• Journal of Semiconductors
  • Vol. 46, Issue 4, 042101 (2025)
Yang Liu1, Kunyuan Lu1, Yujie Zhu1, Xudong Hu1..., Yusheng Li3, Guozheng Shi1, Xingyu Zhou1, Lin Yuan1, Xiang Sun1, Xiaobo Ding1, Irfan Ullah Muhammad1, Qing Shen3, Zeke Liu1,2,* and Wanli Ma1,2,**|Show fewer author(s)
Author Affiliations
  • 1Institute of Functional Nano & Soft Materials (FUNSOM), Joint International Research Laboratory of Carbon-Based Functional Materials and Devices, Soochow University, Suzhou 215123, China
  • 2Jiangsu Key Laboratory of Advanced Negative Carbon Technologies, Soochow University, Suzhou 215123, China
  • 3Faculty of Informatics and Engineering, University of Electro-Communications, Tokyo 182−8585, Japan
  • show less
    DOI: 10.1088/1674-4926/24050026 Cite this Article
    Yang Liu, Kunyuan Lu, Yujie Zhu, Xudong Hu, Yusheng Li, Guozheng Shi, Xingyu Zhou, Lin Yuan, Xiang Sun, Xiaobo Ding, Irfan Ullah Muhammad, Qing Shen, Zeke Liu, Wanli Ma. Colloidal synthesis of lead chalcogenide/lead chalcohalide core/shell nanostructures and structural evolution[J]. Journal of Semiconductors, 2025, 46(4): 042101 Copy Citation Text show less

    Abstract

    Lead chalcohalides (PbYX, X = Cl, Br, I; Y = S, Se) is an extension of the classic Pb chalcogenides (PbY). Constructing the heterogeneous integration with PbYX and PbY material systems makes it possible to achieve significantly improved optoelectronic performance. In this work, we studied the effect of introducing halogen precursors on the structure of classical PbS nanocrystals (NCs) during the synthesis process and realized the preparation of PbS/Pb3S2X2 core/shell structure for the first time. The core/shell structure can effectively improve their optical properties. Furthermore, our approach enables the synthesis of Pb3S2Br2 that had not yet been reported. Our results not only provide valuable insights into the heterogeneous integration of PbYX and PbY materials to elevate material properties but also provide an effective method for further expanding the preparation of PbYX material systems.

    Introduction

    Chalcohalides consisting of mixed chalcogen and halogen anions and metal cations are an emerging family of semiconductors[13]. The rich composition and crystal structure of chalcohalides has provided significant opportunities for fundamental materials research and energy applications[49]. Among them, the expansion of Pb chalcohalides, as an extension of two classic material systems, namely Pb chalcogenides (such as PbS, PbSe) and lead halide perovskites (such as CH3NH3PbI3, CsPbI3), has sparked great interests[1016]. Pb chalcohalides (PbYX) can be understood as binary phase diagrams composed of two phases, PbY−PbX2 (Y = S, Se; X = Cl, Br, I), combined in different proportions[17]. However, early studies found that only a few combinations of phases could exist stably at ambient temperature and pressure, such as Pb5S2I6 and Pb7S2Br10[18]. It was not until recently that the Manna group discovered the synthesis of Pb3S2Cl2, Pb4S3Br2, and Pb4S3I2 nanocrystals (NCs) through chemical methods[19]. These phases do not conform to stable bulk PbY−PbX2 diagrams but can exist stably in the form of NCs at ambient temperature and pressure[17]. Based on the novel material systems, they further find these PbYX exhibits remarkable lattice matching with lead-halide perovskite, making them potential structures to enhance the optoelectronic properties and stability of perovskite materials[2023].

    Furthermore, halide ions have been identified as the optimal surface passivating agents for classical PbS and PbSe NCs, playing a crucial role in enhancing the performance of near-infrared and mid-infrared optoelectronic devices based on PbY NCs[2428]. The formation of Pb−X bonds on the surface can also act as a bridge, connecting them with novel lead halide perovskite materials, thereby achieving superior material and device performance[20, 23]. The successful synthesis of PbXY NCs has deepened our understanding of traditional classical material systems. For instance, passivation with Cl ions significantly improves the luminescence efficiency and stability of PbS and PbSe NCs[29, 30]. In earlier studies, it was speculated that the surface adsorbed only Cl ions or formed a PbClx shell[3133]. Although not confirmed yet, recent research suggests that the surface layer may be a Pb sulfochloride surface layer based on the similarity of reaction conditions[32]. Further exploration of the relation of PbXY with classical PbS and PbSe NCs will deepen our understanding of the structure-property relationships in materials, facilitating the design and synthesis of new materials and structures with superior optoelectronic performance.

    In this study, we have discovered that within the framework of the classical PbS NC synthesis approach, the straightforward introduction of bis(trimethylsilyl) halide (TMS−X, X = Cl, Br, I) precursors enables the synthesis of a series of materials, starting from PbS NCs and progressing to PbS/Pb3S2X2 core-shell structures, ultimately leading to the formation of Pb3S2X2 NCs. We have conducted a detailed investigation of the evolution process and its impact on the optical properties of the NC materials. Notably, we found that the growth of a Pb3S2X2 shell on the surface of PbS effectively enhances its photoluminescence quantum yield (PLQY) from 49.3% to 72.0%. Furthermore, this strategy enables the synthesis of Pb3S2Br2 NCs, which have not been previously reported.

    Experimental

    Chemicals and materials

    Lead acetate trihydrate (PbAc2·3H2O, 99%), hexamethyldisilathiane (TMS−S, 98%), trimethylsilyl chloride (TMS−Cl, 99%), trimethylsilyl bromide (TMS−Br, 99%), trimethylsilyl iodide (TMS−I, 99%), sulfur powder (S, 99.99%), oleylamine hydrochloride (OLA−Cl, 99.5%), oleic acid (OA, 90%), 1-octadecene (ODE, 90%), hexane (Hex, 95%), isopropanol (IPA, 95%), acetone (95%).

    Synthesis of PbS NCs and Pb3S2X2 (X = Cl, Br, I) NCs

    The NC synthesis was performed with a modified recipe previously documented[34]. All operations were performed under a nitrogen atmosphere using standard air-free Schlenk line techniques. Briefly, 1 mmol of PbAc2·3H2O and 0.8 mL of OA were dissolved in 10 mL of ODE in a three-neck flask. The solution was then heated to 100 °C under vacuum for 1 h. The TMS−S (0.25 mmol) and different amounts of TMS−X (X = Cl, Br, I) precursor were dissolved in 1 mL of ODE. The reaction was initiated by rapid injection of the TMS−S and TMS−X precursors into the lead precursor solution at 100 °C under nitrogen. The NCs were grown at this temperature for 10 min and then cooled to room temperature slowly. The solution was then transferred into a nitrogen-filled glove box and purified by precipitation once in hexane/isopropyl alcohol and once in hexane/acetone. The obtained NC solids were stored in a nitrogen-filled glove box.

    For the synthesis of large PbS NCs and Pb3S2X2 NCs, 1.2 mmol of PbAc2·3H2O and 12 mL of OA were added in a three-neck flask by heating the mixture to 100 °C under vacuum for 1 h. The TMS−S (0.2 mmol) and different amounts of TMS−X (X = Cl, Br, I) precursor were dissolved in 1 mL of ODE. The reaction was initiated by rapid injection of the TMS−S and TMS−X precursor into the lead precursor solution at 150 °C under nitrogen. The NCs were grown at this temperature for 5 s and then cooled to room temperature quickly. The solution was then transferred into a nitrogen-filled glove box and purified by precipitation once in hexane/isopropyl alcohol and once in hexane/acetone. The obtained NC solids were stored in a nitrogen-filled glove box.

    Direct synthesis of large Pb3S2Cl2 NCs

    Briefly, 1 mmol of PbAc2·3H2O and 0.8 mL of OA were dissolved in 10 mL of ODE in a three-neck flask by heating the mixture to 100 °C under vacuum for 1 h. The anion precursors were prepared by sonicating 0.5 mmol sulfur powder and 0.8 mmol OLA−Cl in 2 mL dried ODE. The reaction was initiated by rapid injection of the anion precursors into the lead precursor solution at 150 °C under nitrogen. The NCs were grown at this temperature for 30 min and then cooled to room temperature slowly. The solution was then transferred into a nitrogen-filled glove box and purified by precipitation twice in hexane/isopropyl alcohol and once in hexane/acetone. The obtained NC solids were stored in a nitrogen-filled glove box.

    Measurement and characterization

    The absorption spectra of NCs in hexane were recorded at room temperature using an ultraviolet−visible−near infrared (UV−VIS−NIR) spectrophotometer (Lambda950, PerkinElmer). Photoluminescence (PL) spectra and PL lifetime were measured by a FluoroMax-4 spectrofluorometer (HORIBA Scientific). The PLQY was measured using an integrating sphere equipped on FLS1000 fluorescence spectrometer. The excitation wavelength was 750 nm. The PbS NC films for the XPS test were made by spin-coating ~100 nm thick film onto a silicon substrate. The X-ray diffraction (XRD) spectra were obtained on an X-ray diffractometer with a Cu Kα source (PANalytical B.V. Empyrean). TEM measurements were performed by a Tecnai G2 F20 S-Twin system.

    Results and discussion

    The NCs were synthesized based on the typical hot injection method with lead-oleate as the Pb precursor[34]. The reaction was initiated by injecting mixtures of TMS−S and TMS−X with different molar ratios (represented by X : S). Taking TMS−Cl as an example, at low Cl : S values, the reaction proceeded similarly to the synthesis of conventional PbS NCs, where the colorless reaction solution turned red and then rapidly black. As the Cl : S value increased, the transition from red to black gradually slowed down. The obtained XRD pattern of the NCs is presented in Fig. 1(a). The graph reveals that the crystal structure of the NCs obtained at Cl : S values below 0.6 is nearly identical to the cubic phase of PbS, with only a slight shift observed in the (111) plane (Fig. 1(b)). When the Cl : S value exceeded 0.6, the intensity of the (220) peak of PbS NCs gradually decreased and eventually disappeared. Simultaneously, the (111) plane at 25.9° shifted to lower angles, and the (200) plane at 29.7° shifted to higher angles. When the Cl : S value reached 1.8, the two planes settled at 24.4° and 31.8°, respectively, corresponding to the (112) and (013) planes of Pb3S2Cl2. Additionally, when the Cl : S value reached 0.8, a broad peak started to appear around ~35°, and with an increasing Cl : S value, it continued to shift to higher angles, ultimately reaching ~37°, corresponding to the (123) plane of Pb3S2Cl2. This phenomenon is highly reminiscent of the typical evolution from core to core/shell NC structures[3537]. Furthermore, the above XRD diffraction behavior differs significantly from that of a simple mixture of PbS and Pb3S2Cl2 phases (Fig. S1). Therefore, we hypothesize that a PbS/Pb3S2Cl2 core/shell structure is formed under intermediate Cl : S ratios of 0.8−1.6 (with further evidence presented in the subsequent text). When the Cl : S ratio reaches 1.8, pure-phase Pb3S2Cl2 NCs can be obtained. The crystal structure of PbS and Pb3S2Cl2 is shown in Fig. 1(c) and 1(d). The cubic PbS exhibits a regular Pb and S octahedral structure, with the Pb atoms coordinated by six S anions. In contrast, for the Pb3S2Cl2, the Pb is of different chemical environments compared with the PbS. All the Pb atoms occupy the 12a site, which has a point group symmetry of 4 bar and is coordinated by eight anions.

    (Color online) (a) The XRD pattern of the NCs with different feeding ratios of Cl and S. (b) The (111) and (200) peak shift vs. feeding ratio of Cl and S. (c) and (d) The crystal structure of PbS and Pb3S2Cl2.

    Figure 1.(Color online) (a) The XRD pattern of the NCs with different feeding ratios of Cl and S. (b) The (111) and (200) peak shift vs. feeding ratio of Cl and S. (c) and (d) The crystal structure of PbS and Pb3S2Cl2.

    To ascertain the exact nature of surface chemistry after introducing TMS−Cl, we conducted X-ray photoelectron spectroscopy (XPS) measurement. The well-defined Cl 2p peaks in the XPS spectra clearly demonstrate the incorporation of Cl in the obtained NCs (Fig. 2(a)). The Pb 4f7/2 spectra were deconvoluted by fitting the sum of the Lorentz−Gaussian functions. They were found to contain two chemical states, observed at 137.7 and 138.6 eV, respectively (Fig. 2(b)). The main peak at 137.7 eV was assigned to the Pb−S bond, while the peak at 138.6 eV corresponded to Pb−Cl bonds[29]. The Pb−Cl peak becomes stronger as increasing the Cl : S input during the synthesis, providing further evidence for the incorporation of Cl in the NCs. The S 2p spectra become broader as increasing Cl : S ratio, due to the more complicated coordination environment in Pb3S2Cl2 compared to that in PbS crystal (Fig. 2(c)). The element ratio of the NCs is extracted by fitting the peak areas in XPS spectra and presented in Fig. 2(d). The molar ratio of Cl in the final NCs increases with the increment of the Cl : S ratio. When the input Cl : S ratio reaches 1.8, the obtained Pb : S : Cl = 50.3 : 23.6 : 26.1, which is close to the ratio in the chemical formula (Pb3S2Cl2). XPS is a surface-sensitive technique because of the shallow escape depth of photoelectrons, whereas the sampling depth of SEM-EDX measurement is typically above 1 µm. The Cl concentrations measured by XPS for different Cl : S were found to be consistently higher than the values obtained from SEM-EDX analysis, indicative of the presence of a Cl-rich near-surface region within NCs (Fig. 2(e)). The extent of Cl surface enrichment can be estimated by taking the ratio between atomic percent values obtained from XPS and EDX measurements. As shown in Fig. 2(f), the Cl(EDX)/Cl(XPS) ratio monotonically increases and approaches 1 with the increment of the input Cl : S ratio, which indicates that the NCs gradually transform into a homogeneous composition (e.g. Pb3S2Cl2)[38]. The above observation aligns with the previous XRD results, confirming the structural evolution from PbS to PbS/Pb3S2Cl2 core/shell and eventually to pure-phase Pb3S2Cl2 as increasing the input Cl : S ratio.

    (Color online) (a)−(c) High-resolution XPS spectrum in the (a) Cl 2p, (b) Pb 4f, (c) S 2p regions of the NCs with different feeding ratios of Cl and S. (d) The element ratios of the PbS and Pb3S2Cl2 NCs obtained by XPS and SEM-EDS. (e) Cl contents measured by SEM-EDS and XPS the NCs with different feeding ratios of Cl and S. (f) Plot of the ratio between Cl concentrations determined by SEM-EDS and XPS measurements in (e).

    Figure 2.(Color online) (a)−(c) High-resolution XPS spectrum in the (a) Cl 2p, (b) Pb 4f, (c) S 2p regions of the NCs with different feeding ratios of Cl and S. (d) The element ratios of the PbS and Pb3S2Cl2 NCs obtained by XPS and SEM-EDS. (e) Cl contents measured by SEM-EDS and XPS the NCs with different feeding ratios of Cl and S. (f) Plot of the ratio between Cl concentrations determined by SEM-EDS and XPS measurements in (e).

    We further investigated the influence of the aforementioned structural evolution on the optical properties of the NCs. As depicted in Fig. 3(a) and 3(b), a slight red shift is observed in the absorbance and PL peaks when the input Cl : S ratio is 0.2. Subsequently, as the molar ratio of Cl : S increased, the absorbance and PL peaks blueshifted, PLQY gradually increased and Stokes shift decreased, reaching a maximum at Cl : S = 0.8, with a minimum Stokes shift of 104 nm and a PLQY of 72.0% (Fig. 3(c), 3(e) and Fig. S2). Further increase in the Cl : S ratio leads to a decrease in PLQY and an increase in Stokes shift. Interestingly, as the Cl : S ratio reaches 1.2, a new peak located at ~680 nm appears, which can be assigned to the emission from the Pb3S2Cl2 phase (Fig. S3). As the Cl : S ratio reaches 1.6, the exciton absorption peak blue shifts to ~710 nm from the ~906 nm of initial PbS NCs. The color of the obtained NC solution gradually transitions from brownish to red (Fig. 3(d)). Meanwhile, the NCs lose PL entirely in the near-infrared region. Further increasing the Cl : S ratio to 1.8 results in the vanishing of both PL and exciton absorption peaks. To understand the variation of optical properties, we further tested their size and morphology through transmission electron microscopy (TEM) characterization. As shown in Fig. S4 and Fig. S5, with the increase of input Cl/S ratio, the size of NCs gradually decreases. Under low concentration (Cl : S = 0.2), the size reduction effect of Cl may not be significant, but the excitons in the PbS core can delocalize to the shell, resulting in a slight redshift of absorption and PL[39]. As Cl content increases, the binding of Cl and Pb precursors limits the growth of NCs and causes a decrease in the size of the PbS core, resulting in a gradual blueshift of the absorption and PL peaks. The simultaneous appearance of PL peaks from both PbS and Pb3S2Cl2 confirms the formation of a PbS/Pb3S2Cl2 core/shell structure[40, 41]. The time-resolved PL (TRPL) data shows that increasing the amount of Cl can enhance the PL lifetime of the NCs when the Cl : S ratio is below 0.8 (2.01 to 3.70 μs), (Fig. 3(f) and Table S1), which is consistent with the increase of PLQY, demonstrating the passivation effect at low input Cl : S ratios. Further increasing the Cl : S ratio can significantly reduce PL lifetime, with the value as low as 8.68 ns for pure-phase Pb3S2Cl2 NCs (input Cl : S = 1.8), which is also consistent with PL lifetime reported previously (Fig. 3(g), Table S2). Similar structural and optical evolutions can be observed with large PbS NCs (first exciton peak at 1627 nm) as starting materials (Fig. S6−Fig. S8). The dominant size of the finial pure Pb3S2Cl2 can reach 3.2 nm, which was larger than the size of pure Pb3S2Cl2 NCs (1.5 nm) triggered from small PbS NCs. Additionally, the NCs absorbance band edge of these two methods were all around 600 nm, indicating the size of Pb3S2Cl2 NCs has little effect on their optical characterization due to their negligible quantum confinement effects, which is also consistent with the previous report[19].

    (Color online) (a) and (b) The absorbance (a) and PL (b) of the NCs with different feeding ratios of Cl and S. (c) The peak position of absorbance and PL, as well as Stokes shift. (d) The images of the NC solutions. (e)−(g) The PLQY (e) and TRPL (f) and (g) of the NCs with different feeding ratios of Cl and S.

    Figure 3.(Color online) (a) and (b) The absorbance (a) and PL (b) of the NCs with different feeding ratios of Cl and S. (c) The peak position of absorbance and PL, as well as Stokes shift. (d) The images of the NC solutions. (e)−(g) The PLQY (e) and TRPL (f) and (g) of the NCs with different feeding ratios of Cl and S.

    Seeking to clarify the impact of the Pb3S2Cl2 shell on ensemble photophysics, we further investigate the consequences of Pb3S2Cl2 shelling on PbS NCs by using time-resolved emission spectroscopy (TRES). To avoid the Förster resonant energy transfer and self-absorbance phenomenon, we dilute the PbS NCs solution (1 mg/mL) in hexene. The PbS NC solution displays transient emission spectra with a dynamic spectral red-shift without TMSCl introduction. (Fig. 4(a) and 4(c), monitored via the emission peak). Conversely, the shelled NCs with Pb3S2Cl2 display much less overall spectral red-shifting after photoexcitation and the early (<2 μs) component is effectively eliminated (Fig. 4(b) and 4(d)), which reveals the Pb3S2Cl2 can well passivate PbS NCs[42].

    (Color online) Time-resolved emission spectroscopy (TRES) of PbS NCs. (a) and (b) TRES map of control and shelled. (c) and (d) Time-slices of the emission spectra of control and shelled. (e) Summary of dynamic red-shifts of the NC photoluminescence peak after excitation for control and shelled PbS NCs. All experiments are conducted in dilute (1 mg/mL) hexane solution.

    Figure 4.(Color online) Time-resolved emission spectroscopy (TRES) of PbS NCs. (a) and (b) TRES map of control and shelled. (c) and (d) Time-slices of the emission spectra of control and shelled. (e) Summary of dynamic red-shifts of the NC photoluminescence peak after excitation for control and shelled PbS NCs. All experiments are conducted in dilute (1 mg/mL) hexane solution.

    We also find this phase and structural evolution from PbS NCs to Pb3S2Cl2 can be extended to Br and I systems. The XRD pattern and optical evolution Br system show similar behavior as that in the Cl case (Fig. S9). The solution of Pb3S2Br2 and Pb3S2I2 NCs also exhibits a red color (Fig. 5(a)). The extracted bandgap values are 2.2, 2.1, and 1.9 eV for Pb3S2Cl2, Pb3S2Br2 and Pb3S2I2, respectively. To validate the crystal structure, we compared the XRD patterns of Pb3S2Cl2 and Pb3S2Br2 NCs with those of the bulk structure published by Ni et al. The XRD analysis confirmed a good match between the NCs and the bulk structures (Fig. 5(b))[43, 44]. It should be noted that we have discovered that Pb3S2Br2 NCs have not yet been reported. However, this method cannot deliver pure-phase Pb3S2I2 NCs, the XRD pattern from the PbS phase can also be observed even at input I : S ratio up to 1.0 in large PbS NCs (Fig. S10). The obtained Pb3S2X2 all exhibit sphere shapes with sizes around 3 nm (Fig. 5(c)−5(e)). Additionally, we found that by reducing the reactivity of the anion precursors, using ODE−S and oleylamine hydrochloride (OLACl) to replace TMS−S and TMSCl, larger-sized Pb3S2Cl2 NCs can be synthesized, with size up to ~10 nm (Fig. S11).

    (Color online) (a) The absorbance of Pb3S2Cl2, Pb3S2Cl2, and Pb3S2I2 (The photo of Pb3S2X2 solution is shown in the inset). (b) The XRD pattern of Pb3S2Cl2, Pb3S2Cl2, and Pb3S2I2 NCs. The peak marked with an asterisk represents unreacted PbS phase. (c)−(e) The TEM images of Pb3S2Cl2, Pb3S2Cl2, and Pb3S2I2 NCs.

    Figure 5.(Color online) (a) The absorbance of Pb3S2Cl2, Pb3S2Cl2, and Pb3S2I2 (The photo of Pb3S2X2 solution is shown in the inset). (b) The XRD pattern of Pb3S2Cl2, Pb3S2Cl2, and Pb3S2I2 NCs. The peak marked with an asterisk represents unreacted PbS phase. (c)−(e) The TEM images of Pb3S2Cl2, Pb3S2Cl2, and Pb3S2I2 NCs.

    Conclusion

    In this work, we reported the synthesis of PbS/Pb3S2X2 (X = Cl, Br, I) core-shell structures for the first time. Compared to pure PbS NCs, this core-shell structure effectively enhances the optical properties, increasing the PLQY from 49.8% to 72.0% and the fluorescence lifetime from 2.01 to 3.70 μs. Under conditions of high halogen incorporation, we obtained pure-phase Pb3S2Cl2 and Pb3S2Br2 NCs. Notably, our approach enables the synthesis of stable Pb3S2Br2 that had not yet been reported. This work provides a novel synthetic approach to achieve heterogeneous integration of lead chalcohalides (PbYX, X = Cl, Br, I; Y = S, Se) with lead chalcogenides (PbY) and explores their impact on the structure and properties, potentially encouraging further exploration of other metal chalcohalides and heterostructure synthesis and applications.

    Appendix A. Supplementary material

    References

    [1] U V Ghorpade, M P Suryawanshi, M A Green et al. Emerging chalcohalide materials for energy applications. Chem Rev, 123, 327(2023).

    [2] E Wlaźlak, A Blachecki, M Bisztyga-Szklarz et al. Heavy pnictogen chalcohalides: The synthesis, structure and properties of these rediscovered semiconductors. Chem Commun, 54, 12133(2018).

    [3] D Mihailovic. Inorganic molecular wires: Physical and functional properties of transition metal chalco-halide polymers. Prog Mater Sci, 54, 309(2009).

    [4] B Xu, T L Feng, M T Agne et al. Manipulating band structure through reconstruction of binary metal sulfide for high-performance thermoelectrics in solution-synthesized nanostructured Bi13S18I2. Angew Chem Int Ed, 57, 2413(2018).

    [5] D Quarta, S Toso, R Giannuzzi et al. Colloidal bismuth chalcohalide nanocrystals. Angew Chem Int Ed, 61, e202201747(2022).

    [6] S Johnsen, Z F Liu, J A Peters et al. Thallium chalcohalides for X-ray and γ-ray detection. J Am Chem Soc, 133, 10030(2011).

    [7] R M Nie, J Im, S I Seok. Efficient solar cells employing light-harvesting Sb0.67Bi0.33SI. Adv Mater, 31, 1808344(2019).

    [8] R M Nie, K S Lee, M M Hu et al. Heteroleptic tin-antimony sulfoiodide for stable and lead-free solar cells. Matter, 3, 1701(2020).

    [9] R M Nie, B Kim, S T Hong et al. Nanostructured heterojunction solar cells based on Pb2SbS2I3: Linking lead halide perovskites and metal chalcogenides. ACS Energy Lett, 3, 2376(2018).

    [10] Q A Akkerman, G Rainò, M V Kovalenko et al. Genesis, challenges and opportunities for colloidal lead halide perovskite nanocrystals. Nat Mater, 17, 394(2018).

    [11] Y Liu, G Z Shi, Z K Liu et al. Toward printable solar cells based on PbX colloidal quantum dot inks. Nanoscale Horiz, 6, 8(2021).

    [12] G H Carey, A L Abdelhady, Z J Ning et al. Colloidal quantum dot solar cells. Chem Rev, 115, 12732(2015).

    [13] R Saran, R J Curry. Lead sulphide nanocrystal photodetector technologies. Nat Photonics, 10, 81(2016).

    [14] O Voznyy, B R Sutherland, A H Ip et al. Engineering charge transport by heterostructuring solution-processed semiconductors. Nat Rev Mater, 2, 17026(2017).

    [15] D P Li, J R Ma, W B Liu et al. Enhancing performance of inverted quantum-dot light-emitting diodes based on a solution-processed hole transport layer via ligand treatment. J Semicond, 44, 092603(2023).

    [16] X W Qu, X W Sun. Impedance spectroscopy for quantum dot light-emitting diodes. J Semicond, 44, 091603(2023).

    [17] J Fenner, A Rabenau, G Trageser. Solid-state chemistry of thio-, seleno-, and tellurohalides of representative and transition elements. Advances in Inorganic Chemistry and Radiochemistry. Amsterdam: Elsevier, 329(1980).

    [18] B Krebs. Die kristallstrukturen von Pb4SeBr6, Pb5S2J6 and Pb7S2Br10. Z Für Anorg Und Allg Chem, 396, 137(1973).

    [19] S Toso, Q A Akkerman, B Martín-García et al. Nanocrystals of lead chalcohalides: A series of kinetically trapped metastable nanostructures. J Am Chem Soc, 142, 10198(2020).

    [20] M Imran, L C Peng, A Pianetti et al. Halide perovskite–lead chalcohalide nanocrystal heterostructures. J Am Chem Soc, 143, 1435(2021).

    [21] S K Dutta, L C Peng, B Hudait et al. Halide perovskite cluster precursors: A paradigm for obtaining structure- and color-tunable light-emitting nanocrystals. ACS Energy Lett, 7, 3177(2022).

    [22] H W Qiu, F Li, S He et al. Epitaxial CsPbBr3/CdS Janus nanocrystal heterostructures for efficient charge separation. Adv Sci, 10, 2206560(2023).

    [23] S Toso, M Imran, E Mugnaioli et al. Halide perovskites as disposable epitaxial templates for the phase-selective synthesis of lead sulfochloride nanocrystals. Nat Commun, 13, 3976(2022).

    [24] M Vafaie, J Z Fan, A Morteza Najarian et al. Colloidal quantum dot photodetectors with 10-ns response time and 80% quantum efficiency at 1, 550nm. Matter, 4, 1042(2021).

    [25] M X Liu, O Voznyy, R Sabatini et al. Hybrid organic-inorganic inks flatten the energy landscape in colloidal quantum dot solids. Nat Mater, 16, 258(2017).

    [26] W S Shen, Y Liu, L Grater et al. Thickness-variation-insensitive near-infrared quantum dot LEDs. Sci Bull, 68, 2954(2023).

    [27] S Pradhan, M Dalmases, N Taghipour et al. Colloidal quantum dot light emitting diodes at telecom wavelength with 18% quantum efficiency and over 1 MHz bandwidth. Adv Sci, 9, 2200637(2022).

    [28] Y Liu, Y Y Gao, Q Yang et al. Breaking the size limitation of directly-synthesized PbS quantum dot inks toward efficient short-wavelength infrared optoelectronic applications. Angew Chem Int Ed, 62, e202300396(2023).

    [29] J Y Woo, J H Ko, J H Song et al. Ultrastable PbSe nanocrystal quantum dots via in situ formation of atomically thin halide adlayers on PbSe(100). J Am Chem Soc, 136, 8883(2014).

    [30] I Moreels, Y Justo, B De Geyter et al. Size-tunable, bright, and stable PbS quantum dots: A surface chemistry study. ACS Nano, 5, 2004(2011).

    [31] P B Green, Z Q Li, M W B Wilson. PbS nanocrystals made with excess PbCl2 have an intrinsic shell that reduces their stokes shift. J Phys Chem Lett, 10, 5897(2019).

    [32] P B Green, F Y Villanueva, K Z Demmans et al. PbS nanocrystals made using excess lead chloride have a halide-perovskite-like surface. Chem Mater, 33, 9270(2021).

    [33] S Brittman, A E Colbert, T H Brintlinger et al. Effects of a lead chloride shell on lead sulfide quantum dots. J Phys Chem Lett, 10, 1914(2019).

    [34] M A Hines, G D Scholes. Colloidal PbS nanocrystals with size-tunable near-infrared emission: Observation of post-synthesis self-narrowing of the particle size distribution. Adv Mater, 15, 1844(2003).

    [35] W K Bae, K Char, H Hur et al. Single-step synthesis of quantum dots with chemical composition gradients. Chem Mater, 20, 531(2008).

    [36] W N Nan, Y Niu, H Y Qin et al. Crystal structure control of zinc-blende CdSe/CdS core/shell nanocrystals: Synthesis and structure-dependent optical properties. J Am Chem Soc, 134, 19685(2012).

    [37] R T Lechner, G Fritz-Popovski, M Yarema et al. Crystal phase transitions in the shell of PbS/CdS core/shell nanocrystals influences photoluminescence intensity. Chem Mater, 26, 5914(2014).

    [38] Z K Liu, Y X Zhong, I Shafei et al. Tuning infrared plasmon resonances in doped metal-oxide nanocrystals through cation-exchange reactions. Nat Commun, 10, 1394(2019).

    [39] L K Sagar, W Walravens, Q Zhao et al. PbS/CdS core/shell quantum dots by additive, layer-by-layer shell growth. Chem Mater, 28, 6953(2016).

    [40] B Fisher, J M Caruge, D Zehnder et al. Room-temperature ordered photon emission from multiexciton states in single CdSe core-shell nanocrystals. Phys Rev Lett, 94, 087403(2005).

    [41] L Jin, G Sirigu, X Tong et al. Engineering interfacial structure in "Giant" PbS/CdS quantum dots for photoelectrochemical solar energy conversion. Nano Energy, 30, 531(2016).

    [42] C J Imperiale, F Yarur Villanueva, E Nikbin et al. Direct synthesis of ultrasmall PbS nanocrystals passivated with a metal-halide-perovskite-like monolayer. Chem Mater, 36, 4121(2024).

    [43] D R Ni, S Guo, K M Powderly et al. A high-pressure phase with a non-centrosymmetric crystal structure in the PbSe–PbBr2 system. J Solid State Chem, 280, 120982(2019).

    [44] A Rabenau, H Rau. Über sulfidhalogenide des bleis und das Pb4SeBr6. Z Für Anorg Und Allg Chem, 369, 295(1969).

    Yang Liu, Kunyuan Lu, Yujie Zhu, Xudong Hu, Yusheng Li, Guozheng Shi, Xingyu Zhou, Lin Yuan, Xiang Sun, Xiaobo Ding, Irfan Ullah Muhammad, Qing Shen, Zeke Liu, Wanli Ma. Colloidal synthesis of lead chalcogenide/lead chalcohalide core/shell nanostructures and structural evolution[J]. Journal of Semiconductors, 2025, 46(4): 042101
    Download Citation