Introduction

The development of organic-metal halide perovskite materials with the formula ABX3, where A is generally either methylammonium (MA) or formamidinium (FA), B is Pb or Sn, and X is I, Br, or Cl, has rapidly progressed in recent years. These materials have been employed in a wide array of applications, such as in solar cells1,2,3, photodetectors4, light-emitting diodes (LEDs)5, transistors6, and lasers7, owing to their excellent optoelectronic performances, such as adjustable bandgaps8, high absorption coefficients9, and long charge-carrier diffusion lengths10. The power conversion efficiencies of perovskite solar cells have significantly increased in recent years, and a high power conversion efficiency of 25.2% was obtained last year for a solar cell based on a structurally optimized organic-metal halide perovskite11. Recently, halide alloying engineering has been developed as a critically important strategy for promoting the performance of optoelectronic devices based on organic-metal halide perovskite materials12. This strategy has been particularly applied to mixed-halide APbI3xClx perovskite materials13,14,15,16, and progress in this area has rapidly developed. This strategy has been demonstrated to not only improve the optoelectronic properties of the resulting perovskite materials but also enhance their structural stability and even be applicable to bandgap modulation17,18. However, despite this promising progress, a number of fundamental issues regarding the halide alloying engineering process remain unresolved. A particularly important issue involving how Cl atoms are incorporated into MAPbI3xClx materials as a halide alloying agent is a subject of strong debate19,20.

On one side of the above-discussed debate, many reports have provided evidence indicating that no Cl ions are embedded in the inorganic PbI3 lattice. This suggests that Cl ions influence the optoelectronic properties of MAPbI3xClx perovskites only in an indirect way, and numerous indirect enhancement effects have been proposed. For example, Chen et al.19 attributed the optoelectronic performance enhancement of MAPbI3xClx perovskites to the improved heterojunction interfaces induced by the Cl ion incorporation process. Lian et al.20 demonstrated that Cl ion incorporation could improve the crystallinity of MAPbI3xClx perovskite materials and in turn increase the crystal growth rate. However, no Cl ions were detected in the crystal lattice. Fan et al.21 proposed that residual PbCl2 passivated defects in MAPbI3xClx perovskite thin films and thereby improved their electrical properties. On the other side of the debate, a similar number of reports have provided evidence that Cl ions can be embedded in the inorganic PbI3 lattice. For example, Zhang et al.22 demonstrated the presence of Cl ions in the inorganic lattice of MAPbI3xClx nanofibers fabricated using a solution-assembly method. Zhang et al.23 reported that Cl can be maximally doped into the MAPbI3xClx perovskite crystal lattice using an ion exchange−decomposition process. Min et al.24 demonstrated that Cl ions can be introduced into the FAPbI3 crystal lattice through the substitution of FA sites with methylenediammonium dichloride ions. In addition, the possible crystal structures of MAPbI3xClx perovskite materials have been investigated by ab initio computer simulation studies, and the potential optoelectronic properties of the proposed structures have been evaluated25,26. However, previous studies mainly focused on perovskite films, whose lack of stability and the difficult distinction between the surface and bulk are great challenges for accurate detection by various instruments. Moreover, few studies have rigorously examined the I/Cl alloying structures and their long-term stability.

The present study seeks to address this long-standing controversy by presenting a detailed study of the I/Cl alloying structures observed in a series of MAPbI3xClx (x = 0, 0.05, 0.1, 0.3, 0.4, and 0.5) single crystal perovskite samples over time. Here, single-crystal samples are much more ideal than powder samples for the proposed study owing to their greatly reduced crystal defect concentrations and fewer crystal boundaries10,20. The structure of the single crystals is observed to change from tetragonal to cubic with increasing x. Characterization of the samples by means of X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), 207Pb nuclear magnetic resonance (NMR), and 2H NMR reveals that a small amount of Cl can be incorporated into the iodide sites of the inorganic PbI3 lattice, leading to a phase transition from the tetragonal to cubic phase and a series of anomalous cation dynamics evolutions including a slower rotation rate and a more disordered dynamics tendency. Analyses based on time-dependent XRD, 207Pb NMR, and 2H NMR indicate that the I/Cl alloying structures of the MAPbI3xClx crystals are metastable and disappear over time at room temperature. In addition, we explored the photoelectric properties of MAPbI3xClx single-crystal photodetectors and found that the photocurrent response obviously decreases over time, even though Cl incorporation helps boost the photocurrent response. This long-term instability indicates that different conclusions regarding the effect of Cl ions on the properties of MAPbI3xClx perovskites may arise over different time scales, which provides a plausible explanation for the long-standing controversy regarding the effect of Cl ions on the performance of optoelectronic devices based on MAPbI3xClx.

Materials and methods

Materials

All reagents were used as received without further purification: PbI2 (99.999%, Sigma-Aldrich), methylamine hydroiodide (MAI, 99.5%, Xi’an Polymer Light Technology Corp), methylamine hydrochloride (MACl, 99.5%, Xi’an Polymer Light Technology Corp), methylamine hydrobromide (MABr, 99.5%, Xi’an Polymer Light Technology Corp), deuterated methylamine hydroiodide (CD3NH3I, 99%, Cambridge Isotope Laboratories Inc), deuterated methylamine hydroiodide (MACl, 99.5%, Cambridge Isotope Laboratories Inc), gamma-butyrolactone, (GBL, 99.99%, Sigma-Aldrich), dimethylformamide (DMF, 99.99%, Sigma-Aldrich), and dimethylsulfoxide (DMSO, 99.99%, Sigma-Aldrich).

Crystallization of CH3NH3PbI3–xClx/CD3NH3PbI3–xClx single crystals

MAI, MACl, and PbI2 were mixed and dissolved at different molar ratios into GBL as the precursor solution. DMF was chosen as the second solvent when the concentration of MACl exceeded 0.2 M because of the low solubility of MACl in GBL. During dissolution, the precursor solution was ultrasonicated and heated to 60 °C. Next, a filter (0.2 µm) was used to filter invisible impurities. After the above solution was heated and kept at 110 °C for 12 h, MAPbI3–xClx single crystals were harvested. The X-ray 2θ scan on the maximal facet of the typical MAPbI3–xClx single crystals showed two sharp diffraction peaks, suggesting a well-structured single-crystalline nature (Supplementary Fig. S1). Deuterated MAPbI3–xClx was synthesized using a similar approach to that mentioned above, except that CD3NH3I and CD3NH3Cl were used as cation sources.

It is noteworthy that the crystals could be obtained by the above method only for x ≤ 0.5. For x > 0.5, no crystal was synthesized. However, we found that a new trace MAPbCl3 phase grew together with the double halogen perovskite component when x was above 0.3, as proven by the XRD pattern in Supplementary Fig. S2 and 207Pb NMR spectrum in Supplementary Fig. S3. As a result, this article focuses only on the samples with x ≤ 0.3.

Crystallization of MAPbI3–xBrx single crystals

CH3NH3I, CH3NH3Br, and PbI2 were mixed and dissolved at different molar ratios into GBL as the precursor solution, and a filter (0.2 µm) was used to filter invisible impurities. After the above solution was heated and kept at 100 °C for 12 h, MAPbI3–xBrx single crystals were harvested. The static 207Pb NMR spectra of the MAPbI3xBrx (x = 0, 0.2, 1, 3) samples are shown in Supplementary Fig. S4.

Crystallization of MAPbBr3 and MAPbCl3/CD3NH3PbI3 single crystals

To grow perovskite CH3NH3PbBr3 single crystals, MABr and PbBr2 (1:1) were mixed and dissolved in DMF, with the concentration controlled at 1.0 M. A filter (0.2 µm) was used to filter invisible impurities. After the above solution was heated and kept at 80 °C for 6 h, MAPbBr3 single crystals were harvested. To grow the single-crystalline perovskite CH3NH3PbCl3, MACl, and PbCl2 (1:1) were mixed and dissolved in a mixed solution of DMSO and DMF (1:1), with the concentration controlled at 2.0 M. A filter (0.2 µm) was used to filter invisible impurities. After the above solution was heated and kept at 80 °C for 6 h, MAPbCl3 single crystals were harvested. Deuterated MAPbI3 was synthesized using a similar approach to that mentioned above except that CD3NH3I and CD3NH3I were used as cation sources.

Powder X-ray diffraction (XRD) patterns

XRD measurements were carried out on a Rigaku Smart Lab (X-ray source: Cu Kα; λ = 1.54186 Å) operated at 40 kV and 30 mA. MAPbI3–xClx powders were made by grinding a large piece of crystal into a fine powder.

XPS spectra

XPS spectra were measured by an X-ray photoelectron spectrometer (PHI Quantro SXM, ULVAC-PHI) with an Ar ion gun sputtering the surface. Before measuring the sample, the crystal surface was polished in a glove box filled with inert nitrogen gas to expose a fresh surface.

Static 207Pb solid-state NMR

207Pb NMR experiments were carried out on a 300 M NMR instrument operating at 62.73 MHz for 207Pb. A Bruker two-channel static PE probe with a homemade 4 mm coil was used to record the spectra. The 207Pb spectra were acquired using a single pulse excitation pulse sequence. The excitation pulse in the experiments was 2.5 μs at an RF field strength of 100 kHz. The recycle delay was 1 s. All 207Pb NMR spectra were referenced to PbMe4 (δ (207Pb) = 0 ppm) by setting the 207Pb peak of nonspinning solid MAPbCl3 measured at 293 K to −647.5 ppm.

1H solid-state MAS NMR

1H NMR spectra were recorded with a 600 M NMR instrument operating at 600.44 MHz. The spinning rate was set to 10 kHz. The recycle delay was set to 2 s. All 1H chemical shifts were calibrated using adamantane (δ = 1.85 ppm).

Static 2H solid-state NMR

2H NMR experiments were carried out on a 300 M NMR instrument operating at 46.07 MHz for 2H. A Bruker two-channel static PE probe with a homemade 4 mm coil was used to record the spectra. The 2H spectra were acquired using the solid echo sequence. The 2H pulse width was 2.5 μs at an RF field strength of 100 kHz. The 2H patterns were simulated via the weblab (http://weblab.mpip-mainz.mpg.de/weblab/).

Results and discussion

Preparation and characterization of MAPbI3 xClx single crystals

The materials and methods employed for preparing the MAPbI3xClx single crystal samples used in the present study are outlined in detail in the Materials and Methods. The preparation procedures are briefly described as follows. The inverse temperature crystallization (ITC) method was used to prepare MAPbI3 (i.e., x = 0) single crystal perovskites27,28. For MAPbI3xClx single crystal samples, x was varied by adjusting the proportions of methylamine hydrochloride (MACl) and methylamine hydroiodide (MAI) employed during synthesis while retaining a constant concentration of PbI2. Here, MACl was first added into a pure gamma-butyrolactone (GBL) solvent. However, this often yielded a turbid solution due to the poor solubility of MA in GBL, as shown in Fig. 1A. Nonetheless, the solution became transparent again once MAI and PbI2 were added, as shown in Fig. 1B. Second, the solution was heated to 383 K to promote the growth of single crystals. Representative single-crystal MAPbI3xClx samples are shown in Fig. 1C and Supplementary Fig. S1 against a cm2 square inscribed background to indicate the sample sizes. We note that MAPbCl3 impurity phases are synthesized together with MAPbI3xClx crystals for x > 0.3 (Supplementary Figs. S2 and S3). Hence, x = 0.3 was the largest value considered in the present study. Notably, reflecting the influence of Cl doping on the crystal structure, the shapes of the single crystal samples change with increasing Cl concentration. The changes in the crystal shape do not indicate a change in the crystal symmetry or the formation of a polycrystalline component. Similar phenomena have been observed in MAPbI3 single crystals29.

Fig. 1: Precursor solution growth process and structural characterization of MAPbI3−xClx single crystal samples (x = 0, 0.05, 0.1, 0.2, and 0.3).
figure 1

A Undissolved MACl in a GBL solution; B MACl solids completely dissolved in a GBL solution after adding MAI and PbI2; C photos of single-crystal samples; D XRD patterns of the samples in powder form; E enlarged XRD patterns from D, E showing the (004) and (220) diffraction peaks; F XPS Cl 2p core-level spectra; G Cl/Pb atomic ratios obtained from XPS results with respect to x.

Figure 1D, E presents the XRD patterns and enlarged views of the patterns in the 2θ range from ~28 to 29°, respectively, for MAPbI3xClx (x = 0, 0.05, 0.1, and 0.3) single crystal samples at room temperature. We note the presence of a clear splitting of the two peaks indicative of the (004) and (220) facets of the MAPbI3 sample at 2θ values ranging from 28 to 29°. This is a characteristic feature of the diffraction peaks obtained for MAPbI3 perovskite single crystals in the tetragonal phase at room temperature30. These (220) and (004) peaks are observed to shift gradually toward increasing 2θ values with increasing x in the range of 0.05–0.3. This gradual shift indicates that Cl doping causes the dimensions of the crystal lattice cell to progressively decrease with increasing x. We also note that the (220) and (004) diffraction peaks gradually merge with increasing x, indicating that the MAPbI3xClx single crystal lattice structure gradually changes from the tetragonal phase to the cubic phase with increasing x, as verified in past studies31.

The atomic concentrations of Cl atoms in the MAPbI3xClx single crystal samples were evaluated using high-resolution XPS. Prior to conducting XPS analysis, the surface of each single crystal sample was mechanically polished to avoid the influence of residual solvent on the XPS results. Figure 1F presents the XPS Cl 2p core-level spectra of the Cl atoms in the samples, which are composed of the Cl 2p1/2 and Cl 2p3/2 doublet states. We note that the Cl 2p signal intensity clearly increases with increasing x. The atomic Cl concentrations in the samples were determined relative to the atomic Pb concentrations from the areas under the Cl 2p and Pb 4f curves of the XPS core-level spectra obtained at multiple points of each sample surface and scaled according to their respective sensitivity factors. According to the chemical formula MAPbI3xClx, the atomic ratio of Pb/Cl should be equal to x. The Cl/Pb atomic ratios obtained through four repeated tests on the same batch of samples are plotted in Fig. 1G with respect to x. Clearly, the Cl/Pb atomic ratio increases with increasing x, as expected. However, the Cl/Pb atomic ratio in the samples does not scale according to x, and only a Cl/Pb atomic ratio of 0.138 ± 0.021 is obtained for x = 0.3, which is much less than the expected value. This indicates that a substantial proportion of Cl ions failed to enter the single crystal and were accumulated in the precursor during crystal growth, further inhibiting crystal growth. The results also explain why the sample size decreases with increasing x (Fig. 1C).

Static solid-state 207Pb and 2H NMR characterization of MAPbI3 xClx materials

Figure 2A presents the 207Pb solid-state NMR (ssNMR) spectra (solid black lines) of MAPbI3 and MAPbI2.7Cl0.3 single crystal samples. The 207Pb ssNMR results reveal clear changes in the local Pb environments with increasing x. As the central atom in the PbX6 octahedra, the changes in the local environment of Pb can be associated with variations in the chemical bonds between Pb and the surrounding halogen atoms. For comparison, the 207Pb ssNMR spectrum of a MAPbI2.9Br0.1 single crystal sample, synthesized by the same procedure as MAPbI3xClx, is also given in Fig. 2A. The spectra of the Cl- and Br-doped materials include curves (solid red lines) based on the two Gaussian curves (dashed black lines) given in each spectrum. The 207Pb ssNMR spectrum of the MAPbI3 crystal sample presents a very broad symmetrical peak with a chemical shift (δ) of 1423 ppm. In contrast, the spectrum of the MAPbI2.7Cl0.3 sample has a main peak consistent with that of the MAPbI3 spectrum, and a clear shoulder peak is observed at 1117 ppm. The origin of this shoulder peak can be assigned based on a comparison with the shoulder peak in the 207Pb ssNMR spectrum observed for the MAPbI2.9Br0.1 single crystal sample at 1176 ppm, which was assigned to a PbI5Br alloying structure in a past study32,33. Here, the PbI5Br alloying structure includes five iodine atoms and one bromine atom connected to a Pb atom to form an octahedron. We thus tentatively assign the shoulder peak of the MAPbI2.7Cl0.3 spectrum to the PbI5Cl unit. If this assignment is valid, then it would indicate that Cl atoms can replace I atoms in the crystal lattice in a manner similar to that observed in the MAPbI2.9Br0.1 crystal. Details regarding the 207Pb ssNMR spectrum analysis are outlined in Section S2, in conjunction with Supplementary Figs. S46.

Fig. 2: Inorganic lattice and organic cation dynamics characterization of MAPbI3−xClx by ssNMR.
figure 2

A Static 207Pb ssNMR spectra obtained at room temperature (solid black lines) for MAPbI3, MAPbI2.7Cl0.3, and MAPbI2.9Br0.1 single crystal samples. The spectra of the Cl- and Br-doped materials include curves (solid red lines) based on the two Gaussian curves (dashed black lines) given in each spectrum. B Static 2H ssNMR spectra at room temperature (solid black lines) along with the simulated spectra (solid red lines). C Model depicting the motion of MA cations in MAPbI3−xClx used in the spectral simulations. The samples used in A and B were the as-synthesized samples and were ground into powder form before the experiments.

An alternative method of detecting changes in the lattice structure of MAPbI3 perovskite materials is to probe the dynamic changes in the MA cations embedded in the inorganic lattice by means of 2H NMR analysis34. This approach relies on the strong coupling between the MA cations and the surrounding crystal lattice35. Here, small changes in the inorganic lattice cell have been reported to induce significant changes in the dynamics of the embedded MA cations26,36. For this purpose, MAPbI3xClx (x = 0, 0.05, 0.1, 0.2, and 0.3) samples were synthesized in which the methyl group was fully deuterated (Supplementary Fig. S7). In addition, powder samples were used in the 2H NMR experiments to ensure an adequate signal. Figure 2B presents the solid-state 2H NMR spectra (solid black lines) of the deuterated samples. The figure also includes spectra (solid red lines) simulated according to a method discussed later. For MAPbI3, the 2H pattern exhibits a typical Pake doublet lineshape, and the lineshapes of the signals vary significantly with increasing x, indicating gradual changes in the dynamics of the MA cations in the samples, which, as discussed, can be ascribed to changes in the inorganic lattice cell environment of the MA cations.

The motion of MA cations in organic-metal halide perovskite crystals has been studied by several groups34,37,38. These past studies have demonstrated that details regarding the dynamic changes in the MA cations in these perovskite samples can be obtained by analyzing the lineshapes of 2H NMR spectra. Therefore, we applied the MA cation motion model proposed by Franssen et al.37 to model the motion of MA cations in the MAPbI3xClx samples based on the lineshapes presented in Fig. 2B. Details regarding the lineshape analysis are outlined in section S3, in conjunction with Supplementary Figs. S811 and Supplementary Table S1. This process is illustrated here in Fig. 2C and described briefly as follows. An MA cation in the tetragonal phase of MAPbI3 can undergo rotation about two axes, including intrinsic axial rotation (rotation 1) about the Cn axis along with the C–N bond formed between the CH3 and NH3 components of the MA cation and secondary axial rotation (rotation 2) at a rate Ω about the Rc axis, which lies at a specific angle θ with respect to the Cn axis. The θ values are assumed to be distributed about a mean value θ0 according to a Gaussian distribution with standard deviation σ, as indicated by the equation and graph in Fig. 2C. Applying this model with 2σ = 1.5° to the 2H spectrum obtained for the MAPbI3 sample in Fig. 2B yields a θ0 of 58.9°, and the difference between the doublet peaks of the Pake lineshape (vQS) is 4 kHz. We note that the simulated 2H NMR spectrum is in good agreement with the actual spectrum. For MAPbI2.95Cl0.05, the observed 2H NMR spectrum lineshape undergoes obvious changes, where the separation between the two peaks of the Pake doublet lineshape decreases significantly. Applying the model with 2σ = 1.5° to the 2H spectrum MAPbI2.95Cl0.05 yields a θ0 of 57.5°. The slight reduction in θ0 with increasing x stems from the slight phase transition tendency from the tetragonal phase to the cubic phase revealed by XRD in Fig. 1E. With a further increase in x, we note that the 2H patterns gradually transform into a single broad peak for MAPbI3xClx (x = 0.1, 0.2, and 0.3). The lineshape analysis results indicate that the appearance of a broad peak can be attributed to a decrease in Ω, accompanied by an increase in σ. Based on the strong coupling between the MA cations and the inorganic lattice, we can conclude that these changes in the motion of MA cations are associated with the lattice contraction and lattice distortion caused by the random and partial substitution of iodine ions with Cl ions in the inorganic lattice. We also note that compared with the XRD and 207Pb NMR results, the 2H NMR spectra appear to exhibit a much greater sensitivity to structural changes in the inorganic framework, indicating the great potential of this technique for probing structural changes in MAPbI3 perovskite materials.

Time-dependent structural characterization of MAPbI3 xClx materials

Figure 3A presents the time-dependent XRD patterns (solid black lines) of the MAPbI2.7Cl0.3 single crystal sample obtained over a 2θ range of 27.6° to 28.8° at 0, 5, 10, and 20 days after synthesis. The samples were stored in a nitrogen-filled glove box between measurements to eliminate the influence of water vapor and oxygen as much as possible. The solid red line in the figure corresponds to the XRD pattern of the MAPbI3 single crystal sample originally shown in Fig. 1E. In comparison with the XRD pattern of the MAPbI3 single crystal sample, we note that the XRD pattern of the originally cubic MAPbI2.7Cl0.3 single-crystal shown in Fig. 3A gradually transitions over the 20-day period to an XRD pattern very similar to that observed for the MAPbI3 single crystal sample with a tetragonal structure. This strongly indicates that the MAPbI2.7Cl0.3 crystal underwent a gradual solid-phase transition from the cubic phase to the tetragonal phase.

Fig. 3: Long-term stability characterization of the MAPbI2.7Cl0.3 perovskite material at room temperature observed 0, 5, 10, and 20 days after synthesis (solid black lines).
figure 3

A XRD patterns of the samples; B 207Pb ssNMR spectra; C 2H ssNMR spectra. The samples used in this figure were from the same batch. All of the samples were ground into powder form before the experiments and were stored in a nitrogen-filled glove box at all times between the different experiments.

Figure 3B presents the time-dependent 207Pb NMR results obtained for the MAPbI2.7Cl0.3 single crystal sample at 0, 5, 10, and 20 days after synthesis. As before, the samples were stored in a nitrogen-filled glove box at all times between testing. The solid red line in the figure corresponds to the 207Pb NMR spectrum of the MAPbI3 single crystal sample originally shown in Fig. 2A. In comparison with the 207Pb NMR spectrum of the MAPbI3 single crystal sample, we note that the 207Pb NMR spectrum of the MAPbI2.7Cl0.3 single-crystal shown in Fig. 3B gradually transitions over the 20-day period to a symmetrical 207Pb NMR spectrum very similar to that observed for the MAPbI3 single crystal sample, with equivalent peak shifts of 1423 ppm. Here, the shoulder observed in the as-synthesized sample was greatly diminished after 5 days and completely disappeared after 10 days. As discussed above, the shift in the main peak and the appearance of the shoulder can be related to the formation of PbI5Cl alloying structures. Accordingly, the reverse shift of the main peaks and the disappearance of the shoulder thus strongly indicate that the PbI5Cl alloying structures gradually changed and disappeared over time.

Figure 3C presents the time-dependent 2H NMR results obtained for the deuterated MAPbI2.7Cl0.3 powder sample at 0, 5, 10, and 20 days after synthesis. As before, the samples were stored in a nitrogen-filled glove box at all times between testing. The solid red line in the figure corresponds to the 2H NMR spectrum of the MAPbI3 deuterated powder sample originally shown in Fig. 2B. In comparison with the 2H NMR spectrum of the MAPbI3 sample, we note that the 2H NMR spectrum of the MAPbI2.7Cl0.3 single-crystal shown in Fig. 3C gradually transitions over the 20-day period to a standard Pake doublet lineshape similar to that observed for the MAPbI3 single crystal sample. Here, the broad peak observed in the as-synthesized sample became considerably less broad after 5 days, and Pake doublet peaks began to emerge after 10 days. However, we note that the lineshape of the 2H NMR spectrum acquired after 20 days failed to fully return to that of the MAPbI3 spectrum, indicating that some residual distortion remains in the inorganic lattice of the sample.

The 2H NMR spectra of the as-synthesized MAPbI2.7Cl0.3 sample were also acquired at different temperatures (i.e., 294, 310, 330, and 340 K; see Supplementary Fig. S12 in Supporting Information). It was observed that the lineshapes of the 2H NMR spectra remained nearly unchanged before and even after exceeding the phase transition temperature of MAPbI3 (337 K). This result indicates that the MAPbI2.7Cl0.3 structure does not undergo a similar phase transition as that of MAPbI3. This may also be associated with the doping of Cl atoms in the inorganic lattice.

The combined observations above clearly reveal many intriguing details regarding the structural transformations of MAPbI2.7Cl0.3 over time. As discussed, the time-dependent XRD results indicate that the MAPbI2.7Cl0.3 crystal undergoes a gradual solid-phase transition from the cubic phase to the tetragonal phase within 20 days. In addition, while the time-dependent changes in the chemical shift and lineshape of the 207Pb spectrum strongly indicate that the PbI5Cl alloying structures present in the as-synthesized sample gradually disappear over time, the 207Pb NMR results cannot reveal whether the PbI5Cl alloying structures have completely disappeared after 20 days due to the limited sensitivity of 207Pb NMR signals. The time-dependent XRD results suffer from a similar limitation in that XRD is not sensitive to local changes in the crystalline structure. Fortunately, this issue is clarified to some extent by the fact that the lineshape of the 2H NMR spectrum acquired after 20 days did not return fully to that of the MAPbI3 spectrum. As discussed, this indicates that some residual distortion remained in the inorganic lattice of the sample. Accordingly, this suggests that some amount of alloying structures remains in the sample but not at a sufficiently high level to sustain the cubic phase of the crystal or to greatly affect the local Pb environment overall. Clearly, 2H NMR represents a more sensitive technique for probing changes in the inorganic lattice of MAPb3 halide perovskite crystals due to Cl doping.

Photoresponse of photodetectors based on MAPbI3 xClx single-crystal absorbers

The influence of Cl doping on the optoelectronic properties of MAPbI3xClx materials was illustrated by fabricating five planar-type photodetectors with appropriately shaped and finished MAPbI3xClx (x = 0, 0.05, 0.1, 0.2, and 0.3) single crystal absorbers. An image of a representative photodetector device is shown in Fig. 4A. Here, interdigitated Au wire electrodes (0.5 mm in length, 40 µm in width, and approximately 100 nm thick) were evaporated on the well-polished (100) plane of each single crystal. The effective absorber area of each detector was 0.1 mm2. The current responses were measured under a bias voltage of 1 V with illumination provided by a 405 nm wavelength continuous laser beam with an optical power density of 0 to 100 mW/cm2. Figure 4B shows the photocurrent responses of the photodetectors employing the as-synthesized single-crystal absorbers as a function of the incident light intensity. The results demonstrate that the photocurrent increased almost linearly with increasing incident light intensity for each single crystal photodetector. Meanwhile, the photocurrent of the photodetectors at a given incident light intensity was positively correlated with the Cl doping ratio (i.e., x), where the photocurrent increased with increasing x. Figure 4C presents the photocurrents of the devices obtained with respect to time under continuous laser illumination operating with an optical power density of 70 mW/cm2. As anticipated, all of the photodetectors exhibited a very good photoresponse, where the photocurrent immediately increased and decreased at the beginning and end of each laser pulse, respectively. We also note that the photocurrent obtained at each laser pulse noticeably increased with increasing x. For example, the photocurrent of the device based on MAPbI3 was only 3 µA, whereas that of the device based on MAPbI2.7Cl0.3 was 23 µA.

Fig. 4: Photoresponse measurement of photodetectors based on MAPbI3−xClx single crystals.
figure 4

A Image of a representative photodetector based on a MAPbI3−xClx single-crystal absorber. The interdigitated Au wire electrodes were evaporated onto the well-polished (100) plane of the single crystal absorber. B Plots of the photocurrent with respect to the optical power density of the incident illumination for photodetectors based on as-synthesized MAPbI3−xClx (x = 0, 0.03, 0.1, 0.2, and 0.3) single crystal absorbers. C Plots of the photocurrent with respect to time under pulsed laser illumination (70 mW/cm2) for photodetectors based on as-synthesized MAPbI3−xClx single-crystal absorbers. Plots of the photocurrent with respect to the optical power density of the incident illumination for photodetectors based on as-synthesized MAPbI3−xClx single-crystal absorbers (filled black squares) and the same photodetectors after storage in a nitrogen-filled glove box for 20 days (filled red circles): D x = 0; E x = 0.1; F x = 0.3. In all experiments, the bias voltage of the devices was 1 V, and a laser operating at a wavelength of 405 nm was used as the light source.

However, a comparison of the photocurrents with respect to the optical power density of the incident illumination for photodetectors based on the as-synthesized MAPbI3, MAPbI2.95Cl0.05, and MAPbI2.7Cl0.3 single-crystal absorbers, shown by the filled black squares in Fig. 4D, E, and F, respectively, and the same photodetectors after storage in a nitrogen-filled glove box for 20 days (filled red circles) demonstrates that the optoelectronic properties of the Cl-doped perovskite materials were not stable and degraded after 20 days of storage to photocurrents essentially commensurate with that of MAPbI3. The photoresponse data obtained for photodetectors based on MAPbI2.9Cl0.1 and MAPbI2.8Cl0.2 behaved similarly (Supplementary Fig. S13). In contrast, the photoresponse data obtained for the MAPbI3-based photodetector exhibited no obvious changes after storage for 20 days, which demonstrates the stability of the MAPbI3 structure. The above results support the previous time-dependent analyses based on XRD, 207Pb NMR, and 2H NMR, which demonstrated that the crystal lattices of the MAPbI3xClx (x = 0.05, 0.1, 0.2, and 0.3) single crystals changed from cubic to tetragonal after 20 days, and the Cl alloying structures in the Cl-doped materials gradually disappeared.

As discussed, the roles of Cl ions in MAPbI3xClx perovskites are a very controversial issue. A number of theoretical studies have demonstrated that the substitution of Cl for I in the inorganic lattice can have an inherently positive effect on the electronic properties of MAPbI3xClx perovskites by reducing electron–phonon coupling and electron−hole recombination25, enhancing the intensity of the p orbital of I ions in the HOMO state39, or enhancing electron localization and the charge-carrier diffusion length26. However, other studies have demonstrated that Cl substitution may have a negative effect on the properties of MAPbI3xClx perovskites19,22. For example, the disassociation of Cl atoms from the inorganic lattice can generate a large amount of anionic defects in the lattice, which serve as deep-level defects that may strongly hinder the generation and diffusion of charge carriers40. Here, the long-term instability of MAPbI3xClx alloying structures reported in the present work sheds some light on this controversial issue. In the as-synthesized state of MAPbI3xClx single crystals, the substitution of Cl for I contributes to enhanced electronic properties of the Cl-doped perovskites and explains the origin of their enhanced optoelectronic properties. Meanwhile, the Cl atoms gradually dissociate from the inorganic lattice over time and may result in the generation of anionic vacancies in the lattice. As a result, the optoelectronic properties of the MAPbI3xClx single crystals deteriorate over time. This long-term instability indicates that different conclusions regarding the effect of Cl ions on the optoelectronic properties of MAPbI3xClx perovskites may arise over different time scales, which provides a plausible explanation for the long-standing controversy regarding the effect of Cl ions on the performance of optoelectronic devices based on MAPbI3xClx.

In addition, the stability issues of mixed-halide perovskites, particularly APbI3xBrx materials, have attracted some attention recently41,42. The results revealed that APbI3xBrx can reversibly segregate into I-rich and Br-rich domains under continuous visible illumination43,44. For MAPb3xClx, new Cl-rich domains, such as MAPbCl3, were not discovered in either the XRD patterns or the 207Pb NMR spectra. We thus hesitate to directly correlate the instability of MAPb3xClx over time to the halide segregation present in APbI3xBrx. However, the halide segregation phenomenon is possibly related to some metastable states present only under illumination. Therefore, an in situ study of MAPb3xClx is worth investigating if one wants to probe the halide segregation phenomenon. Such an in situ study is ongoing in our laboratory.

Conclusion

In this article, we report a detailed study of MAPbI3xClx (x = 0.0 to 0.3) single crystal structures based on a combination of XRD, XPS, and solid-state NMR. The results revealed that a small amount of Cl atoms could substitute for iodide atoms in the inorganic lattice of Cl-doped crystals. This substitution distorted the inorganic lattice, resulting in a transformation from the tetragonal phase to the cubic phase and a cation dynamics change with increasing x. However, the Cl/I alloying structures were demonstrated to be unstable over a long term of ~20 days. Time-dependent XRD and solid-state NMR results demonstrated that the Cl atoms gradually dissociated from the Cl/I alloying structures. This Cl dissociation induced a counter transformation from the cubic phase to the tetragonal phase. The photocurrents obtained for photodetectors based on as-synthesized MAPbI3xClx (x = 0, 0.05, 0.1, 0.2, and 0.3) absorbers linearly increased with increasing x, but the photocurrents obtained for the same photodetectors after 20 days of storage degraded to photocurrents commensurate with that of MAPbI3, which supported the results of analyses based on the time-dependent structural transformations of MAPbI3xClx single crystals occurring over long time periods. This indicates that the optoelectronic properties of the Cl-doped perovskite materials are strongly correlated with the presence of Cl/I alloying structures. Moreover, the observed long-term instability indicates that different conclusions regarding the effect of Cl ions on the optoelectronic properties of MAPbI3xClx perovskites may arise over different time scales, which provides a plausible explanation for the long-standing controversy regarding the effect of Cl ions on the performance of optoelectronic devices based on MAPbI3xClx.