The luminescence and charge transport properties of inorganic CsPbX3 perovskite nanocrystals (NCs) make them attractive candidates for various optoelectronic applications, such as lasing, X-ray imaging, light communication, and light-emitting diodes (LEDs). However, to realize cutting-edge device performance, high-quality NCs with high photoluminescence quantum yields (PLQYs) are essential. Therefore, substantial efforts and progress have been made to attain superior design/engineering and optimization of the inorganic NCs with a focus on surface quality, reduced nonradiative charge carrier recombination centers, and improved colloidal stabilities. Metal-ion doping has been proven to have a robust influence on the electronic band structure, PL behavior, and charge carrier recombination dynamics. Thus, in this perspective, we summarize the recent progress of the significant impact of metal cation doping on the optical properties, including the PL enhancement of CsPbCl3, CsPbBr3, and CsPbI3 perovskite NCs. Moreover, we shed light on the mechanism behind such improved properties. We conclude by recommending possible aspects and strategies to be further explored and considered for better utilization of these doped NCs in thin-film optoelectronic and energy conversion devices.
I. INTRODUCTION
All-inorganic lead-halide perovskite nanocrystals (NCs), typically CsPbX3 (X = Cl−, Br−, and I−), are being considered as next-generation light emitters due to their desirable photophysical properties, such as high photoluminescence quantum yield (PLQY: e.g., near unity), high color-purity with tunable emission, high defect tolerance, and good thermal and moisture stabilities.1–9 Because of the ionic nature and low crystallization temperature, colloidal inorganic perovskite NCs can be easily synthesized by low-cost solution-processed methods.4 Such superlative properties render CsPbX3 nanocrystals widely utilizable in optoelectronic applications such as solar cells,10–13 light-emitting diodes (LEDs),14–21 lasers,22–25 photodetectors,26–28 single-photon sources,29,30 X-ray scintillators,31,32 visible-light communications,33 and high-definition displays.34,35 Despite the great promise for various applications, several studies indicate that solution-processed perovskite NCs suffer from the existence of a substantial level of trap-mediated nonradiative losses including point defects (excessive Pb2+ atoms or halide vacancies) and structural disorders, which act as parasitic nonradiative recombination centers, leading to photoluminescence quenching and poor thermal-stability and photo-stability.17,26,36–39 Recently, several promising steps have been taken to break the present bottleneck and tackle these issues, including (i) surface treatment with different ligands or metal ion salts,17,42–46 (ii) employing halide-rich synthesis conditions or controlling the lead and halogen ratio during the synthesis,37,46,47 and (iii) doping with targeted metal halide salts.40,41 In semiconductor nanocrystals, dopant engineering is a promising approach for tailoring and modulating the luminescence efficiency as well as the stability of the NCs.48,49 In general, the incorporation of dopant states either imparts extra electrons (i.e., n-type doping) or extra holes (i.e., p-type doping), which allows for fine control over the recombination processes and accordingly enhances the optical properties of the host, or induces new emissions from dopants.50,51 In addition, judicious doping enables introducing new electronic energy levels within the bandgap and corresponding low-energy electronic transitions without changing the crystal structure or the intrinsic optical features of the host materials.41,52 In the case of the inorganic perovskite system, the intrinsic flexibility of the ionic crystalline scaffold (APbX3) makes it more convenient and versatile to achieve targeted doping compared to the conventional metal chalcogenide semiconductor nanocrystals, such as PbS and CdSe.41 Moreover, introducing B-site (i.e., the site of Pb2+) dopants could precisely tune the luminescence and optical properties and bandgaps, stabilize crystal structures, create new emission characteristics, and reduce the defect state density.41,53 In the perovskite NC systems, the doping can be realized by adding the metal-ion salts into the precursor solutions during the hot-injection or room temperature synthesis.50,54,55 In addition, postsynthetic cation exchange reactions have been shown to be another efficient way for tuning the compositional and optical properties of the NCs.56,57 Recently, a wide array of metal ions, including transition metals, rare earth elements, and lanthanides, have been successfully incorporated into perovskite NCs and led to a substantial augmentation of the NCs’ optical properties.41 For instance, the doping of Mn2+ and lanthanide ions (e.g., Yb3+, Er3+, La3+, …) into perovskite NCs can generate a high PLQY—approaching near-unity in some cases—of metal ion-related emission as a result of the efficient energy transfer from the perovskite NC host to the dopants.58–62 However, another variety of monovalent, divalent, and heterovalent metal ion dopants (Ni2+, Cd2+, Cu2+, Ca2+, Sr2+, Na+, Ce3+, and Sb3+) has led to the improved PLQY of the perovskite NCs without introducing new emission bands.50,54,55,63–66 On the other hand, it was also noted that some dopants (i.e., Bi3+, Al3+, and Ag+) tend to introduce trap states in the bandgap that quench the photoluminescence.67–69 Despite these impressive strides in the area of perovskite doping, the exact influence of dopant ions on the optical properties and their behavior inside the perovskite matrix remains unclear and hard to predict.40 In this perspective, we summarize the recent studies on the B-site metal ion doping of CsPbX3 (X = Cl, Br, and I) NCs that lead to the improved optical characteristics (i.e., a higher PLQY and longer PL lifetime after doping). We present the recent works and progress made on the metal doping of CsPbCl3, CsPbBr3, and CsPbI3 and give an overview of the current understanding of the inherent doping mechanisms that lead to PL enhancement and how dopants control the optical properties of inorganic perovskite materials. Finally, we discuss the present limitations and challenges to be solved to further expand the potential integration of doped perovskite NCs into optoelectronic devices for better efficiency and operational stability.
A. Metal-ion doped CsPbCl3 nanocrystals
The CsPbCl3 NC system is the only system in the Cs-Pb-X perovskite family that can achieve deep-blue emission due to its intrinsically large bandgap (∼3.0 eV). However, point defects (such as Cl vacancies, with the lowest defect formation energy occurring under Cl-poor synthetic conditions26) can introduce local structural distortions and lattice disorders. Such defects can induce intragap defect states and nonradiative recombination centers that degrade the optical properties of the pristine NCs.63 Therefore, several research efforts have been recently devoted to increasing the PLQY of CsPbCl3 NCs in order to expand their practical optoelectronic applications. Here, we focus on the recent developments related to metal-ion incorporation in pristine CsPbCl3 NC hosts that improve the optical properties and PLQY without introducing new emission bands from metal ions. In the literature, several dopants have been incorporated into the CsPbCl3 NC hosts, especially alkaline earth (AE2+) metals and transition metals. For instance, due to the comparable ionic radii of AE2+ metal ions [i.e., Ba2+, Sr2+, Ca2+, and Mg2+ (135, 118, 100, and 72 pm, respectively) relative to Pb2+ (119 pm)],70 Chen and co-workers have carefully explored the effects of doping AE2+ metal ions into CsPbCl3 NCs (via the hot injection method) on suppressing the structural defects and improving the performance of the LED devices.65 As shown in Fig. 1(a), they found that the photophysical properties were correlated with the type and amount of AE metal ions: (i) upon incorporating Sr2+, Ca2+, and Mg2+, a significant increase in the PLQY was observed, from ∼0.8% to 36.6% (Mg2+), to 50.7% (Sr2+), and to 77% (Ca2+); (ii) an extremely low PLQY was observed in the case of Ba2+ (0.9%); and (iii) increasing the doping concentration of Ca2+ from 2% to 11% shows a remarkable increase in the PL lifetime and PLQY [see the time-resolved PL spectra in Fig. 1(a), lower panel]. In addition, the measured nonradiative decay rates clearly suggest a successful suppression of the surface defect states and/or vacancies.
In the case of doping with transition metal ions, Yong et al. have studied the effect of incorporating Ni ions into CsPbCl3 NCs using the hot injection method, as Ni can strongly coordinate with halide ions.63 From the steady-state absorption and PL spectra of CsPbCl3 NCs before and after Ni2+-doping [Fig. 1(b)], there is virtually no change in the absorption spectra of CsPbCl3 NCs after Ni2+-doping, but a substantial PL enhancement of CsPbCl3 NCs was observed. Interestingly, a remarkable increase in the PLQY (up to 96.6%) was achieved, suggesting negligible nonradiative recombination in Ni2+-doped NCs when the doping concentration reached 11.9 mol. %. The time-resolved PL measurements indicated that the PL decay of CsPbCl3 NCs becomes slower after Ni-doping (i.e., an increased PL lifetime), providing another strong piece of experimental evidence for the efficient suppression of the nonradiative recombination channels [Fig. 1(b), left panel]. Importantly, the crystal structural characterizations assisted by both X-ray absorption fine structure (EXAFS) and magic angle spinning (MAS) measurements revealed that Ni2+ ions can substantially suppress the structural formation of defects, especially the Cl vacancy, in the CsPbCl3 NCs. Therefore, the obtained near-unity PLQY of CsPbCl3 can be attributed to the improved crystal lattice order when the Ni2+ ions are introduced into the host NCs.
Samanta and co-workers have recently studied the doping of Cd2+ metal ions into CsPbCl3 NCs using room-temperature postsynthetic cation exchange/treatment.64 Similar to the Ni2+-doping case,63 the Cd2+ ions induced no PL spectral shift despite that the radii of Cd2+ ions (93 pm) and Ni2+ ions (69 pm) are much smaller than that of Pb2+ ions [Fig. 1(c)]. Surprisingly, the PLQY was improved, reaching near-unity (96% ± 2%) after Cd2+ doping. The Cd2+-doped CsPbCl3 NCs were found to exhibit improved stability under the continuous illumination in ambient conditions. The time-resolved PL measurements [Fig. 1(c), left panel] suggested that the doped NCs exhibit a much longer lifetime as compared to that of the pristine NCs. In addition, the ultrafast transient absorption and the temperature-dependent PL measurements along with the calculated nonradiative decay rates revealed that the Cd2+ ions can introduce a shallow energy level and dramatically suppress the nonradiative channels, which could be one possible reason for the increased PLQY.
Very recently, De et al. have investigated another transition metal ion (i.e., Cu+) doped into CsPbCl3 NCs by using the one-pot hot injection method.54 The PLQY increased by almost ∼120 times upon Cu+ doping (from ∼0.5% up to 60%). Similar to the previous cases using transition metal ions, no PL shift was observed in Cu+-doped CsPbCl3 NCs. The time-resolved PL measurements showed that the Cu+-doped NCs show much longer PL lifetime as compared to undoped NCs; furthermore, the femtosecond transient absorption measurements suggest the suppression of the ultrafast carrier trapping processes in the doped CsPbCl3 NCs.
Due to the fast ion exchange between perovskite NCs and metal ion dopants, diffusion doping can also be realized via postsynthetic surface treatment. We have recently reported a simultaneous dual-surface doping/passivation using YCl3 metal ion salt on the surface of CsPbCl3 NCs.26 We achieved an enhancement in the PLQY (from 1% to 60%) of CsPbCl3 NCs with improved surface stability [Fig. 1(d)]. From time-resolved PL spectra, we also observed a significant increase in the PL lifetime from 0.4 to 1.7 ns, which suggests that the nonradiative decay rates were substantially reduced after YCl3 treatment. This was also supported by density functional theory (DFT) calculations, which show that Y3+ and Cl− ions efficiently occupy the surface Pb–Cl ion vacancies, enriching the density of states in the conduction band without creating any mid-gap states.
B. Metal-ion doped CsPbBr3 nanocrystals
Among Cs-Pb-X compounds, CsPbBr3 NCs are considered to be the most promising green light emitters for LED applications17,18 because of their relative thermal stability (compared to CsPbI3 and CsPbCl3), low moisture sensitivity, and high PLQY.1,67,71 However, there is still a need to substantially improve the external quantum efficiencies (EQEs) of the CsPbBr3 NC-based LED devices before their commercial applications are realized. Hence, doping could also become a key strategy to increase the EQE.72 Yao et al. have recently doped CsPbBr3 NCs with Ce3+ ions by using the hot injection method, taking into account that the Ce3+ ion is not too dissimilar in ionic radius (103 pm) from the Pb2+ ion (119 pm).50 They found that CsPbBr3 NCs when doped with a substantive amount of Ce3+ ions remarkably showed a twofold enhancement in the PLQY (reaching up to 89%) as compared to the pristine CsPbBr3 NCs. The steady-state absorption and PL showed that the Ce3+ dopants could induce a slight blue shift in the absorption edge and a large enhancement in the PL intensity [Fig. 2(a)]. Under ambient conditions, the CsPbBr3 NCs showed significant improvement in stability due to the lattice contraction after Ce3+-doping.50 Interestingly, through the comparison of the average PL lifetimes and the PLQY of CsPbBr3 NCs at different dopant concentrations, it was found that the PL enhancement was accompanied by a decrease in the average PL lifetimes as the dopant concentration increased [Fig. 2(a), left side]. They further explained such an unexpected relationship based on the ultrafast transient absorption measurements: the near band-edge states from Ce3+ dopants provided more radiative channels for the PL enhancement, and meanwhile, the enhanced couplings between these states as well as between the higher-lying and lowest excitonic states can promote the PL decay. In addition, the fabricated LED devices based on the Ce3+-doped CsPbBr3 NCs showed a 3-fold increase in the EQE from 1.6% to 4.4%.
Very recently, Li et al. have reported a doping strategy of univalent Na+ ions into CsPbBr3 NCs at room-temperature.55 The doped CsPbBr3 NCs showed better color purity, improved stability, and a higher PLQY (∼85%) compared to the pristine NCs. As shown in Fig. 2(b), the steady-state absorption and PL spectra of the colloidal Na+-doped CsPbBr3 NCs with different Na/Pb atom ratios show that upon Na+ doping both absorption and emission peaks were gradually blue shifted by increasing the doping concentration, which indicates that such doping can slightly alter the bandgap of CsPbBr3. The temperature-dependent PL measurements suggest that the emission intensity of the doped NCs decreased by increasing the temperature, which results from the thermally activated nonradiative recombination. Moreover, the time-resolved PL measurements showed that the average PL lifetime became much longer and the calculated nonradiative recombination rates dropped dramatically after Na+-doping, highlighting that Na+ ions substantially suppressed the formation of charge carrier trapping states in the CsPbBr3 NCs. Interestingly, the white light-emitting devices (WLEDs) employing the Na+-doped NCs as solid-state phosphors exhibited a stable high quality white-light and a notably high power efficiency (67.3 l m/W).
Zhang et al.73 have reported the room-temperature doping of Sb3+ ions (92 pm ionic radius) into ultra-small blue emissive CsPbBr3 NCs. They found that the doping of Sb3+ ions could reduce the surface energy, boost the PLQY of blue emission up to 73.8%, and enhance the stability even at elevated temperatures in solution (40–100 °C). The steady-state absorption and PL spectra in Fig. 2(c) demonstrate a slight reduction in the full width at half maximum (FWHM) due to the increased exciton binding energy. The PLQY of the doped NCs [sample C in Fig. 2(c)] increased from a 50% baseline to as high as 73%. The PL decay kinetics suggested that the Sb3+ doping could passivate the defect states below the bandgap and increase the probability of radiative recombination. It was noted that the CIE coordinates of the doped CsPbBr3 NCs were very close to the primary blue color according to the National Television System Committee (NTSC) TV color standard, which bodes well for the applications of these NCs in display.
C. Metal-ion doped CsPbI3 nanocrystals
CsPbI3 NCs are very promising red-light emitters. Although pristine CsPbI3 NCs can achieve a high PLQY (>70%), these NCs suffer from severe phase instability. The black phase of CsPbI3 exhibits an undesirable tolerance factor that leads to poor phase stability, i.e., it undergoes a spontaneous room temperature phase transition to a nonphotoactive yellow phase when exposed to humidity or ambient conditions.74,75 Thus, tremendous efforts have been devoted for developing various synthetic approaches to stabilize the black phase of CsPbI3 NCs while preserving a high PLQY.2,76,77 Due to the similar ionic radii of Sr2+ and Pb2+ ions (118 pm vs 119 pm), Lu et al.66 have studied the doping effects of Sr2+ on the photo/electroluminescence efficiency and the phase stability of CsPbI3 NCs. In their study, the SrCl2 coprecursor introduced a simultaneous Sr2+ ion doping and surface Cl− ion passivation. The steady-state absorption and PL spectra of CsPbI3 NCs with different amounts of SrCl2 showed that both the absorption peak and PL peak exhibited a slight blue-spectral shift when increasing the SrCl2:PbI2 ratio. Following the SrCl2 addition, a significant increase in the PLQY of CsPbI3 NCs from 65% to 84% was observed and the improved stability of the cubic -phase of CsPbI3 NCs was achieved [Fig. 3(a)]. It is worth noting that the enhanced phase stability of CsPbI3 NCs can be attributed to the direct increase in the formation of energy and the modifications of the tolerance factor of the CsPbI3 crystal structure. The time-resolved PL kinetics suggest that the average PL lifetimes dramatically increased when employing SrCl2, which demonstrates that the surface defects were efficiently reduced and the surface of CsPbI3 NCs became Cl− enriched—reducing the likelihood of uncoordinated Pb atoms—resulting in the increased PLQY and a much-prolonged charge carrier lifetime. The authors achieved a 13.5% EQE of LED devices by employing the Sr2+-doped CsPbI3 NCs.
Shen et al. have studied the partial substitution of Pb2+ ions into CsPbI3 NCs through Zn2+ alloying.78 As the ionic radius of Zn2+ (74 pm) is much smaller than that of Pb2+ (119 pm), the reduced lattice constant of the NCs suggests that Zn2+ cations efficiently replace Pb2+ cations. The steady-state absorption and PL showed that both the absorption and the PL peaks of the NCs were blue shifted by increasing the amount of Zn2+ ions in CsPbI3 hosts. The authors attributed these spectral shifts to the observations that the lattice contraction and the emission intensity increased at first and then decreased with the increasing concentration of Zn2+ in the NCs [see Fig. 3(b)]. Interestingly, the PLQY of the alloyed NCs dramatically increased up to 98.5% when changing the Zn2+ concentration. In addition, the radiative decay rates showed a 4-fold increase, while a 10-fold decrease was seen in the nonradiative decay rates, indicating an effective suppression of the defect states in alloyed NCs. The stability of the NCs was also improved as a result of modulating the tolerance factor of the lattice structure. Interestingly, the nanocrystals switched from n-type due to I− vacancies to nearly ambipolar for the alloyed NCs. More importantly, alloying Zn2+ into NCs can achieve an enhanced EQE of up to 15.1% in the LED-based device.
Very recently, Bera et al. have found that the incorporation of Sb3+ ions into the CsPbI3 NCs using the hot injection method could stabilize the crystal phase under ambient atmosphere [Fig. 3(c)].79 They observed that excess doping changed the shape of the cubic NCs and induced a 2D structure assembly, forming platelet-shaped NCs, and consequently quenched the emission efficiency and reduced the overall stability, as shown in the atomic model of the crystal structure. Since the ionic radius of Sb3+ (92 pm) is smaller than that of Pb2+ (119 pm), the Sb3+ ions can increase the tolerance factor of the crystal (i.e., an indicator for the stability and distortion of perovskite structures, which is defined as , where rA is the radius of the A-site cation, rB is the radius of the B-site cation, and r0 is the radius of the anion) and therefore improve the crystal stability. However, replacing Pb (ii) with Sb (iii) ions could lead to severe charge imbalance. The steady-state absorption and PL exhibited an apparent redshift by increasing Sb3+ concentration. The time-resolved PL measurements show that the PL lifetime of Sb3+-doped CsPbI3 NCs increased relative to the pristine NCs, but still within the same nanosecond order. The PLQY of the doped CsPbI3 NCs (doping concentration ∼10%) was 83%, and this value was practically decreased by increasing the Sb3+ amount deviated from ∼10%. DFT calculations suggested that upon Sb3+ ion incorporation, an increase in the crystal cohesive energy and a decrease in the bandgap were observed. Remarkably, the authors achieved a 9.4% power conversion efficiency (PCE) by employing these doped NCs in solar cell devices.
Mir et al. have recently studied the effect of postsynthetic Mn2+ doping on the surface and lattice energy of CsPbI3 NCs.80 They found that Mn2+ ions can passivate the NC surface and restrict the growth of the NCs and stabilize the black phase of CsPbI3 NCs. Therefore, the postsynthetic Mn-doped NCs showed improved PL ambient stability as compared to the undoped NCs [see Fig. 3(d)]. The time-resolved PL decays showed that a radiative lifetime was about the same before and after doping, respectively, suggesting that Mn-doping did not introduce any additional nonradiative decay channels to the CsPbI3 NC host. Due to the large difference between the Mn–I bond (283 kJ mol−1) and the Pb–I bond (194 kJ mol−1) dissociation energies, the lattice contraction stabilizes the lattice and the surface of Mn-doped NCs relative to the undoped NCs. It is worth highlighting that different from the cases of CsPbCl3 and CsPbBr3 NCs, the energy transfer into the d-states of Mn2+ from the host would not occur and there is no emission from Mn2+ ions in the doped CsPbI3 NCs because the Mn d-d transient energy level (∼2.12 eV) is higher than the bandgap of CsPbI3 NCs (∼1.8 eV).
Very recently, Guvenc et al. studied the effects of Gd3+ doping on the optical properties and the phase stability of α-CsPbI3 NCs. Since the ionic radius of the Gd3+ ion (93 pm) is smaller than that of the Pb2+ ion (119 pm), Gd3+ incorporation into the perovskite lattice increased the Goldschmidt tolerance factor of the CsPbI3 crystal and accordingly led to a more stable perovskite structure.81 The authors attributed the prolonged phase stability to several factors, such as (i) increased tolerance factor of the perovskite structure, (ii) distorted cubic symmetry, and (iii) decreased defect density in nanocrystals. Interestingly, Gd3+ doping (10 mol. %) of α-CsPbI3 led to an increase in both the PLQY from 70% to 80% and the fluorescence lifetime from 47.4 ns to 64.4 ns, which stem from the promoted radiative recombination due to the reduced defect state density. In addition, similar to the Sb3+ doping case,79 the DFT calculations revealed that Gd3+ doping increases the cohesive energy per atom from 3.22 eV to 3.32 eV for neat and Gd-doped CsPbI3, respectively, which indicates that the formation of Gd3+-doped CsPbI3 NCs is more favorable than neat CsPbI3 NCs. Moreover, the electronic band structure of Gd-doped CsPbI3 crystals showed an n-type semiconductor behavior without any substantial modifications in the band structure.
D. Understanding the doping effects of perovskite nanocrystals by density functional theory (DFT) calculations
Several theoretical tools, especially density functional theory (DFT), have been used to understand the doping mechanisms and their relationship to the optical properties in inorganic perovskite materials at the atomic level. For CsPbCl3, Chen et al. have found that in the case of Ca2+ and Sr2+doped CsPbCl3 NCs,65 (i) compared to foreign atoms (dopant ions), Pb2+ atoms are involved at an early stage of NC formation; (ii) once the NCs were formed, the Pb2+ atoms were depleted and uncoordinated surface atoms were generated; and (iii) since the Ca2+ ions were abundant in the solution, the Ca2+ ions can strongly attach to the surface of NCs because of the large bond dissociation energy of Ca–Cl (409 kJ/mol). The authors also suggested that Ca2+ (or Sr2+) isovalently occupies the Pb2+ crystallographic sites, resulting in the formation of a passivation layer at the NC’s surface, as was concluded from XPS measurements. The formation of such a layer decreases the concentration of point defects caused by uncoordinated constituent atoms at the NC’s surface (e.g., the uncoordinated Pb2+ ions) as shown in Fig. 4(a), leading to the improved short-range order of the lattice as suggested by EXAFS measurements. However, for the larger alkaline earth metal ions, such as Ba2+, the bond dissociation energy of Ba–Cl (443 kJ/mol) is so large that the Ba2+ ions can enter the lattice of NCs even at the early stage of NC growth. Although, due to Ba2+ having a larger ionic radius than Pb2+ (135 pm vs 119 pm), doping of Ba2+ ions can increase the local strain of the lattice, giving rise to a higher probability of forming atomic point defects as suggested by DFT calculations. Such structural defects may further trigger the incorporation of more Ba2+ ions into NCs, thus leading to the successful doping of Ba2+ in the core region as well as further lattice expansion. Therefore, the PLQY of Ba2+-CsPbCl3 NCs is very low. Different from the case of Ba2+ dopants, which can introduce deep trap states that quench the PL of CsPbCl3 NCs, Modal et al. have shown that both Cd2+ and Cl− play a significant role in suppressing the halide vacancies by using CdCl2 for increasing the PLQY. In addition, De et al. have attributed the high PLQY of Cu+-doped CsPbCl3 NCs to the suppression of carrier trapping and showed that the majority of the carriers follow a longer recombination time as indicated by transient absorption measurements.
Yong et al. achieved a near-unity PLQY by doping Ni2+ ions into CsPbCl3, and their experimental results suggest that the doping of Ni2+ ions can substantially suppress the formation of structural defects such as chloride vacancies and results in improved short-range order of the perovskite lattice and more ordered lattice.63 Their DFT calculations revealed that Ni2+-doped CsPbCl3 NCs show a much larger defect formation energy of Pb vacancy (VPb) than those of VCl and VCs, suggesting that VCl and VCs could be the dominant defects in CsPbCl3 NCs and Ni doping into CsPbCl3 favors the suppression of those defects. It is worth mentioning that VCl and Ni-VCl can introduce deep defect levels, which act as carrier trapping centers for nonradiative recombination [Fig. 4(b)].
For CsPbBr3 NCs, Li et al. have explained the PL enhancement and the higher PLQY in the Na+-doped CsPbBr3 NCs due to the decrease in the carrier trapping centers.55 Their DFT calculations suggested that the formation energy of the antisite NaPb defect is the lowest regardless of the position of the Fermi level. Moreover, when the Fermi level is close to the conduction band minimum (CBM), the formation energy of the NaPb defect is still the lowest among all the studied defects. The formation energy calculations demonstrate that NaPb is the dominant defect in Na+-doped orthorhombic-phase CsPbBr3. To understand the mechanism of the enhanced band-edge emission in the case of Ce3+ ions doping, Yao et al. have conducted femtosecond transient absorption measurements to study the ultrafast exciton dynamics of Ce3+ doped and undoped samples.50 They found that the Ce3+ ions can introduce a new band-edge state near the conduction band, leading to the increase in the lowest excitonic state density, facilitating the coupling between the higher-lying excitonic states and the lowest excitonic state as well as the coupling between the lowest excitonic state and the bandgap trap states. In addition, in our DFT calculations, the calculated doping charge transition levels suggest that most of the Ce3+ dopant related defects could form the shallow transition levels within the bandgap in both cubic-phase and orthorhombic-phase Ce3+-doping CsPbBr3.82 These shallow levels can preserve the bulk electronic structure and do not degrade the optoelectronic properties. From the projected density of states (PDOSs) for cubic-phase CsPbBr3 after introducing the antisite CePb and interstitial Cei dopants, we found that (i) for pristine CsPbBr3, the valence band maximum (VBM) and CBM of CsPbBr3 are mainly composed of Br-4p, Pb-6s, and Pb-6p orbitals and show delocalized electronic distributions; (ii) upon Ce3+ doping, the bandgaps of CsPbBr3 in both phases remain almost unchanged after introducing the Ce3+ ions at the Pb2+ site, and the charge densities for the CBM retain the delocalized feature; and (iii) the interstitial Ce3+ has a small contribution to the conduction band edge in the PDOS of cubic-phase CsPbBr3, and the same electronic feature can be found in the orthorhombic case with an interstitial Ce3+. Therefore, we have confirmed that Ce3+ can be doped into CsPbBr3 NCs without introducing additional trap states and can slightly increase the electron density in the conduction band, which modulates the exciton relaxation and recombination of CsPbBr3 NCs.
In the case of CsPbI3 NCs, Bera et al. have highlighted that Sb3+ doped CsPbI3 showed an increase in the cohesive energy along with a decrease in the bandgap due to Sb incorporation based on the DFT calculations.79 Lu et al. have also highlighted that the mechanism for the PL enhancement in Sr-doped CsPbI3 NCs is due to the reduction of surface defects when using SrCl2 as a coprecursor.66 When using SrCl2 as dopants, the surface of CsPbI3 NCs becomes enriched with Cl−, which eventually passivates the uncoordinated Pb atoms, resulting in the increased PLQY and PL lifetime.
In summary, with the help of the DFT calculations, several aspects of the doping effects in CsPbX3 NCs can be reasonably explained from atomic level interactions. From the calculated dopant formation energies, we can predict the probable atomic positions of metal ion dopants; the dopants can occupy the Cs, Pb, or halide sites if they have small intersite formation energies; otherwise, they can act as interstitials in the perovskite host. For the former case, we need to further determine the charge transition levels of dopants so that we can predict whether the dopants introduce trap states that quench the PL or increase the conduction band states, thereby enhancing the PL. In the case of interstitial substitution, on the other hand, significant energy transfer from the perovskite host to the dopants can promote the emission from dopants instead of the perovskite host.
II. CONCLUSION AND REMARKS
Despite the great promise that perovskite NCs hold for light emission applications and optoelectronic devices, optimizing the perovskite material’s quality and stability remains as major impediments. Doping the perovskite NC lattice with metal ions presents a viable approach to overcome those impediments. The near-unity PLQY was achieved in several cases after metal ion doping as a result of the effective suppression of the defect states within the host NCs as well as on their surface. However, several outstanding issues need to be addressed before the full potential of doping strategies is realized, including (i) the precise atomic positions and distribution of metal ion dopants inside the perovskite lattice are still poorly understood. For instance, it has been reported that some dopants can induce deep trap state levels while others merely introduce shallow defect states. In this case, high-level DFT calculations (e.g., the combination of spin-orbit couplings and hybrid functionals) are needed to accurately predict the formation energies and charge transition levels of the dopants in the host; (ii) the energy transfer and energy level alignment between the perovskite nanocrystal host and metal ion guest need to be further explored and deciphered in order to understand the observed PL quenching or enhancement in the same materials when different metal ion dopants are used; and (iii) although a near-unity PLQY in perovskite NCs can be achieved upon metal ion doping in the solution, it is still challenging to obtain spin-coated thin films of perovskite NCs with a high PLQY and practical thermal-stability and photo-stability. The effect of dopants on the morphology of NC thin films should be accorded more attention. Moreover, a proper surface treatment of the doped perovskite NCs may help us to further improve the quality of their thin films and facilitate the fabrication of higher efficiency devices.
ACKNOWLEDGMENTS
The authors acknowledge funding support from KAUST.