Introduction

Over the past few years, inorganic lead halide perovskites have attracted great attention as a promising optoelectronic material for light-emitting devices, photodetectors, and low-threshold lasers, owing to their high photoluminescence quantum yield (PL QY), narrow emission, and tunable band gap1,2,3,4. The basic building block in these material is PbX6 octahedra (where X is a halogen), whose diverse connectivity can produce structures with various dimensionalities, ranging from three-dimensional (3D) to zero-dimensional (0D)5,6,7. The 3D inorganic perovskites, with general formula APbX3 (A = Cs, Rb, and X = Cl, Br, or I), consist of an extended network of corner-sharing PbX6 octahedra with cavities occupied by A ions8,9. Despite being the most explored material, the poor chemical stability of 3D perovskites against moisture, inherent phase transformation, and ion migration make their low-dimensional counterparts more favorable for optoelectronic applications4,6,10.

In particular, 0D inorganic perovskite-like Cs4PbX6 system have crystal structure in which the PbX6 octahedra are decoupled from each other by the surrounding Cs+ ions. The complete isolation of octahedra leads to strong quantum confinement and exciton–phonon interactions, which in turn can result in exciton localization, self-trapping, and polaron formation10. The optical features of 0D Cs4PbX6 are governed by transitions between the electronic states of Pb2+ ions, and its broad ultraviolet (UV) emission has been assigned to the radiative decay of Frenkel excitons at Pb2+ sites11,12,13,14,15. However, the origin of their PL in the visible range is still under debate because of mixed views on the efficient green luminescence (QY of 45%) of Cs4PbBr6, which some studies have been attributed to the minor 3D CsPbBr3 nanoscale impurity3,16,17,18.

Nikl et al.12 reported a UV emission band at 355 nm for Cs4PbCl6 single crystals, assigning to Pb2+ ion emission, originating from the optical transitions of 3P0,11S0 in the isolated PbCl6 octahedra, similar to Pb2+ doping in the alkali halide hosts12. In addition, the coexistence of 3D CsPbCl3-like impurity was noted by 414 nm emission in Cs4PbCl6. While, Mohammed’s group19 reported Cs4PbBr6 characteristic with two broad UV emissions at 340 nm (high-energy) and 400 nm (low-energy) attributing to Pb2+ ion and charge-transfer band denoted as “D-state”, respectively. D-state emission originates via the transfer of excited electrons from (PbBr6)4 octahedra to the Pb2+ occupying Cs sites (D-states), through their strong coupling in Cs4PbX6 host19. Therefore, 0D Cs4PbX6 behave as an ideal host–guest system, with periodically doped individual PbX6 species in a wide band gap matrix. However, though many reports attribute the green emission (512 nm) to the intrinsic property of Cs4PbBr6, we believe it arise from the coexistence of CsPbBr3 impurity, similar to 414 nm emission of Cs4PbCl612. Nevertheless, 0D cesium lead halide perovskites still represent largely unexplored and intriguing perovskite-like material, expected to exhibit interesting optoelectronic properties, due to their strongly localized excitons and high exciton binding energies (150 to 380 meV)3,20,21.

The coexistence of 3D CsPbX3 impurity has precluded a detailed understanding on the optical properties of Cs4PbX6, leading to a growing demand for their synthesis in pure form and their phase stabilization. Solvent washing has been recently reported for the synthesis of Cs4PbBr6 free from CsPbX3 impurity, which, however, is inadequate to treat the chloride compositions due to their poor solubility3. As the inevitable coexistence of CsPbX3 phase represents a major drawback, new synthetic strategies are of utmost interest for obtaining pure Cs4PbX6 nanocrystals with reduced CsPbX3 impurity.

Recently, the molecule-like behavior of Cs4PbBr6 has been discussed in terms of charge carrier transport and polaron formation, arising from weak interactions between isolated octahedra10. Thus, understanding the fundamental optoelectronic behavior of 0D Cs4PbX6 would promote the development of bulk perovskite-based devices. Incorporation of transition metal ions such as manganese (Mn2+) in the semiconductor nanocrystal leads to interesting optoelectronic properties by modulating the electronic properties of the host, providing new routes for designing solid-state lighting and light-harvesting devices22,23. Hence, Mn2+ ions can serve as the sensitive probe for investigating the local structure of the host and altering its optical and electronic behavior. Thus far, the potential effect of isovalent cation doping on the isolated octahedral units of Cs4PbX6 remains unexplored, particularly in terms of phase stabilization and optical properties.

In this work, we probe the optical behavior of isolated octahedra in the Cs4PbX6 (X = Br, Cl, Br/Cl) by introducing Mn2+ at the octahedral sites and explore its phase stabilization over commonly coexisting CsPbX3 impurity. The luminescence measurements show that Mn2+ dopant significantly alter the optical properties of different host emissive states in the Cs4PbX6. The dopant emission is controlled by adjusting the band gap of Cs4PbX6 through halide identity, facilitating energy transfer for enhancing the dopant emission efficiency in both colloidal and solid forms. The quantum-confined structure of 0D Cs4PbX6 further improves the stability of the host and efficiency of the dopant emission, particularly in their solid form.

Results

Mn2+-doped Cs4PbX6 nanocrystals

A series of 0D Cs4PbX6 (X = Br, Br/Cl, and Cl) nanocrystals, doped with manganese, were synthesized via modified reverse microemulsion method (Supplementary Figure 1 and Supplementary Note 1)20. A short-chain alkylamine (octylamine) surfactant was employed instead of the previously reported long-chain oleylamine to control size and morphology20. The exact composition of the synthesized undoped and Mn2+-doped Cs4Pb(Br/Cl)6 (denoted as mixed halide) analogs are Cs4PbBr2Cl4 and Cs4Pb1xMnxBr2–2xCl4+2x, respectively, with x varying from 0.05 (5% Mn) to 0.80 (80% Mn). The Pb/Mn ratios were estimated from inductively coupled plasma−optical emission spectroscopy (Supplementary Table 1). For simplicity, the synthesized samples were represented as Cs4PbX6:x% Mn, where X = (Br/Cl), Cl, and Br; and x = percentage of Mn concentration with respect to displaced lead ions.

Local and bulk structure

The crystal structure of Cs4PbX6 consists of caged PbX6 octahedra, isolated by interspersed Cs–X bridges21. Two types of Cs sites are present: the Cs(1) site form an alternating octahedra with PbX6, whereas the Cs(2) trigonal prisms share one triangular face to form infinite [CsPbX6]n3 chains along the [001] direction24. The synthesized samples were highly crystalline matching with the rhombohedral Cs4PbX6 phase (X = Br, Br/Cl, and Cl; space group: R\(\bar 3\)c) as presented in Fig. 1a. However, in the undoped Cs4PbX6, the presence of 3D CsPbX3 impurity were clearly apparent in all the halide analogs (Fig. 1a, Supplementary Figure 2, and Supplementary Note 2). Despite using high Cesium content (Cs:Pb = 5:1), formation of undesired CsPbX3 phase were observed in the undoped Cs4PbX6 probably due to localized occurrence of Cs-deficient regions.

Fig. 1
figure 1

Local and bulk structures of Mn2+-doped Cs4PbX6. a X-ray diffraction patterns of Mn2+-doped Cs4PbX6 series with different doping concentrations showing the rhombohedral structure is preserved after Mn-doping. Selected 2θ range of (22 to 24° and 29 to 31°) XRD patterns were presented. A monotonic shift of the XRD peak at Bragg’s angle at 23.4° towards higher 2θ is the result of progressive lattice contraction as the Mn2+ concentration increases, due to the substitutional replacement of Pb2+ with isovalent but smaller Mn2+ ions. The additional impurity cubic phases of 3D CsPb(Br/Cl)3, 3D CsPbCl3, CsBr, Cs(Br/Cl), and CsCl were denoted as hash, asterisk, triangle, open circle, and closed circle, respectively. It is to be noted that the Cs:Pb precursor ratio was fixed at 5:1 for all the samples of Cs4PbX6 (Br, Cl, and Br/Cl). b Phase fraction of the desired 0D Cs4PbX6 perovskite with respect to 3D CsPbX3 perovskite phase and segregation of CsX, with varying Mn concentration in the Cs4PbX6 synthesis. c Octahedral tilting/distortion without and with Mn doping (10% Mn) causing the phase stabilization of Cs4Pb(Br/Cl)6 structure. Electron density distribution profile derived from Maximum Entropy method for the d undoped Cs4PbCl6, e Cs4PbCl6:10% Mn, and f Cs4PbCl6:20% Mn, along the (010) plane. g Schematic representation of phase stabilization of Cs4PbX6 perovskite structure upon Mn2+ incorporation

Upon Mn incorporation, the Bragg’s peak at 23° monotonically shifts to higher 2θ, suggesting the substitution of Pb2+ with smaller Mn2+ ion in Cs4PbX6 (X = Br, Br/Cl, Cl), which is also evidenced from the decrease in lattice parameters (Table 1, Supplementary Tables 2 to 4)24,25. Lattice contraction of 0.3% and volume change of 2%, after Mn (20% Mn) doping, resulted in the shortening of lead halide (Pb–Br/Cl) bond lengths and similar feature was also noted in chloride analog (Supplementary Table 5, 6). The strong bonding nature upon Mn doping is evidenced from high bond-dissociation energy of Mn–Cl (361 kJ mol−1) than Pb–Br bonds (247 kJ mol−1)26. The valence states and substitution of Mn2+ ions at the isolated octahedral PbX6 sites was reaffirmed from X-ray photoelectron spectroscopy and electron spin resonance results, respectively (Supplementary Figures 3a to 3f and Supplementary Note 3)22,27,28. Quantification of multiple crystalline phases (Cs4PbX6 and CsPbX3) in the synthesized samples were estimated using Rietveld refinement of X-ray diffraction (XRD) results. Interestingly, Mn doping resulted in a decrease in the phase fraction (mass%) of 3D CsPbX3 impurity, which further reduced with Mn content in both Br/Cl and Cl analogs, as shown in Fig. 1b. In other words, the phase fraction of 0D Cs4Pb(Br/Cl)6 with respect to the total (0D+3D) phase gradually increased from 86 to 100% upon increasing Mn concentration from 0 to 20%. This increase in the phase fraction of Cs4PbX6 was also apparent in the pure chloride and bromide analog (Supplementary Figure 2), suggesting that Mn2+ incorporation in the framework favors the formation of Cs4PbX6 phase, while preventing the undesirable CsPbX3 phase. A CsPbX3 impurity-free Cs4Pb(Br/Cl)6 phase was obtained with 10% Mn content. Despite the complete suppression of CsPbX3, additional CsX (8%) phase tends to precipitate at higher Mn (above 10% Mn) content in both Cs4Pb(Br/Cl)6 and Cs4PbCl6 (Fig. 1b). Since the presence of Mn2+ in the lattice favors the Cs4PbX6 formation, the excess Cs precursor (5:1) which is used to favor the 0D phase, tend to segregate as CsX at higher Mn concentration. The CsX formation could be further prevented by decreasing the Cs precursor (Cs:Pb = 4.5:1) in the Mn2+-substituted samples, leading to pure Cs4PbX6 nanocrystals (Supplementary Figure 4). Our attempt to incorporate higher Mn content above 20% Mn, resulted in an additional hexagonal CsMnCl3 phase, along with Cs4Pb(Br/Cl)6 (Supplementary Figure 5 and Supplementary Note 4).

Table 1 Structural parameters of Cs4Pb(Br/Cl)6 with varying concentration of Mn, calculated using the Rietveld refinement

Phase stabilization of Cs4PbX6

The versatility of the Cs–Pb–X ternary compounds (particularly Cs4PbX6 and CsPbX3) lies in their ability to interconversion via post-synthetic physical and chemical treatments29. The post-synthetic conversion of cubic CsPbX3 to rhombohedral Cs4PbX6 has been demonstrated via amine- and thiol-mediated extraction of PbBr230, while the reverse conversion from Cs4PbX6 to CsPbX3 has been reported by the insertion of PbX2, stripping of CsX, excess oleic acid, and by heat treatment (90 to 180 °C)18. These post-synthetic conversion mechanisms are inadequate to explain the simultaneous formation of Cs4PbX6 and CsPbX3 phases, during synthesis. It has been demonstrated that the phase stabilization of perovskite semiconductors could be achieved by controlling the octahedral tilting via partially substituting Pb2+ with other metal ions. For instance, substitution of smaller Mn2+ stabilizes the cubic α-CsPbI3 phase by reducing the octahedral rotation or tilting via a decrease in the bond angle of Pb–I–Pb below 180°, which is along the interconnected neighboring PbX6 octahedra through the bridging halide ion31. In the present case, a similar octahedral tilting (enhanced) results in the phase stabilization of low-symmetry Cs4Pb(Br/Cl)6 upon Mn2+ incorporation. A decrease in the Pb–X bond length after Mn doping (10% Mn) to 2.950 Å, than undoped Cs4Pb(Br/Cl)6 (2.991 Å), resulted in an enhanced octahedral tilting, preventing the undesired high-symmetry cubic CsPbX3 phase (Fig. 1c). Here, the octahedral tilting in the Cs4PbX6 (X = Br/Cl) is described based on the X–Pb–Cs1 bond angle along c axis of the unit cell, which changes from 54.94° (undoped) to 54.76° upon Mn incorporation. This octahedral tilting would distort the PbX6 octahedra, which subsequently lowers the symmetry of crystal structure. It is worth mentioning that only two stable compositions exist in the phase diagram of mixed CsX-PbX2, under Cs-rich or Pb-deficient conditions viz., low-symmetry rhombohedral Cs4PbX6 and high-symmetry cubic CsPbX332. Therefore, enhanced octahedral tilting upon Mn substitution, result in the formation of only possible low-symmetry Cs4PbX6 phase; at the same time, destabilizing the cubic CsPbX3 phase under Cs-rich condition. This phase stabilization of 0D Cs4PbX6 via Mn substitution, might favor deeper investigation to their unexplored and intriguing properties.

Electronic structure

The changes in the electronic structure namely bonding nature and electron density distribution in the Cs4PbCl6 were analyzed with Mn incorporation via maximum entropy method (MEM) using XRD results. In the Cs4PbCl6, the electron density around Pb and Cl is anisotropic and highly resolved with no feature of electron sharing between Pb and Cl, resembling electrostatic attractive/repulsive forces (Fig. 1d). Upon Mn substitution a significant change in the local charge density around Pb atom was apparent, suggesting that Mn modulates the electronic structure of the 0D Cs4PbCl6 (Fig. 1e, f). With increasing Mn content, high electron density distribution between Pb and Cl atoms, suggests covalent bonding nature of (Pb/Mn)–Cl bonds (Supplementary Figure 6). This also supports the stronger (Pb/Mn)–Cl bonds or shorter bond lengths of Pb–Cl upon Mn doping, as observed from XRD. The change in local electron density after Mn-doping results in the lattice distortion (also evidenced from octahedral tilting, Fig. 1c) and could lead to electron-phonon (lattice) coupling and formation of polaron33. Recently, the formation of polarons in the Cs4PbBr6 has been reported from density functional theory calculation and transient absorption measurements10. In the present case, at variance with the undoped Cs4PbX6, a strong lattice distortion upon Mn incorporation could eventually enhance the electron-phonon coupling, leading to polaron formation. Therefore, it could be concluded that Mn2+ incorporation modulates the electronic structure and favors the phase stabilization of Cs4PbX6, preventing the formation of undesirable CsPbX3 impurity (Fig. 1g). Though the phase stabilization of Cs4PbX6 were concluded based on XRD results, this technique has its own limitation of insensitivity to detect any minor crystalline phase below 5%34. Henceforth, to further support the phase stabilization, optical characterization was exploited owing to its high sensitivity to detect even trace of fluorescent CsPbX3 impurity.

The excitation and photoluminescence (PL) properties of Cs4PbX6:Mn colloids were presented in Fig. 2. The undoped Cs4Pb(Br/Cl)6 exhibited broad emission at 432 nm (ranging 415 to 470 nm) under 365 nm excitation (Fig. 2a). The broad emission arises from the contribution from “D-states” of the host and 3D CsPb(Br/Cl)3 impurity (443 nm emission, Supplementary Figure 7), where the former dominates the latter. The presence of CsPb(Br/Cl3) impurity was also evident from the broad absorption ranging 360 to 400 nm in the undoped sample (Fig. 2b). It is to be noted that D-state emission arise from the charge transfer from halide ions of the PbX6 octahedra to Pb2+ ion occupying mismatched neighboring Cs+ sites, resulting in the creation of Frenkel excitons localized at the vicinity of Pb ions19,35. Upon Mn incorporation (5% Mn), a narrow emission band at 420 nm was observed, corresponding to the D-state of Cs4Pb(Br/Cl)6 phase, with feeble emission from CsPb(Br/Cl)3 impurity. The inclusion of chlorine during Mn doping (MnCl2 precursor) may contribute meagerly to the D-state emission compared to undoped one. Further increase in Mn content (above 10% Mn), eliminated the CsPb(Br/Cl)3 impurity as evidenced from its negligible absorption (360–400 nm), absorption narrowing, and unaltered D-state emission (Fig. 2a, b). The above results reaffirm the strong phase stabilization of Cs4PbX6 in presence of Mn, further corroborating the XRD results (Fig. 1a)36,37. The additional orange emission at 599 nm, is assigned to the forbidden Mn2+ dd transitions22, indicating the presence of exchange coupling between charge carriers of Cs4Pb(Br/Cl)6 and Mn2+ ions. Further, the single-halide Cs4PbX6 analogs were investigated to understand the role of Mn2+ on the optoelectronic properties and distinguish them from halide effect, which is a limitation in the mixed halides. The undoped Cs4PbBr6 exhibited two UV emission bands and intense green emission (512 nm) arising from Pb2+/D-state emission and CsPbBr3 impurity, respectively (Supplementary Figure 8 and Supplementary Note 5). While in the undoped Cs4PbCl6, the 414 nm emission corresponds to the band-edge emission of CsPbCl3 impurity12 and the hump ranging 450 to 480 nm is assigned to the quantum size effect of CsPbCl3. Upon Mn incorporation, a new emission at 400 nm arises from the D-state of Cs4PbCl6, with less contribution from CsPbCl3, compared to undoped one, which completely disappeared with higher Mn (20% Mn) content (Supplementary Figure 4), reaffirming the phase stabilization of Cs4PbX6 upon Mn incorporation.

Fig. 2
figure 2

Luminescent properties of Mn2+-doped Cs4PbX6. a PL spectra of Mn2+-doped Cs4PbX6 colloids were measured under 365 nm excitation, where X = Br/Cl and Cl, with Mn contents denoted as 0% Mn, 5% Mn, 10% Mn, and 20% Mn and the insets shows the respective samples under 365 nm UV lamp. b UV-Vis spectra of Mn2+-doped Cs4PbX6 colloids. c PL spectra of Mn2+-doped Cs4PbX6 colloids were measured under 290 nm excitation and the insets shows the respective samples under 254 nm UV lamp. The PL spectra presented are the PL intensities corrected using optical density. The asterix mark denote the 3D CsPb(Br/Cl)3 phases. FE-SEM images of d undoped Cs4Pb(Br/Cl)6 and e Cs4Pb(Br/Cl)6:10% Mn, and f, g HR-TEM images of Cs4Pb(Br/Cl)6:10% Mn. A lattice spacing of 0.31 nm corresponding to (131) plane of rhombohedral structure of Cs4Pb(Br/Cl)6

The undoped and Mn2+-doped Cs4PbCl6 displayed blue and pinkish (combination of D-state ‘blue’ and Mn2+ ‘orange’) emission under 365 nm, respectively (Fig. 2a, inset). The dual UV absorption bands at 218 and 308 nm, are the characteristic features of Cs4Pb(Br/Cl)6. The band gaps estimated from the absorption maxima are 3.83, 4.02, and 4.25 eV, for Cs4PbBr6, Cs4Pb(Br/Cl)6, and Cs4PbCl6, respectively (Supplementary Figure 8a), which agrees with the earlier reports18. It is interesting to note the increase in Cs4PbX6 host emission after Mn doping (Fig. 2a–c) in both mixed halide and chloride analog, similar to the perovskite nanocrystals38. The host emission increment is more prominent in the mixed halide (Br/Cl) analog than pure chloride, which is presumed to the passivation effect of chloride ion, on the pre-existing defects of the host38. Therefore, Mn2+ doping significantly alters the competitive kinetics between the radiative and non-radiative relaxation process of the Cs4PbX6 host.

The PL properties were investigated, by exciting at energies higher than band gap of host (290 nm), due to their strong absorption. Mn2+-doped (Br/Cl) samples exhibited dual UV emissions at 372 nm and a hump at 324 nm, arising from optical 3P0,1,21S0 transitions of Pb2+ ion35,39 and henceforth denoted as ‘Pb2+ ion’ emission (Fig. 2c) . Unlike at 365 nm excitation, the D-state (420 nm) emission of both undoped and Mn2+-doped was either absent or buried in the broad 300 to 450 nm spectral range under 290 nm excitation. A similar, broad Pb2+ emission (354 nm) was observed for Cs4PbCl612, without the D-state emission, when excited at 290 nm. With Mn doping, an apparent drop in the 3D CsPb(Br/Cl)3 emission (ranging 430 to 450 nm) further demonstrates the phase stabilization of Cs4PbX6 (Fig. 2c). In addition, despite the unaltered Pb2+ emission, an enhanced Mn2+ emission in the mixed halides, was observed with increasing Mn concentration, due to efficient energy transfer from Pb2+ to Mn2+. The unusual red-shifted emission from 599 to 635 nm for Cs4Pb(Br/Cl)6:20% Mn sample, is attributed to the formation of Mn-to-Mn pairs at higher Mn concentration, which reduces the 4T1 - 6A1 energy gap, resulting in the red shift of Mn2+ emission40. Hence, the phase stabilization of Cs4PbX6 (X = Cl, Br/Cl) with Mn incorporation were demonstrated in the chlorine-dominant system (above 60%), through XRD and optical results.

The Mn incorporation was also investigated in bromide analog (Cs4PbBr6) for monitoring phase stabilization via XRD and PL results (Supplementary Figures 2 and 8). With increasing Mn, the green emission (512 nm) of CsPbBr3 impurity decreased gradually, which completely disappeared above 70% Mn, suggesting the prevention of CsPbBr3 impurity in presence of Mn, agreeing with the XRD results. However, unlike Br/Cl and Cl analogs, the bromide counterpart necessitates higher Mn2+ concentration (above 50% Mn) to stabilize the Cs4PbBr6 structure, which is ascribed to the low substitution ratio of Mn2+ observed in the bromine-dominated Cs–Pb–Br perovskite-like system41. In addition, this leads to residual segregation of CsBr, at higher Mn concentration (above 50% Mn). The above results further validate the phase stabilization of zero-dimensional Cs4PbX6 upon Mn incorporation in all the bromine, Br/Cl, and chlorine analogs.

The synthesized Cs4Pb(Br/Cl)6:10% Mn exhibited cubic morphology with particle size ranging between 20 and 40 nm (Fig. 2d–g) which is much smaller than undoped one (50 to 120 nm). Further increasing the Mn content did not affect the particle size. It is worthy to note that the emission property of Cs4PbX6 host is independent of the particle size due to its intrinsic structural confinement and extremely localized emissive centers (isolated octahedra). However, the dependence of size and morphology of Cs4PbX6 on the length of alkyl chains of amine surfactant were clearly apparent. Unlike long-chain amines (oleylamine), which lead to a narrow particle size distribution in the 26 ± 4 nm range (mixture of spherical, hexagonal, and cubic morphologies)20, the use of shorter-chain octylamine in the present work favored a cubic shape with a relatively wide size distribution.

Probing the host emissive states

The excitation-dependent PL properties of Cs4Pb(Br/Cl)6 were explored to get insight of multiple luminescent states of the host. A range of excitation wavelengths from 280 to 365 nm (4.40 to 3.40 eV) were employed including excitation energies higher than the band gap of Cs4Pb(Br/Cl)6 host (4.00 eV). At higher excitation energies (λex of 280 nm), dominant D-state emission (432 nm) was accompanied by Pb2+ emission (324 and 375 nm) for undoped Cs4Pb(Br/Cl)6 (Fig. 3a). It is to be noted that 432 nm emission is denoted here as D-state emission, disregarding the less contribution from CsPb(Br/Cl)3 impurity (Fig. 2a), for the ease of emphasizing on dependence of host emissions (Pb2+ and D-state) with varying excitation energies. As excitation wavelength increased from 280 to 310 nm, the D-state emission gradually decreased, where Pb2+ emission remains dominant. While above 310 nm excitation, Pb2+ emission disappeared due to insufficient energy to excite its 3P1 levels and only D-state is retained. Upon Mn doping, the D-state emission completely disappeared in the excitation wavelengths ranging, 280 to 310 nm, exhibiting only Pb2+ emission (Fig. 3b, c). The absence of D-state emission after Mn2+ doping is probably due to suppression of charge-transfer process from PbX6 octahedra to Pb2+ occupying Cs+ site, primarily at high excitation energies (above 4.0 eV), where both Pb2+ and D-states are excited. Therefore, the presence of Mn2+ plays a key role in regulating the optical behavior of molecule-like isolated PbX6 octahedra in Cs4PbX6 nanocrystals.

Fig. 3
figure 3

Probing the host states at the isolated octahedron with Mn2+ ion. Colloidal PL spectra measured at different excitation wavelengths ranging 280 to 365 nm for the a undoped Cs4Pb(Br/Cl)6 and b Cs4Pb(Br/Cl)6:5% Mn. c Emission intensities of Pb2+ and D-state emission with varying excitation energies for the undoped and Mn-doped Cs4Pb(Br/Cl)6:5% Mn, d excitation spectra of Cs4Pb(Br/Cl)6:5% Mn colloids by monitoring at 375 nm (Pb2+ ion), and 420 nm (D-state), e PL intensity ratio of Mn2+ to the host emission with respect to varying excitation energies, f PL intensity ratio of Mn2+ to the host states with varying Mn concentration for Cs4PbX6 colloids, where X = (Br/Cl) and Cl, g origin of host emissions namely Pb2+ and D-state under high excitation energies, without and with Mn doping at high excitation energies. The solid, dashed, and curved arrows are denoted for optical excitation, emission, and energy transfer processes, respectively. h Strength of energy transfer from the host states namely Pb2+ and D-state to the Mn2+ emission are depicted. The upward, downward, and curved arrows are denoted for the optical excitation, emission, and energy transfer processes, respectively

The energy transfer and direct excitation of 3P1 levels of Pb2+ (375 nm) and D-states (420 nm), were explored from excitation spectrum of 5% Mn2+-doped Cs4Pb(Br/Cl)6 (Fig. 3d). The spectrum reveals Pb2+ ion can be excited at high excitation energies above the band gap of Cs4Pb(Br/Cl)6 host (below 310 nm excitation); while the D-state is excited at lower excitation energies typically below the host band gap, (above 310 nm excitation). A direct excitation of D-state at 280 nm, is also apparent along with Pb2+ ion35. However, the energy transfer between the Pb2+ and D-state of the host cannot be ruled out for the undoped sample, due to the occurrence of D-state emission in the excitation wavelengths (290 to 310 nm) where only Pb2+ state is excited, as reported earlier35, and is consistent with Cs4PbX6 solids (Supplementary Figure 9 and Supplementary Note 6).

Moreover, Mn2+ emission is well-known to be sensitized by the host via energy transfer, where Pb2+ and D-state sensitizes at high and low excitation energies, respectively. Therefore, Mn2+ emission at varying excitation energies is determined from the absorption of the corresponding host state and efficiency of energy transfer from host states. The Mn2+ emission relative to the total host emission (i.e., Pb2+ + D-state) was the highest at 280 nm, attributing to the efficient energy transfer from the Pb2+ state, which then gradually decreased at lower excitation energies, apart from 320 nm (Fig. 3e). Although the direct excitation of both Pb2+ and D-states is possible at 280 nm, Mn2+ is sensitized only through Pb2+ state, owing to the suppression of D-state in presence of Mn. The, dominant Mn2+ emission at 320 nm excitation, ascribed to the Mn2+ sensitization from both D-state and partially from Pb2+ ion. In addition, energy transfer from Pb2+ ion to Mn2+ was more efficient than from D-state (Fig. 3f, Supplementary Figure 10, and Supplementary Note 7). Moreover, the presence of Mn2+ considerably suppresses the D-state emission by blocking the charge-transfer process from isolated octahedra to the D-state, especially at high excitation energies (about 4.3 eV, 280 nm excitation), where both the D-state and Pb2+ could be directly excited (Fig. 3g). This may be caused by the weak coupling between electronically decoupled (Pb/Mn)X6 octahedra and D-state (Pb2+ at Cs site) which deteriorates the charge transfer between them, due to strong Mn–Cl bond formation. The sensitization of Mn2+ is predominantly occurs from Pb2+ than the D-state, due to larger energy difference between Mn2+ and Pb2+ states (Fig. 3h). Moreover, the efficient energy transfer to Mn2+ dopant in the wide band gap host, due to large energy difference between host and Mn2+ levels, is well documented38. The above results demonstrate that the Mn2+ incorporation significantly affects the optical properties of host state in Cs4PbX6 nanocrystals.

Luminescence efficiency

The potential application of perovskite-like Cs4PbX6 as emitting layer in color-converting devices has been explored in colloidal and solid forms (thin films and powders)16 and is expected to have high conversion efficiency similar to other color converter viz. phosphor42. High PL QY has been reported in perovskite colloids by employing ligands to stabilize the nanocrystals1,43, and thin films (solids) of micron-sized grains44,45. Here, although PL QY of undoped Cs4Pb(Br/Cl)6 nanocrystal is low (5%), the overall QY substantially increased to 32% upon Mn2+ doping, under 365 nm excitation (Table 2, Supplementary Figure 11a, and Supplementary Note 8). Highest PL QY of 29% was obtained for Cs4Pb(Br/Cl)6:10% Mn and Cs4PbCl6:10% Mn (Table 2), which is among the best values reported for Mn2+ emission (Supplementary Table 8)46. In addition, Mn2+ luminescence was the highest at lower (365 nm) and higher excitation energy (290 nm), for Cs4Pb(Br/Cl)6 and Cs4PbCl6, respectively, due to their suitable band gaps (determined by halide species), enabling efficient energy transfer to Mn2+. The overall superior PL QY of Mn2+-doped Cs4PbX6 colloids (X = Br/Cl, Cl) is attributed to the spatially confined zero-dimensional structure, and presence of dopant at the isolated octahedra favoring enhanced dopant emission through the localized excitons and high exciton binding energy (Supplementary Figure 11b to 11e). The reported exciton binding energies of Cs4PbBr6 and Cs4PbCl6 are 3533 and 153 eV47, respectively, compared to CsPbBr3 (19 to 62 meV) and CsPbCl3 (72 meV)48,49,50. These results suggest that the high luminescent Cs4PbX6 nanocrystals doped with Mn2+ could represent promising emitting materials for efficient solid-state lighting applications.

Table 2 PL quantum yield of the Cs4PbX6 colloids in n-octane medium

In the Cs4PbX6 solids, a similar low PL QY was observed for undoped sample, but the overall PL QY increased markedly upon Mn doping in both mixed halide and chloride samples (Supplementary Figure 11e). PL QY of 21% was obtained for Mn2+ emission in Cs4PbCl6 solids, under 290 nm excitation (Supplementary Table 9). The bright orange emission of Mn2+-doped Cs4PbX6 in their solid form, is a unique feature among Mn2+-doped perovskite solids (Fig. 4c). This remarkable PL QY and retention of dopant emission in solid form is the attributes of spatial confinement of low-dimensional Cs4PbX6 induced by isolated octahedra, eventually favoring a stable dopant emission in solids.

Fig. 4
figure 4

Lifetime of Mn2+-doped Cs4PbX6. Time-resolved PL spectra of Cs4Pb(Br/Cl)6 colloids excited at 375 nm and monitoring the D-state and Mn2+ emission at a 420 and b 600 nm, respectively. c Images of Cs4PbX6 perovskite solids under 254 and 365 nm UV lamp, which are denoted by their respective halides and Mn concentrations

To further elucidate the emission properties of Cs4PbX6 host and Mn2+, time-resolved PL results were performed by monitoring the D-state (420 nm) and Mn2+ emission (600 nm), exciting at 375 nm (Fig. 4a, b, Supplementary Figures 12 to 14, and Supplementary Note 9). A nominal decrease in the average lifetime of D-state emission in Cs4PbX6 (X = Br/Cl and Cl) was observed with increasing Mn content (Supplementary Figure 13). For instance, the average lifetime of D-state decreased from 1.08 ns (undoped) to 0.82 ns (20% Mn), corroborating the Mn2+ sensitization by the Cs4Pb(Br/Cl)6 host (Table 3). A long lifetime was obtained for the Mn2+ emission in Cs4Pb(Br/Cl)6 and Cs4PbCl6 of 1.20 ms and 1.39 ms, respectively (Fig. 4b and Supplementary Figure 13b). The lifetime of D-states of Cs4Pb(Br/Cl)6 converged with three components namely 3.7 (slow), 2.6, and 0.40 ns (last two combined is fast), similar to the previous report3. The first slow (radiative) component is assigned to D-state emission and the fast (non-radiative) component to the energy transfer process to Mn2+. A rise in the contribution of slow component with increasing Mn2+ content, agrees with the increase in D-state emission (Supplementary Figure 12). This suggests that the inclusion of chlorine species with Mn-doping (MnCl2 precursor), passivates the pre-existing defects in the Cs4Pb(Br/Cl)6 host, resulting to an enhanced host PL efficiency38. Hence, the passivation effect of chlorine in the Cs4Pb(Br/Cl)6, would contribute greatly to the PL efficiency of D-state rather than energy transfer to Mn2+ at 375 nm excitation. On the contrary, superior contribution of fast component in Cs4PbCl6, illustrates the dominant energy transfer from D-state to Mn2+ (Supplementary Figures 13 and 14) at 375 nm excitation. The chloride passivation effect in the Cs4PbCl6 may be negligible as chloride content remain unchanged with increasing dopant concentration. Therefore, the D-state emission in mixed halide Cs4Pb(Br/Cl)6, is predominated by the passivation effect of chlorine over energy transfer to Mn2+, while in the Cs4PbCl6, energy transfer to Mn2+ dominated over passivation effect.

Table 3 Average lifetime of D-state emission of Cs4Pb(Br/Cl)6 and Cs4PbCl6 monitored at 420 nm and 400 nm, respectively

The mechanism of energy transfer from host states (Pb2+ and D-state) to Mn2+ in the mixed halide perovskite is further illustrated in the Fig. 5. The high PL QY of Mn2+-doped Cs4PbX6 is the attributes of structure-induced quantum confinement with high exciton binding energy favoring enhanced emission, compared to 3D CsPbX3 perovskite (Fig. 5a). The superior PL QY of Mn2+-doped Cs4Pb(Br/Cl)6 and Cs4PbCl6 were observed at low (365 nm) and high excitation energy (290 nm), due to high absorption in their respective band gaps arising from halide identity. The Pb2+ ions are excited at high excitation energies (above 4.0 eV), leading to 375 nm emission, while D-states are excited at low excitation energies (below 4.0 eV), and both these states eventually generate Mn2+ emission through energy transfer (Fig. 5b). The presence of Mn2+ suppresses the charge transfer from isolated octahedra to the D-state, due to the weak coupling with the regular crystal lattice sites particularly at higher excitation energy. The enhanced overall PL QY of Mn2+-doped Cs4PbX6 in both colloidal and solid form is attributed to the synergistic effect of structural quantum confinement effect from the isolated octahedra and high exciton binding energies, facilitating dopant emission through enhanced energy transfer process.

Fig. 5
figure 5

Mechanism of energy transfer from host states to Mn2+. a Illustration of high exciton binding energy (at above 150 meV) due to the localized excitons formed in the isolated octahedra of 0D Cs4PbX6 (X = Cl, Br/Cl) perovskite, compared to low binding energies (less than 60 meV) of 3D CsPbX3 perovskite, where filled and hollow circles represent electrons and holes, respectively. b Configurational coordinate representation of Cs4Pb(Br/Cl)6 perovskite depicting Pb2+-based (3P0,1  →1S0 transition) and D-state emissions, and their energy transfer to the Mn2+ energy levels. The long dash and dotted arrows depicts the excitation and energy transfer process leading to emission (solid arrows)

Discussion

In conclusion, we rationalized the exceptional optical behavior of isolated octahedra in the zero-dimensional cesium lead halide structure by introducing Mn2+ dopants. Structural, PL, and lifetime results confirmed the incorporation of Mn2+ in the 0D Cs4PbX6 lattice. The Mn2+ incorporation stabilized the Cs4PbX6 structure via enhanced octahedral tilting and limited compositional variation of Cs–Pb salts, suppressing the formation of CsPbX3 impurity phase. The PL QY at a given excitation energy was determined by the energy transfer from the host states to Mn2+ and band gap of the host. A high PL QY for Mn2+ emission was achieved in the colloidal (29%) and solid (21%, powder) forms; making the solid form of Cs4PbX6 as a promising emissive layer for light-emitting devices. The enhanced PL QY of the Mn2+ emission was attributed to the synergistic effect of structure-induced spatial confinement of 0D Cs4PbX6 and electronically decoupled PbX6 octahedra favoring dopant emission via localized exciton. The present work provides deep insight into the structure and molecular behavior of 0D perovskite-like Cs4PbX6 material and opens an avenue to design low-dimensional perovskites with photo and chemical stability for high-performance optoelectronic applications.

Methods

Materials

All the reagents were used without any purification. Cesium carbonate (Cs2CO3, 99.999% Kojundo), lead (II) bromide (PbBr2, 99%, Sigma-Aldrich), lead (II) chloride (PbCl2, 98%, Sigma-Aldrich), manganese (II) bromide (MnBr2, 98%, Sigma-Aldrich) manganese (II) chloride tetrahydrate (MnCl2.4H2O, 99%, Sigma-Aldrich), oleic acid (OA, 90%, Fluka), octylamine (OcA, 97%, Sigma-Aldrich), N,N-dimethylformamide (DMF, 99.8%, Sigma-Aldrich), n-hexane (99.8%, Junsei, Japan), n-octane (90%, Sigma-Aldrich), 1-octadecene (95%, Sigma-Aldrich), hydrochloric acid (37%, Sigma-Aldrich), and hydrobromic acid (48%, Sigma-Aldrich) were employed for the synthesis.

Synthesis of undoped and Mn2+-doped Cs4Pb(Br/Cl)6 colloids

The exact compositions of the synthesized undoped and Mn2+-doped Cs4Pb(Br/Cl)6 series are Cs4PbBr2Cl4 and Cs4Pb1xMnxBr2–2xCl4+2x with x = 0.05 (5% Mn), 0.10 (10% Mn), and 0.20 (20% Mn), respectively. Mn2+-doped Cs4Pb(Br/Cl)6 nanocrystals were synthesized by a reverse microemulsion method at room temperature, as reported earlier20. PbBr2 precursor was used as the lead precursor. A simultaneous anion (Cl) and cation (Mn2+) inclusion were effected by employing MnCl2.4H2O precursor for the synthesis of Mn2+-doped Cs4Pb(Br/Cl)6. An optimal concentration of octylamine (OcA) and oleic acid (OA) surfactants were used in facilitating the solubilization of Cs, Pb, and Mn precursors. The mixed MnCl2.4H2O and PbBr2 precursors, and the Cs-oleate precursor were synthesized separately. First, a mixture of 2.25 g of Cs2CO3 and 21.5 mL of OA was stirred and degassed at 130 °C, which is further maintained under vacuum for 1 h to generate a yellowish stock of Cs-oleate precursor. Second, 0.2 mL Cs-oleate precursor, 10 mL of n-hexane, 5 mL of OA were loaded into a 50-mL three-neck flask, followed by mild degassing (vacuum) for 5 min and Argon purging for 5 min. Third, a mixture of PbBr2 and MnCl2.4H2O (DMF, 1 mL), HCl (38 wt%, 19 μL), 0.10 mL OA, and 0.05 mL OcA was swiftly injected into the flask, under vigorous stirring. Upon stirring for 10 min, white color crystals were formed suggesting the formation of Mn2+-doped Cs4Pb(Br/Cl)6. The as-synthesized nanocrystals were collected via centrifugation at 7000 rpm for 5 min followed by dispersion in 2 mL of n-octane for further characterization. The contents of Cs-oleate, HCl, and surfactants (OA, OcA) were maintained the same for varying concentration of Mn precursors.

The undoped Cs4Pb(Br/Cl)6 was synthesized following the same procedure, in the absence of MnCl2.4H2O. A color change from a pale-white to green was observed in 10 min upon the addition of Pb precursor into the Cs-oleate, which is then centrifuged and dispersed in n-octane for further characterization. N-octane was employed as the solvent for the Mn2+-doped colloids for its superior dispersibility over hexane and toluene.

Synthesis of undoped and Mn2+-doped Cs4PbX6 (X = Br, Cl)

Cs4PbX6 samples were synthesized using stoichiometric quantity of PbX2 and HX (15 μL) as the lead and halide precursors along with Cs-oleate, where X = Br, Cl. The limited solubility of PbCl2 in DMF solvent at room temperature was overcome by warming the contents until it gets dissolved. Mn2+-doped Cs4PbX6 samples were synthesized following the same procedure along with manganese precursors of MnCl2.4H2O and MnBr2, for the chloride and bromide analog, respectively. 

Synthesis of CsPbX3 (X = Br, Br/Cl, Cl) perovskites

CsPbX3 (X = Br, Br/Cl, Cl) nanocrystals were synthesized as described by Palazon et al.29. In a typical synthesis, PbX2 (0.2 mmol; X = Br, Cl, Br/Cl), 0.47 mL oleic acid, 5.0 mL octadecene, and 0.50 mL oleylamine were loaded in a 25 mL 3-neck flask and dried under vacuum at 100 °C for 15 min. The temperature was raised to 190 °C, after degassing and 0.5 mL of Cs-oleate solution (made from 0.4 g of Cs2CO3 dissolved in 11.75 g of octadecene, and 1.55 g oleic acid at 100 °C under vacuum), which was pre-heated on a hot-plate at 100 °C was swiftly injected. Five seconds after the injection, the nanocrystal solution was quickly cooled to room temperature with an ice bath and by the addition of 5 mL of toluene (at 100 °C). The nanocrystals were collected by centrifugation (2500 rpm/2 min.) and re-dispersed in 5 mL toluene. The composition of (Br/Cl) analog is CsPb(Br0.4/Cl0.6)3, where a mix of 0.08 mmol of PbBr2 and 0.12 mmol of PbCl2 precursor were employed for the synthesis.

Synthesis of zero-dimensional Cs4PbX6 solids

The as obtained colloidal samples dispersed in n-octane were centrifuged and re-dispersed in tert-butanol and centrifuged at 7000 rpm for 5 min. The samples were then dried at 80 °C overnight to get respective Cs4PbX6  solids. The body color of the zero-dimensional solids were pure white for the Cs4PbCl6, Cs4PbCl6:Mn, and Cs4Pb(Br/Cl)6:Mn, except Cs4Pb(Br/Cl)6 (pale greenish white), and Cs4PbBr6 (bright yellow).

Characterization

The synthesized zero-dimensional Cs4PbX6 nanocrystals were characterized by powder X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), electron spin resonance spectra (ESR), inductive coupled plasma-optical emission spectroscopy (ICP-OES), field-emission scanning electron microscope (FE-SEM), and high-resolution tunneling electron microscope (HR-TEM). The XRD was performed using a Philips X’Pert diffractometer with Cu Kα radiation, in the range of 10° to 120° with a step size of 0.026° with the synthesized powders. The structural information was derived from Rietveld refinement using the General Structure Analysis System (GSAS) software suite. The VESTA program was used to draw the crystal structure51. The phase fractions of multiple phases were estimated using Rietveld refinement of XRD results considering full refinement of crystallographic and instrumental parameters as implemented in the GSAS program suite. The FE-SEM and HR-TEM images were recorded using a Hitachi S-4700 and FEI Tecnai F20 (200 kV), respectively, at Korea Basic Science Institute (KBSI), Gwangju, South Korea. The sample for TEM measurements were prepared by dispersing the nanocrystals in the ethanol medium. It is to be noted that the undoped and Mn2+-doped Cs4PbX6 perovskite were unstable upon exposure to high-energy electron beam and tend to damage within short period of time. XPS was performed on the powder samples using VG Multilab 2000 instruments to identify the chemical state of the elements on the sample surface (below 10 nm), employing Al as the target anode for the X-ray generation. The energy of the X-ray photon (Al Kα) is 1480 eV with a line width of 0.5 eV. The chemical compositions of the synthesized samples were analyzed using ICP-OES (Perkin-Elmer, OPTIMA 8300), at KBSI, Gwangju, South Korea. ESR measurement was carried out using JEOL JES-FA200 with an X-band microwave frequency of 9.17 GHz, with sample dispersion in n-octane. The liquid samples were taken in an ESR tube and frozen in liquid nitrogen at 173 K. All the spectra were recorded using the following ESR parameters: microwave power, 1 mW; modulation amplitude, 0.2 mT; modulation frequency, 100 kHz; and sweep time, 120 s.

Photoluminescence was measured using a Hitachi F-4500 fluorescence spectrophotometer over the wavelength range of 200 to 750 nm, using n-octane dispersions in a 1×1 cm quartz cuvette. UV-vis spectra were obtained using Optizen POP UV/Vis spectrophotometer. The lifetime was measured using a time-correlated single photon counting (TCSPC) system on FL920 Edinburgh instruments at Korean Advance Institute of Science and Technology (KAIST), South Korea. Pulsed laser irradiation with 300 nm was used as the excitation source. Internal quantum efficiency was measured with 365 and 290 nm excitation using a xenon laser (Hamamatsu C9920-02) at the Korea Photonics Technology Institute (KOPTI), South Korea. Time-resolved photoluminescence measurement for Mn2+ (600 nm) and D-state emission (420 nm for Cs4Pb(Br/Cl)6 and 400 nm for Cs4PbCl6) was carried out using single-mode pulsed diode laser as the excitation source (λmax = 375 nm) at Korea Basic Science Institute (KBSI), Daegu Center, South Korea.