Abstract
Doping magnetic transition metal ions (e.g., Mn2+) into colloidal quantum dots endows novel optical and magnetic properties to the host materials. CsPbBr3 quantum dots (QDs) are emerging light-emitting materials with high structural and chemical flexibility in the visible spectral regime. However, efficiently doping Mn2+ ions in CsPbBr3 QDs remains challenging, especially when size confinement and ensemble uniformity are needed for understanding the underexplored exciton-dopant exchange interaction. Here, we introduce a doping mechanism based on electrostatic surface Mn2+ adsorption that enables efficient Mn2+ incorporation in strongly confined CsPbBr3 QDs. The resultant QDs are found to have a Cs-deficient stoichiometry compared to their undoped counterparts. A redox reaction-based purification method was developed to remove Mn2+ cations that are tightly adsorbed on the surface to determine the concentration of lattice-incorporated Mn2+. Our synthesis enables a Mn2+ doping/alloying concentration of up to ∼44% with a Mn2+ photoluminescence efficiency exceeding 90%. This allows for the determination of the intrinsic exciton-to-dopant energy transfer rate.


Introduction
Incorporating impurities such as manganese(II) in colloidal quantum dots (QDs) has been demonstrated as a versatile way to impart new optical and magnetic properties to the host material. − Over the past decade, lead halide perovskite QDs have been explored as a new family of host materials for their highly efficient photoluminescence (PL) and facile synthesis. − In the all-inorganic CsPbX3 (X = Cl, Br, I) family, Mn2+-doped CsPbCl3 QDs were first demonstrated. , Although CsPbCl3 QDs often have low photoluminescence quantum yield (PLQY), benefiting from the fast exciton-to-Mn Dexter-type energy transfer, Mn2+-doped CsPbCl3 nanocrystals exhibit intense and broad emission from 4T1g to 6A1g d-d transitions rather than weak and sharp emission from excitons. − The improved emission efficiency in the visible spectral range has enabled the application of CsPbCl3 for solar concentrators and down-converters. −
Compared to CsPbCl3, the band gap of CsPbBr3 falls into the visible spectral regime and is more suitable for solar energy harvesting and light-emitting applications. However, direct doping of Mn2+ ions into CsPbBr3 perovskite nanocrystals has been surprisingly more challenging than CsPbCl3 nanocrystals. Alternatively, Mn2+ doping in mixed-halide (Cl/Br) perovskite nanocrystals can be obtained either by introducing multiple halides in precursors during the synthesis or by converting presynthesized Mn-doped CsPbCl3 into CsPb(Br/Cl)3 through postsynthesis Br– exchange. Such approaches often result in the loss of Mn2+ dopants and their PL emissions when the Br– composition increases. Recent studies have also revealed that Mn–Cl bonds, rather than Mn–Br bonds, are preferred in mixed-halide nanocrystals. The uncertainties in material composition and the chemical environment of Mn2+ add additional barriers to understanding the Mn incorporation mechanisms and exciton-Mn interaction in pure CsPbBr3 lattices.
Another method to introduce Mn2+ into the host lattices is cation exchange. While the approach has not been very fruitful for 0D CsPbBr3 QDs, strong Mn2+ PL emission can be obtained by cation-exchange in anisotropic CsPbBr3 nanoplatelets (NPLs), nanowires (NWs), and 2D nanoclusters. − The large surface area and surface defects facilitate the Mn2+ attachment and the diffusion of Mn2+ into the lattices. In addition, confinement effects due to reduced dimensionality increase the bandgap, promoting exciton-to-dopant energy transfer for enhanced Mn emission. However, such postsynthesis treatments often yield significant variations in particle lateral sizes, which can be intensified by Ostwald ripening during the treatment. Additionally, the Mn2+ dopant can still migrate or be expelled from the nanoparticles. It is worth noting that the one-pot synthesis of Mn-doped low-dimensional nanostructures at room temperature has also been demonstrated. ,− Despite the successes of Mn2+ doping in these anisotropic nanostructures, the lack of ensemble uniformity and uncertainties on dopant distribution in each particle have hindered the understanding of exciton-dopant exchange interaction from the ensemble-averaged optical properties.
Direct hot-injection synthesis of 0D Mn-doped CsPbBr3 QDs usually produces weakly quantum-confined or irregularly shaped nanoparticles. The relatively small host bandgap leads to a low Mn2+ emission efficiency. The bandgap of CsPbBr3 QDs can be expanded to ∼ 2.7 eV by reducing the QD size. − Furthermore, strong 3D confinement forces the spatial overlap between the wave functions of the Mn dopants and the exciton, promoting Mn2+ PL efficiencies. , Unfortunately, there have been limited attempts to directly dope Mn2+ in strongly confined CsPbBr3 QDs adopting hot-injection synthesis for high ensemble uniformity. A pioneering study suggests that the high Mn–Br bond dissociation energy makes the decomposition of Mn–Br precursors difficult, subsequently reducing doping efficiency. In addition, incorporating Mn2+, a hard Lewis acid, into QD lattices composed of soft Lewis acid and base (Pb2+ and Br–) will be less preferred than doping Mn2+ in CsPbCl3 with a hard Lewis base (Cl–). The thermodynamics of Mn-X bonds have certainly affected the ability to synthesize Mn2+-doped CsPbBr3 QDs. To date, improved Mn2+ PL emission in directly hot-injection synthesized Mn2+-doped CsPbBr3 QDs (∼6.5 nm) has been demonstrated as a side product through the conversion of intermediate Mn2+-doped 2D L2PbBr4 plates into Mn2+-doped QDs and NPLs. Most recently, room temperature Mn2+ doping in 5 nm CsPbBr3 QDs has also been demonstrated. Nevertheless, the size regulation effect and Mn2+ incorporation efficiency or dopant density are still limited. Therefore, a generalized doping method is needed to produce monodispersed QDs with high doping efficiency in order to understand the photophysics of exciton-Mn interaction in CsPbBr3 QDs.
In this work, we developed a synthesis under a bromide-rich environment with high Mn2+ ionic strength for efficient Mn2+ doping in size-confined CsPbBr3 QDs. The resulting Mn2+-doped CsPbBr3 QDs (∼4 nm) exhibit efficient (90 ± 10% PLQY) Mn2+ emission with a doping/alloying concentration up to ∼44%. These Mn2+-doped QDs are highly nonstoichiometric, featuring Cs-deficient regions near the QD surface. The Mn2+ doping efficiency increases with the extent of the Cs deficiency. Nuclear magnetic resonance (NMR) spectroscopy and elemental analysis suggest that a large quantity of Mn2+ ions is tightly adsorbed on the QD surface, which can be thoroughly removed by chemical redox-reaction-based purification using H2O2 and HBr without affecting the Mn2+ ions incorporated in the QD lattices. Finally, the exciton-to-dopant energy transfer rate is measured by transient absorption (TA) spectroscopy. Our study provides a facile way to dope Mn2+ in perovskite QDs, providing new insights to apply perovskite QDs to light-harvesting, spintronics, and hot electron production applications.
Results
Mn2+-doped CsPbBr3 QDs were prepared using manganese(II) acetate tetrahydrate as the Mn2+ precursor in a one-pot synthesis (details in Experimental Section). The bromide-rich environment is created by adding hydrobromic acid (HBr aqueous solution) to the system, leading to a bromide-terminated QD surface. , HBr also reacts with Mn acetate tetrahydrate (Mn(Ac)2·4H2O) to facilitate the decomposition of Mn2+ precursors by producing acetic acid that can be removed from the system after extended evacuation. The Mn2+ incorporation efficiency can be tuned by solely varying the loading amount of Mn(Ac)2·4H2O/HBr (more details of the synthesis conditions and their effects on Mn2+ doping are provided in the Discussion section, Supporting Information Table S1, Supporting Information Note 1, and Figure S1).
Absorption and PL spectra of the Mn2+-doped and undoped (control) QDs at room temperature are plotted in Figure a and b. The overall PLQY of Mn2+-doped CsPbBr3 QDs (PL emissions from both exciton and Mn) is generally above 90% (Figure S2a), and the size of both Mn2+-doped and undoped QDs is ∼4 nm (Figure c and d, Figure S3) with a tight size distribution (±7.4% and ± 6%, respectively). Compared to undoped QDs, Mn2+-doped QDs tend to be more irregularly shaped. The Mn2+-doped QDs show an almost completely quenched exciton PL. By integrating the PL peaks from Mn2+ (∼605 nm), the highest Mn2+ emission PLQY is determined to be >90%. The exciton PL exhibits a blueshift of ∼50 meV in doped QDs compared to undoped QDs sharing similar sizes (Figures S4 and S5). This phenomenon has been observed in II–VI QDs and lead halide perovskite NCs, which can be attributed to the possible increase in bandgap resulting from Mn2+ alloying and local lattice periodicity breaking. − Additionally, Mn2+incorporation significantly blueshifts (∼186 meV) the absorption spectrum of the QDs (see Figure a) and changes the profile of higher-order exciton absorptions (black vertical arrows), suggesting that Mn2+ alloying may have resulted from efficient dopant incorporation. No anisotropic nanostructures, such as NPLs and NWs, are produced during the synthesis, implying a doping mechanism different from the previously reported monolayer perovskite-mediated doping.
1.
(a) Absorption spectra of Mn2+-doped and undoped CsPbBr3 QDs measured at room temperature. Vertical arrows denote higher-order exciton peaks. (b) PL spectra of Mn2+-doped and undoped CsPbBr3 QDs (Room temperature, excited using a 385 nm LED). (c) High-angle Annular Dark-field Scanning Transmission Electron Microscopy (HAADF-STEM) images of Mn2+-doped CsPbBr3 QDs (purified by the GPC-chemical-GPC approach, as discussed in the main text and details given in the Experimental Section), and (d) undoped CsPbBr3 QDs.
Inductively coupled plasma mass spectrometry (ICP-MS) elemental analysis is typically used to determine the chemical composition and doping concentration of QDs. Accurate determination of chemical composition in perovskite QDs is usually challenging due to the inefficient removal of unreacted precursors. Following previous reports, we purified the Mn2+-doped QD colloids using gel permeation chromatography (GPC) columns , after traditional antisolvent precipitation/resuspension cycles (Figure S6). The cationic composition (atomic ratios, normalized to Pb2+) of QDs is then obtained from ICP-MS (Supporting Information Table S2). The Br/Pb ratio of QD was determined using energy-dispersive X-ray spectroscopy (EDS) (Figure S7 and S8) and X-ray photoelectron spectroscopy (XPS). The chemical composition (not including Mn2+) is Cs0.58–0.66PbMn x Br3.5–4.8 for doped QD and Cs0.73–0.9PbBr3.2–4 for undoped QDs. The bromide-rich QD composition in both QDs is expected since an anion-rich synthesis environment is used.
The Cs/Pb ratio of Mn2+-doped QDs with various Mn2+ PLQYs and undoped QDs is plotted in Figure a. The Cs/Pb stoichiometry of undoped QDs (0.7–0.9) matches that of the reported model structure with the QD’s surface terminated by [PbBr2/ABr] (A = cationic ligand) (Supporting Information Note 2). In this model, surface undercoordinated Cs+ ions were replaced by cationic ligands, such as oleylammonium, resulting in a Cs/Pb atomic ratio of less than one as the QD size decreases. The oleylammonium bromide surface passivation is supported by the NMR spectrum (Figure S9). In contrast, the Cs/Pb ratio in Mn2+-doped QDs is consistently reduced to approximately 0.6 (as low as 0.55) with increased Mn2+ doping, indicating a Cs deficiency in the QD lattices. Note that the sizes of all Mn2+-doped QDs with high Mn PLQYs (>40%) are similar (4–4.5 nm). Therefore, the size effect on the Cs/Pb ratio in these QDs is negligible. For doped QDs with lower Mn PLQY (<30%), the size can be slightly larger, which may contribute to a slightly higher Cs/Pb ratio. It is also worth noting that a considerable number of Pb2+ ions are replaced by Mn2+ in heavily doped QDs. Therefore, the actual Cs+ loss can be underestimated in heavily doped QDs by calculating the Cs/Pb atomic ratios.
2.
(a) Mn2+ PLQY of Mn2+-doped and undoped CsPbBr3 QDs with respect to the Cs/Pb ratio. Decreasing the Cs/Pb atomic ratio enhances the nonstoichiometry and Mn2+ emission (All QDs have overall PLQYs > 70%). Additional photographs are provided in Figure S11. (b) Schematic representation of the distribution of Mn2+ ions and ligands in relation to the amount of Cs. (c) Zeta-potential (ζ-potential) and (d) Dynamic light scattering (DLS) intensity-based hydrodynamic size distribution (nm) for Mn2+-doped and undoped CsPbBr3 QD colloids (purified by antisolvent only). DLS correlation functions are provided in Supporting Information Figure S12. (e) Expanded region of 1H NMR spectrum of Mn2+-doped and undoped CsPbBr3 QDs in chloroform-d (CDCl3). (f) 133Cs NMR signals for Mn2+-doped CsPbBr3 QDs and their fits showing core (green) and intermediate (purple) Cs+ species, alongside undoped CsPbBr3 QDs display core (green), intermediate (purple), and surface (yellow) Cs+ species.
Elemental analyses also showed that the Mn/Pb atomic ratio is 10–50 in almost all Mn2+-doped QD samples, even after GPC purification. Given the sophisticated purification processes, such an unexpectedly high Mn2+ concentration in QD colloids cannot be simply attributed to free-standing Mn2+ precursors. Instead, it suggests that many Mn2+ ions are tightly attached to QD surfaces, and some of them are probably bound to the surface as Z-type ligands. The extremely large excess Mn2+ concentration (up to ∼17,000 Mn2+ per QD) is beyond the capacity of typical surface ligand coverage, suggesting the majority of the excess Mn2+ ions are physically adsorbed on the surface of the QD. EDS and XPS measurements also confirm the large Mn/Pb ratio in the doped QDs (Figures S8 and S10). Additionally, the XPS results indicate that the doped QD sample contains a mixture of long-chain hydrocarbon moieties (oleylammoniums and oleates), with their quantity larger than that of the undoped QDs. This suggests that many excess Mn2+ ions exist as Mn carboxylates, such as Mn-oleates. As illustrated in Figure b, an electric double layer structure is proposed, in which QDs are charged due to the nonstoichiometry, and Mn2+ ions can electrostatically adsorb on the charged QD surface, forming a diffusion layer that balances the charge on the QDs and stabilizes the QD colloids.
The electric double layer structure is evidenced by a ζ-potential of ∼21 mV in Mn2+-doped QD colloids (Figure c). The positive ζ-potential provides further evidence that many surface-physiosorbed Mn2+ ions are firmly attached to QDs and cannot be easily removed by conventional purification methods. In comparison, the undoped QDs have a negligible ζ-potential value (Figure c). To better study the colloidal structure of our Mn2+-doped QDs, we measured their hydrodynamic diameters (D h) using dynamic light scattering (DLS). For undoped QDs, D h is slightly smaller but close to the physical size of QD-ligands (∼4 nm, Figure d). The difference between D h and the size measured using STEM imaging is attributed to the uncertainties in refractive indices when estimating the size distribution using the correlation function shown in Figure S12. In stark contrast, the D h of Mn2+-doped QDs is ∼15 nm with a small population at ∼130 nm. The increased D h is attributed to electric double layers and QD aggregations induced by the QD surface charges and lack of direct organic ligand passivation in Mn2+-doped QDs. STEM images also confirm that the Mn2+-doped QDs can aggregate (Figure S13). The surface Mn2+ adsorption on QDs plays an important role in facilitating Mn2+ incorporation in CsPbBr3 lattices, given that perovskite QD surface ions are often under dynamic solubility equilibria. Details of the proposed mechanism for efficient Mn2+ doping are described in the Discussion section.
1H NMR spectroscopy was used to study the chemistry of Mn2+-doped QD surfaces. Figure e (blue) shows that undoped QDs are capped by alkylammonium cations, which exhibit broadened peaks (indicated by the dashed vertical lines) at 3.34 ppm (alpha protons) and 6.87 ppm (ammonium protons) from oleylammoniums, in good agreement with previous studies. − However, the characteristic peak of alkylammonium alpha-protons nearly disappears in the 1H NMR spectrum of Mn2+-doped QDs. Additionally, the peak from the alkenyl protons at 5.4 ppm in oleate anions and oleylammonium cations is significantly broadened (Figure e (red)). The line broadening is attributed to the effect of the paramagnetic Mn2+ ions adsorbed on the QD’s surface, forming ion pairs with organic oleates. The oleates are counterions for surface-adsorbed Mn2+, imparting the colloidal stability of doped QDs in organic solvents such as toluene and hexanes. The interaction of the paramagnetic Mn2+ ions with surrounding molecules shortens the spin–spin relaxation time (T2), which in turn leads to the broadening of the peaks (given that the line width is inversely proportional to T2 ( ). The NMR results agree well with our QD surface structural model.
The nonstoichiometry of Mn2+-doped QDs is further supported by the 133Cs NMR spectrum (Figure f, top). In the undoped QD (Figure f, bottom), the 133Cs signal is broadened because Cs+ ions are distributed in different regions in the QDs with various lattice disordering. The spectrum can be fitted with a minimum of three Gaussian peaks, corresponding to three groups of Cs+ ions depending on their locations in the QDs and on the degree of lattice disorder: the lattices in the core (least disordered), the lattices on or very close to the surface (most disordered), and the lattices in the intermediate region. This is in good agreement with a previous study using well-passivated CsPbBr3 QDs. The 133Cs signal of the Mn2+-doped QDs is much narrower (Figure f, top) due to, in particular, the absence of the downfield (yellow-shaded) component associated with the Cs+ ions located in the surface lattices. This suggests that the nonstoichiometric Mn2+-doped QD has Cs deficient lattices on and near the surface, promoting surface Mn2+ adsorption. The 1H and 133Cs NMR spectra of Mn2+-doped QDs with lower Mn2+ PLQY show less peak narrowing and a surface Cs component smaller than undoped QDs (Figure S14). The line width changes observed in the 133Cs NMR spectrum corroborate the correlation between the extent of Cs deficiency and Mn2+ doping efficiency.
Excess Mn2+ ions in the QD colloid must be removed to accurately quantify the Mn2+ incorporation. Since the surface-adsorbed Mn2+ is not incorporated into the crystal lattices, we employed a chemical purification method by oxidizing Mn2+ using H2O2 and removing the generated MnO x using hydrobromic acid (Details in the Experimental Section). Figure a shows the absorption and PL spectra of a Mn2+-doped QD sample before chemical purification, after H2O2 oxidation, and after HBr treatment. No noticeable spectral shifts or Mn2+ PL intensity changes are found, indicating that the surface-adsorbed Mn2+ ions are not emissive, and the lattice incorporated Mn2+ ions are not oxidized. However, after H2O2 treatment, the QD solution becomes darker in color (Figure b) due to the generation of MnO x , which tends to be detached from the QD surface. HBr aqueous solution was then added to the QD solution to reduce MnO x and extract the produced Mn2+ into the aqueous phase. After cleaning, the QD solution returns to its light-yellow color. It is worth noting that extensive iterations of H2O2/HBr purification will decrease the Mn2+ emission intensity and compromise the QD colloidal stability (Figure S15). After chemical purification, the QDs were purified again using GPC to remove any possible free-standing Mn2+.
3.
(a) Absorption and PL spectra of Mn2+-doped QD and surface purification with H2O2 and HBr. The intensity of the PL spectra in (a) is normalized with respect to the samples’ absorbance values at 385 nm. (b) Schematic diagrams illustrating Mn2+-doped QDs before, during the oxidation of Mn2+, and after the removal of surface-adsorbed Mn2+ ions, accompanied by photographs taken in daylight (left column) and under UV light (right column). (c) XRD patterns of the Mn2+-doped (red curve) and undoped (blue curve) CsPbBr3 QDs. A noticeable shift of the (200) XRD peak to higher 2θ values results from lattice contraction as Mn2+ ion concentration increases, expected from substituting larger Pb2+ ions with smaller Mn2+ ions. (d) EPR spectra of Mn2+-doped CsPbBr3 QDs before and after GPC-chemical-GPC purifications at room temperature.
Electron paramagnetic resonance (EPR) spectroscopy was also employed to investigate the different chemical environments of Mn2+ ions in the Mn2+-doped QD colloids (Figure d). The EPR spectrum of Mn2+-doped QD colloids after only antisolvent purifications exhibits strong EPR signal with a full width at half-maximum (fwhm) of ∼235 G (Figure d, red curve). No hyperfine structure was observed. Based on the elemental analyses above, the EPR response of Mn2+-doped QD samples before purification is likely dominated by the surface-adsorbed Mn2+. In fact, a very similar EPR spectrum is obtained from a Mn-oleate solution (Figure S16), also aligning with a previous report. Interestingly, after GPC-chemical-GPC purification, the EPR signal of the Mn2+-doped QD colloids almost vanished (Figure d, green curve). The absence or weakness of EPR responses from lattice-incorporated Mn2+ in Mn2+-doped CsPbBr3 NCs was attributed to strong antiferromagnetic coupling in linear Mn–Br–Mn bonds. Given the high lattice incorporated Mn2+ density in our Mn2+-doped QDs, it is reasonable that the average distance between Mn2+ ions is very small, leading to very broad or nearly nondetectable EPR signals. In lightly doped QDs where dopants are more separated, weak and broad EPR responses with a fwhm of ∼600 G were found after decoupling the sharp EPR signals of surface-adsorbed Mn2+ from the overall EPR responses (this sample is not chemically purified due to the low concentration of surface-adsorbed Mn2+, which makes it unstable in harsh chemical environments, Figure S17). In addition, we note that similar broad EPR signals were reported for 0.5% and 3% Mn-doped CsPbI3 QD samples, which were also attributed to substantial antiferromagnetic coupling of the Mn2+ ions.
To test the effectiveness of physiosorbed Mn2+ removal through the GPC-chemical-GPC purification, QD samples synthesized using the highest Mn-precursor loading ratio were used for low-temperature EPR analysis (Figure S18). After the GPC-chemical-GPC purification, QDs exhibit continuously undetectable EPR responses even at low temperatures (120 K). In contrast, similar QDs, only after antisolvent precipitation-resuspension cycles, show strong EPR signals at both room and low temperatures (Figure S18). It has also been reported that reasonably solvated Mn2+ species should show readily detectable EPR signals at low temperatures. The EPR results strongly suggest the thorough removal of surface-adsorbed Mn2+ species. This is also corroborated by ICP-MS analysis of GPC-chemical-GPC purified QDs, which shows a significantly decreased Mn/Pb atomic ratio of less than 1 compared to 40–45 before the GPC-chemical-GPC purification. The Cs/Pb ratio remained almost unchanged (0.58), indicating that the structure of the QDs remains nearly intact after purification. Furthermore, we have quantified the spin concentration in the QD samples before GPC-chemical-GPC purification (Figure S19). The resultant 2.9 × 1020 spins/mL corresponded to a Mn2+ concentration that matches very well with the concentration of removed Mn2+ in this sample determined by ICP-MS (details in Supporting Information Note 3). These results corroborated the conclusion that almost all spins in the EPR spectrum came from surface physiosorbed Mn2+, which were nearly all removed by the GPC-chemical-GPC purification.
Knowing that the surface physiosorbed Mn2+ can be fully removed, a correlation between Mn2+ PLQY and doping concentration in Mn2+-doped QDs was established by performing ICP-MS analyses on a series of Mn2+-doped QD samples purified by the GPC-chemical-GPC cycles (Figure S20). The doping/alloying concentrations of QDs, assuming only Pb2+ ions were replaced by Mn2+, range from ∼6.5% to ∼44%. The Mn2+ PLQY increases sublinearly with the doping concentration, implying Mn2+ dopants are not optically equivalent when the doping/alloying concentration is high. The efficient incorporation of Mn2+ in CsPbBr3 QDs was also confirmed using X-ray diffraction (XRD) measurements. As shown in Figure c, the (200) reflection from the XRD pattern of Mn2+-doped QDs experienced a 0.60° shift to higher angles due to the lattice constraint induced by replacing Pb2+ with smaller Mn2+. According to Vegard’s law (Supporting Information Note 4), the Mn2+ doping concentration is estimated to be 19%, falling into the range of the estimated doping concentration from the ICP-MS analysis.
Surface-adsorbed Mn2+ plays an important role in the luminescence stability of QDs. Studies show that lattice-incorporated Mn2+ tends to diffuse to the surface and leave the crystal. The PL spectra of an Mn2+-doped QD colloid (Figure a) show virtually no intensity drops after five iterations of precipitation-resuspension purification cycles using methyl acetate as the antisolvent. In comparison, an undoped QD colloid experiences a PLQY efficiency drop from ∼80% to <20% after two iterations of purification (Figure S21a). Additionally, the Mn2+-doped QD showed no noticeable PLQY decrease over 120 days of storage (Figure S21b). The outstanding Mn2+ PL stability can be attributed to the protection of surface Mn2+ adsorption, which suppresses the diffusion loss of doped Mn2+. Indeed, after GPC-chemical-GPC purification, Mn2+-doped QDs exhibit reduced but adequate Mn2+ PL stability (Figure S22a and S22b). To better reveal the function of surface-adsorbed Mn2+ and the stability of lattice-incorporated Mn2+ ions, we have studied the thermal stability of Mn2+ PL emission using QDs before and after GPC-chemical-GPC purifications. Specifically, the PL spectra of QD colloids were monitored over time at 80 °C (Figure S22c and S22d). A QD solution with surface-adsorbed Mn2+ exhibits an ∼20% Mn2+ PL intensity drop at 600 nm, whereas purified QD experienced a ∼50% Mn2+ PL intensity drop. While the PL drop can be partially attributed to heat-induced nonradiative processes that compete with exciton-to-dopant ET, the additional PL loss in the purified sample can be explained by the loss of Mn2+, presumably located in the surface lattices (or close to) of the QDs. Interestingly, the Mn2+ PL intensity drop of purified QDs became very slow after 5 min, indicating that some Mn2+ ions incorporated in QD lattices are more stable than others. The two categories of Mn2+ dopants coincide with the two different types of lattices in QDs: surface lattices with Cs deficiency and center/intermediate lattices with normal stoichiometry. Further thoughts on the potential correlations between Mn2+ doping locations and their chemical environments are given in the Discussion section.
4.
(a) PL spectra of Mn2+-doped QDs after 1×, 2×, and 5× antisolvent precipitation-dissolution iterations using methyl acetate as the antisolvent. (b) Absorption (dotted red curve) and PLE spectra of Mn2+-doped QD were monitored at 582 nm (green curve) and 473 nm (blue curve), respectively. (c) The decay curve of exciton emission for Mn2+-doped CsPbBr3 QDs. (d) Phosphorescence decay curves recorded at 600 nm show the Mn2+ emission from Mn2+-doped CsPbBr3 QDs (Mn2+ PLQY ∼75%). The solid lines in (c) and (d) are fits using the biexponential decay functions. (e) 2D contour plot of TA spectra of Mn2+-doped CsPbBr3 QDs with pump intensities of ∼18.8 μJ/cm2. (f) Bleach recovery dynamics in TA spectra of undoped (blue, monitored at 470 nm) and Mn2+-doped (orange, monitored at 450 nm) CsPbBr3 QDs.
Photoluminescence excitation (PLE) spectra obtained at two different wavelengths (473 and 582 nm) from the exciton and Mn2+ PL overlap with the absorption spectrum of Mn2+-doped QDs, suggesting all Mn2+ emission bands are sourced from excitons in QDs (Figure b). To understand the recombination dynamics of excitons and Mn2+ dopants in CsPbBr3 QDs, we performed time-resolved PL intensity measurements. The PL decay trace of the exciton emission can be fitted by a biexponential decay function with a ∼370 ps fast component (convoluted with the instrument response function) and a 5.9 ns slow component (Figure c). Given the high PLQY of Mn2+-doped QDs, we attribute the fast decay to energy transfer (ET) or possible electron transfer, and the slower decay to exciton radiative recombination from a small fraction of undoped QDs or QDs with insufficient dopant incorporation in the ensemble. The temporal evolution of the Mn2+ emission intensity is plotted in Figure d. Considering the two categories of Mn2+ in lattices near the surface and in the core, the Mn2+ PL decay curve was fit by a biexponential decay function with time constants of 188 and 386 μs (the fitting parameters and a single exponential fit are provided in Figure S23 and Table S3). Such a biexponential PL decay is more evident in QDs with lower doping levels (Figure S24). The two decay components are tentatively attributed to Mn2+ dopants in two regions with different degrees of interdopant interaction. In comparison, the Mn2+ emission lifetime can be as long as ∼2 ms and 300–400 μs when the dopants are diluted in methylammonium (MA) lead bromide single crystals and CsPbBr3 NPLs, respectively. ,, Given that the heavy atom effect should be similar in these two crystals, the shortened lifetime can be attributed to the Mn–Mn couplings. Indeed, reducing the Mn2+ doping level leads to an overall slower Mn2+ PL decay (Figure S24). However, two decay components can still be observed, indicating that the two possible categories of Mn2+ local environments are likely inherited from the doping mechanism.
Although reported for Mn2+-doped CsPbCl3 NCs, intrinsic ET rates remain elusive in 0D Mn2+-doped CsPbBr3 QDs. TA spectroscopy was used to determine the ET rate in Mn2+-doped QDs. Figure e and f show the TA spectra and the bleach recovery dynamic traces monitored at the peak position for undoped and Mn2+-doped CsPbBr3 QDs, respectively. Both samples exhibit a PLQY greater than 75%, indicating that the exciton trapping process has a minimal impact on the bleach recovery dynamics. The fast bleach recovery in Mn2+-doped QDs is therefore attributed to the exciton-to-dopant ET. It is worth noting that possible exciton-to-dopant charge transfer (CT) is not ruled out in this study. CT from excitons to Mn2+ has been discovered in Mn-doped lead halide perovskites. ,, We will, however, focus our discussions on ET since a recent study has revealed that CT is more preferred in lead chloride perovskites. The apparent ET rate is then extracted using reported methods to obtain a time constant of ∼94 ps. Unlike large NCs and 2D NPLs with uncertain degrees of exciton Mn2+ wave function overlapping, all lattice-incorporated Mn2+ are presumably able to couple with the exciton in these strongly confined QDs. From the estimated doping concentration of the sample (103–188 Mn2+ ions per QD, considering uncertainties in PLQY measurements, see Figure S20), the intrinsic exciton-to-Mn ET rate in Mn2+-doped CsPbBr3 is calculated to be 0.06–0.1 ns–1 per Mn2+. This rate, even enhanced by strong quantum confinement, is 30–50 times slower than that found in Mn2+-doped CsPbCl3 and >100 times slower than that in Mn2+-doped CdS and CdSe QDs. ,,
The relatively slow ET rate in Mn2+-doped CsPbBr3 QDs is unlikely to be caused by Mn-to-exciton back ET, given the large difference between the exciton and Mn2+ PL energies (∼630 meV). Although the ET rate can be influenced by the high ionicity of Mn2+ in perovskite lattices and the lower exciton energy, the significantly slower ET rate in Mn2+-doped CsPbBr3 QDs compared to that in Mn2+-doped CsPbCl3 QDs still suggests that the Mn-exciton exchange interaction is relatively weak in Mn2+-doped CsPbBr3 QDs. A previous study also indicates that the Mn-exciton exchange coupling in Mn2+-doped CsPbI3 QDs is unexpectedly weak. Several factors can contribute to the slow average ET rate in CsPbBr3 QDs. First, the potential Mn2+ dopant clustering can also introduce MnBr4 2–-rich domains inside the QD, promoting exciton localizations and decreasing exciton-Mn wave function overlapping. Second, the recently reported , lattice periodicity-breaking effect in Mn2+-doped CsPbBr3 QDs can also localize excitons to nondopant sites. Last but not least, Mn2+ dopants experience different lattice stoichiometries, and they may not share the same exciton-dopant interaction. To provide some insights on this, the ET rate is also determined for Mn2+-doped CsPbBr3 QDs exhibiting ∼53% and ∼7% Mn2+ PLQYs (Figure S25). Interestingly, a faster per Mn2+ ET rate is found in the lightly doped sample. This observation is qualitatively consistent with the observation that the Mn2+ PLQY increases sublinearly with doping concentration (Figure S20), which suggests that Mn2+ in different lattice environments may not exhibit the same ET kinetics. (More details are provided in Supporting Information Note 5.)
Discussion
The doping of QDs is inherently hindered by the “self-purification” mechanism, whereby impurities are repelled toward the surface during nanocrystal growth. Self-purification makes smaller QDs even harder to dope, given that the impurity formation energy increases with increasing quantum confinement. In the case of perovskites, highly ionic lattices lead to large ion migration mobilities. , This will accelerate the exclusion of Mn2+ dopant in perovskites. It has been well-established that the doping efficiency of QDs highly depends on the surface adsorption of dopant ions. To incorporate cationic substitutional impurities, an anion-rich host will facilitate surface adsorption of cationic impurities for enhanced doping efficiency. While demonstrated in conventional II–VI QDs, such a surface-adsorption facilitated doping process has not been well understood in perovskite QDs, presumably due to their dynamic surfaces that are unlikely to stabilize surface-adsorbed ions.
We propose a doping mechanism considering the surface adsorption of Mn2+ and stoichiometry deviations in our Mn2+-doped QDs. As illustrated in Figure , the bromide-rich surface will attract cations near the surface at the early stage of QD growth. Previous studies show that Pb2+ tends to form anionic complexes such as PbBr3 – and PbBr4 2–. , Therefore, Mn2+ and Cs+ are likely the main cationic species in the reaction mixture. When the divalent Mn2+ ions are readily available in large quantities (Mn2+/Cs+ ∼3.6, molar ratio), their surface adsorption is electronically favored, and the Mn2+ incorporation becomes efficient. In contrast, monovalent Cs+ surface adsorption is not favored. Therefore, further growth of the QD will result in Cs-deficient lattices close to the surface of QDs, as evidenced by 133Cs NMR spectroscopy. This Cs deficiency will subsequently encourage the surface physisorption of Mn2+ for charge neutrality. This can also explain why our QDs have firmly adsorbed Mn2+ ions that are >10-fold equivalent to the Pb2+ ions that constitute the QDs.
5.
Diagram showing the suggested growth mechanism of Mn2+-doped CsPbBr3 QDs. Schemes are not in-scale.
To provide some experimental insights into the proposed doping mechanism, we have systematically investigated the effect of changes in synthesis conditions on the structural and optical properties of the obtained Mn2+-doped QDs (Table S1). First, the Mn(Ac)2·4H2O and HBr reaction is crucial for achieving efficient doping. Reducing the amount of Mn(Ac)2·4H2O or withdrawing the 230 °C heating process (Figure S26) will decrease the Mn2+ doping efficiency. This is consistent with the surface adsorption-driven mechanism of Mn2+ incorporation. Second, reducing the bromide concentration (HBr) will lead to low Mn2+ incorporation, which indicates the necessity of bromide-rich conditions. Third, increasing the Cs precursor concentration will not significantly compromise the Mn2+ doping and the degree of Cs deficiency in the Mn2+-doped QDs. This is expected in the proposed electrostatic surface adsorption model since Mn2+ possesses a higher charge density than Cs+. Additionally, reducing the Cs precursor concentration will only decrease the reaction yield of the synthesis but not the Mn2+ incorporation. This insensitivity to the amount of Cs precursors supports the Mn2+ surface adsorption model. Finally, the QD growth can be suppressed by loading additional Mn(Ac)2·4H2O/HBr during synthesis. This allows for the synthesis of Mn2+-doped QDs with a size of only 3 nm (±0.3 nm) (Figure S27). This strongly suggests the efficient surface Mn2+ adsorption, which could cover the surface and suppress the QD growth if additional Mn2+ ions are introduced.
The Mn2+ surface adsorption model suggests that Mn2+ ions incorporated into lattices near the QD surface may experience a different chemical environment. Due to the lack of Cs+, increased lattice distortion is expected, which could affect the Mn coordination environment. This hypothesis can explain the above observation that some Mn2+ ions are less thermally stable than others, as well as the revelation that there are two different Mn2+ PL decay time constants. Given that the nonstoichiometric surface lattices can offer additional flexibility in Mn2+ incorporation, it should be noted that the possibility of Mn2+ doping through occupying Cs+ sites or even interstitial doping cannot be ruled out. In fact, in Figure c, the (110) XRD peak of doped QDs broadens and shifts toward smaller angles compared to that of undoped QDs. This suggests crystal structural changes, such as increased lattice inhomogeneity and lattice expansion induced by possible interstitial dopants. , It is also worth noting that the interstitially doped Mn2+ can also exhibit the typical d-d PL emission. The potential Cs+-substitutional or interstitial Mn2+ doping can lead to a higher apparent dopant concentration estimated by elemental analysis. Finally, this doping strategy can also be applied to CsPbCl3 QDs. Although successful Mn2+ doping has been demonstrated in CsPbCl3 QDs, our method can dope CsPbCl3 QDs more efficiently, manifested as a significantly red-shifted Mn2+ PL band (from ∼609 to 650 nm) due to strong Mn–Mn coupling (Figure S28). In principle, the nonstoichiometric perovskite QDs can lift the barrier of the highly dynamic surface of perovskite QDs for a large variety of impurity doping.
Conclusion
In conclusion, we have demonstrated that size-confined CsPbBr3 QDs can be efficiently doped with Mn2+. The Mn2+-doped QDs exhibit 90 ± 10% Mn2+ emission PLQY and extraordinary PL stability. These doped QDs are nonstoichiometric as they are Cs-deficient. Detailed structural characterization shows that the surface of the doped QDs is covered with a large concentration of Mn2+ that cannot be removed using traditional purification methods such as antisolvent precipitation and GPC column. A chemical redox reaction-based purification is developed to thoroughly remove the surface-adsorbed Mn2+, enabling the detailed quantification of lattice-incorporated Mn2+ in CsPbBr3 QDs. Using QDs with simultaneously high PLQY and quantum confinement, the exciton-to-dopant energy transfer rate was measured for the first time in Mn2+-doped CsPbBr3 QDs. The ET rate of ∼0.06 ns–1 suggests an intrinsic suppressed exchange interaction between the excitons and Mn2+ in CsPbBr3 compared to CsPbCl3. The nonstoichiometric QD facilitated surface adsorption promises facile and efficient doping and alloying cationic impurities into traditionally “almost undopable” CsPbBr3 QDs.
Experimental Section
Materials
The following chemicals were used as received: cesium carbonate (Cs2CO3, puratronic, 99.994% metals basis, Alfa Aesar), lead(II) bromide (PbBr2, puratronic, 99.999% metals basis, Alfa Aesar), manganese acetate tetrahydrate (Mn(CH3COO)2·4H2O, ACROS Organics), oleylamine (OAm, technical grade, 70%, Sigma-Aldrich), oleic acid (OA, technical grade, 90%, Sigma-Aldrich), 1-octadecene (ODE, technical grade, 90%, Sigma-Aldrich), hydrobromic acid (HBr, 48 wt % in water, Acros), hydrogen peroxide solutions (H2O2, 30 wt., ACS grade, Sigma-Aldrich), nitric acid (HNO3, 67–70% w/w, VWR Chemicals), rhodamine 6G perchlorate (99%, Aldrich Chemistry), acetone (certified ACS, Fisher), hexanes (HPLC grade, Millipore), toluene (ACS grade, Fisher), methyl acetate (ReagentPlus, 99%, Sigma-Aldrich) acetonitrile (C2H3N, >99.9%, Sigma-Aldrich) and tetrabutylammonium hexafluorophosphate (TBAPF6, Sigma-Aldrich). Bio-Beads S-X1 GPC medium was sourced from Bio-Rad Laboratories.
Cs-Precursor Preparation
The Cs-oleate precursor was prepared by dissolving 300 mg of Cs2CO3 in a mixture of 1.2 mL of OA and 3.2 mL of ODE in a 50 mL three-necked round-bottom flask. This mixture was degassed under a vacuum at room temperature for 3–5 min while stirring vigorously. The mixture was then heated to 130 °C under vacuum. Following this, the flask was filled with nitrogen and cooled to 120 °C for subsequent use.
Mn2+-Doped CsPbBr3 QD Synthesis
60 mg of PbBr2 (0.16 mmol) and 255 mg of Mn (CH3COO)2·4H2O (Mn(Ac)2·4H2O, 1.04 mmol) were added in a three-neck round-bottom flask, followed with 1 mL of OA, 1 mL of OAm, 5 mL of ODE, and 0.5 mL of HBr. The mixture was heated at 150 °C under vacuum for 40 min. The flask was then filled with nitrogen, and an additional 1 mL of dried OA and 1 mL of dried OAm were injected to solubilize the unreacted solids. To obtain dried OA and OAm, both ligands were dried under vacuum in separate flasks at 150 °C for 15 min. The solution was next heated to 200 °C for 40 min, after which the temperature was increased to 230 °C and held for 5–10 min until stabilized. Then, the reaction mixture was allowed to decrease to 173 °C, and 0.7 mL of the Cs-precursor solution (containing 0.3 mmol CsOA) was injected. Following a short reaction time of ∼5 s, the reaction was rapidly quenched with an ice bath. The crude solution was centrifuged at 7800 rpm for 7 min, and the precipitate was discarded. The QDs were precipitated from supernatant by adding acetone in a 4:1 volume ratio, followed by centrifugation at 7800 rpm for 7 min. The supernatant was discarded, and the precipitate was then resuspended in toluene. QDs with various Mn2+ PLQYs can be achieved by varying the amount of Mn(CH3COO)2·4H2O and injection temperature; details can be found in Table S1. Undoped QDs were prepared by adopting a previously reported method. For synthesizing QDs with Mn2+ PLQY above ∼50%, the doping could be tuned by simply changing the amount of Mn(Ac)2·4H2O/HBr precursors (up to 255 mg and 0.5 mL, respectively) without adjusting other conditions. For synthesizing QDs with lower Mn2+ PLQYs, the injection temperature was increased (up to 200 °C), and the amount of Mn(Ac)2·4H2O/HBr precursors was decreased.
Antisolvent Purification
Mn2+-doped QDs colloid was then purified by adding methyl acetate in a 3:1 volume ratio and centrifuged at 7800 rpm for 7 min to reprecipitate QDs. The collected precipitate was redispersed in hexanes or toluene for storage and further characterizations. QDs with various Mn2+ PLQYs can be achieved by varying the amount of Mn(CH3COO)2·4H2O and injection temperature; details can be found in Table S1. Undoped QDs were prepared by adopting a previously reported method.
GPC Purification of Mn2+-Doped QDs
Mn2+-doped QDs were purified by GPC using the method reported by B. Greytak et al. Briefly, 4–5 g of Bio-Beads were washed three times and soaked overnight using toluene before use. Next, ∼5 mL of toluene was added to a glass column, which was carefully blocked with glass wool to prevent any bead leakage. The soaked Bio-Beads were then transferred to the column to reach a bed height of 10 cm. Once the column was fully packed, it was thoroughly rinsed with toluene until no free polystyrene was detected in the eluent, as confirmed by UV–vis absorption and fluorescence. Then, 1 mL of QD colloid was carefully loaded into the GPC column. The collected QDs were loaded into another GPC column after chemical purification for further purification iterations. The eluted QDs were collected and stored for elemental analysis and spectroscopic measurements.
Chemical Purification
20 μL of H2O2 solution (30 wt %) was added to the Mn2+-doped CsPbBr3 QD toluene solution. The solutions were then shaken to initiate Mn2+ oxidation. The Mn2+-doped CsPbBr3 QD solution will turn a dark brown color and start bubbling. The organic phase was collected, 20 μL of HBr was added, and the container was shaken for 15 s. After the organic phase became clear, the solution was centrifuged to accelerate phase separation, and the organic phase (QD colloid) was collected for further analysis. It is worth noting that the Mn2+ PL stability can vary depending on the relative amount of QDs and H2O2 solutions.
Elemental Analyses
The composition of QDs was determined using ICP-MS (Agilent 7850). The QD solutions using various purification methods were digested in concentrated nitric acid for ICP-MS analyses. The Br content in the halide perovskite was determined by STEM and EDS using Titan Themis 300 S/TEM. QD samples for XPS analysis were prepared by drop-casting the purified QD solution onto clean silicon substrates. An Omicron DAR 400 equipped with a CN10 charge neutralizer was used to collect the XPS data.
Structural Characterizations
Electron microscope images of all nanocrystals were obtained using a Titan Themis 300 TEM microscope operated at 300 kV. DLS analysis of CsPbBr3 NCs and Mn2+-doped CsPbBr3 QDs suspended in hexanes was recorded using a pUNk v1.0.0.3 DLS system equipped with a red laser (660 nm). The light scattering intensity was measured as a photon count rate in units of kilocounts per second (kcps). To account for varying scattered light intensities from nanoparticles of different sizes, the instrument automatically adjusted the incident laser beam power to achieve an optimal photon count rate. A built-in attenuator was used to set the laser power to specific levels as required. ζ-potential measurements were performed using the Anton Paar Litesizer 500. A suspension of Mn2+-doped or undoped CsPbBr3 nanocrystals (NCs) in toluene was introduced into a Univette equipped with a quartz cuvette containing 700 μL acetonitrile and 5–20 μL of 10 mM tetrabutylammonium hexafluorophosphate (TBAPF6) in acetonitrile as the supporting electrolyte. XRD patterns for CsPbBr3 NCs and Mn2+-doped CsPbBr3 NCs were collected on a Rigaku SmartLab X-ray diffractometer with a Cu–Kα source. The samples were prepared by drop-casting purified NCs onto a zero-background sample holder to ensure precise detection of XRD signals. Measurements were performed at room temperature with scans recorded in the 10–40° (2θ) range. Data analysis was conducted using Rigaku’s PDXL2 software package.
NMR Experiments
The 1H and 133Cs NMR spectra were acquired at a probe temperature of 20 °C on a 500 MHz JEOL ECZL NMR instrument (operating at 500.16 MHz for proton and 65.60 MHz for cesium) equipped with a 5 mm gradient inverse broadband Royal probe. The 1H NMR spectrum was recorded with single pulse excitation, a spectral width of 7508 Hz, and a pulse width of 3.95 μs (45° flip angle). A scan number of 8 transients and a repetition rate of 4.19 s (3.19 s for the acquisition time and 1 s for the relaxation delay) were used. The raw data (FIDs) for protons were processed without any apodization function prior to Fourier transformation. The 133Cs NMR spectrum was recorded with a spectral width of 19723.9 Hz with 40K data points and a pulse width of 7.02 μs (45° flip angle). A scan number of 4096 transients and a repetition rate of 3.66 s (1.66 s for the acquisition time and 2 s for the relaxation delay) were used. Exponential weighting with a line-broadening function of 25 Hz was applied before the Fourier transformation. 133Cs is the only naturally occurring stable isotope of cesium with a nonzero nuclear spin (I) of 7/2. Since it has a very small quadrupole moment (−3 × 10–3) (eQ) × 10–28 C m–2, a high natural abundance (100%), and a low magnetogyric ratio (3.5277 × 107 rad T–1 s–1) and has an excellent receptivity relative to the value of 269 for 13C. 1H and 133C NMR chemical shifts for protons are reported in parts per million (ppm) and are referenced to residual protons in the solvent peak CHCl3 at 7.26 ppm. 133Cs NMR chemical shifts are referenced using the cesium peak from Cs2CO3 at 0.0 ppm as the external standard.
EPR Experiments
The EPR spectra were acquired at room temperature using a commercial Bruker Biospin EMX spectrometer operating at a frequency of 9.51 GHz. The Mn2+-doped CsPbBr3 NCs were suspended in an EPR-inactive hexane or toluene solvent, and a few milliliters of the solution were put in 4 mm OD low-loss quartz tubes. The tubes were then inserted in the middle of the cylindrical microwave cavity for the EPR measurements. Typical microwave powers of 1–2 mW with 3 G modulation amplitude and 100 kHz field modulation were employed for these experiments. The Zeeman splitting g-values were calibrated using a DPPH (2,2-diphenyl-1-picrylhydrazyl) standard. The magnetic field resonance values and FHWM line widths were determined from fits to the spectra using Lorentzian line shapes (i.e., the first derivative of the Lorentzian-shaped microwave absorption curves).
Low-temperature EPR spectra were recorded with a Bruker/ColdEdge ESR900 WaveGuide cryostat, operating at a frequency of 9.35 GHz. This setup included a liquid helium flow system, allowing temperature control between 3 and 300 K. The Mn2+-doped CsPbBr3 NCs were dissolved in an EPR-inactive toluene solvent, with several milliliters of the mixture placed into 4 mm OD thin-wall Suprasil EPR tubes. The experiments were conducted using standard microwave power levels ranging from 0.2 to 2 mW, featuring a modulation amplitude of 3 G and a field modulation frequency of 100 kHz. Calibration of the Zeeman splitting g-values was performed using BDPA (α,γ-bisdiphenylene-β-phenylallyl) standard.
Optical Spectroscopy Measurements
An Ocean Insight Maya 2000 spectrometer was used to record absorption and PL spectra. A 385 nm LED was used to excite the samples for PL measurements. The PLQY of QDs was measured with respect to rhodamine 6G standard (PLQY 95%) as well as a reference QD sample (Mn2+-doped CdS/ZnS, PLQY 60.4% (Figure S2b), both excited with OBIS LX SF 405 nm continuous-wave laser (Coherent). QDs and Rhodamine 6G were suspended/dissolved in hexane and ethanol, respectively, and the difference in solvent refractive indices was accounted for when calculating the PLQY. Absolute PLQY was measured using an integrating sphere (Labsphere) coupled with a computer-controlled spectrometer (Ocean Optics QE Pro). The light sources used were a 405 nm laser diode (LDM405, Thorlabs, 4.0 mW) and a 415 nm fiber-coupled LED excitation source (M415F3, Thorlabs, 14.4 mW). PLE measurements were performed on colloidally suspended, highly Mn2+-doped CsPbBr3 nanocrystals in hexanes, using a quartz cuvette as the sample holder. The measurements were conducted with a HORIBA Jobin Yvon Fluorolog-3 spectrofluorometer equipped with a xenon lamp. Data were collected using the two-curve method over a wavelength range of 250–750 nm.
Pump–probe TA measurements were conducted at room temperature using a HELIOS TA spectrometer from Ultrafast Systems. The 405 nm, 2 kHz output from an Apollo-Y Optical Parametric Amplifier (OPA) served as the pump beam, with a fluence of ∼18.8 μJ/cm2. The probe beam, with a white-light spectrum, was generated in a sapphire crystal within the HELIOS spectrometer by focusing the 1064 nm, 2 kHz output from a Hyperion femtosecond amplified laser. Both the pump and probe pulses had a duration of approximately 350 fs.
Time-Resolved Photoluminescence (TRPL) Spectroscopy
The exciton PL lifetimes were measured using a time-correlated single-photon counting technique. The colloidal sample in a cuvette was excited with a 405 nm pulsed laser (Picoquant LDH-D-C-405) driven by a Picoquant Sepia PDL828 module at a 5 MHz repetition rate. The PL emission from the Mn2+-doped CsPbBr3 NCs was collected with an achromatic lens, sent through a set of long-pass (425 nm, Edmund Optics #84-742) and bandpass (466 ± 20 nm, Edmund Optics #86-352) filters to remove the scattered laser and the Mn2+ emission, and detected using a single-photon avalanche photodiode (Hamamatsu C11202-100). The photon arrival time was recorded using a Picoquant HydraHarp 400 correlator. The Mn2+ PL lifetimes were measured by exciting the sample using the 375 nm, 80 MHz output of a pulsed laser (Beckl & Hickl). The average excitation power was 30 μW, focused to a beam diameter of around 200 μm. PL was collected in free-space, backscattering geometry, and spectrally resolved with an HRS-300 Acton Spectrometer coupled to a Teledyne PIXIS400 CCD camera. The temporal resolution of the PL decay was achieved using a Time-Correlated Single Photon Counting (TCSPC) system by Becker & Hickl, comprising an IDQ-id100 fast avalanche photodiode and an SPN 130, capable of simultaneously measuring both nanosecond (fluorescence) and millisecond (phosphorescence) decay components. The phosphorescence decay time is measured in a triggered accumulation multichannel scaler (TA-MCS) mode, where the high-frequency laser output is modulated at lower frequencies. For an in-depth description of the electronics behind the TA-MCS mode, please refer to the Beckl & Hickl available online free of charge.
Supplementary Material
Acknowledgments
This material is based upon work supported by the National Science Foundation under Award No. CHE-2316919. We thank the TAMU Materials Characterization Facility (RRID: SCR_022202) for TEM and XPS measurements. We thank Dr. Preston Larson and the Samuel Noble Microscopy Laboratory at the University of Oklahoma (OU) for conducting XRD measurements. We thank Rishav Mukerjee for assistance with GPC purification and Nazmiye Gökçe Altınçekiç and Dr. Hyunho Noh for their assistance with ζ-potential measurements. We thank Dr. Philip C. Bourne for his assistance with DLS measurements at the OU Protein Production and Characterization Core facility, which is supported by Institutional Development Awards (IDeA) from the National Institute of General Medical Sciences of the National Institutes of Health (Grants P20GM103640 and P30GM145423), the OU Vice President for Research and Partnerships, and the OU College of Arts and Sciences. The photoluminescence quantum yield (PLQY) measurements performed at OU were supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Science, under Award No. DE-SC0021158. The absolute PLQY measurements performed at The University of Chicago were supported: Y.C. is supported by the National Science Foundation under Award No. CHE-2404291. Y.L. and D.V.T. are supported by the University of Chicago Materials Research Science and Engineering Center, supported by the National Science Foundation under Award Number DMR-2011854. We thank Dr. Alex N. Frickenstein, Shakya Sankalpani Gunasena Wije Munige, and Dr. Steven Foster for assistance with the ICP-MS measurements at OU Mass Spectrometry, Proteomics & Metabolomics Core Facility. EPR measurements performed at the U.S. Naval Research Laboratory were supported by the Office of Naval Research. EPR measurements performed at OU by N. Mohamed-Raseek and J. W. Peters were supported by the U.S. Department of Energy (DOE), Office of Science, Office of Basic Energy Sciences (BES), Chemical Sciences, Geosciences and Biosciences Division, Solar Photochemistry Program.
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.5c12086.
Additional optical measurements (UV–vis, PL, PLQY, TRPL, TA), elemental analyses (EDS, ICP-MS, XPS), STEM imaging, NMR, EPR, synthesis conditions, additional photographs of QD samples, and TRPL fitting parameters. (PDF)
The authors declare no competing financial interest.
References
- Pradhan N.. Mn-Doped Semiconductor Nanocrystals: 25 Years and Beyond. J. Phys. Chem. Lett. 2019;10(10):2574–2577. doi: 10.1021/acs.jpclett.9b01107. [DOI] [PubMed] [Google Scholar]
- Pradhan N., Das Adhikari S., Nag A., Sarma D. D.. Luminescence, Plasmonic, and Magnetic Properties of Doped Semiconductor Nanocrystals. Angew. Chem., Int. Ed. 2017;56(25):7038–7054. doi: 10.1002/anie.201611526. [DOI] [PubMed] [Google Scholar]
- Norris D. J., Efros A. L., Erwin S. C.. Doped Nanocrystals. Science. 2008;319(5871):1776–1779. doi: 10.1126/science.1143802. [DOI] [PubMed] [Google Scholar]
- Bryan J. D., Gamelin D. R.. Doped Semiconductor Nanocrystals: Synthesis, Characterization, Physical Properties, and Applications. Prog. Inorg. Chem. 2005;54:47–126. doi: 10.1002/0471725560.ch2. [DOI] [Google Scholar]
- Dey A., Ye J., De A., Debroye E., Ha S. K., Bladt E., Kshirsagar A. S., Wang Z., Yin J., Wang Y.. et al. State of the Art and Prospects for Halide Perovskite Nanocrystals. ACS Nano. 2021;15(7):10775–10981. doi: 10.1021/acsnano.0c08903. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Protesescu L., Yakunin S., Bodnarchuk M. I., Krieg F., Caputo R., Hendon C. H., Yang R. X., Walsh A., Kovalenko M. V.. 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]
- Kovalenko M. V., Protesescu L., Bodnarchuk M. I.. Properties and potential optoelectronic applications of lead halide perovskite nanocrystals. Science. 2017;358(6364):745–750. doi: 10.1126/science.aam7093. [DOI] [PubMed] [Google Scholar]
- Liu W., Lin Q., Li H., Wu K., Robel I., Pietryga J. M., Klimov V. I.. Mn2+-Doped Lead Halide Perovskite Nanocrystals with Dual-Color Emission Controlled by Halide Content. J. Am. Chem. Soc. 2016;138(45):14954–14961. doi: 10.1021/jacs.6b08085. [DOI] [PubMed] [Google Scholar]
- Parobek D., Roman B. J., Dong Y., Jin H., Lee E., Sheldon M., Son D. H.. Exciton-to-Dopant Energy Transfer in Mn-Doped Cesium Lead Halide Perovskite Nanocrystals. Nano Lett. 2016;16(12):7376–7380. doi: 10.1021/acs.nanolett.6b02772. [DOI] [PubMed] [Google Scholar]
- De A., Mondal N., Samanta A.. Luminescence tuning and exciton dynamics of Mn-doped CsPbCl3 nanocrystals. Nanoscale. 2017;9(43):16722–16727. doi: 10.1039/C7NR06745C. [DOI] [PubMed] [Google Scholar]
- Das Adhikari S., Dutta S. K., Dutta A., Guria A. K., Pradhan N.. Chemically Tailoring the Dopant Emission in Manganese-Doped CsPbCl3 Perovskite Nanocrystals. Angew. Chem., Int. Ed. 2017;56(30):8746–8750. doi: 10.1002/anie.201703863. [DOI] [PubMed] [Google Scholar]
- Guria A. K., Dutta S. K., Adhikari S. D., Pradhan N.. Doping Mn2+ in Lead Halide Perovskite Nanocrystals: Successes and Challenges. ACS Energy Lett. 2017;2(5):1014–1021. doi: 10.1021/acsenergylett.7b00177. [DOI] [Google Scholar]
- Mir W. J., Jagadeeswararao M., Das S., Nag A.. Colloidal Mn-Doped Cesium Lead Halide Perovskite Nanoplatelets. ACS Energy Lett. 2017;2(3):537–543. doi: 10.1021/acsenergylett.6b00741. [DOI] [Google Scholar]
- Wei Q., Li M., Zhang Z., Guo J., Xing G., Sum T. C., Huang W.. Efficient recycling of trapped energies for dual-emission in Mn-doped perovskite nanocrystals. Nano Energy. 2018;51:704–710. doi: 10.1016/j.nanoen.2018.06.073. [DOI] [Google Scholar]
- Yuan X., Ji S., De Siena M. C., Fei L., Zhao Z., Wang Y., Li H., Zhao J., Gamelin D. R.. Photoluminescence Temperature Dependence, Dynamics, and Quantum Efficiencies in Mn2+-Doped CsPbCl3 Perovskite Nanocrystals with Varied Dopant Concentration. Chem. Mater. 2017;29(18):8003–8011. doi: 10.1021/acs.chemmater.7b03311. [DOI] [Google Scholar]
- Choe H., Jin H., Lee S. J., Cho J.. Machine Learning-Directed Predictive Models: Deciphering Complex Energy Transfer in Mn-Doped CsPb(Cl1-yBry)3 Perovskite Nanocrystals. Chem. Mater. 2023;35(14):5401–5411. doi: 10.1021/acs.chemmater.3c00731. [DOI] [Google Scholar]
- Cai T., Wang J., Li W., Hills-Kimball K., Yang H., Nagaoka Y., Yuan Y., Zia R., Chen O.. Mn2+/Yb3+ Codoped CsPbCl3 Perovskite Nanocrystals with Triple-Wavelength Emission for Luminescent Solar Concentrators. Adv. Sci. 2020;7(18):2001317. doi: 10.1002/advs.202001317. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Meinardi F., Akkerman Q. A., Bruni F., Park S., Mauri M., Dang Z., Manna L., Brovelli S.. Doped Halide Perovskite Nanocrystals for Reabsorption-Free Luminescent Solar Concentrators. ACS Energy Lett. 2017;2(10):2368–2377. doi: 10.1021/acsenergylett.7b00701. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Q., Zhang X., Jin Z., Zhang J., Gao Z., Li Y., Liu S. F.. Energy-Down-Shift CsPbCl3:Mn Quantum Dots for Boosting the Efficiency and Stability of Perovskite Solar Cells. ACS Energy Lett. 2017;2(7):1479–1486. doi: 10.1021/acsenergylett.7b00375. [DOI] [Google Scholar]
- Ji S., Yuan X., Zheng J., Cao S., Ji W., Li H., Zhao J., Zhang H.. Near-unity blue-orange dual-emitting Mn-doped perovskite nanocrystals with metal alloying for efficient white light-emitting diodes. J. Colloid Interface Sci. 2021;603:864–873. doi: 10.1016/j.jcis.2021.06.138. [DOI] [PubMed] [Google Scholar]
- Kumawat N. K., Yuan Z., Bai S., Gao F.. Metal Doping/Alloying of Cesium Lead Halide Perovskite Nanocrystals and their Applications in Light-Emitting Diodes with Enhanced Efficiency and Stability. Isr. J. Chem. 2019;59(8):695–707. doi: 10.1002/ijch.201900031. [DOI] [Google Scholar]
- Liu Y., Zhao Y., Zhu Y., Zhan Y., Matras-Postolek K., Yang P.. Ion exchange derived CsPbBr3:Mn nanocrystals with stable and bright luminescence towards white light-emitting diodes. Mater. Res. Bull. 2022;153:111915. doi: 10.1016/j.materresbull.2022.111915. [DOI] [Google Scholar]
- De Siena M. C., Sommer D. E., Creutz S. E., Dunham S. T., Gamelin D. R.. Spinodal Decomposition During Anion Exchange in Colloidal Mn2+-Doped CsPbX3 (X = Cl, Br) Perovskite Nanocrystals. Chem. Mater. 2019;31(18):7711–7722. doi: 10.1021/acs.chemmater.9b02646. [DOI] [Google Scholar]
- Qiao T., Parobek D., Dong Y., Ha E., Son D. H.. Photoinduced Mn doping in cesium lead halide perovskite nanocrystals. Nanoscale. 2019;11(12):5247–5253. doi: 10.1039/C8NR10439E. [DOI] [PubMed] [Google Scholar]
- Chen M., Han X., Xing K., Song Y., Cao S., Zou B., Zheng J., Zhao J.. Exciton-to-Dopant Energy Transfer Dynamics in Mn2+ Doped CsPbBr3 Nanowires Synthesized by Diffusion Doping. J. Phys. Chem. Lett. 2023;14(51):11543–11549. doi: 10.1021/acs.jpclett.3c03036. [DOI] [PubMed] [Google Scholar]
- Zhao Y., Xie C., Zhang X., Matras-Postolek K., Yang P.. Mn:CsPbBr3 Nanoplatelets for Bright White-Emitting Displays. ACS Appl. Nano Mater. 2021;4(6):6223–6230. doi: 10.1021/acsanm.1c01016. [DOI] [Google Scholar]
- Huang G., Wang C., Xu S., Zong S., Lu J., Wang Z., Lu C., Cui Y.. Postsynthetic Doping of MnCl2 Molecules into Preformed CsPbBr3 Perovskite Nanocrystals via a Halide Exchange-Driven Cation Exchange. Adv. Mater. 2017;29(29):1700095. doi: 10.1002/adma.201700095. [DOI] [PubMed] [Google Scholar]
- Wu L., Wang Y., Kurashvili M., Dey A., Cao M., Döblinger M., Zhang Q., Feldmann J., Huang H., Debnath T.. Interfacial Manganese-Doping in CsPbBr3 Nanoplatelets by Employing a Molecular Shuttle. Angew. Chem., Int. Ed. 2022;61(15):e202115852. doi: 10.1002/anie.202115852. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xu K., Vickers E. T., Luo B., Allen A. L. C., Chen E., Roseman G., Wang Q., Kliger D. S., Millhauser G. L., Yang W.. et al. First Synthesis of Mn-Doped Cesium Lead Bromide Perovskite Magic Sized Clusters at Room Temperature. J. Phys. Chem. Lett. 2020;11(3):1162–1169. doi: 10.1021/acs.jpclett.9b03700. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mohata H., Nasipuri D., Dutta S. K., Pradhan N.. Limiting Halide Exchange and Doping Mn(II) in Vertex-Oriented Cube-Connected Patterned Lead Halide Perovskite Nanorods. Chem. Mater. 2025;37(1):497–506. doi: 10.1021/acs.chemmater.4c02908. [DOI] [Google Scholar]
- VanOrman Z. A., Cárdenes Wuttig M., Reponen A.-P. M., Kim T.-S., Casaday C. E., Cui D., Deshpande T., Jöbsis H. J., Schouwink P., Oveisi E.. et al. Metal Doping of Strongly Confined Halide Perovskite Nanocrystals under Ambient Conditions. J. Am. Chem. Soc. 2025;147(19):16536–16544. doi: 10.1021/jacs.5c03629. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Justice Babu K., Kaur G., Shukla A., Kaur A., Goswami T., Ghorai N., Ghosh H. N.. Concurrent Energy- and Electron-Transfer Dynamics in Photoexcited Mn-Doped CsPbBr3 Perovskite Nanoplatelet Architecture. J. Phys. Chem. Lett. 2021;12(1):302–309. doi: 10.1021/acs.jpclett.0c03267. [DOI] [PubMed] [Google Scholar]
- Wang L., Mao H., Li J., Li Y., Li M., Zhu J., Fan B., Liu W., Shao G., Xu H.. Heavy Mn-doped CsPbBr3 nanocrystals synthesized by high energy ball milling with high stability. Nanotechnology. 2022;33(45):455703. doi: 10.1088/1361-6528/ac81d6. [DOI] [PubMed] [Google Scholar]
- He W., Zhang Q., Qi Y., Xiong J., Ray P., Pradhan N. R., Shahbazyan T. V., Han F., Dai Q.. Luminescence properties of CsPbBr3:Mn nanocrystals. J. Nanopart. Res. 2021;23(4):80. doi: 10.1007/s11051-021-05184-7. [DOI] [Google Scholar]
- Dong Y., Qiao T., Kim D., Parobek D., Rossi D., Son D. H.. Precise Control of Quantum Confinement in Cesium Lead Halide Perovskite Quantum Dots via Thermodynamic Equilibrium. Nano Lett. 2018;18(6):3716–3722. doi: 10.1021/acs.nanolett.8b00861. [DOI] [PubMed] [Google Scholar]
- Akkerman Q. A., Nguyen T. P. T., Boehme S. C., Montanarella F., Dirin D. N., Wechsler P., Beiglböck F., Rainò G., Erni R., Katan C.. et al. Controlling the nucleation and growth kinetics of lead halide perovskite quantum dots. Science. 2022;377(6613):1406–1412. doi: 10.1126/science.abq3616. [DOI] [PubMed] [Google Scholar]
- Ye J., Gaur D., Mi C., Chen Z., Fernández I. L., Zhao H., Dong Y., Polavarapu L., Hoye R. L. Z.. Strongly-confined colloidal lead-halide perovskite quantum dots: from synthesis to applications. Chem. Soc. Rev. 2024;53(16):8095–8122. doi: 10.1039/D4CS00077C. [DOI] [PubMed] [Google Scholar]
- Chen H.-Y., Maiti S., Son D. H.. Doping Location-Dependent Energy Transfer Dynamics in Mn-Doped CdS/ZnS Nanocrystals. ACS Nano. 2012;6(1):583–591. doi: 10.1021/nn204452e. [DOI] [PubMed] [Google Scholar]
- Parobek D., Dong Y., Qiao T., Son D. H.. Direct Hot-Injection Synthesis of Mn-Doped CsPbBr3 Nanocrystals. Chem. Mater. 2018;30(9):2939–2944. doi: 10.1021/acs.chemmater.8b00310. [DOI] [Google Scholar]
- Qiao T., Parobek D., Son D. H.. Photons and charges from colloidal doped semiconductor quantum dots. J. Mater. Chem. C. 2019;7(47):14788–14797. doi: 10.1039/C9TC05150C. [DOI] [Google Scholar]
- Atteberry M. L., Mi C., Chandra S., Hidayatova L., Dong Y.. Unraveling the Growth Mechanism of Strongly Confined CsPbBr3 Perovskite Quantum Dots under Thermodynamic Equilibrium Control. Chem. Mater. 2024;36(9):4521–4529. doi: 10.1021/acs.chemmater.4c00160. [DOI] [Google Scholar]
- Vlaskin V. A., Barrows C. J., Erickson C. S., Gamelin D. R.. Nanocrystal Diffusion Doping. J. Am. Chem. Soc. 2013;135(38):14380–14389. doi: 10.1021/ja4072207. [DOI] [PubMed] [Google Scholar]
- Feldmann S., Gangishetty M. K., Bravić I., Neumann T., Peng B., Winkler T., Friend R. H., Monserrat B., Congreve D. N., Deschler F.. Charge Carrier Localization in Doped Perovskite Nanocrystals Enhances Radiative Recombination. J. Am. Chem. Soc. 2021;143(23):8647–8653. doi: 10.1021/jacs.1c01567. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ahmed G. H., Liu Y., Bravić I., Ng X., Heckelmann I., Narayanan P., Fernández M. S., Monserrat B., Congreve D. N., Feldmann S.. Luminescence Enhancement Due to Symmetry Breaking in Doped Halide Perovskite Nanocrystals. J. Am. Chem. Soc. 2022;144(34):15862–15870. doi: 10.1021/jacs.2c07111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Abiodun S. L., Pellechia P. J., Greytak A. B.. Effective Purification of CsPbBr3 Nanocrystals with High Quantum Yield and High Colloidal Stability via Gel Permeation Chromatography. J. Phys. Chem. C. 2021;125(6):3463–3471. doi: 10.1021/acs.jpcc.1c00207. [DOI] [Google Scholar]
- Maes J., Balcaen L., Drijvers E., Zhao Q., De Roo J., Vantomme A., Vanhaecke F., Geiregat P., Hens Z.. Light Absorption Coefficient of CsPbBr3 Perovskite Nanocrystals. J. Phys. Chem. Lett. 2018;9(11):3093–3097. doi: 10.1021/acs.jpclett.8b01065. [DOI] [PubMed] [Google Scholar]
- Hills-Kimball K., Yang H., Cai T., Wang J., Chen O.. Recent Advances in Ligand Design and Engineering in Lead Halide Perovskite Nanocrystals. Adv. Sci. 2021;8(12):2100214. doi: 10.1002/advs.202100214. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ravi V. K., Santra P. K., Joshi N., Chugh J., Singh S. K., Rensmo H., Ghosh P., Nag A.. Origin of the Substitution Mechanism for the Binding of Organic Ligands on the Surface of CsPbBr3 Perovskite Nanocubes. J. Phys. Chem. Lett. 2017;8(20):4988–4994. doi: 10.1021/acs.jpclett.7b02192. [DOI] [PubMed] [Google Scholar]
- De Roo J., Ibáñez M., Geiregat P., Nedelcu G., Walravens W., Maes J., Martins J. C., Van Driessche I., Kovalenko M. V., Hens Z.. Highly Dynamic Ligand Binding and Light Absorption Coefficient of Cesium Lead Bromide Perovskite Nanocrystals. ACS Nano. 2016;10(2):2071–2081. doi: 10.1021/acsnano.5b06295. [DOI] [PubMed] [Google Scholar]
- Aebli M., Kaul C. J., Yazdani N., Krieg F., Bernasconi C., Guggisberg D., Marczak M., Morad V., Piveteau L., Bodnarchuk M. I.. et al. Disorder and Halide Distributions in Cesium Lead Halide Nanocrystals as Seen by Colloidal 133Cs Nuclear Magnetic Resonance Spectroscopy. Chem. Mater. 2024;36(6):2767–2775. doi: 10.1021/acs.chemmater.3c02901. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qiao T., Liu X., Rossi D., Khurana M., Lin Y., Wen J., Cheon J., Akimov A. V., Son D. H.. Magnetic Effect of Dopants on Bright and Dark Excitons in Strongly Confined Mn-Doped CsPbI3 Quantum Dots. Nano Lett. 2021;21(22):9543–9550. doi: 10.1021/acs.nanolett.1c03114. [DOI] [PubMed] [Google Scholar]
- Bailey R. E., Nie S.. Alloyed Semiconductor Quantum Dots: Tuning the Optical Properties without Changing the Particle Size. J. Am. Chem. Soc. 2003;125(23):7100–7106. doi: 10.1021/ja035000o. [DOI] [PubMed] [Google Scholar]
- Guner T., Akbali B., Ozcan M., Topcu G., Demir M. M., Sahin H.. Monitoring the Doping and Diffusion Characteristics of Mn Dopants in Cesium Lead Halide Perovskites. J. Phys. Chem. C. 2018;122(21):11543–11549. doi: 10.1021/acs.jpcc.8b02580. [DOI] [Google Scholar]
- Zhou C., Tian Y., Khabou O., Worku M., Zhou Y., Hurley J., Lin H., Ma B.. Manganese-Doped One-Dimensional Organic Lead Bromide Perovskites with Bright White Emissions. ACS Appl. Mater. Interfaces. 2017;9(46):40446–40451. doi: 10.1021/acsami.7b12456. [DOI] [PubMed] [Google Scholar]
- Yang Z., Yuan X., Song Y., Chen M., Xing K., Cao S., Zheng J., Zhao J.. Thickness-Dependent Photoluminescence Properties of Mn-Doped CsPbBr3 Perovskite Nanoplatelets Synthesized at Room Temperature. J. Phys. Chem. C. 2023;127(43):21227–21234. doi: 10.1021/acs.jpcc.3c05639. [DOI] [Google Scholar]
- Rossi D., Parobek D., Dong Y., Son D. H.. Dynamics of Exciton-Mn Energy Transfer in Mn-Doped CsPbCl3 Perovskite Nanocrystals. J. Phys. Chem. C. 2017;121(32):17143–17149. doi: 10.1021/acs.jpcc.7b06182. [DOI] [Google Scholar]
- Ricciarelli D., Meggiolaro D., Belanzoni P., Alothman A. A., Mosconi E., De Angelis F.. Energy vs Charge Transfer in Manganese-Doped Lead Halide Perovskites. ACS Energy Lett. 2021;6(5):1869–1878. doi: 10.1021/acsenergylett.1c00553. [DOI] [PMC free article] [PubMed] [Google Scholar]
- K, R. P. ; Viswanatha, R. . Mechanism of Mn emission: Energy transfer vs charge transfer dynamics in Mn-doped quantum dots. APL Materials 2020, 8 (2). 10.1063/1.5140888 (acccessed 7/9/2025). [DOI] [Google Scholar]
- Singh R., Liu W., Lim J., Robel I., Klimov V. I.. Hot-electron dynamics in quantum dots manipulated by spin-exchange Auger interactions. Nat. Nanotechnol. 2019;14(11):1035–1041. doi: 10.1038/s41565-019-0548-1. [DOI] [PubMed] [Google Scholar]
- Erwin S. C., Zu L., Haftel M. I., Efros A. L., Kennedy T. A., Norris D. J.. Doping semiconductor nanocrystals. Nature. 2005;436(7047):91–94. doi: 10.1038/nature03832. [DOI] [PubMed] [Google Scholar]
- Dalpian G. M., Chelikowsky J. R.. Self-Purification in Semiconductor Nanocrystals. Phys. Rev. Lett. 2006;96(22):226802. doi: 10.1103/PhysRevLett.96.226802. [DOI] [PubMed] [Google Scholar]
- Solari S. F., Poon L.-N., Wörle M., Krumeich F., Li Y.-T., Chiu Y.-C., Shih C.-J.. Stabilization of Lead-Reduced Metal Halide Perovskite Nanocrystals by High-Entropy Alloying. J. Am. Chem. Soc. 2022;144(13):5864–5870. doi: 10.1021/jacs.1c12294. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Torma A. J., Li W., Zhang H., Tu Q., Klepov V. V., Brennan M. C., McCleese C. L., Krzyaniak M. D., Wasielewski M. R., Katan C.. et al. Interstitial Nature of Mn2+ Doping in 2D Perovskites. ACS Nano. 2021;15(12):20550–20561. doi: 10.1021/acsnano.1c09142. [DOI] [PubMed] [Google Scholar]
- Becker, W. bh TCSPC handbook, 9th ed.; 2023. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.





