolate QDs from the environment. MOFs hold great potential among various materials due to their tunable nanoscale pores, effectively isolating QDs from ambient conditions. The pore size and metal nodes of MOFs also provide ideal nucleation sites for perovskite precursors, minimizing aggregation and defects during the growth process [47]. Furthermore, the spatial confinement within MOFs significantly mitigates the influence of external factors such as moisture and oxygen, thereby enhancing both the stability and optical properties of the embedded perovskite. Recent studies have verified the advantages and feasibility of perovskite–MOF systems. For example, Wan et al. reported that amino-functionalized UiO-66(NH2) can effectively anchor CsPbBr3 QDs, utilizing stable coordination between Pb2+ and zirconium nodes to improve photocatalytic performance [48]. Xie et al. demonstrated the in situ confined growth of ultra-small CsPbBr3 within UiO-66 pores, highlighting how appropriate pore sizes and defect structures suppress non-radiative recombination and enhance luminescence [49]. Additionally, Shi et al. found that embedding MAPbBr3 into UiO-66 yields uniformly dispersed QDs that maintain stable emission in encryption and decryption applications, indicating that the MOF framework serves as a physical barrier against moisture and oxygen [50]. MOFs significantly reduce environmental degradation of perovskites and preserves their intrinsic optical performance. In this study, we adopt UiO-66 (Zr6(μ3-O)4(μ3-OH)4(BDC)6, BDC = terephthalate) with missing-linker defects as our template [51]. It exhibits outstanding chemical stability, maintaining a stable structure during the synthesis process and its emission wavelength falls outside the visible light spectrum, thus does not affect the PL signal of the sample. Previous studies have examined CsPbBr3 embedded within UiO-66, highlighting its potential for applications such as electrochemical luminescence, photocatalysis, and temperature sensing [48], [49], [50], [51], [52], [53], [54], [55], [56], [57]. Their optical properties, such as exciton-phonon scattering, exciton binding energy, and applications in optoelectronic devices have yet to be fully explored. We examine the stability and optical properties of CsPbBr3@UiO-66 through temperature-dependent PL and TRPL and demonstrate strong exciton–polariton coupling in a hybrid microcavity, highlighted by anti-crossing behavior in dispersion curves – an important advancement for polaritonic devices.
2. Materials and methods
The procedures for synthesizing CsPbBr3@UiO-66 QDs are depicted in Figure 1(a). UiO-66 powder with missing-linker defects was first synthesized by following the reported procedure [58]. Thereafter, the MOF was subjected to a self-limiting solvothermal deposition in MOF (SIM) method to coordinate spatially dispersed Pb2+ ions on hexa-zirconium nodes of UiO-66 [58], [59], [60], [61], resulting in Pb-UiO-66 powder. It is worth mentioning that during the SIM process, transition metal ions such as Pb2+ ions were coordinated on the hexa-zirconium cluster of the MOF where the terminal –OH/OH2 pairs were initially located, by forming the metal-oxygen bonds between the guest metal ion and the cluster; the formation of such structures has been confirmed both computationally and experimentally for a range of guest metal ions [59], [60], [61]. A CsBr precursor solution was subsequently added to the Pb-UiO-66 powder mixture. After this procedure, a solid with obvious photoluminescent property under the irradiation of a handheld UV lamp was obtained, implying the formation of bulk CsPbBr3 QDs within the MOF. This observation also indicates that the Pb–O bonds in Pb-UiO-66 were broken upon the exposure to the CsBr precursor so that the perovskite, CsPbBr3, could be generated; the similar two-step formation of pore-confined nanoparticles in Zr-based MOFs has been reported previously [62], [63]. Figure 1(b) shows the CsPbBr3@UiO-66 power immersed in water, emitting strong green light for over 180 min of storage, and the accelerated aging test is detailed in Figure S1, Supporting Information. Scanning electron microscope (SEM) micrographs (Figure 1(c)) show that the UiO-66 particles exhibit a uniform morphology with an average diameter of approximately 100 nm. Transmission electron microscopy (TEM) micrographs (Figure 1(d)) reveal the presence of CsPbBr3 quantum dots (QDs) within the pores of UiO-66. The high-resolution TEM image shows distinct lattice fringes with an interplanar spacing of approximately 0.58 nm, which corresponds to the (100) plane of CsPbBr3, as supported by the X-ray diffraction (XRD) pattern in Figure 2(d). The CsPbBr3 QDs in TEM images appear larger (beyond 1–2 nm) due to aggregation or overlap within UiO-66. While the UiO-66 pores provide spatial confinement, localized assembly of multiple QDs can result in the apparent size increase. Additional TEM images presented in Table S1 further confirm the crystallinity of the CsPbBr3 QDs, showing consistent lattice plane indexing and interplanar spacing. The observed aggregation does not compromise the framework integrity, as evidenced by the consistent lattice fringes and the uniform distribution of Zr in the framework. These results collectively confirm the successful synthesis of CsPbBr3 QDs within UiO-66, maintaining both crystallinity and structural integrity. For the convenience of subsequent measurements and experiments, the remaining CsPbBr3@UiO-66 is stored in vials, as shown in Figure 1(e). High-angle annular dark-field scanning transmission electron microscope (HAADF-STEM) and energy-dispersive X-ray (EDX) element mapping images of a typical CsPbBr3@UiO-66 are shown in Figure 1(f) and (g). The HAADF-STEM images provide visual confirmation of the spatial confinement of CsPbBr3 QDs within the UiO-66 framework. The contrast between the CsPbBr3 QDs and the UiO-66 framework demonstrates that the QDs are well-dispersed within the MOF pores, maintaining their dispersion within the framework. EDX element mapping (Figure 1(g)) complements the HAADF-STEM observations by providing quantitative compositional analysis. The EDX mapping shows the uniform distribution of Cs, Pb, and Br elements within the UiO-66 framework, further validating the successful embedding of CsPbBr3 QDs. The consistent presence of Zr across the framework confirms the structural integrity of UiO-66 during the synthetic process. The combination of HAADF-STEM and EDX mapping thus demonstrates the spatial confinement of CsPbBr3 QDs and the uniform incorporation of the elements, ensuring the stability and integrity of the composite material.
Figure 1:
Synthesis and characterization of CsPbBr3@UiO-66. (a) Schematic representation of the synthetic procedure for preparing CsPbBr3@UiO-66. (b) Photograph of fresh CsPbBr3@UiO-66 immersed in water under UV light illumination. (c) SEM and (d) TEM images of CsPbBr3@UiO-66. (e) CsPbBr3@UiO-66 is stored in vials and preserved in a dry box. (f) HAADF-STEM and (g) EDX elemental mapping images of a typical CsPbBr3@UiO-66.
Figure 2:
Spectroscopic verification demonstrating the crystallinity and pore size of the CsPbBr3@UiO-66. (a) Adsorption isotherm of UiO-66, Pb-UiO-66, and CsPbBr3@UiO-66. As the pores are filled, the adsorption quantity decreases, suggesting that the available pore space is limited. (b) Pore size distribution. The synthesized UiO-66 has a pore size of approximately 1∼2 nm, consistent with the TEM observations. (c) Powder XRD spectra of synthesized UiO-66, Pb-UiO-66, and CsPbBr3@UiO-66. (d) XRD spectra comparing the experimental results of CsPbBr3@UiO-66 with powdered monoclinic CsPbBr3 QDs and the simulated results of UiO-66.
Understanding the changes in porosity during the sequential incorporation of Pb2+ ions and CsPbBr3 QDs into the UiO-66 framework is essential. If confirms the successful confinement of QDs and evaluating the structural integrity of the MOF. To evaluate the structural integrity and porosity of the synthesized materials, nitrogen adsorption-desorption measurements were performed at 77 K to investigate the porosity of each material. From the isotherm in Figure 2(a), it can be inferred that UiO-66 belongs to the category of typical microporous materials for both Pb-UiO-66 and PQD@UiO-66, exhibiting Type I isotherms [54]. At low relative pressures, the adsorption quantity increases rapidly; as the relative pressure increases, the micropores become fully filled, approaching near-saturation adsorption capacity and resulting in a flat curve. During the experimental process, UiO-66 is sequentially incorporated with Pb and CsPbBr3. Due to the larger volume of Pb atoms and the crystallization of CsPbBr3 within the pores, the volume of the micropores becomes occupied, leading to a decrease in adsorption capacity. Brunauer–Emmett–Teller (BET) surface area of UiO-66 estimated from the isotherm is 1,510 m2/g, similar to that reported previously for defective UiO-66 [64]. After the incorporation of Pb2+ ions and the formation of pore-confined CsPbBr3, the BET surface areas decrease to 970 m2/g and 320 m2/g, respectively. Figure 2(b) presents the pore size distributions extracted from Figure 2(a) with the use of the linear density functional theory (DFT) and a slit-pore model. The result shows that the pore size of UiO-66 is distributed in the range of approximately 1–2 nm, originating from both the structurally defined pore of UiO-66 with the size of around 1.2 nm and defective cavities with sizes ranging from 1.4 to 2.0 nm [58], [61]. The pore size of UiO-66 observed here aligns well with the diameter of representative CsPbBr3 QD found in the TEM image of CsPbBr3@UiO-66 (Figure 1(d)), confirming that we have successfully synthesized QDs and incorporated them into the pores of UiO-66, effectively filling the pores. We next sought to investigate the crystallinity of both CsPbBr3 and UiO-66. XRD measurements were performed on UiO-66, Pb-UiO-66, and CsPbBr3@UiO-66 to analyze their respective diffraction patterns and evaluate how the incorporation of QDs impacts the framework’s structural characteristics. As shown in Figure 2(c), the characteristic low-angle diffraction peaks of UiO-66 are present in all three XRD patterns, confirming that the overall structure of UiO-66 is retained after the SIM process and the subsequent formation of pore-confined CsPbBr3 QDs. It is noted that the intensity of UiO-66 diffraction peaks decreases in the CsPbBr3@UiO-66 pattern compared to pure UiO-66. This reduction in intensity does not indicate the destruction of the UiO-66 framework but rather reflects changes in the material’s composition and the reduced relative content of UiO-66 in the composite. The decrease in peak intensity can be attributed to the incorporation of CsPbBr3 QDs into the UiO-66 framework, which partially occupies the pores of the MOF and reduces its overall proportion in the composite sample. This observation is also corroborated by nitrogen adsorption–desorption measurements, which show that the BET surface area of the composite remains approximately 320 m2/g following the formation of pore-confined CsPbBr3. Given that QDs contribute minimally to the surface area, this result strongly supports the conclusion that the UiO-66 framework retains both its structural integrity and porosity throughout the synthesis process. Therefore, while the XRD peak intensity decreases, this change is consistent with the expected compositional shift due to the incorporation of CsPbBr3 QDs and does not reflect collapse or destruction of the UiO-66 framework. The combined XRD and BET results provide strong evidence that the structural integrity of the UiO-66 framework is maintained during and after the pore-confined synthesis of CsPbBr3 QDs. To confirm the crystallinity of CsPbBr3 synthesized within UiO-66, we employed an indirect approach by comparing experimental XRD patterns with simulated lattice arrangements of UiO-66 and orthorhombic CsPbBr3 QDs. We found that the CsPbBr3 can crystallize within UiO-66 while maintaining a crystalline structure similar to that of CsPbBr3 QDs (COD code 4510745). The characteristic peaks located at 26.4°, 26.5°, 27.46°, 28.62°, 30.43°, 30.74° correspond to the (220), (022), (131), (221), (040) and (202) planes. The interplanar spacing calculated from XRD is consistent with the lattice fringes observed in the TEM micrographs. Detailed interplanar spacings for these planes are presented in Table S1 of the Supporting Information. From these interplanar spacing values and their corresponding planes, we derived the lattice constants of CsPbBr3@UiO-66 as a = 8.2437 Å, b = 11.7405 Å, and c = 8.1982 Å, demonstrating the preservation of crystallinity and structural integrity in the synthesized composite material.
FTIR spectra of UiO-66, Pb-UiO-66 and CsPbBr3@UiO-66 were collected and plotted in Figure S2. Three major characteristic peaks of UiO-66, including those associated with the asymmetric vibration of O–C–O in coordinated carboxylate groups at 1,582 cm−1, the C=C in the aromatic ring of the linker at 1,506 cm−1 and the symmetric vibration of O–C–O at 1,395 cm−1, can be clearly observed in all spectra. These features agree well with those of defective UiO-66 reported previously [65], [66]. Only a tiny peak located at around 1,700 cm−1, attributed to the uncoordinated carboxylic groups of linkers, can be observed in the FTIR spectrum of UiO-66, again confirming the successful synthesis of the MOF. The intensity of this peak is almost the same in both spectra of Pb-UiO-66 and CsPbBr3@UiO-66, which implies that the installation of Pb2+ ions and the further formation of pore-confined perovskite QDs did not significantly alter the coordination of the porous framework. As revealed by FTIR data, no chemical bonds were formed between the organic moieties of MOF and perovskite, suggesting that the CsPbBr3 QDs were physically confined within the MOF pore.
3. Results and discussion
Temperature-dependent PL spectra of CsPbBr3@UiO-66 were measured from 4 K to 300 K (Figure 3(a)). The integrated PL intensity was analyzed using an Arrhenius equation (Equation (1)):
(1) |
where I 0 is the PL spectrum area at 0 K, E b (meV) is the exciton binding energy, T (K) is temperature. A is fitting parameter. The fitting results yield E b ≈ 59 ± 0.8 meV, higher than values reported for bulk CsPbBr3 and thin films (Table S5). The exciton binding energy depends on the spatial confinement of charge carriers. In bulk materials or thin films, the continuous density of states (DOS) and spatial delocalization weaken the Coulomb interaction between electrons and holes. In contrast, the nanoscopic pores (∼1–2 nm) of UiO-66 confine CsPbBr3 QDs, limiting carrier delocalization and increasing spatial overlap, which enhances electron-hole interaction and raises the energy required to dissociate excitons into free carriers. Unlike bulk and thin-film perovskites, CsPbBr3@UiO-66 exhibited a slight blue shift in PL peak positions with temperature increase. This indicates CsPbBr3@UiO-66’s thermal and structural dynamics differ from conventional perovskites. To further investigate the underlying mechanisms, we used a one-oscillator model (Equation (2)) to analyze the temperature dependence of the PL peak energy:
(2) |
where E g (eV) is the bandgap, E g,0 (eV) is the bandgap without thermal and phonon contributions, C th (eV/K) is the coefficient of thermal expansion, C ep (eV) is the coefficient of electron-phonon interaction and ℏω (eV) is the average optical phonon energy. The details of the fitting results refer to Table S3. In the low-temperature range (4–150 K), the PL peak energy shows a linear increase, which is well described by the thermal expansion term. The extracted coefficient C th = is significantly smaller than that typically reported for CsPbBr3 [64], [67], [68], suggesting that the thermal effects on lattice expansion are suppressed by the confinement within the UiO-66 framework. At higher temperatures (T > 150 K), the slope of the PL peak energy curve begins to gradually decrease. This behavior arises from thermally activated electron–phonon interactions, which lead to bandgap renormalization (orange dash line). The fitting yields a negative electron–phonon coupling coefficient , indicating that electron–phonon interactions reduce the bandgap energy as the thermal energy increases. The fitted average phonon energy ℏω ≈ 47 ± 2 meV indicates that the porous UiO-66 structure amplifies this effect by introducing surface/interface-induced defects, intensifying phonon–electron coupling.
Figure 3:
Temperature-dependent PL characteristics of CsPbBr3@UiO-66. (a) Temperature-dependent PL spectrum of CsPbBr3@UiO-66 measured across a temperature range from 4 K to 300 K. The spectrum exhibits distinct temperature-dependent behavior, where the emission peak progressively shifts towards shorter wavelengths as the temperature increases, accompanied by a broadening of the FWHM. This blue-shifting phenomenon is indicative of the reduction in exciton binding energy with rising thermal energy and is further analyzed in 3(c). (b) Integrated PL intensity plotted as a function of temperature. The data closely follow an Arrhenius model, accounting for the exciton binding energy, as described by Equation (1). (c) Temperature-dependent blueshift of the PL emission peak energy, fitted using the one-oscillator model (Equation (2)). The shift in emission energy is primarily attributed to thermal expansion effects at lower temperatures (<150 K), while electron-phonon interactions dominate at higher temperatures (>150 K). The fitting results, shown by the solid curve, highlight the contribution of thermal lattice expansion (C th ) and electron–phonon coupling (C ep ), with the observed blueshift further corroborating the excellent thermal stability of the CsPbBr3@UiO-66 system. (d) TRPL measurement of QDs, CsPbBr3@UiO-66 and TF. The results reveal that CsPbBr3@UiO-66 exhibits τ r comparable to QDs, yet its τ nr is significantly longer than that of TF. This highlights the dual role of the UiO-66 framework in preserving radiative properties while effectively suppressing non-radiative recombination through reduced bulk defects and enhanced surface passivation. (e) Absorption spectra of CsPbBr3@UiO-66, with the experimental data and the theoretical fit. The E b measured from this spectrum is similar with the value derived from Arrhenius model in Figure 3(b).
To quantitatively analyze the temperature-dependent PL linewidth broadening, we fitted the measured full width at half maximum (FWHM) using the independent Boson model (3) [69], [70]:
(3) |
where Γ (meV) is linewidth, Γ0 (meV) is inhomogeneous broadening, ℏω LO (meV) is longitudinal optical (LO) phonon energy and ΓLO (meV) is the contribution of electron–optical phonon interaction. The details of the fitting results refer to Table S4 and Figure S3. The LO phonon energy extracted from the linewidth analysis is ℏω LO ≈ 29 ± 3, whereas the average phonon energy obtained from Equation (2) yields ℏω ≈ 47 ± 2 meV. This difference arises because the average phonon energy ℏω is an effective value encompassing contributions from multiple phonon modes and electron–phonon interactions across the entire electron population. The clear identification of a dominant LO phonon mode from the independent Boson model suggests that electron–phonon coupling in our CsPbBr3@UiO-66 system is predominantly mediated through LO phonons. This finding indicates that Fröhlich-type coupling to LO phonons – rather than acoustic phonon scattering – serves as the principal mechanism underlying the homogeneous PL linewidth broadening in this hybrid material. Such behavior aligns well with previous studies on bulk lead-halide perovskites, where Fröhlich interactions involving polar LO phonons have consistently been identified as the dominant source of PL line broadening at room temperature [71], [72]. The associated Fröhlich scattering rate (1/τ fr ) can be expressed as proportional to [73], [74]:
(4) |
where and are the high-frequency and low-frequency dielectric constants of the medium and is the LO phonon occupation number described by the Bose–Einstein distribution: . As the temperature increases, the LO phonon population rises, enhancing the Fröhlich scattering rate. This intensified electron–phonon scattering at higher temperatures accelerates exciton energy redistribution and momentum relaxation processes, thereby leading to pronounced PL linewidth broadening. The observed temperature-dependent increase in linewidth clearly reflects the thermal broadening mechanisms driven by enhanced electron–phonon coupling.
To examine carrier dynamics influenced by the UiO-66 framework, we performed TRPL measurements and compared them with those of CsPbBr3 thin films (TF) and QDs (Figure 3(d)). PL decay curves were precisely fitted using a double-exponential decay function: , where A 1 and A 2 are the amplitudes of the two exponential decay components. The extracted radiative (τ r ) and non-radiative lifetimes (τ nr ) for CsPbBr3@UiO-66, TF, and QDs are summarized in Table 1 and the average lifetime associated with τ r and τ nr can be described as Equation (5) [75]:
(5) |
Table 1:
Lifetime of QDs, CsPbBr3@UiO-66 and TF.
QDs | CsPbBr3@UiO-66 | TF | |
---|---|---|---|
1.34 | 3.00 | 6.35 | |
4.04 | 16.52 | 15.21 | |
2.50 | 11.43 | 13.04 |
Detailed fitting results of the TRPL measurements for these materials are provided in Table S5 and Figure S4 of the Supporting Information. TRPL measurements of CsPbBr3@UiO-66 reveal distinct recombination behaviors, with radiative recombination lifetimes intermediate between TF and QDs, but with longer non-radiative recombination times.
In perovskite materials, non-radiative recombination pathways typically involve bulk defects, such as vacancies and interstitials, as well as surface traps caused by surface irregularities [76], [77], [78], [79]. Under low pump fluence conditions, we employed a set of differential equations (Supporting Information) to describe the photogenerated carrier density (n c (t)) following PL excitation [80]. This model identifies the presence of two distinct types of trap states in CsPbBr3@UiO-66: bulk traps, characterized by fast trapping times, and surface traps, which exhibit slower trapping dynamics [81], [82], [83]. The bulk trap density is whereas the surface trap density is . The bulk trap density of CsPbBr3@UiO-66 is lower than that of QDs and TF, whereas its surface trap density is significantly higher than both. We propose that the UiO-66 framework influences non-radiative recombination through a dual role. It acts as a protective barrier for CsPbBr3, shielding it from environmental factors like moisture and oxygen, thereby reducing internal crystal defects and passivating bulk traps. Simultaneously, the high surface area and porous structure of UiO-66 create numerous interfaces between CsPbBr3 QDs and the framework, which become hotspots for trap states [84], [85]. Additionally, vibrational modes at these interfaces may couple with the electronic states of CsPbBr3, introducing further non-radiative recombination pathways. This dual influence highlights the interplay between the structural protection offered by UiO-66 and the trap state formation induced at its interfaces. We also measure the absorption spectrum of CsPbBr3@UiO66 at room temperature, and model that in the framework of the Elliot’s theory including hydrogen-like excitonic effects [86], [87], [88]. As shown in Figure 3(e), the absorption spectrum of CsPbBr3@UiO-66 clearly reveals an excitonic peak distinct from the continuum band absorption. The exciton binding energy obtained from the absorption spectra in Figure 3(e) is consistent with the value derived from the Arrhenius equation in Figure 3(b). CsPbBr3@UiO-66 yields an exciton binding energy E b ≈ 60 ± 3 meV, which is larger than that of QDs and TF (E b ≈ 45 ± 5 and 30 ± 4 meV, Table S7). The spectrum’s broad linewidth is attributed to inelastic scattering caused by material defects and phonon interactions. Defects, such as surface traps and bulk vacancies, create localized energy states within the bandgap, broadening absorption pathways. Additionally, strong electron–phonon coupling in the confined CsPbBr3@UiO-66 system intensifies these effects, as phonon-induced lattice vibrations further disrupt electronic states, resulting in the observed broadening of excitonic and continuum absorption features. To assess the air and water stability of CsPbBr3@UiO-66, it was spin-coated onto a sapphire substrate, and the powder was measured individually. Using a 405 nm CW laser at 1 kW/cm2, the normalized PL intensity remained stable under ambient conditions (60 % relative humidity, 25 °C) for over 120 weeks and for 36 months when stored in a dry box (Figure 4(a)). Statistical analysis of the first five weeks showed a median emission wavelength of ∼528 nm (Figure 4(b)), which informs the design of the resonant cavity length model. Additionally, CsPbBr3@UiO-66 retained 70 % of its initial PL intensity after 20 days of water immersion (Figure 4(c)).
Figure 4:
Environmental stability of CsPbBr3@UiO-66 in aqueous and ambient conditions. (a) Long-term stability of CsPbBr3@UiO-66 in ambient conditions (60 % relative humidity, 25 °C). The normalized PL intensity remains unchanged over 120 weeks, demonstrating the material’s remarkable environmental resistance. Additionally, after being stored in a dry box for 36 months, the PL intensity continues to be stable with no significant degradation, indicating superior long-term stability under both ambient and controlled conditions. (b) Long-term stability of CsPbBr3@UiO-66 in ambient conditions (60 % relative humidity, 25 °C). The material exhibits outstanding stability, with the normalized PL intensity remaining constant over a period of 120 weeks, with no signs of degradation even after 36 months in a dry storage environment. (c) Water stability test of CsPbBr3@UiO-66. The normalized PL intensity is measured after submersion in water for over 20 days, retaining 70 % of its initial intensity, demonstrating significant resistance to water-induced degradation. The inset shows a photograph of the sample in its aqueous environment, further highlighting its stability in real-world conditions.
To explore strong light–matter interactions, CsPbBr3@UiO-66 was integrated into a hybrid MC with a distributed Bragg reflector (DBR) at the bottom and an Ag layer on top. The cavity quality factor was measured as Q = 151. The linewidth of the exciton transition is ℏΓ ex = 69.4 meV which is extracted from the PL spectra, as shown in Figure S7(a) and (b). The strong coupling between the cavity modes and excitons in CsPbBr3@UiO-66 is investigated by measuring the angle-resolved photoluminescence (ARPL) spectroscopy in Figure 5(a). The PL spectra reveal the formation of polaritonic states, with distinct lower polariton branches (LPB) and upper polariton branches (UPB). Figure 5(b) shows the PL spectrum measured at a specific angle (9.37°), where two emission peaks corresponding to the LPB and UPB are clearly observed. The peak positions were precisely determined using Lorentzian fitting. The results indicate a clear energy splitting between the polaritonic branches. We reach the strong coupling regime at room temperature where the cavity detuning is = E C − E ex = −35 meV. To directly investigate exciton–photon coupling, the dispersion curve was fitted using the coupled oscillator mode [89]:
(6) |
where E C and E ex correspond to energy levels of cavity modes and excitons coupled by the interaction potential g A , respectively. E indicates the eigenvalues corresponding to the energies of polariton modes. The photon and exciton fraction in each LPB and UPB are given by the amplitude squared of C and X, where |C|2 + |X|2 = 1. They are referred to as the Hopfield coefficients. g A is the exciton–photon coupling strength.
Figure 5:
ARPL mapping and polariton dispersion in CsPbBr3@UiO-66 microcavity. (a) ARPL spectra for the device from 1.45° to 30°. The vertical red dashed line represents the CsPbBr3@UiO-66 exciton energy and the red curves trace the dispersion of microcavity polariton modes. (b) Expanded view of PL spectra features at 9.37°. A significant LPB PL peak is observed around 2.371 eV, as well as a prominent UPB peak at 2.317 eV. The PL spectrum is also fitted to multiple Lorentzian peaks to locate the exact PL peak positions. (c) The energy versus angle dispersion, extracted from the ARPL spectra. The dispersion is fitted to a coupled oscillator model, showing Rabi splitting of ∼45 meV. (d), (e) The corresponding exciton and cavity photon weights of the UPB and LPB dispersions as a function of the angles.
The eigenvalues are given by
(7) |
From Equation (7), two hybrid modes E UPB and E LPB, are observed. The ARPL data in Figure 5(a) were analyzed, and the fitted peak positions are shown in Figure 5(c), revealing the dispersion relationship of E UPB and E LPB. At the emission angle of 24°, the Rabi splitting was calculated as , satisfying the strong coupling condition ℏΩRabi > (ℏΓ ex + ℏΓMC)/2. This confirms the system is in the strong coupling regime. Figure 5(d) and (e) show the exciton and photon components of the polariton states, described by weighing fractions calculated as Hopfield coefficients, plotted as functions of cavity detuning [90].
4. Conclusion
In this study, we synthesized a CsPbBr3@UiO-66 composite by assembling CsPbBr3 QDs within UiO-66 micropores. This approach enhanced the environmental stability of CsPbBr3 QDs, enabling continuous luminescence for over 30 months in ambient air and stable emission in water for several days. This stability addresses a critical limitation of traditional perovskite materials in optoelectronic applications. Temperature-dependent PL and TRPL measurements revealed detailed exciton–phonon interactions, providing insights into the material’s fundamental optical properties. We anticipate that these optical features can be generalized to other MOFs and perovskite QD compositions. We anticipate that these optical features can be generalized to other MOFs and perovskite QD compositions. With the porous MOF to coordinate metal ions such as Pb2+ first, followed by the heterogeneous reaction between the porous solid and another precursor to form the perovskite QDs confined within the MOF pore, the size and optical property of the resulting perovskite QDs should be highly tunable by simply adjusting the pore size and pore structure of MOFs. However, since most MOFs themselves are not chemically stable in water [91], the careful selection of MOFs that are both stable in water as well as capable of coordinating Pb2+ ions prior to the further formation of perovskite is necessary. In addition to UiO-66, such a two-step formation of pore-confined perovskite QDs should also be generalizable to other group (IV) metal-based MOFs such as MOF-808, NU-1000 and PCN-222, which are all chemically robust in water and possess terminal –OH/OH2 pairs on their nodes for the installation of guest metal ions by SIM [92]. Taken together, these findings suggest that the rational selection of MOFs – featuring chemical stability, pore size compatibility, and low background emission – could provide a universal strategy for enhancing the optical performance and durability of perovskite QDs. The strong light–matter interaction in hybrid MC demonstrated the formation of hybrid polaritonic states, positioning the CsPbBr3@UiO-66 composite as a promising candidate for polaritonic devices and strong coupling studies. The combination of superior environmental stability and strong coupling behavior highlights the CsPbBr3@UiO-66 composite as a robust and versatile platform for advanced optoelectronic applications. These results advance perovskite-based technologies in polaritonics [76], low-threshold lasing [77], and emerging photonic applications [78], [79], [93]. Future work could optimize optical properties and explore scalable integration.
Supplementary Material
Supplementary Material Details
Acknowledgments
The authors would like to thank Meng-Cheng Yen and Prof. Ya-Ju Lee for providing the CsPbBr3 QDs in the experiments.
Supplementary Material
This article contains supplementary material (https://doi.org/10.1515/nanoph-2025-0059).
Footnotes
Research funding: This research received partial support from The Pilot Directions for NSTC Grant for the Einstein Program, National Science and Technology Council (NSTC), and Ministry of Science and Technology (MOST) under Contract Numbers MOST 111-2636-M-006-025, NSTC 111-2112-M-006-023-MY3, 112-2223-E-006-003-MY3, 113-2221-E-006-104-MY3 and 113-2112-M-006-031. We also thank the support from Ministry of Education (MOE) of Taiwan under Yushan Young Fellow Program (MOE-112-YSFEE-0005-001-P2). Additionally, it was supported by the Higher Education Sprout Project, Ministry of Education, granted to the Headquarters of University Advancement at National Cheng Kung University (NCKU).
Author contributions: CCL: methodology, investigation, data curation, writing. SCW, CHS, ZLL, YL, ZYW, SCW and CKL: data curation. CWK and HCH: resources, writing - review & editing, funding acquisition. YHC: conceptualization, supervision, resources, writing - review & editing, project administration, funding acquisition. All authors have accepted responsibility for the entire content of this manuscript and consented to its submission to the journal, reviewed all the results and approved the final version of the manuscript.
Conflict of interest: Authors state no conflict of interest.
Data availability: The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.
Associated content: Further experimental details regarding the instrumentation for synthesis and material characterizations were included in the Supporting Information.
References
- [1].Lee K.-H., et al. Highly efficient, color-pure, color-stable blue quantum dot light-emitting devices. ACS Nano . 2013;7(8):7295–7302. doi: 10.1021/nn402870e. [DOI] [PubMed] [Google Scholar]
- [2].Steckel J. S., et al. Quantum dots: the ultimate down-conversion material for LCD displays. J. Soc. Inf. Disp. . 2015;23(7):294–305. doi: 10.1002/jsid.313. [DOI] [Google Scholar]
- [3].Swarnkar A., et al. Quantum dot–induced phase stabilization of α-CsPbI3 perovskite for high-efficiency photovoltaics. Science . 2016;354(6308):92–95. doi: 10.1126/science.aag2700. [DOI] [PubMed] [Google Scholar]
- [4].Wang H.-C., et al. Mesoporous silica particles integrated with all-inorganic CsPbBr3 perovskite quantum-dot nanocomposites (MP-PQDs) with high stability and wide color gamut used for backlight display. Angew. Chem., Int. Ed. . 2016;55(28):7924–7929. doi: 10.1002/anie.201603698. [DOI] [PubMed] [Google Scholar]
- [5].Yoon H. C., Kang H., Lee S., Oh J. H., Yang H., Do Y. R. Study of perovskite QD down-converted LEDs and six-color white LEDs for future displays with excellent color performance. ACS Appl. Mater. Interfaces . 2016;8(28):18189–18200. doi: 10.1021/acsami.6b05468. [DOI] [PubMed] [Google Scholar]
- [6].Li G., Huang J., Zhu H., Li Y., Tang J.-X., Jiang Y. Surface ligand engineering for near-unity quantum yield inorganic halide perovskite QDs and high-performance QLEDs. Chem. Mater. . 2018;30(17):6099–6107. doi: 10.1021/acs.chemmater.8b02544. [DOI] [Google Scholar]
- [7].Wei Y., Cheng Z., Lin J. An overview on enhancing the stability of lead halide perovskite quantum dots and their applications in phosphor-converted LEDs. Chem. Soc. Rev. . 2019;48(1):310–350. doi: 10.1039/C8CS00740C. [DOI] [PubMed] [Google Scholar]
- [8].Dong Y., et al. Bipolar-shell resurfacing for blue LEDs based on strongly confined perovskite quantum dots. Nat. Nanotechnol. . 2020;15(8):668–674. doi: 10.1038/s41565-020-0714-5. [DOI] [PubMed] [Google Scholar]
- [9].Liu Z., et al. Micro-light-emitting diodes with quantum dots in display technology. Light Sci. Appl. . 2020;9(1):83. doi: 10.1038/s41377-020-0268-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [10].Lin J., et al. Perovskite quantum dots glasses based backlit displays. ACS Energy Lett. . 2021;6(2):519–528. doi: 10.1021/acsenergylett.0c02561. [DOI] [Google Scholar]
- [11].Du J., et al. Zn-Cu-In-Se quantum dot solar cells with a certified power conversion efficiency of 11.6% J. Am. Chem. Soc. . 2016;138(12):4201–4209. doi: 10.1021/jacs.6b00615. [DOI] [PubMed] [Google Scholar]
- [12].Veldhuis S. A., et al. Perovskite materials for light-emitting diodes and lasers. Adv. Mater. . 2016;28(32):6804–6834. doi: 10.1002/adma.201600669. [DOI] [PubMed] [Google Scholar]
- [13].Song J., et al. Room-temperature triple-ligand surface engineering synergistically boosts ink stability, recombination dynamics, and charge injection toward EQE-11.6% perovskite QLEDs. Adv. Mater. . 2018;30(30):1800764. doi: 10.1002/adma.201800764. [DOI] [PubMed] [Google Scholar]
- [14].Shen H., et al. Visible quantum dot light-emitting diodes with simultaneous high brightness and efficiency. Nat. Photonics . 2019;13(3):192–197. doi: 10.1038/s41566-019-0364-z. [DOI] [Google Scholar]
- [15].Song J., et al. Over 30% external quantum efficiency light-emitting diodes by engineering quantum dot-assisted energy level match for hole transport layer. Adv. Funct. Mater. . 2019;29(33):1808377. doi: 10.1002/adfm.201808377. [DOI] [Google Scholar]
- [16].Eames C., Frost J. M., Barnes P. R. F., O’Regan B. C., Walsh A., Islam M. S. Ionic transport in hybrid lead iodide perovskite solar cells. Nat. Commun. . 2015;6(1):7497. doi: 10.1038/ncomms8497. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [17].Jeon N. J., et al. Compositional engineering of perovskite materials for high-performance solar cells. Nature . 2015;517(7535):476–480. doi: 10.1038/nature14133. [DOI] [PubMed] [Google Scholar]
- [18].Protesescu L., et al. Nanocrystals of cesium lead halide perovskites (CsPbX3, X = Cl, Br, and I): novel optoelectronic materials showing bright emission with wide color gamut. Nano Lett. . 2015;15(6):3692–3696. doi: 10.1021/nl5048779. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [19].Berhe T. A., et al. Organometal halide perovskite solar cells: degradation and stability. Energy Environ. Sci. . 2016;9(2):323–356. doi: 10.1039/c5ee02733k. [DOI] [Google Scholar]
- [20].Yang W. S., et al. Iodide management in formamidinium-lead-halide-based perovskite layers for efficient solar cells. Science . 2017;356(6345):1376–1379. doi: 10.1126/science.aan2301. [DOI] [PubMed] [Google Scholar]
- [21].Jena A. K., Kulkarni A., Miyasaka T. Halide perovskite photovoltaics: background, status, and future prospects. Chem. Rev. . 2019;119(5):3036–3103. doi: 10.1021/acs.chemrev.8b00539. [DOI] [PubMed] [Google Scholar]
- [22].Kim M., et al. Methylammonium chloride induces intermediate phase stabilization for efficient perovskite solar cells. Joule . 2019;3(9):2179–2192. doi: 10.1016/j.joule.2019.06.014. [DOI] [Google Scholar]
- [23].Kim J. Y., Lee J.-W., Jung H. S., Shin H., Park N.-G. High-efficiency perovskite solar cells. Chem. Rev. . 2020;120(15):7867–7918. doi: 10.1021/acs.chemrev.0c00107. [DOI] [PubMed] [Google Scholar]
- [24].Min H., et al. Perovskite solar cells with atomically coherent interlayers on SnO2 electrodes. Nature . 2021;598(7881):444–450. doi: 10.1038/s41586-021-03964-8. [DOI] [PubMed] [Google Scholar]
- [25].Yoo J. J., et al. Efficient perovskite solar cells via improved carrier management. Nature . 2021;590(7847):20250059. doi: 10.1038/s41586-021-03285-w. [DOI] [PubMed] [Google Scholar]
- [26].Deschler F., et al. High photoluminescence efficiency and optically pumped lasing in solution-processed mixed halide perovskite semiconductors. J. Phys. Chem. Lett. . 2014;5(8):1421–1426. doi: 10.1021/jz5005285. [DOI] [PubMed] [Google Scholar]
- [27].Song J., Li J., Li X., Xu L., Dong Y., Zeng H. Quantum dot light-emitting diodes based on inorganic perovskite cesium lead halides (CsPbX3) Adv. Mater. . 2015;27(44):7162–7167. doi: 10.1002/adma.201502567. [DOI] [PubMed] [Google Scholar]
- [28].Stranks S. D., Snaith H. J. Metal-halide perovskites for photovoltaic and light-emitting devices. Nat. Nanotechnol. . 2015;10(5):391–402. doi: 10.1038/NNANO.2015.90. [DOI] [PubMed] [Google Scholar]
- [29].McClure E. T., Ball M. R., Windl W., Woodward P. M. Cs2AgBiX6 (X = Br, CI): new visible light absorbing, lead-free halide perovskite semiconductors. Chem. Mater. . 2016;28(5):1348–1354. doi: 10.1021/acs.chemmater.5b04231. [DOI] [Google Scholar]
- [30].Zhang X., et al. Stable high efficiency two-dimensional perovskite solar cells via cesium doping. Energy Environ. Sci. . 2017;10(10):2095–2102. doi: 10.1039/c7ee01145h. [DOI] [Google Scholar]
- [31].Zuo L., et al. Polymer-modified halide perovskite films for efficient and stable planar heterojunction solar cells. Sci. Adv. . 2017;3(8):e1700106. doi: 10.1126/sciadv.1700106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [32].Brennan M. C., Draguta S., Kamat P. V., Kuno M. Light-induced anion phase segregation in mixed halide perovskites. ACS Energy Lett. . 2018;3(1):204–213. doi: 10.1021/acsenergylett.7b01151. [DOI] [Google Scholar]
- [33].Niu T., et al. Stable high-performance perovskite solar cells via grain boundary passivation. Adv. Mater. . 2018;30(16):1706576. doi: 10.1002/adma.201706576. [DOI] [PubMed] [Google Scholar]
- [34].Boyd C. C., Cheacharoen R., Leijtens T., McGehee M. D. Understanding degradation mechanisms and improving stability of perovskite photovoltaics. Chem. Rev. . 2019;119(5):3418–3451. doi: 10.1021/acs.chemrev.8b00336. [DOI] [PubMed] [Google Scholar]
- [35].McBean C. L., et al. Generalizable, electroless, template-assisted synthesis and electrocatalytic mechanistic understanding of perovskite LaNiO3 nanorods as viable, supportless oxygen evolution reaction catalysts in alkaline media. ACS Appl. Mater. Interfaces . 2017;9(29):24634–24648. doi: 10.1021/acsami.7b06855. [DOI] [PubMed] [Google Scholar]
- [36].Shi P., et al. Template-assisted formation of high-quality α-phase HC(NH2)2PbI3 perovskite solar cells. Adv. Sci. . 2019;6(21):1901591. doi: 10.1002/advs.201901591. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [37].Vila-Liarte D., et al. Templated-assembly of CsPbBr3 perovskite nanocrystals into 2D photonic supercrystals with amplified spontaneous emission. Angew. Chem., Int. Ed. . 2020;59(40):17750–17756. doi: 10.1002/anie.202006152. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [38].Wang B., et al. Large-scale synthesis of highly luminescent perovskite nanocrystals by template-assisted solid-state reaction at 800 °C. Chem. Mater. . 2020;32(1):308–314. doi: 10.1021/acs.chemmater.9b03804. [DOI] [Google Scholar]
- [39].Favacho V. S. S., Melo D. M. A., Costa J. E. L., Silva Y. K. R. O., Braga R. M., Medeiros R. L. B. A. Perovskites synthesized by soft template-assisted hydrothermal method: a bibliometric analysis and new insights. Int. J. Hydrogen Energy . 2024;78(12):1391–1428. doi: 10.1016/j.ijhydene.2024.06.326. [DOI] [Google Scholar]
- [40].Li B., Wen H.-M., Cui Y., Zhou W., Qian G., Chen B. Emerging multifunctional metal–organic framework materials. Adv. Mater. . 2016;28(40):8819–8860. doi: 10.1002/adma.201601133. [DOI] [PubMed] [Google Scholar]
- [41].Tong L., et al. Recent progress in the preparation and application of quantum dots/graphene composite materials. RSC Adv. . 2017;7(76):47999–48018. doi: 10.1039/C7RA08755A. [DOI] [Google Scholar]
- [42].He X., et al. Synthesis of perovskite CsPbBr3 quantum dots/porous boron nitride nanofiber composites with improved stability and their reversible optical response to ammonia. Inorg. Chem. . 2020;59(2):1234–1241. doi: 10.1021/acs.inorgchem.9b02947. [DOI] [PubMed] [Google Scholar]
- [43].Meng J., et al. Advances in metal–organic framework coatings: versatile synthesis and broad applications. Chem. Soc. Rev. . 2020;49(10):3142–3186. doi: 10.1039/C9CS00806C. [DOI] [PubMed] [Google Scholar]
- [44].Tang Y., Wu H., Cao W., Cui Y., Qian G. Luminescent metal–organic frameworks for white LEDs. Adv. Opt. Mater. . 2021;9(23):2001817. doi: 10.1002/adom.202001817. [DOI] [Google Scholar]
- [45].Bai Y., et al. Metal-organic frameworks nanocomposites with different dimensionalities for energy conversion and storage. Adv. Energy Mater. . 2022;12(4):2100346. doi: 10.1002/aenm.202100346. [DOI] [Google Scholar]
- [46].Sunny F., Kurukkal Balakrishnan S., Kalarikkal N. An overview on lead halide perovskite based composites and heterostructures: synthesis and applications. ChemNanoMat . 2024;10(3):e202300484. doi: 10.1002/cnma.202300484. [DOI] [Google Scholar]
- [47].Xu X., et al. Microwave-assisted in-situ synthesis of low-dimensional perovskites within metal-organic frameworks for optoelectronic applications. Appl. Mater. Today . 2024;40:102418. doi: 10.1016/j.apmt.2024.102418. [DOI] [Google Scholar]
- [48].Wan S., Ou M., Zhong Q., Wang X. Perovskite-type CsPbBr3 quantum dots/UiO-66(NH2) nanojunction as efficient visible-light-driven photocatalyst for CO2 reduction. Chem. Eng. J. . 2019;358(15):1287–1295. doi: 10.1016/j.cej.2018.10.120. [DOI] [Google Scholar]
- [49].Xie Z., et al. In situ confined growth of ultrasmall perovskite quantum dots in metal–organic frameworks and their quantum confinement effect. Nanoscale . 2020;12(32):17113–17120. doi: 10.1039/D0NR04741D. [DOI] [PubMed] [Google Scholar]
- [50].Shi L., Wang J., Zhou L., Chen Y., Yan J., Dai C. Facile in-situ preparation of MAPbBr3@UiO-66 composites for information encryption and decryption. J. Solid State Chem. . 2020;282:121062. doi: 10.1016/j.jssc.2019.121062. [DOI] [Google Scholar]
- [51].Cavka J. H., et al. A new zirconium inorganic building brick forming metal organic frameworks with exceptional stability. J. Am. Chem. Soc. . 2008;130(42):13850–13851. doi: 10.1021/ja8057953. [DOI] [PubMed] [Google Scholar]
- [52].Dirin D. N., et al. Harnessing defect-tolerance at the nanoscale: highly luminescent lead halide perovskite nanocrystals in mesoporous silica matrixes. Nano Lett. . 2016;16(9):5866–5874. doi: 10.1021/acs.nanolett.6b02688. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [53].Zhao G., Huang X., Wang X., Wang X. Progress in catalyst exploration for heterogeneous CO2 reduction and utilization: a critical review. J. Mater. Chem. A . 2017;5(41):21625–21649. doi: 10.1039/C7TA07290B. [DOI] [Google Scholar]
- [54].Abebe B., Murthy H. C. A., Amare E. Summary on adsorption and photocatalysis for pollutant remediation: mini review. J. Encapsulation Adsorpt. Sci. . 2018;8(4):4. doi: 10.4236/jeas.2018.84012. [DOI] [Google Scholar]
- [55].Zhang J., Tian B., Wang L., Xing M., Lei J. Roles and properties of cocatalysts in semiconductor-based materials for efficient CO2 photoreduction. In: Zhang J., Tian B., Wang L., Xing M., Lei J., editors. Photocatalysis: Fundamentals, Materials and Applications . Singapore: Springer; 2018. pp. 275–305. [Google Scholar]
- [56].Liu T., et al. A visual electrochemiluminescence molecularly imprinted sensor with Ag+@UiO-66-NH2 decorated CsPbBr3 perovskite based on smartphone for point-of-care detection of nitrofurazone. Chem. Eng. J. . 2022;429(1):132462. doi: 10.1016/j.cej.2021.132462. [DOI] [Google Scholar]
- [57].Zhang C., et al. Based on Z-scheme heterojunction CsPbBr3/UiO-66 composite photocatalytic degradation. Appl. Organomet. Chem. . 2023;37(7):e7122. doi: 10.1002/aoc.7122. [DOI] [Google Scholar]
- [58].DeStefano M. R., Islamoglu T., Garibay S. J., Hupp J. T., Farha O. K. Room-temperature synthesis of UiO-66 and thermal modulation of densities of defect sites. Chem. Mater. . 2017;29(3):1357–1361. doi: 10.1021/acs.chemmater.6b05115. [DOI] [Google Scholar]
- [59].Noh H., et al. An exceptionally stable metal–organic framework supported molybdenum(VI) oxide catalyst for cyclohexene epoxidation. J. Am. Chem. Soc. . 2016;138(44):14720–14726. doi: 10.1021/jacs.6b08898. [DOI] [PubMed] [Google Scholar]
- [60].Li Z., et al. Fine-tuning the activity of metal–organic framework-supported cobalt catalysts for the oxidative dehydrogenation of propane. J. Am. Chem. Soc. . 2017;139(42):15251–15258. doi: 10.1021/jacs.7b09365. [DOI] [PubMed] [Google Scholar]
- [61].Chuang C.-H., Li J.-H., Chen Y.-C., Wang Y.-S., Kung C.-W. Redox-hopping and electrochemical behaviors of metal–organic framework thin films fabricated by various approaches. J. Phys. Chem. C . 2020;124(38):20854–20863. doi: 10.1021/acs.jpcc.0c03873. [DOI] [Google Scholar]
- [62].Chen Y.-L., et al. Pore-confined cobalt sulphide nanoparticles in a metalorganic framework as a catalyst for the colorimetric detection of hydrogen peroxide. Mater. Adv. . 3(15):6364–6372. doi: 10.1039/D2MA00542E. [DOI] [Google Scholar]
- [63].Gao K., Li H., Meng Q., Wu J., Hou H. Efficient and selective visible-light-driven oxidative coupling of amines to imines in air over CdS@Zr-MOFs. ACS Appl. Mater. Interfaces . 2021;13(2):2779–2787. doi: 10.1021/acsami.0c21007. [DOI] [PubMed] [Google Scholar]
- [64].Yang Z., et al. Engineering the exciton dissociation in quantum-confined 2D CsPbBr3 nanosheet films. Adv. Funct. Mater. . 2018;28(14):1705908. doi: 10.1002/adfm.201705908. [DOI] [Google Scholar]
- [65].Han Y., et al. Facile synthesis of morphology and size-controlled zirconium metal–organic framework UiO-66: the role of hydrofluoric acid in crystallization – CrystEngComm (RSC Publishing) . Apr;15:2025. https://pubs.rsc.org/en/content/articlelanding/2015/ce/c5ce00729a [Online]. Available at: accessed: [Google Scholar]
- [66].Shen C.-H., Chen Y.-H., Wang Y.-C., Chang T.-E., Chen Y.-L., Kung C.-W. Probing the electronic and ionic transport in topologically distinct redox-active metal–organic frameworks in aqueous electrolytes. Phys. Chem. Chem. Phys. . 2022;24(17):9855–9865. doi: 10.1039/D2CP00117A. [DOI] [PubMed] [Google Scholar]
- [67].Zhihai W., et al. Air-stable all-inorganic perovskite quantum dot inks for multicolor patterns and white LEDs. J. Mater. Sci. 2019;54(9):6917–6929. doi: 10.1007/s10853-019-03382-2. [DOI] [Google Scholar]
- [68].Zhang C., et al. Exciton photoluminescence of CsPbBr3@SiO2 quantum dots and its application as a phosphor material in light-emitting devices. Opt. Mater. Express . 2020;10(4):1007–1017. doi: 10.1364/OME.389847. [DOI] [Google Scholar]
- [69].Lee J., Koteles E. S., Vassell M. O. Luminescence linewidths of excitons in GaAs quantum wells below 150 K. Phys. Rev. B . 1986;33(8):5512–5516. doi: 10.1103/PhysRevB.33.5512. [DOI] [PubMed] [Google Scholar]
- [70].Wei K., Xu Z., Chen R., Zheng X., Cheng X., Jiang T. Temperature-dependent excitonic photoluminescence excited by two-photon absorption in perovskite CsPbBr3 quantum dots. Opt. Lett. . 2016;41(16):3821–3824. doi: 10.1364/OL.41.003821. [DOI] [PubMed] [Google Scholar]
- [71].Wright A. D., et al. Electron–phonon coupling in hybrid lead halide perovskites. Nat. Commun. 2016;7(1):11755. doi: 10.1038/ncomms11755. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [72].Zhou X., Zhang Z. Electron–phonon coupling in CsPbBr3 . AIP Adv. . 2020;10(12):125015. doi: 10.1063/5.0017149. [DOI] [Google Scholar]
- [73].Haug H., Koch S. W. Quantum Theory of the Optical and Electronic Properties of Semiconductors . Singapore: World Scientific; 2009. [Google Scholar]
- [74].Snoke D. W. Solid State Physics: Essential Concepts . Cambridge, UK: Cambridge University Press; 2020. [Google Scholar]
- [75].Li S., et al. Water-resistant perovskite nanodots enable robust two-photon lasing in aqueous environment. Nat. Commun. . 2020;11(1):1192. doi: 10.1038/s41467-020-15016-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [76].Dang N. H. M., et al. Tailoring dispersion of room-temperature exciton-polaritons with perovskite-based subwavelength metasurfaces. Nano Lett. . 2020;20(3):2113–2119. doi: 10.1021/acs.nanolett.0c00125. [DOI] [PubMed] [Google Scholar]
- [77].Huang C.-Y., et al. CsPbBr3 perovskite quantum dot vertical cavity lasers with low threshold and high stability. ACS Photonics . 2017;4(9):2281–2289. doi: 10.1021/acsphotonics.7b00520. [DOI] [Google Scholar]
- [78].Guo B., et al. Ultrastable near-infrared perovskite light-emitting diodes. Nat. Photonics . 2022;16(9):637–643. doi: 10.1038/s41566-022-01046-3. [DOI] [Google Scholar]
- [79].Jiang Y., et al. Synthesis-on-substrate of quantum dot solids. Nature . 2022;612(7941):679–684. doi: 10.1038/s41586-022-05486-3. [DOI] [PubMed] [Google Scholar]
- [80].Xing G., et al. Low-temperature solution-processed wavelength-tunable perovskites for lasing. Nat. Mater. . 2014;13(5):476–480. doi: 10.1038/nmat3911. [DOI] [PubMed] [Google Scholar]
- [81].Uhd Jepsen P., et al. Ultrafast carrier trapping in microcrystalline silicon observed in optical pump–terahertz probe measurements. Appl. Phys. Lett. . 2001;79(9):1291–1293. doi: 10.1063/1.1394953. [DOI] [Google Scholar]
- [82].Xing G., et al. Charge transfer dynamics in Cu-doped ZnO nanowires. Appl. Phys. Lett. . 2011;98(10):102105. doi: 10.1063/1.3558912. [DOI] [Google Scholar]
- [83].Xing G., et al. Long-range balanced electron- and hole-transport lengths in organic-inorganic CH3NH3PbI3 . Science . 2013;342(6156):344–347. doi: 10.1126/science.1243167. [DOI] [PubMed] [Google Scholar]
- [84].Cui L., et al. Multifunctional interface modifier ammonium silicofluoride for efficient and stable all-inorganic CsPbBr3 perovskite solar cells. Chem. Eng. J. . 2022;431(1):134193. doi: 10.1016/j.cej.2021.134193. [DOI] [Google Scholar]
- [85].Ray S., Ranjan Sahoo M., Mukherjee S., Perumal A., Nayak S. K., Bhaumik S. Understanding the charge transfer mechanism in CsPbBr3 nanocrystals and nitrogen-doped carbon quantum dot heterostructures: effect of nanocrystal encapsulation. RSC Adv. . 2023;13(50):35551–35561. doi: 10.1039/D3RA06307K. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [86].Saba M., et al. Correlated electron–hole plasma in organometal perovskites. Nat. Commun. . 2014;5(1):5049. doi: 10.1038/ncomms6049. [DOI] [PubMed] [Google Scholar]
- [87].Baranowski M., Plochocka P. Excitons in metal-halide perovskites. Adv. Energy Mater. . 2020;10(26):1903659. doi: 10.1002/aenm.201903659. [DOI] [Google Scholar]
- [88].Yen M.-C., et al. Tamm-plasmon exciton-polaritons in single-monolayered CsPbBr3 quantum dots at room temperature. Adv. Opt. Mater. . 2023;11(4):2202326. doi: 10.1002/adom.202202326. [DOI] [Google Scholar]
- [89].Liu X., et al. Strong light–matter coupling in two-dimensional atomic crystals. Nat. Photonics . 2015;9(1):30–34. doi: 10.1038/nphoton.2014.304. [DOI] [Google Scholar]
- [90].Deng H., et al. Exciton-polariton Bose-Einstein condensation | Rev. Mod. Phys. . Dec;25:2024. https://journals.aps.org/rmp/abstract/10.1103/RevModPhys.82.1489 [Online]. Available at: [Google Scholar]
- [91].Howarth A. J., et al. Chemical, thermal and mechanical stabilities of metal–organic frameworks | Nature Reviews Materials. . Apr;15:2025. https://www.nature.com/articles/natrevmats201518 [Online]. Available at: accessed: [Google Scholar]
- [92].Tsai M.-D., Wu K.-C., Kung C.-W. Zirconium-based metal–organic frameworks and their roles in electrocatalysis. Chem. Commun. . 2024;60(64):8360–8374. doi: 10.1039/D4CC02793K. [DOI] [PubMed] [Google Scholar]
- [93].Sun Y., et al. Bright and stable perovskite light-emitting diodes in the near-infrared range. Nature . 2023;615(7954):830–835. doi: 10.1038/s41586-023-05792-4. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Supplementary Material Details