Next Article in Journal
Bacterial Effect on the Crystallization of Mineral Phases in a Solution Simulating Human Urine
Previous Article in Journal
Historical Perspective on Diffraction Line-Profile Analyses for Crystals Containing Defect Clusters
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Morphologically Controlled Synthesis of Cs2SnCl6 Perovskite Crystals and Their Photoluminescence Activity

Key Laboratory for Ultrafine Materials of Ministry of Education, School of Materials Science and Engineering, East China University of Science and Technology, Shanghai 200237, China
*
Author to whom correspondence should be addressed.
Crystals 2019, 9(5), 258; https://doi.org/10.3390/cryst9050258
Submission received: 17 April 2019 / Revised: 14 May 2019 / Accepted: 16 May 2019 / Published: 18 May 2019
(This article belongs to the Section Inorganic Crystalline Materials)

Abstract

:
The Cs2SnX6 perovskites have attracted much attention due to excellent optoelectronic properties and high stability. In the present work, we have focused on the morphology control and photoluminescence characteristics of the Cs2SnCl6 perovskite crystals. The synthesis process of the Cs2SnCl6 crystals includes two stages composed of the formation of initial crystals and the growth of Cs2SnCl6; the later originated from the oxidization of CsSnCl3. This process has been confirmed by Scanning electron microscope (SEM) and X-rays diffraction (XRD). By controlling the concentration of the initial reactants and hydrochloric acid in the solution to change the supersaturation of the solution, different crystal morphologies, such as truncated octahedron, octahedron, hexapod, quasi-sphere, have been obtained. In relatively a low supersaturation solution, the amount of growth units dominates the crystal growth process to obtain the hexapod and self-assembly crystals. In contrast, in relatively high supersaturation solution, nucleation predominates to yield small size truncated octahedrons and near-spherical Cs2SnCl6 crystals. The synthesized Cs2SnCl6 crystals have shown a wide emission band peaking at 450 nm with full width at half maximum (FWHM) 63 nm due to the defects introduced by Sn2+. The photoluminescence intensities of crystals synthesized at various conditions exhibited considerable difference, which was about 60 times between the highest and the lowest.

1. Introduction

In recent years, all-inorganic metal halide perovskite nanocrystals (generally CsPbX3) have attracted scientific interest due to their attractive optoelectronic properties together with low cost and facile solution synthesis process, which have great potential in photovoltaic devices and light emitting devices [1,2,3]. In spite of their outstanding performance, all-inorganic lead halide perovskite still has to face the thorny issue that lead has intrinsic toxicity [4]. In view of non-toxic Sn2+ ions, tin (Sn)-based perovskite CsSnX3 has been a very good alternative to be used in solar cells, infrared light emitting diodes and lasers [5,6,7,8,9]. Unfortunately, the CsSnX3 perovskites still suffer from the oxidation of Sn2+ to more stable Sn4+, leading to a high sensitivity to the ambient atmosphere (oxygen, moisture, etc.) [5,6,10,11].
As a defect variant of the traditional cubic ABX3 perovskite, Cs2SnX6 has a similar crystal structure with CsPbX3. Cs2SnX6 is considerably stable to oxygen and moisture due to containing Sn4+ rather than Sn2+, therefore it is promising for large-scale production [12]. Recently, numerous publications have reported their preparation methods and potential applications. Wang et al. [13] synthesized Cs2SnI6 nanocrystals and demonstrated that Cs2SnI6 based field effect transistors exhibit the property of P-type semiconductor with high hole mobility (>20 cm2/(Vs)). Dolzhnikov et al. [14] employed a thermal injection method to prepare Cs2SnI6 crystals with different sizes by controlling the reaction time and temperature. Kaltzoglou et al. [12] synthesized Cs2SnCl6, Cs2SnBr6, Cs2SnI6 and Cs2SnI3Br3 crystals at a high temperature by using different solvents [15]. These crystals were employed as hole transport materials in dye-sensitized solar cell with the maximum power conversion efficiency of 4.23% under one solar irradiation [15]. Saparov et al. [4] obtained an n-type Cs2SnCl6 semiconductor film by annealing SnI4 vapor to a glass substrate covered by a CsI film. Xiaofeng Qiu et al. [6] oxidized the CsSnI3 film to Cs2SnI6 one as a light absorbing layer in a solar cell. In addition, Bi-doped Cs2SnCl6 has exhibited high efficient blue light emission originated from [BiSn + VCl] defect complex introduced by Bi3+ [16]. However, these studies have mainly focused on the particle size, optoelectronic characteristics and application of the Cs2SnX6 crystals in the field of optoelectronic devices. To our knowledge, rarely work has been done on their morphology control and photoluminescence characteristics.
Owing to the low solubility product constant of Cs2SnCl6 in aqueous hydrochloric acid, the reaction between cesium chloride and tin chloride is rapid, as a result, Cs2SnCl6 rapidly nucleates and grows. In this work, the growth process of Cs2SnCl6 crystals was studied by controlling the concentration of the reactants and hydrochloric acid. The Cs2SnCl6 crystals have shown interesting morphologies such as truncated octahedron, octahedron, hexapod and quasi-sphere. In addition, we have investigated the photoluminescence (PL) properties of Cs2SnCl6 crystals. They exhibited a similar wide emission band peaking at 450 nm with different intensities under 350 nm excitation.

2. Experiments

Tin (Sn, 99.9%, Aladdin, Shanghai, China), hydrochloric acid (HCl, 36-38%, Titan Technology Co., Ltd., Shanghai, China), cesium chloride (CsCl, 99.99%, Aladdin, Shanghai, China), deionized water (H2O), absolute alcohol (C2H6O, Titan Technology Co., Ltd., Shanghai, China).
Firstly, tin powder was dissolved in a corresponding amount of HCl by a mole ratio of 1:4 to obtain a SnCl2 hydrochloric acid solution. The corresponding amount of CsCl was added to a beaker, then deionized water and HCl (the amount of added HCl expressed as the concentration of hydrochloric acid in the solution) were added according to the calculated value and finally, 1.5 mmol SnCl2 solution was added. The mixture was stirred in the air at room temperature for 7 h and then the precipitate was collected by centrifugation and washed with anhydrous ethanol. Finally, the product naturally dried in an ambient environment. Typically, 0.178 g Sn powders were dissolved in 0.125 mL HCl to prepare SnCl2 solution, 0.505 g CsCl, 26.5 mL deionized water and additional 9 mL HCl were used to prepare the sample at the reaction condition of 0.042 mol/L reactants and 3 mol/L HCl.
The morphology and particle size of the samples were measured with a Hitachi S-4800 scanning electron microscopy (SEM). The photoluminescence (PL) and photoluminescence excitation (PLE) spectra were determined by a Fluorolog-3-P UV-vis-NIR fluorescence spectrophotometer (Jobin Yvon, France). The absorption spectrum was obtained by a UV-2550 UV-VIS spectrophotometer (Shimadzu, Japan).

3. Result and Discussion

3.1. The Crystal Growth and Morphologies of the Cs2SnCl6 Crystals in Different Synthesis Conditions

In this work, SnCl2 solution was prepared by dissolving Sn in HCl with a molar ratio of 1:4. In such a dissolution process, the oxygen from the air partially oxidized SnCl2 to SnCl4 in the presence of excessive HCl, leading to the coexistence of SnCl2 and SnCl4 in the obtained SnCl2 solution (still expressed as SnCl2 for convenience). Therefore, the formation process of the Cs2SnCl6 crystals includes two stages as follows:
Stage 1:
SnCl4 + 2HCl → H2[SnCl6]
H2[SnCl6] + 2CsCl → Cs2SnCl6↓ + 2HCl
SnCl2 + 2HCl → H[SnCl3]
H[SnCl3] + CsCl → CsSnCl3↓ + HCl
Stage 2:
CsSnCl3 → CsCl + SnCl2
SnCl2 + 2HCl + O2 → SnCl4 + H2O
SnCl4 + 2HCl → H2[SnCl6]
H2[SnCl6] + 2CsCl → Cs2SnCl6↓ + 2HCl
In stage 2, an overall balanced reaction will be:
CsSnCl3 + 2HCl + CsCl + 1/2O2 → Cs2SnCl6↓ + H2O
Such a process has been examined by XRD patterns of the samples at different reaction time (shown in Figure S1, see Supplementary materials). CsSnCl3 and Cs2SnCl6 crystals firstly coexisted in the samples and then the diffraction peaks of CsSnCl3 gradually disappeared and those of Cs2SnCl6 gradually increased with increasing the reaction time; finally, only the diffraction peaks of Cs2SnCl6 crystals were observed after a reaction time larger than 6 h. Therefore, the final crystal formation process consisted of two stages above mentioned. The first one was the fast nucleation and growth process of the initial Cs2SnCl6 and CsSnCl3 crystals according to reactions (1)–(4), which were related to the concentration of HCl and initial reactants. The second one was the growth process of Cs2SnCl6 oxidized from CsSnCl3 (see the Equations (5)–(8) about the oxidization and precipitation process) on the initial Cs2SnCl6 crystals during the first stage, that is, the initial Cs2SnCl6 crystals acted as the nuclei of the second growth stage, the growth rate also depended on the concentration of HCl and reactants. As the overall content of Cs2SnCl6 crystals was constant when the same initial reactants content had been employed, the ratio of Cs2SnCl6 crystals between the first and second stage remarkably impaired the crystal morphology.
Figure 1 presents the SEM images of the Cs2SnCl6 crystals (determined by XRD patterns in Figure S2 in Supplementary materials) synthesized by using different amounts of hydrochloric acid for 0.042 mol/L SnCl2 solution. At 3 mol/L HCl, the complete octahedral crystals can be observed in Figure 1a. When the HCl concentration was increased to 5 mol/L, the crystals showed a uniform hexapod structure containing six thin and long carambola-like branches (Figure 1b). Interestingly, when the HCl concentration was increased to 9 and 10 mol/L, the hexapod crystals at different growth periods (Figure 1c,d), with an expanded volume and enlarged branches, were present clearly, in which some regular octahedrons with eight holes in eight faces of the octahedron can be seen apparently. These regular octahedrons were obviously composed of six small octahedral crystals growing along three orthogonal directions. The holes among these small octahedrons were gradually filled up into a complete smooth octahedron structure. To clearly be seen, the typical morphologies of the Cs2SnCl6 crystals at different HCl concentration have been shown in Figure 2 accompanied by the inset simulated diagram. For a face-centered cubic crystal [12], according to the law of Bravais [17], the normal sequence of crystal faces appearing is {111}, {100}, {110}. Therefore, the initial Cs2SnCl6 crystal nuclei are normal octahedral structure with eight {111} faces (Figure 2a). The growth process of the crystal changes from the octahedral spiral growth to the continuous growth of the branched crystal as the driving force (that is, the amount of the growth units Cs2SnCl6) increases [18], as a result, the morphologies of obtained Cs2SnCl6 crystals changed from the octahedron to the hexapod (Figure 2b–d) and then to the self-assembled bigger regular octahedron (Figure 2e,f). Such morphology changing process is akin to the evolution of the Cu2O crystal morphologies reported in the literature, which is called self-assembly [19].
Figure 3 shows the morphologies of the Cs2SnCl6 crystals (determined by XRD, see Figure S3 in Supplementary materials) obtained by using different amounts of HCl for 0.083 mol/L SnCl2 solution. At 3 mol/L HCl, there were some complete octahedrons (about 20 μm) accompanied by incomplete regular octahedrons (about 20–30 μm) with holes, which were similar to those in Figure 1. At 4 mol/L HCl, a uniform carambola-like structure with particle size about 20 μm was formed by rapidly growing incomplete octahedrons (Figure 3b). Six carambola-like particles may assemble to a hexapod. By increasing the HCl concentration, the length of the hexapod branch decreased gradually leading to a smaller and smaller particle size. Finally, the morphologies varied from the hexapod to the truncated octahedron and then to the quasi-sphere.
These changes in morphology and size with increasing HCl content are related to the homogeneous nucleation and growth process of the crystals. The amount and morphology of the initial Cs2SnCl6 crystals depend on the supersaturation of the solution in the first stage of the initial crystal formation, the higher supersaturation results in larger amounts of nuclei as well as the amount of the initial Cs2SnCl6 crystals. As the initial Cs2SnCl6 crystals act as the nuclei of the second growth stage, the final particle size and morphology are related to the amount of Cs2SnCl6 oxidized from Cs2SnCl3 at stage 2 (see reactions (5–9)). This suggested that the ratio of Cs2SnCl6 crystals (defined as R1/2) between the first and second stage plays a key role in the crystal morphology as mentioned above. During the two crystal formation stages, the final complete octahedron can be obtained if the amount of the initial Cs2SnCl6 crystals matches that during the second growth stage, in this case, we define R1/2 as b (means balance). R1/2 larger than b means the larger amount of the initial crystals, contrarily the larger amount of Cs2SnCl6 produced in the second growth stage as the growth units. The crystals with the different dendritic degree form if R1/2 > b because of insufficient Cs2SnCl6 growth units. Increasing the crystal growth units, the crystals grow from the dendritic to the self-assembly structure, and then to the complete smooth octahedron due to the continual growth of the growth units around the initial dendritic surface, in line with the morphology changing process in the case of the 0.042 mol/L SnCl2 solution (Figure 2). In contrast, at the reactant concentration of 0.083 mol/L, the larger supersaturation led to more nuclei during the first stage, thus the growth units comparatively decreased, resulting in R1/2 > b. As a result, the morphologies of the Cs2SnCl6 crystals gradually changed from the fast-growing dendritic crystals to the fast-nucleating and slow-growing quasi-spherical particles, and the size decreased accordingly.
The difference of morphology changing between 0.042 and 0.083 mol/L reactants originated from the larger difference in the reactant concentration. Figure 4 presents the SEM images of the crystals with the different reactant concentrations at 5 mol/L HCl (For convenient comparison, the images of 0.042 and 0.083 mol/L have also shown in this figure). When the concentration of the reactants increased to 0.165 mol/L, the crystals exhibited a truncated octahedron structure with a particle size of only 1–2 μm as well as increasing the HCl concentration to 0.083 mol/L. At higher concentrations, nanosize Cs2SnCl6 crystals with impurities (the XRD patterns shown in Figure S4) were obtained (determined by Transmission electron microscope (TEM), shown in Figure S5 in Supplementary materials). According to the above discussion, the effect of the reaction conditions on the morphology of the Cs2SnCl6 crystals has been summarized in Figure 5 and Figure S6.

3.2. Photoluminescence Characteristics

As a kind of perovskite, Cs2SnCl6 also exhibited highly efficient blue light emission originated from [BiSn + VCl] defect complex introduced by Bi3+ [16], however, no report about the photoluminescence of the Cs2SnCl6 crystal without dopants has been published except for the recent work [20] to the best of our knowledge. Figure 6a shows the PL spectra of the crystals prepared at 0.083 mol/L SnCl2 solution and different HCl concentrations under 350 nm excitation; the PLE and PL spectra of the crystals in Figure 3g are shown in Figure 6b. The PL spectrum of the Cs2SnCl6 crystals has a wide emission band peaking at 450 nm with FWHM 63 nm and a large Stokes shift about 100 nm. The absorption spectra (Figure S7) have shown weak and wide absorption band about 350 nm when the crystals exhibited strong photoluminescence characteristics, reversely, this absorption band could not be distinguished from the non-photoluminescence crystals. This is similar to the literature [16], indicating that such emission spectra also originate from some kinds of defects, which are VSn(IV) and VCl due to Sn2+ in the crystals (see XPS spectra in Figure S8). It suggests the existence of the characteristic 3d3/2 (495.7 eV) and 3d5/2 (487.2 eV) peaks of Sn2+ in addition to Sn4+ 3d3/2 (496.1 eV) and 3d5/2 (487.6 eV) [16,19]. The intensity of the PL emission peak depends on the number of defects in the crystals, and that of high purity and uniform crystals is relatively weaker [21,22], the highest one is 60 times larger than the lowest one as shown in Figure 6a. However, the relationship between the photoluminescence intensity and the crystal morphology is still irregular. Besides, the huge difference in photoluminescence intensity is significantly associated with the number of defects owing to the different reaction condition. These have been discussed in detail in our recent work [22].

4. Conclusions

In summary, the different crystal morphologies of Cs2SnCl6, such as truncated octahedron, octahedron, hexapod, quasi-sphere, have been obtained by controlling the initial reactants and HCl concentration in the solution. The amount of crystal nucleation and growth units increases as increasing the initial reactants HCl concentration. When the amount of nucleation is more than the growth units, nucleation controls the crystal growth to obtain crystals with different dendrites. Conversely, when the number of growth units is larger, the driving force controls the crystal growth to obtain the hexapod and self-assembled crystals. Under 350 nm excitation, the Cs2SnCl6 crystals exhibit a wide emission band peaking at 450 nm with an FWHM of 63 nm and a large Stokes shift of 100 nm. Due to the influence of the defects number, the highest one is 60 times larger than the lowest one.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4352/9/5/258/s1, Figure S1: XRD patterns of the Cs2SnCl6 crystals prepared at different time, with a hydrochloric acid concentration of 5 mol/L and an SnCl2 concentration of 0.083 mol/L, Figure S2: XRD patterns of the Cs2SnCl6 crystals synthesized at different HCl concentrations (3, 5, 9, 10 mol/L) in the aqueous solution with SnCl2 concentration 0.042 mol/L, Figure S3: XRD patterns of the Cs2SnCl6 crystals synthesized at different HCl concentrations in the aqueous solution with SnCl2 concentration 0.083 mol/L, Figure S4: XRD patterns of the Cs2SnCl6 crystals synthesized at different SnCl2 concentration in the aqueous solution with a hydrochloric acid concentration of 5 mol/L, Figure S5: HR-TEM images of the Cs2SnCl6 crystals prepared at normal temperature in the aqueous solution with a hydrochloric acid concentration of 5 mol/L and a SnCl2 concentration of (a, b) 0.208 mol/L, (c, d) 0.250 mol/L, Figure S6: Summary of crystal morphology and synthesis conditions, Figure S7: UV-vis absorption spectroscopy of the Cs2SnCl6 crystals prepared with 0.083 mol/mL SnCl2 and different concentration of HCl. The illustration is a partial magnification between 300–400 nm, Figure S8: (a) XPS survey spectrum of Cs2SnCl6 crystals prepared with 0.083 mol/mL SnCl2 and different concentration of HCl. (b) High-resolution XPS spectra and peak fitting for Sn 3d.

Author Contributions

Investigation, H.Z. and L.Z.; Resource, J.C., L.C. and C.L.; Writing—Original Draft Preparation, H.Z.; Writing—Review and Editing, S.Y.

Conflicts of Interest

We declare that we have no financial and personal relationships with other people or organizations that can inappropriately influence our work, there is no professional or other personal interest of any nature or kind in any product, service and/or company that could be construed as influencing the position presented in, or the review of, the manuscript entitled.

References

  1. Palazon, F.; Di Stasio, F.; Akkerman, Q.A.; Krahne, R.; Prato, M.; Manna, L. Polymer-free Films of Inorganic Halide Perovskite Nanocrystals as UV-to-White Color-conversion Layers in LEDs. Chem. Mater. 2016, 28, 2902–2906. [Google Scholar] [CrossRef]
  2. Wang, Y.; Li, X.; Song, J.; Xiao, L.; Zeng, H.; Sun, H. All-inorganic Colloidal Perovskite Quantum Dots: A New Class of Lasing Materials with Favorable Characteristics. Adv. Mater. 2015, 27, 7101–7108. [Google Scholar] [CrossRef]
  3. Ramasamy, P.; Lim, D.H.; Kim, B.; Lee, S.H.; Lee, M.S.; Lee, J.S. All-inorganic cesium lead halide perovskite nanocrystals for photodetector applications. Chem. Commun. 2016, 52, 2067–2070. [Google Scholar] [CrossRef] [PubMed]
  4. Saparov, B.; Sun, J.P.; Meng, W.; Xiao, Z.; Duan, H.S.; Gunawan, O.; Shin, D.; Hill, I.G.; Yan, Y.; Mitzi, D.B. Thin-film Deposition and Characterization of a Sn-deficient Perovskite Derivative Cs2SnI6. Chem. Mater. 2016, 28, 2315–2322. [Google Scholar] [CrossRef]
  5. Lee, B.; Stoumpos, C.C.; Zhou, N.; Hao, F.; Malliakas, C.; Yeh, C.Y.; Marks, T.J.; Kanatzidis, M.G.; Chang, R.P. Air-stable Molecular Semiconducting Iodosalts for Solar Cell Applications: Cs2SnI6 as a Hole Conductor. J. Am. Chem. Soc. 2014, 136, 15379–15385. [Google Scholar] [CrossRef]
  6. Qiu, X.; Cao, B.; Yuan, S.; Chen, X.; Qiu, Z.; Jiang, Y.; Ye, Q.; Wang, H.; Zeng, H.; Liu, J.; et al. From Unstable CsSnI3 to Air-stable Cs2SnI6: A Lead-free Perovskite Solar Cell Light Absorber with Bandgap of 1.48eV and High Absorption Coefficient. Sol. Energ. Mat. Sol. C. 2017, 159, 227–234. [Google Scholar] [CrossRef]
  7. Noel, N.K.; Stranks, S.D.; Abate, A.; Wehrenfennig, C.; Guarnera, S.; Haghighirad, A.A.; Sadhanala, A.; Eperon, G.E.; Pathak, S.K.; Johnston, M.B.; et al. Lead-free Organic–inorganic Tin Halide Perovskites for Photovoltaic Applications. Energ. Environ. Sci. 2014, 7, 3061–3068. [Google Scholar] [CrossRef]
  8. Xing, G.; Kumar, M.H.; Chong, W.K.; Liu, X.; Cai, Y.; Ding, H.; Asta, M.; Gratzel, M.; Mhaisalkar, S.; Mathews, N.; et al. Solution-processed Tin-based Perovskite for Near-infrared Lasing. Adv. Mater. 2016, 28, 8191–8196. [Google Scholar] [CrossRef]
  9. Lai, M.L.; Tay, T.Y.; Sadhanala, A.; Dutton, S.E.; Li, G.; Friend, R.H.; Tan, Z.K. Tunable Near-infrared Luminescence in Tin Halide Perovskite Devices. J. Phys. Chem. Lett. 2016, 7, 2653–2658. [Google Scholar] [CrossRef]
  10. Qiu, X.; Jiang, Y.; Zhang, H.; Qiu, Z.; Yuan, S.; Wang, P.; Cao, B. Lead-free Mesoscopic Cs2SnI6 Perovskite Solar Cells Using Different Nanostructured ZnO Nanorods As Electron Transport Layers. Phy. Status Solidi-RRL. 2016, 10, 587–591. [Google Scholar] [CrossRef]
  11. Xiao, Z.; Zhou, Y.; Hosono, H.; Kamiya, T. Intrinsic Defects in a Photovoltaic Perovskite Variant Cs2SnI6. Phys. Chem. Chem. Phys. 2015, 17, 18900–18903. [Google Scholar] [CrossRef] [PubMed]
  12. Kaltzoglou, A.; Antoniadou, M.; Kontos, A.G.; Stoumpos, C.C.; Perganti, D.; Siranidi, E.; Raptis, V.; Trohidou, K.; Psycharis, V.; Kanatzidis, M.G.; et al. Optical-vibrational Properties of the Cs2SnX6(X = Cl, Br, I) Defect Perovskites and Hole-transport Efficiency in Dye-sensitized Solar Cells. J. Phys. Chem. C 2016, 120, 11777–11785. [Google Scholar] [CrossRef]
  13. Wang, A.; Yan, X.; Zhang, M.; Sun, S.; Yang, M.; Shen, W.; Pan, X.; Wang, P.; Deng, Z. Controlled Synthesis of Lead-free and Stable Perovskite Derivative Cs2SnI6 Nanocrystals via a Facile Hot-injection Process. Chem. Mater. 2016, 28, 8132–8140. [Google Scholar] [CrossRef]
  14. Dolzhnikov, D.S.; Wang, C.; Xu, Y.; Kanatzidis, M.G.; Weiss, E.A. Ligand-free, Quantum-confined Cs2SnI6 Perovskite Nanocrystals. Chem. Mater. 2017, 29, 7901–7907. [Google Scholar] [CrossRef]
  15. Kaltzoglou, A.; Antoniadou, M.; Perganti, D.; Siranidi, E.; Raptis, V.; Trohidou, K.; Psycharis, V.; Kontos, A.G.; Falaras, P. Mixed-halide Cs2SnI3Br3 Perovskite as Low Resistance Hole-transporting Material in Dye-sensitized Solar Cells. Electrochim. Acta 2015, 184, 466–474. [Google Scholar] [CrossRef]
  16. Tan, Z.; Li, J.; Zhang, C.; Li, Z.; Hu, Q.; Xiao, Z.; Kamiya, T.; Hosono, H.; Niu, G.; Lifshitz, E.; et al. Highly Efficient Blue-emitting Bi-doped Cs2SnCl6 Perovskite Variant: Photoluminescence Induced by Impurity Doping. Adv. Funct. Mater. 2018, 28, 1801131. [Google Scholar] [CrossRef]
  17. Donnay, D.J.D.; Harker, D. A New Law of Crystal Morphology Extending the Law of Bravais. Am. Mineral. 1937, 22, 446–467. [Google Scholar]
  18. Sunagawa, I. Growth and Morphology of Crystals. Forma-Tokyo 1999, 14, 147–166. [Google Scholar]
  19. Giuseppe, A.; Nuccioo, B.; John, R.F.; Antta, F.; Giovanna, I.; Giovanni, P.; Maria, V.R. Mössbauer, Far-infrared, and XPS Investigations of SnCl2 and SnCl4 Introduced in Polyconjugated Monosubstituted Acetylene Matrices. Appl. Spectrosc. 1995, 49, 237–240. [Google Scholar]
  20. Zhang, H.; Zhu, L.; Cheng, J.; Chen, L.; Liu, C.; Yuan, S. Photoluminescence Characteristics of Sn2+ and Ce3+-doped Cs2SnCl6 Double-perovskite Crystals. Materials 2019, 12, 1501. [Google Scholar] [CrossRef] [PubMed]
  21. Ho, J.Y.; Huang, M.H. Synthesis of Submicrometer-sized Cu2O Crystals with Morphological Evolution from Cubic to Hexapod Structures and Their Comparative Photocatalytic Activity. J. Phys. Chem. C 2009, 113, 14159–14164. [Google Scholar] [CrossRef]
  22. Zhang, B.; Guo, F.; Yang, L.; Jia, X.; Liu, B.; Xie, Z.; Chen, D.; Lu, H.; Zhang, R.; Zheng, Y. Shape-evolution Control of Hybrid Perovskite CH3NH3PbI3 Crystals via Solvothermal Synthesis. J. Cryst. Growth 2017, 459, 167–172. [Google Scholar] [CrossRef]
Figure 1. The SEM images of the Cs2SnCl6 crystals prepared with (a) 3 mol/L, (b) 5 mol/L, (c) 9 mol/L, (d) 10 mol/L HCl for 0.042 mol/L SnCl2 solution.
Figure 1. The SEM images of the Cs2SnCl6 crystals prepared with (a) 3 mol/L, (b) 5 mol/L, (c) 9 mol/L, (d) 10 mol/L HCl for 0.042 mol/L SnCl2 solution.
Crystals 09 00258 g001
Figure 2. The self-assembly process of Cs2SnCl6 crystal (A) octahedron, (BD) Hexapod, (E,F) Incomplete octahedron.
Figure 2. The self-assembly process of Cs2SnCl6 crystal (A) octahedron, (BD) Hexapod, (E,F) Incomplete octahedron.
Crystals 09 00258 g002
Figure 3. The SEM images of the Cs2SnCl6 crystals prepared by (a) 3 mol/L, (b) 4 mol/L, (c) 5 mol/L, (d) 6 mol/L, (e) 7 mol/L, (f) 8 mol/L, (g) 9 mol/L, (h) 10 mol/L HCl for 0.083 mol/L SnCl2.
Figure 3. The SEM images of the Cs2SnCl6 crystals prepared by (a) 3 mol/L, (b) 4 mol/L, (c) 5 mol/L, (d) 6 mol/L, (e) 7 mol/L, (f) 8 mol/L, (g) 9 mol/L, (h) 10 mol/L HCl for 0.083 mol/L SnCl2.
Crystals 09 00258 g003
Figure 4. The SEM images of the Cs2SnCl6 crystals prepared at 5 mol/L HCl concentration for different SnCl2 concentrations: (a) 0.042 mol/L, (b) 0.083 mol/L, (c) 0.167 mol/L.
Figure 4. The SEM images of the Cs2SnCl6 crystals prepared at 5 mol/L HCl concentration for different SnCl2 concentrations: (a) 0.042 mol/L, (b) 0.083 mol/L, (c) 0.167 mol/L.
Crystals 09 00258 g004
Figure 5. The summary flowchart of the crystal morphologies and synthesis conditions.
Figure 5. The summary flowchart of the crystal morphologies and synthesis conditions.
Crystals 09 00258 g005
Figure 6. (a) The PL spectra of the Cs2SnCl6 crystals in Figure 3. (b) The absorption, PL and PLE spectra of the Cs2SnCl6 crystals in Figure 3g.
Figure 6. (a) The PL spectra of the Cs2SnCl6 crystals in Figure 3. (b) The absorption, PL and PLE spectra of the Cs2SnCl6 crystals in Figure 3g.
Crystals 09 00258 g006

Share and Cite

MDPI and ACS Style

Zhang, H.; Zhu, L.; Cheng, J.; Chen, L.; Liu, C.; Yuan, S. Morphologically Controlled Synthesis of Cs2SnCl6 Perovskite Crystals and Their Photoluminescence Activity. Crystals 2019, 9, 258. https://doi.org/10.3390/cryst9050258

AMA Style

Zhang H, Zhu L, Cheng J, Chen L, Liu C, Yuan S. Morphologically Controlled Synthesis of Cs2SnCl6 Perovskite Crystals and Their Photoluminescence Activity. Crystals. 2019; 9(5):258. https://doi.org/10.3390/cryst9050258

Chicago/Turabian Style

Zhang, Hongdan, Ludan Zhu, Jun Cheng, Long Chen, Chuanqi Liu, and Shuanglong Yuan. 2019. "Morphologically Controlled Synthesis of Cs2SnCl6 Perovskite Crystals and Their Photoluminescence Activity" Crystals 9, no. 5: 258. https://doi.org/10.3390/cryst9050258

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop