Synthesis and characterization
Rare earth RE3+ (RE = Ho, Er, Tm, Nd) ions doped Cs4Mn1 − wCdwBi2Cl12 crystals were synthesized by the hydrothermal method as well as by changing the concentration of RE3+ and stoichiometry of Mn/Cd in the precursor. Details can be found in the Supporting Information (SI). Considering the same valence (+ 3) and similar ionic radii of Bi3+ and Ho3+, Tm3+, Er3+, Nd3+, these RE3+ ions tend to occupy the lattice of Bi3+ ions in the rhombohedral structure of Cs4MnBi2Cl12, forming the [Ho/Tm/Er/NdCl6] octahedra (Fig. 1a). In this structure, each [BiCl6]3− octahedron is surrounded by three Mn2+-centered octahedra. The crystal structures of Cs4MnBi2Cl12 and Ho3+-doped Cs4MnBi2Cl12 were examined using high-resolution transmission electron microscopy (HRTEM), selected-area electron diffraction (SAED), and X-ray diffraction (XRD). The HRTEM images of Cs4MnBi2Cl12 and Ho3+-doped Cs4MnBi2Cl12 display clear lattice fringes with lattice spacing of 3.82 Å and 3.90 Å, respectively, which could be indexed as crystal plane (2, -1, 0) of rhombohedra of Cs4MnBi2Cl12 phase (Fig. 1b and 1c). The increase in the interplanar spacing value of the (2, -1, 0) plane from 3.82 Å to 3.90 Å further confirms the lattice expansion after Ho3+ introducing into host. The SAED patterns of Cs4MnBi2Cl12 and Ho3+-doped Cs4MnBi2Cl12 show very sharp diffraction spots (Fig. 1d and 1e), proving the single crystal characteristics for the studied samples. XRD pattern of Ho3+-doped Cs4MnBi2Cl12 proves the phase purity (Figure S1). The actual RE3+ doping concentration in Cs4MnBi2Cl12:xRE3+ (RE = Ho, Er, Tm, Nd) can be estimated by using inductively coupled plasma optical emission spectrometer (ICP-OES). The doping concentrations were found to deviate considerably from the nominal content of the Cs4MnBi2Cl12 precursor (Table S1), as was previously reported for RE3+-doped compounds.31,38,39 Scanning electron microscopy mapping was applied to analyze the composition of each compound (Figure S2). All the elements are evenly overspread in the images, with no element enrichment, indicating that there is no phase separation in the as-synthesized perovskite materials. This result implies the successful doping of RE3+ ions into the Cs4MnBi2Cl12 host lattice. The transparent Cs4MnBi2Cl12 and Ho3+-doped Cs4MnBi2Cl12 single crystals exhibit intense orange and orange-red emission upon UV light illumination (Fig. 1f and 1g). Figure 1h depicts the normalized Ho L3-edge extended X-ray absorption fine structure (EXAFS) spectra of Ho3+-doped Cs4MnBi2Cl12. The EXAFS oscillations are clearly observed by the function spectrum k3χ(k) (Fig. 1i). Figure 1j shows the Fourier transformed Ho L3-edge EXAFS spectra k3χ(k). By fitting within the allowable margin of error, the bond length and coordination information of Ho3+ in Cs4MnBi2Cl12 were obtained. EXAFS fitting parameters at the Ho L3-edge is provided in the Table S4. As expected, the coordination number of Ho is ~ 6, which coordinated with six chlorine atoms. The bond length of Ho-Cl is ~ 2.58 Å, which is closer to the Bi-Cl bond length, confirming that Ho3+ is replacing Bi3+ rather than Mn2+ in Cs4MnBi2Cl12. Similarly, Er3+, Tm3+, Nd3+ ions can also be doped into this host, and Raman spectroscopy shows that the doping does not induce phase transitions and additional vibrational modes (Figure S3). From the electron paramagnetic resonance (EPR) spectra of Cs4Mn0.23Cd0.77Bi2Cl12, Cs4MnBi2Cl12 and Ho3+-doped Cs4MnBi2Cl12, a single broad line is observed (centered around 3480 G, effective g value g ~ 2.0, Figure S4) due to the high Mn concentration. This can be attributed to the fact that Mn ions are coupled by dipolar or exchange magnetic interactions. In this case, the hyperfine sextet disappears due to fluctuations of the dipolar or exchange interactions.
Steady-state absorption and PL properties
The similarity of absorption spectra (250–400 nm) of Cs4CdBi2Cl12, Cs4MnBi2Cl12, Ho3+-doped Cs4MnBi2Cl12 confirms that three compounds can be considered as insulating hosts. Their optical properties are controlled mainly by local electronic transitions associated with isolated [BiCl6]3− centers (Fig. 2a). Interestingly, Ho3+-doped and undoped Cs4MnBi2Cl12 give rise to intense new absorption with peak at ~ 430 (6A1g→4T2g) and ~ 520 nm (6A1g→4T2g) compared with Cs4CdBi2Cl12. The new absorption should arise from the contribution of Mn2+ orbitals at the band edges. The obtained light-orange crystals of Ho3+-doped Cs4MnBi2Cl12 show a bright orange-red emission upon 365 nm UV light excitation, similar to pure Cs4MnBi2Cl12 (Fig. 2b). Figure 2b and 2c show PL spectra of Cs4CdBi2Cl12, Ho3+-doped and undoped Cs4MnBi2Cl12 in VIS and NIR region, respectively. Besides the broad-band emission derived from Mn2+ (4T2g→6A1g), there is a sharp emission peak located at 645 nm, corresponding to the 5F8→5I8 transition of Ho3+. In addition to the characteristic emission in VIS region, RE3+ ions usually present rich PL spectra in NIR region. Obvious emission around 985, 1194 and 1488 nm correspond to the transitions of Ho3+ 5F5→5I7, 5I6→5I8, and 5F5→5I6, respectively (Fig. 2c). From the PL decay curves of Ho3+-doped Cs4MnBi2Cl12 in Fig. 2d, the average PL lifetime are 0.058, 1.246, 1.192 ms monitored at 596, 645, 985 nm, respectively.
After monitoring the PL peak at 596, 645, 985 nm, the Ho3+-doped Cs4MnBi2Cl12 presents similar PLE spectrum profile (Fig. 2e). This is composed of a strong broad excitation from 250 to 400 nm and two distinct excitation peaks at ∼430 and ∼520 nm associated with spin-forbidden d − d transition of octahedrally coordinated high-spin Mn2+. In fact, the above PLE spectra resemble the PLE spectrum monitored at 596 nm without Ho3+ emission, indicating the existence of an energy transfer channel from Cs4MnBi2Cl12 to Ho3+. Beyond that the PL intensity of RE ions with f–f transition is usually sensitive to the excitation wavelength. However, with the excitation wavelength varying from 310 to 430 nm, the relative PL intensity between Ho3+ and Mn2+ emission is almost constant (Figure S5). Accordingly, the indirect excitation may be the dominant excitation resource for RE3+ ions, with an optimal excitation wavelength of ~ 370 nm. Ho3+ ions are not directly excited but excited via an energy transfer channel from the Cs4MnBi2Cl12 host. Because f–f transition of RE3+ ions is forbidden and thus incapable to compete with the excitation of perovskite host. The fact that the radiation light only excites the host but the characteristic emission of Ho3+ ions directly confirm the existence of energy transfer. On the contrary, no characteristic emissions are observed in Ho3+-doped Cs4CdBi2Cl12, indicating that Ho3+ will emit red-NIR light when Mn2+ acts as a bridge to transfer energy to Ho3+, as shown in Fig. 2f.
Er3+/Tm3+/Nd3+ - singly doped Cs4MnBi2Cl12 samples also exhibit their f-f characteristic emissions in the VIS and NIR region. Figure S6 illustrates the PLE/PL spectra of RE3+-doped Cs4MnBi2Cl12 (RE = Er, Tm, Nd) and AM 1.5 solar spectrum over wavelengths of 280–1650 nm. The PLE spectrum of Er3+/Tm3+/Nd3+-doped Cs4MnBi2Cl12 by monitoring the PL peak at VIS and NIR range are similar to that of monitored at 596 nm without Er3+/Tm3+/Nd3+ emission, implying the existence of the energy transfer channel from Cs4MnBi2Cl12 to Er3+/Tm3+/Nd3+. From the PL spectrum of Er3+-doped Cs4MnBi2Cl12, besides the broad emission from Mn2+, the sharp emission peaks located at 672, 983, 1158, 1540 nm, correspond to the transitions 4F9/2, 4I9/2, 4I11/2, 4I13/2→4I15/2 of Er3+, respectively. The sharp emission peaks at 694, 802, 1211, 1432 nm from Tm3+-doped Cs4MnBi2Cl12 are ascribed to the transitions 3F3, 3H4, 3H5→3H6, 3H4→3F4 of Tm3+, respectively. The sharp emission peaks located at 886, 1061, 1344 nm from PL spectrum of Nd3+-doped Cs4MnBi2Cl12, correspond to the transitions 4F3/2→4I9/2, 4I11/2, 4I13/2 of Nd3+, respectively. Moreover, the PLE spectrum consists of several bands in the UV–VIS range, which is consistent with the maximum photon flux region of the solar spectrum. In addition, RE3+-doped Cs4MnBi2Cl12 (RE = Ho, Er, Tm, Nd) give an intense orange-red emission in 500–780 nm and NIR emission from 800 to 1650 nm, matching well with the optimal spectral response of solar cells. Therefore, RE3+-doped Cs4MnBi2Cl12 materials are promising solar spectral converters by achieving efficient NIR emission for solar cells. RE3+-doped Cs4MnBi2Cl12 (RE = Ho, Er, Tm) with different concentrations were also prepared for further investigated. XRD patterns of RE3+-doped Cs4MnBi2Cl12 demonstrate pure phases for all syntheses (Figure S7a-S7c). As RE3+ ions concentration increases, the characteristic emission bands of RE3+ ions gradually become obvious, which not only supplement the red-light component, but also achieve full-spectrum NIR emission (Figure S7d-S7e). The photoluminescence quantum yield (PLQY) for Cs4MnBi2Cl12 is ~ 26% under 370 nm UV light excitation, and RE3+ doping can increase the PLQY to 35–59% owing to the superposition of the additional RE3+ emissions (Figure S7f-S7i). A schematic variation of emission color in the Commission Internationale de l´Eclairage (CIE) diagram by a change in the Ho3+ concentration depicts a color tuning in orange-red region (Figure S8a). The decreased PL lifetime decay curves of Mn2+ emission in Ho3+-doped Cs4MnBi2Cl12 with different Ho3+ concentrations substantially prove the existence of energy transfer (Figure S8b). Energy transfer from halide perovskite hosts to RE ions may provide new avenues for achieving some exciting optical properties and goals.
Woodward et al. reported that replacing Cd2+ with Mn2+ to form Cs4Mn1 − wCdwBi2Cl12 inhibits energy migration to defect sites where nonradiative decay can occur.40 The PL intensity in the Cs4Mn1 − wCdwBi2Cl12 prepared via precipitation reaction reaches a maximum near w = 0.73, where each [BiCl6]3− sensitizer is approximately on average adjacent to one Mn2+ activator, and the PLQY reaches 57%. Here, Cs4Mn1 − wCdwBi2Cl12 (w = 0–1) prepared by hydrothermal method exhibits a maximum PLQY of 93% when w = 0.77. XRD patterns indicate pure phases for all syntheses (Figure S9a). The broadening of 6s2→6s1p1 transition is attributed to the increase of electronic dimensionality with enhancing Mn incorporation (Figure S9b). The bands at 430 and 520 nm are associated with the spin-forbidden 6A1(S)→4T2(G) and 6A1(S)→4T1(G) transitions, respectively. These transitions are forbidden in octahedral coordination, but magnetic coupling between Mn2+ pairs can relax the parity selection rules. The PLE intensity of the both bands steadily increase as increases Mn content due to the larger amount of Mn2+-coupled pairs. The overlap of absorption spectra indicates that Mn2+ emission originates from the absorption of host (Figure S9d). The PL spectra of the studied Cs4Mn1 − wCdwBi2Cl12 (w = 0–1) samples reach a maximum near 596 nm (Figure S9c). The emission position hardly changes with increasing w, and the maximum PL intensity is observed when 77% Mn2+ ions were replaced by Cd2+. In Cs4MnBi2Cl12, each Bi3+ ion has three Mn2+ adjacencies. The relatively low PLQY of ∼26% measured for the w = 0 sample suggests that energy transfer to the killer site is too easy in Cs4MnBi2Cl12 end member (Figure S9e-S9f). Incorporation of Cd2+ acts as electron-isolation of Mn − Bi − Mn network, which increases radiative recombination at Mn2+ site. It is worth noting that w = 0.67 is quite close to the replacement level where the PL is strongest. On average, this means that each Mn2+-centered octahedron neighbors one Bi3+-centered octahedron. These results could help us to conclude that the optimal structure puts a single Mn2+ ion next to each [BiCl6]3− octahedron.
Optical and electronic bandgap tailoring
To search for a suitable metal halide perovskite for achieving excellent optoelectronic properties in the VIS and NIR region, the optical properties of RE3+ doping in the Cs4Mn1 − wCdwBi2Cl12 have been investigated in detail. XRD patterns reveal the pure phase products for all syntheses (Figure S10). Variations in the PL spectra of Cs4Mn1 − wCdwBi2Cl12:RE3+ (w = 0.1, 0.3, 0.5, 0.7, 0.9) samples depending on the w values (different Cd/Mn ratios) are shown in Fig. 3a-3c for RE = Ho, Er, Tm, respectively. The PL intensity of RE3+ characteristic emission bands for all the samples gradually decrease with increasing w values. While the PL intensity of Mn2+ reaches a maximum when w = 0.7, which is consistent with the observation in RE3+-undoped Cs4Mn1 − wCdwBi2Cl12. The gradual weakening of PL intensity of RE3+ characteristic emission bands ascribed to the indirect-to-direct bandgap transition is observed. The optical bandgap properties of Cs4Mn1 − wCdwBi2Cl12 were measured by UV-VIS diffuse reflectance (DR) spectroscopy. The experimentally obtained indirect and direct bandgaps are shown in Fig. 3d. Cs4MnBi2Cl12 has the characteristics of an indirect band gap semiconductor, with a shallow absorption at 2.72 eV (indirect band gap) and a large absorption at 3.00 eV (direct band gap). Moreover, the linear region of the Tauc plot of (F(R)hv]1/2 versus hv reveals the expected phonon-assisted processes, with phonon absorption and emission transitions occurring at 2.72 and 3.00 eV, respectively. The indirect band gap of this material is estimated as 2.72 eV. With introducing Cd2+ into Cs4MnBi2Cl12, the absorption edge becomes larger. Cs4CdBi2Cl12 possesses a characteristic of direct band gap with absorption edge at 3.40 eV, according to the Tauc plot. Schematic illustration of possible electronic dual-bandgap structure for Cs4CdBi2Cl12 and Cs4MnBi2Cl12 is displayed in Figure S11. To investigate the electrical and optical contributions of the two end members in Cs4Mn1 − wCdwBi2Cl12, we performed the first-principle calculations for Cs4MnBi2Cl12 and Cs4CdBi2Cl12. Figure 3e and 3f show the calculated band structures of original Cs4CdBi2Cl12 and Cs4MnBi2Cl12. For Cs4CdBi2Cl12, the valence band maximum (VBM) mainly consists of the Cl 3p orbitals, the conduction band minimum (CBM) is mainly composed of Bi 6p orbitals. They are both located at B point, leading to a direct band gap of 3.0 eV, which is smaller than experimental measurements of 3.40 eV. The energy difference between experimental observations and theoretical calculations is understandable because density functional theory-Perdew-Burke-Ernzerhof (DFT-PBE) calculations underestimate the band gap. Nevertheless, such difference does not affect our qualitative analysis on the electronic and optical properties of Cs4Mn1 − wCdwBi2Cl12. For Cs4MnBi2Cl12, the VBM and CBM mainly originate from the Mn 3d and Bi 6p orbitals, respectively (Fig. 3f). The p–d repulsive force between the Mn 3d and Bi 6p orbitals pushes the CBM state at G point to a lower energy, leading to an indirect band gap of 1.9 eV.
The projected density of states (PDOS) for Cs4CdBi2Cl12 shows overlap of Cl 3p in the deeper valence and conduction band states (Figure S12a). A large overlap of Bi 6p orbitals in the conduction band range can be observed. For Cs4MnBi2Cl12, the states near both the VBM and CBM are both contributed by the Mn states (Figure S12b). Such change in the band edge states is expected to change the properties of optical transitions close to the band gap energy. The CBM states in Ho3+-doped Cs4MnBi2Cl12 (Fig. 3g) are mainly composed of Bi 6p, Mn 3d and Ho 3d states, while the CBM states in Ho3+-doped Cs4CdBi2Cl12 (Fig. 3h) are mainly contributed by the Bi 6p orbitals with neglectable amount of Ho 3d states. Actually, the Bi 6p and Mn 3d states are mixed into the unoccupied Ho 3d state, while the Cl 3p is mixed into the occupied Ho 3d state. Thus, the Bi 6p and Mn 3d states can contribute to the PL process by the optically allowed transitions to the Cl 3p state. On the contrary, for Ho3+-doped Cs4CdBi2Cl12, the optical transition from the Bi 6p state mixing with unoccupied Ho 3d to the Cl 3p state mixed into the occupied Ho 3d state is neglectable because of its parity forbidden nature (Fig. 3h). Moreover, the amount of Ho 3d state mixed into the Bi 5p state is very small, resulting in the loss of PL intensity (Fig. 3a). For Cd/Mn solid solution, the CBM is mainly contributed by Mn 3d rather than Cd 5s, meaning that the optical properties of solid solution are mainly dominated by Mn constituent.
Energy transfer mechanisms
The luminescence mechanisms of Cs4Mn1 − wCdwBi2Cl12 have been discussed in previous reports.27,41,42 Du et al. considered that the weak PL of Cs4CdBi2Cl12 is far from the corresponding band-edge emission suggesting the defect emission.43 Kuang et al. recently reported the crystal structure of Cs4MnBi2Cl12, which presents a typical red emission of Mn2+ at 610 nm.44 Each BiCl63− octahedron is surrounded by three Mn2+-centered octahedra. Diluting the concentration of Mn2+ by alloying with Cd2+ in Cs4Mn1 − wCdwBi2Cl12 results in a significant enhancement of PL intensity. Cd: Mn = 2:1, namely, one Mn2+-centered octahedron neighbors each Bi3+-centered octahedron, is quite close to the substitution level with the strongest PL intensity. Schematic diagram of local structure and luminescence mechanism in RE3+-doped Cs4Mn1 − wCdwBi2Cl12 is proposed in Fig. 4a. Although the strongest PL intensity of Mn2+ is the sample at Cd:Mn = 2:1, the transferred energy to the RE3+ is very small due to the less Mn2+ energy bridge, resulting in a weak RE3+ luminescence. In RE3+-doped Cs4MnBi2Cl12, [BiCl6]3− will efficiently transfer energy to RE3+ through the more bridge function of [MnCl6]4−. Figure 4b describes the process of optical absorption and emission in RE3+-doped Cs4MnBi2Cl12. [BiCl6]3− octahedron plays an important role as a UV light photosensitizer in exciton production. The Mn2+ in [MnCl6]4− octahedron is both the exciton acceptor and emission center to afford the intense PL via 4T1g→6A1g transition. The exciton transfer process from [BiCl6]3− octahedron to [MnCl6]4− octahedron effectively increases the exciton density in 4T1g, and increasing PL quantum yield. The excitation energy is then transferred to the excited state energy level of RE3+. RE3+ f-electrons then de-excite via multiple transitions and give transition emission. Generally, defects in the Cs4CdBi2Cl12 capture energy from [BiCl6]3−, and most of the energy is released through non-radiative transitions, which makes it impossible to further transfer the energy to RE3+ ions. In addition, the energy levels of [BiCl6]3− and [CdCl6]4− do not match the energy levels of RE3+, resulting in RE3+ ions with forbidden f-f transitions cannot emit light in the Cs4CdBi2Cl12 host. On the contrary, as Mn2+ replaces Cd2+, the energy level of Mn2+ in [MnCl6]4− octahedron can efficiently match the energy level of RE3+, which can better absorb the energy from Mn2+ to produce light. Under n-UV light excitation, the electrons in the ground state levels of [BiCl6]3− are excited to the excited state levels, then the excited electrons release energy through two energy transfer processes. Firstly, the energy in excited energy levels could be transferred to 4T1g energy levels of Mn2+ and return back to the ground 6A1g energy level with 596 nm orangish-red light. Secondly, the energy in excited 4T1g energy levels of Mn2+ could be transferred to 5F5, 4F9/2, 3F3 energy levels of RE3+ (RE = Ho, Er, Tm). Compared with the energy level of [CdCl6]4−, the lowest energy level of Mn2+ easily transfers energy to 5F5, 4F9/2, 3F3 energy levels of RE3+ (RE = Ho, Er, Tm), as a result, the [BiCl6]3−→Mn2+→RE3+ energy transfer occurs by Mn2+ energy bridge. Moreover, as the Mn2+ content increases, the energy transfer efficiency is becoming higher and reaches the highest when it is completely replaced (w = 0), and the strongest RE3+ luminescence is achieved.
Steady-state temperature-dependent PL properties
Photophysical mechanisms of RE3+-doped Cs4MnBi2Cl12 were further experimentally investigated via steady-state temperature-dependent PL spectra. To better illustrate the temperature-dependence of emission, the contour plots are shown in Fig. 5a-5c and Figure S13. Generally, the PL intensity of Mn2+ (4T1g→6A1g) decreases with increasing temperatures for three RE3+-doped Cs4MnBi2Cl12 samples because of an enhancement of non-radiative transition probability. This is ascribed to the enhanced thermal vibration of the matrix lattice at high temperature, the increase of thermally activated phonons and the strengthened interaction between electrons and phonons.45,46 Based on previous studies, the relationship between PL intensity (Ii) and temperature (T) is expressed as follows:47,48
here N(T) is the excited states’ population at a certain temperature T, τ denotes the lifetime of excited states, P represents the phonon number, ℏω is the phonon energy, R and WNR are the radiative and non-radiative transition rate, respectively. After Eq. 3 is merged into Eq. 1, the equation is expressed as:
where ΔE is the energy gap between the excited and ground states. It is clear from Eq. 4 that the PL intensity is mainly affected by two processes: non-radioactive decay and thermal agitation. In the low temperature range, the thermal distribution rate is high while the non-radiation rate is low, resulting in an increase of PL intensity with increasing temperatures. Nevertheless, the non-radiation rate rapidly increases in the high temperature range and the radiation probability significantly decreases. Thus, the PL intensity decreases at high temperatures. In addition, there is a clearly observable difference in temperature-dependence between these RE3+ and Mn2+ emission peaks. The RE3+ emission perform a weaker thermal quenching compared with Mn2+ emission. This suggests that the 4f electron of RE3+ ions is well screened from the surrounding defect sites, through the closed 5s25p6 outer shell electrons of RE3+.32 The specific thermal quenching trend of these emission peaks at VIS and NIR regions is described in Fig. 5d-5f. For Ho3+-doped Cs4MnBi2Cl12, the integrated PL intensity of Ho3+ emission (5F6→5I8, 5F6→5I7) increases with maximum PL intensity at 250 K and then decreases throughout the temperature range of 100 to 400 K due to the associative effects of nonradioactive decay and thermal agitation. For Er3+-doped Cs4MnBi2Cl12, there are three energy transitions of NIR emission for Er3+: 4I9/2→4I15/2, 4I11/2→4I15/2 and 4I13/2→4I15/2. Their integrated PL intensity does not change much in the temperature range from 100 to 250 K, and then declines rapidly with increasing temperatures. There is an obvious temperature-dependent difference between NIR and VIS emission peaks, and VIS emission exhibits a stronger thermal quenching. This is largely attributed to the photon energy at VIS region is much higher than that at NIR region. According to Eqs. 1 − 4, the non-radiative decay and quenching activation energy are more significant at VIS region. As a result, the PL intensity at VIS region reduces more strongly with increasing temperatures. For Tm3+-doped Cs4MnBi2Cl12, a temperature-dependent difference among these NIR emission peaks is clearly observed. The PL thermal quenching from 3H5→3H6 is particularly strong. This is attributed to the phenomenon of phonon-assisted population inversion for Tm3+ in Cs4MnBi2Cl12, increasing the population of the 3H5 state with the increase of temperature (100–450 K). In addition, the energy transfer should be strongly dependent on the temperature due to the energy difference between excited states of interacting ions. Mn2+-Tm3+ energy transfer makes the weaker thermal quenching from 3H4→3H6. The integrated PL intensity of Tm3+ emission (3H4→3H6) increases with maximum intensity at 300 K and then decreases due to the cooperative effects of nonradioactive decay and thermal agitation. Moreover, RE3+-doped Cs4MnBi2Cl12 single crystals exhibit outstanding stability toward ambient conditions. Their excellent structural stability is proved by no distinct decompositions in XRD patterns (Figure S14a) under storage at ambient conditions for 200 days. The stability of these single crystals in various organic solvents was also examined. As shown in Figure S14b, they remained highly stable in ethanol (EtOH), isopropanol (IPA), toluene (PhMe), Acetone (ACE) and hexyl hydride (HDI), and even for 100 days, which is confirmed by the XRD patterns (Figure S14c). Due to the ionic nature of Cs4MnBi2Cl12, the emission faded when exposed in dimethyl sulfoxide (DMSO) and N,N-dimethylformamide (DMF) for 10 h.
Fabrication and performance of pc-LEDs
In order to investigate the practical performances in LED applications of RE3+-doped Cs4MnBi2Cl12, pc-LEDs were fabricated by assembling the optimized samples with commercially available 400 nm LED chips. Firstly, RE3+-doped and undoped Cs4MnBi2Cl12 (RE = Ho, Er, Tm) samples with and without the commercial blue BAM:Eu2+ and green (BaSr)2SiO4:Eu2+ phosphors were constructed. The single-color pc-LEDs of RE3+-doped and undoped Cs4MnBi2Cl12 emit intense orange-red/red light (Fig. 6a) with the CIE xy coordinates of (0.576, 0.399), (0.589, 0.394), (0.579, 0.391) and (0.574, 0.399) for without and with RE = Ho, Er, Tm doping, respectively (yellow symbol in Figure S15). Four fabricated pc-WLEDs of RE3+-doped and undoped Cs4MnBi2Cl12 produce warm white light with full-spectrum covering the whole VIS region (Fig. 6b). The CIE xy coordinates are (0.349, 0.394), (0.384, 0.395), (0.358, 0.392) and (0.352, 0.376), respectively (white symbol in Figure S15).
These four fabricated pc-WLEDs exhibit low correlated color temperature (CCT) of 4999, 4033, 4732, 4870 K and high CRI (Ra = 83.3, 93.0, 87.9, 88.3) for undoping, doping with RE = Ho, Er, Tm, respectively. It is worth noting that the RE3+ doping results in the decreased CCT value and increased CRI value, because the contribution of red component is greater than that of orange-red component in pc-WLEDs. As shown in the images, the working devices based on the mixed luminescent materials emit bright warm white light, which demonstrates that RE3+-doped Cs4MnBi2Cl12 samples are promising for lighting applications. Secondly, the full-spectrum from VIS to short wavelength infrared (VIS-SWIR) LED was designed by RE3+-codoped Cs4MnBi2Cl12 (RE = Ho, Er, Tm) samples and the PL spectrum is given in Fig. 6c. This PL emission spectrum is composed of a broad emission band at 596 nm due to the 4T2g→6A1g transition of Mn2+ and other narrow emission band at 645, 802, 985, 1158, 1432, 1540 nm due to 5F5→5I8 (Ho3+), 3H4→3H6 (Tm3+), 5F5→5I7 (Ho3+), 4I11/2→4I15/2 (Er3+), 3H4→3F4 (Tm3+), 4I13/2→4I15/2 (Er3+) transitions. Due to the unique spectral characteristics of SWIR light, such as invisible to the naked eye and special penetration ability, the fabricated SWIR LED can be used for night vision surveillance and penetration applications. Under a fluorescent light, the part of the emission hand covered by the 750 nm filter is invisible on viewing by a visible camera (Fig. 6d-i). When the NIR pc-LED is added and turned on, the obscured hand is clear in the NIR camera. The NIR emission from the pc-LED can pass through the filter. Furthermore, the NIR camera can also detect the tissue damage in fruits, which is difficult to detect by the naked eye (Fig. 6d-ii). NIR spectroscopy is considered as an effective method for qualitative and quantitative analysis in the food and agricultural industries due to the different NIR optical absorption responses.49 To demonstrate the application prospect of NIR spectral analysis, the NIR PL spectra penetrated by NIR light emitted by RE3+-codoped Cs4MnBi2Cl12 (RE = Ho, Er, Tm) were measured with water and ethanol as representatives (Fig. 6e). Clearly, this NIR-emitting phosphor shows a characteristic NIR absorption response. Three absorption band ranges located at around 950–1000, 1150–1250 and 1400–1550 nm can be detected (Fig. 6f). They correspond to the characteristic overtones of the O–H stretch and C-H stretch according to the absorption spectrum of water and ethanol, respectively.