research communications\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Crystal structure of strontium perchlorate anhydrate, Sr(ClO4)2, from laboratory powder X-ray diffraction data

CROSSMARK_Color_square_no_text.svg

aDaegu Gyeongbuk Institute of Science and Technology (DGIST), Daegu 42988, Republic of Korea
*Correspondence e-mail: st.hong@dgist.ac.kr

Edited by M. Weil, Vienna University of Technology, Austria (Received 16 January 2019; accepted 8 March 2019; online 15 March 2019)

The crystal structure of strontium perchlorate anhydrate, Sr(ClO4)2, was determined and refined from laboratory powder X-ray diffraction data. The material was obtained by dehydration of Sr(ClO4)2·3H2O at 523 K for two weeks. It crystallizes in the ortho­rhom­bic space group Pbca and is isotypic with Ca(AlD4)2 and Ca(ClO4)2. The asymmetric unit contains one Sr, two Cl and eight O sites, all on general positions (Wyckoff position 8c). The crystal structure consists of Sr2+ cations and isolated ClO4 tetra­hedra. The Sr2+ cation is coordinated by eight O atoms from eight ClO4 tetra­hedra. The validity of the crystal structure model for Sr(ClO4)2 anhydrate was confirmed by the bond valence method.

1. Chemical context

The alkaline earth metal ions (Mg2+, Ca2+, Sr2+ and Ba2+) have received attention as ion carriers for next-generation batteries (Wang et al., 2013[Wang, R. Y., Wessells, C. D., Huggins, R. A. & Cui, Y. (2013). Nano Lett. 13, 5748-5752.]), and their perchlorates are used as inorganic salts of conventional nona­queous electrolytes for electrochemical cells in Mg- and Ca-ion batteries (Whittingham et al., 2018[Whittingham, M. S., Siu, C. & Ding, J. (2018). Acc. Chem. Res. 51, 258-264.]; Tchitchekova et al., 2017[Tchitchekova, D. S., Monti, D., Johansson, P., Bardé, F., Randon-Vitanova, A., Palacín, M. R. & Pnrouch, A. (2017). J. Electrochem. Soc. 164, A1384-A1392.]; Padigi et al., 2015[Padigi, P., Goncher, G., Evans, D. & Solanki, R. (2015). J. Power Sources, 273, 460-464.]). It is crucial to obtain anhydrous salts to achieve high electrochemical cell performance since hydrated salts can cause unwanted side reactions as a result of increased water content in the nona­queous electrolyte. Strontium perchlorate is highly hygroscopic and exists in several hydrated forms. So far, Sr(ClO4)2·3H2O, Sr(ClO4)2·4H2O and Sr(ClO4)2·9H2O have been identified by single-crystal X-ray diffraction (Hennings et al., 2014[Hennings, E., Schmidt, H. & Voigt, W. (2014). Acta Cryst. E70, 510-514.]). However, the crystal structure of the anhydrous phase has not been reported to date because of the difficulty in growing single crystals. Previously, we have determined the structures of anhydrous magnesium, barium and calcium perchlorate from laboratory powder X-ray diffraction data (Lim et al., 2011[Lim, H.-K., Choi, Y. S. & Hong, S.-T. (2011). Acta Cryst. C67, i36-i38.]; Lee et al., 2015[Lee, J. H., Kang, J. H., Lim, S.-C. & Hong, S.-T. (2015). Acta Cryst. E71, 588-591.], 2018[Lee, D., Bu, H., Kim, D., Hyoung, J. & Hong, S.-T. (2018). Acta Cryst. E74, 514-517.]). Using the same techniques for the Sr salt, we were able to determine and refine the crystal structure of strontium perchlorate anhydrate.

2. Structural commentary

The crystal structure of anhydrous strontium perchlorate, Sr(ClO4)2, is isotypic with Ca(AlD4)2 (Sato et al., 2009[Sato, T., Sørby, M. H., Ikeda, K., Sato, S., Hauback, B. C. & Orimo, S. (2009). J. Alloys Compd. 487, 472-478.]) and Ca(ClO4)2 (Lee et al., 2018[Lee, D., Bu, H., Kim, D., Hyoung, J. & Hong, S.-T. (2018). Acta Cryst. E74, 514-517.]). Compared with Ca(ClO4)2, the unit-cell parameters a, b and c of Sr(ClO4)2 are increased by 3.0, 2.9 and 3.4%, respectively, because Sr2+ (1.26 Å for eight-coordination) has a larger ionic radius than Ca2+ (1.12 Å for eight-coordination; Shannon, 1976[Shannon, R. D. (1976). Acta Cryst. A32, 751-767.]).

There are one Sr, two Cl and eight O sites in the asymmetric unit, all on general positions 8c. The crystal structure (Fig. 1[link]) is composed of Sr2+ cations and isolated ClO4 tetra­hedra. The isolated ClO4 tetra­hedra are slightly distorted and exhibit a range of 105.4 (7)–113.5 (7)° for the O—Cl—O angles. The local environment around the Sr2+ cation is presented in Fig. 2[link]. It is coordinated by eight O atoms from eight ClO4 tetra­hedra, with an average Sr—O distance of 2.582 Å (Table 1[link]). The latter is inter­mediate between those of Ca—O (2.476 Å; Lee et al., 2018[Lee, D., Bu, H., Kim, D., Hyoung, J. & Hong, S.-T. (2018). Acta Cryst. E74, 514-517.]) and Ba—O (2.989 Å; Lee et al., 2015[Lee, J. H., Kang, J. H., Lim, S.-C. & Hong, S.-T. (2015). Acta Cryst. E71, 588-591.]) polyhedra, and in good agreement with the sum of the ionic radii of the respective alkaline earth metal and oxygen ions (Shannon, 1976[Shannon, R. D. (1976). Acta Cryst. A32, 751-767.]).

Table 1
Selected bond lengths (Å)

Sr1—O1i 2.512 (5) Cl1—O5 1.436 (5)
Sr1—O2ii 2.591 (5) Cl1—O6 1.436 (5)
Sr1—O3iii 2.546 (5) Cl1—O7 1.426 (5)
Sr1—O4iv 2.622 (5) Cl1—O8 1.423 (5)
Sr1—O5v 2.650 (5) Cl2—O1 1.469 (5)
Sr1—O6ii 2.540 (5) Cl2—O2 1.414 (5)
Sr1—O7vi 2.590 (5) Cl2—O3 1.425 (5)
Sr1—O8 2.604 (8) Cl2—O4 1.422 (5)
Symmetry codes: (i) [x+{\script{1\over 2}}, -y+{\script{1\over 2}}, -z+1]; (ii) -x+1, -y+1, -z+1; (iii) [x+{\script{1\over 2}}, y, -z+{\script{1\over 2}}]; (iv) [-x+{\script{1\over 2}}, -y+1, z-{\script{1\over 2}}]; (v) [-x+1, y+{\script{1\over 2}}, -z+{\script{1\over 2}}]; (vi) [x, -y+{\script{1\over 2}}, z-{\script{1\over 2}}].
[Figure 1]
Figure 1
The local environment of the Sr2+ cation (yellow sphere) surrounded by ClO4 tetra­hedra (purple). Symmetry codes refer to Table 1[link].
[Figure 2]
Figure 2
The crystal structure of Sr(ClO4)2 in two different viewing directions, i.e. approximately along (a) [001] and (b) [010]. Sr2+ cations are yellow and ClO4 tetra­hedra are purple.

Empirical bond valence sums (BVSs) can be used to check structure models (Brown, 2002[Brown, I. D. (2002). In The Chemical Bond in Inorganic Chemistry. Oxford University Press.]). In this regard, the BVSs for the ions in the crystal structure of Sr(ClO4)2 were calculated with the program Valence (Brown & Altermatt, 1985[Brown, I. D. & Altermatt, D. (1985). Acta Cryst. B41, 244-247.]; Brese & O'Keeffe, 1991[Brese, N. E. & O'Keeffe, M. (1991). Acta Cryst. B47, 192-197.]; Hormillosa et al., 1993[Hormillosa, C., Healy, S., Stephen, T. & Brown, I. D. (1993). Bond Valence Calculator. Version 2.0. McMaster University, Canada. https://www.CCP14.ac.uk/solution/bond_valence/.]). The expected charges of the ions match well with the obtained BVS values (given in valence units), thus confirming the validity of the crystal structure: Sr 2.18, Cl1 6.99, Cl2 6.96, O1 1.91, O2 2.08, O3 2.06, O4 2.03, O5 1.96, O6 2.02, O7 2.03 and O8 2.04.

3. Synthesis and crystallization

Anhydrous strontium perchlorate was obtained by dehydration of Sr(ClO4)2·3H2O (98%, Alfa Aesar). The hydrated Sr(ClO4)2 powder was ground thoroughly in an agate mortar and added to a glass bottle. The bottle was placed in an oven at 523 K for two weeks under atmospheric conditions and then transferred to a glove-box under an argon atmosphere. For the powder X-ray diffraction measurements, anhydrous Sr(ClO4)2 was again ground in an agate mortar and placed in a tightly sealed dome-type X-ray sample holder commercially available from Bruker. The dome was double-sealed with vacuum grease to prevent hydration during measurement.

4. Refinement details

Details of the crystal data collection and structure refinement are summarized in Table 2[link]. Powder X-ray diffraction (PXRD) data for anhydrous Sr(ClO4)2 were collected from a Bragg–Brentano diffractometer (PANalytical Empyrean) using Cu Kα1 radiation, a focusing primary Ge(111) monochromator (λ = 1.5406 Å) and a position-sensitive PIXcel 3D 2×2 detector. The angular range was 10 ≤ 2θ ≤ 130°, with a step of 0.0131° and a total measurement time of 8 h at room temperature. The PXRD pattern was indexed using the TREOR90 algorithm (Werner, 1990[Werner, P. E. (1990). TREOR90. University of? Stockholm, Sweden.]) run in CRYSFIRE (Shirley, 2002[Shirley, R. (2002). The Crysfire 2002 System for Automatic Powder Indexing: User's Manual. Guildford, UK: The Lattice Press.]) through the positions of 23 reflections, resulting in an ortho­rhom­bic unit cell. Systematic reflection conditions suggested the space group Pbca. The crystal structure was determined by a combination of the powder profile refinement program GSAS (Larson & Von Dreele, 2000[Larson, A. C. & Von Dreele, R. B. (2000). General Structure Analysis System (GSAS). Report LAUR 86-748. Los Alamos National Laboratory, New Mexico, USA.]) and the single-crystal structure refinement program CRYSTALS (Betteridge et al., 2003[Betteridge, P. W., Carruthers, J. R., Cooper, R. I., Prout, K. & Watkin, D. J. (2003). J. Appl. Cryst. 36, 1487.]). For a three-dimensional view of the Fourier electron-density maps, MCE was applied (Rohlícek & Husák, 2007[Rohlíček, J. & Hušák, M. (2007). J. Appl. Cryst. 40, 600-601.]). Initially, a structural model with only one dummy atom at an arbitrary position in the unit cell was used. Structure factors were extracted from the powder data and then direct methods were applied to calculate the initial solution of the crystal structure using SHELXS97 (Sheldrick, 2008[Sheldrick, G. M. (2008). Acta Cryst. A64, 112-122.]) run in CRYSTALS, which yielded the Sr site as a starting atomic position. The initial dummy atom model was then replaced with the partial model, and this data was adopted for a Le Bail fit in GSAS. Improved structure factors were then extracted, which were used for the refinement in CRYSTALS. Such processes were iterated until a complete and satisfactory structural model was obtained. Finally, Rietveld refinement in GSAS was employed to complete the structure model, resulting in reasonable isotropic displacement parameters and agreement indices. For the final Rietveld refinement with GSAS, the Sr—O and Cl—O bond lengths were restrained with a tolerance value of 2% with respect to the distances determined from CRYSTALS, which matched reasonably well with the radii sums of Shannon (1976[Shannon, R. D. (1976). Acta Cryst. A32, 751-767.]). The final Rietveld plot is displayed in Fig. 3[link].

Table 2
Experimental details

Crystal data
Chemical formula Sr(ClO4)2
Mr 286.52
Crystal system, space group Orthorhombic, Pbca
Temperature (K) 298
a, b, c (Å) 14.18206 (10), 9.78934 (11), 9.37624 (10)
V3) 1301.73 (2)
Z 8
Radiation type Cu Kα1, λ = 1.5405 Å
Specimen shape, size (mm) Flat sheet, 24.9 × 24.9
 
Data collection
Diffractometer PANalytical Empyrean
Specimen mounting Packed powder
Data collection mode Reflection
Scan method Step
2θ values (°) 2θmin = 10.009, 2θmax = 129.991, 2θstep = 0.013
 
Refinement
R factors and goodness of fit Rp = 0.086, Rwp = 0.125, Rexp = 0.096, R(F2) = 0.14871, χ2 = 1.716
No. of parameters 40
No. of restraints 23
Computer programs: X'Pert Data Collector (PANalytical, 2011[PANalytical (2011). X'Pert Data Collector and X'Pert HighScore Plus. PANalytical BV, Almelo, The Netherlands.]), GSAS (Larson & Von Dreele, 2000[Larson, A. C. & Von Dreele, R. B. (2000). General Structure Analysis System (GSAS). Report LAUR 86-748. Los Alamos National Laboratory, New Mexico, USA.]), X'Pert HighScore Plus (PANalytical, 2011[PANalytical (2011). X'Pert Data Collector and X'Pert HighScore Plus. PANalytical BV, Almelo, The Netherlands.]), SHELXS97 (Sheldrick, 2008[Sheldrick, G. M. (2008). Acta Cryst. A64, 112-122.]), CRYSTALS (Betteridge et al., 2003[Betteridge, P. W., Carruthers, J. R., Cooper, R. I., Prout, K. & Watkin, D. J. (2003). J. Appl. Cryst. 36, 1487.]) and VESTA (Momma & Izumi, 2011[Momma, K. & Izumi, F. (2011). J. Appl. Cryst. 44, 1272-1276.]).
[Figure 3]
Figure 3
PXRD Rietveld refinement profiles for anhydrous Sr(ClO4)2 measured at ambient temperature. Black dots mark experimental data, the solid red line represents the calculated profile and the solid green line is the background. The bottom trace presents the difference curve (blue) and the ticks denote the expected Bragg reflection positions (magenta).

Supporting information


Computing details top

Data collection: X'Pert Data Collector (PANalytical, 2011); cell refinement: GSAS (Larson & Von Dreele, 2000); data reduction: X'Pert HighScore Plus (PANalytical, 2011); program(s) used to solve structure: SHELXS97 (Sheldrick, 2008) and CRYSTALS (Betteridge et al., 2003); program(s) used to refine structure: GSAS (Larson & Von Dreele, 2000); molecular graphics: VESTA (Momma & Izumi, 2011); software used to prepare material for publication: GSAS (Larson & Von Dreele, 2000).

Strontium perchlorate anhydrate top
Crystal data top
Sr(ClO4)2Z = 8
Mr = 286.52F(000) = 1088.0
Orthorhombic, PbcaDx = 2.925 Mg m3
Hall symbol: -P_2ac_2abCu Kα1 radiation, λ = 1.5405 Å
a = 14.18206 (10) ÅT = 298 K
b = 9.78934 (11) Åwhite
c = 9.37624 (10) Åflat_sheet, 24.9 × 24.9 mm
V = 1301.73 (2) Å3Specimen preparation: Prepared at 298 K
Data collection top
PANalytical Empyrean
diffractometer
Data collection mode: reflection
Radiation source: sealed X-ray tube, PANalytical Cu Ceramic X-ray tubeScan method: step
Specimen mounting: packed powder2θmin = 10.009°, 2θmax = 129.991°, 2θstep = 0.013°
Refinement top
Least-squares matrix: fullProfile function: CW Profile function number 4 with 18 terms Pseudovoigt profile coefficients as parameterized in P. Thompson, D.E. Cox & J.B. Hastings (1987). J. Appl. Cryst.,20,79-83. Asymmetry correction of L.W. Finger, D.E. Cox & A. P. Jephcoat (1994). J. Appl. Cryst.,27,892-900. Microstrain broadening by P.W. Stephens, (1999). J. Appl. Cryst.,32,281-289. #1(GU) = 0.000 #2(GV) = 0.000 #3(GW) = 0.000 #4(GP) = 9.252 #5(LX) = 0.900 #6(ptec) = 0.00 #7(trns) = 0.00 #8(shft) = -3.8100 #9(sfec) = 0.00 #10(S/L) = 0.0208 #11(H/L) = 0.0005 #12(eta) = 0.7500 #13(S400 ) = 0.0E+00 #14(S040 ) = 7.8E-04 #15(S004 ) = 1.5E-04 #16(S220 ) = 3.7E-04 #17(S202 ) = 6.1E-04 #18(S022 ) = -1.1E-03 Peak tails are ignored where the intensity is below 0.0010 times the peak Aniso. broadening axis 0.0 0.0 1.0
Rp = 0.08640 parameters
Rwp = 0.12523 restraints
Rexp = 0.096(Δ/σ)max = 0.05
R(F2) = 0.14871Background function: GSAS Background function number 1 with 34 terms. Shifted Chebyshev function of 1st kind 1: 118.082 2: -166.900 3: 123.865 4: -59.7925 5: 10.6865 6: 21.3336 7: -31.0975 8: 23.2200 9: -6.96687 10: -9.51726 11: 20.8794 12: -23.8022 13: 18.4347 14: -9.14997 15: -1.10995 16: 9.35323 17: -13.3633 18: 13.2873 19: -9.61569 20: 4.08246 21: 1.61524 22: -5.79316 23: 6.77390 24: -5.01271 25: 2.27833 26: 0.646733 27: -2.78842 28: 3.78393 29: -3.23100 30: 2.18997 31: -0.908158 32: -0.401332 33: 0.778547 34: -0.792308
9139 data pointsPreferred orientation correction: March-Dollase AXIS 1 Ratio= 0.79858 h= 1.000 k= 0.000 l= 0.000 Prefered orientation correction range: Min= 0.71363, Max= 1.96360
Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
Sr10.60125 (9)0.46444 (17)0.2136 (2)0.0206 (3)*
Cl10.4370 (3)0.2794 (4)0.4905 (4)0.0254 (3)*
Cl20.1592 (2)0.3965 (4)0.6866 (4)0.0275 (3)*
O10.1833 (5)0.2549 (6)0.7237 (14)0.0271 (14)*
O20.2200 (4)0.4878 (11)0.7587 (12)0.0335 (14)*
O30.1636 (7)0.4175 (13)0.5363 (6)0.0364 (14)*
O40.0665 (4)0.4250 (10)0.7362 (11)0.0225 (14)*
O50.3759 (6)0.2145 (7)0.3888 (11)0.0356 (14)*
O60.3829 (6)0.3692 (11)0.5799 (10)0.0361 (14)*
O70.4728 (6)0.1764 (11)0.5835 (12)0.0315 (14)*
O80.5132 (6)0.3514 (12)0.4267 (12)0.0305 (14)*
Geometric parameters (Å, º) top
Sr1—Cl13.931 (4)Cl2—Sr1ix3.945 (3)
Sr1—Cl1i3.937 (4)Cl2—Sr1ii3.778 (3)
Sr1—Cl1ii3.779 (4)Cl2—Sr1x3.746 (4)
Sr1—Cl1iii3.669 (4)Cl2—Sr1xi3.898 (4)
Sr1—Cl2iv3.945 (3)Cl2—O11.469 (5)
Sr1—Cl2ii3.778 (3)Cl2—O21.414 (5)
Sr1—Cl2v3.746 (4)Cl2—O31.425 (5)
Sr1—Cl2vi3.898 (4)Cl2—O41.422 (5)
Sr1—O1v2.512 (5)O1—Sr1x2.512 (5)
Sr1—O2ii2.591 (5)O1—Cl21.469 (5)
Sr1—O3vi2.546 (5)O2—Sr1ii2.591 (5)
Sr1—O4iv2.622 (5)O2—Cl21.414 (5)
Sr1—O5iii2.650 (5)O3—Sr1xi2.546 (5)
Sr1—O6ii2.540 (5)O3—Cl21.425 (5)
Sr1—O7i2.590 (5)O4—Sr1ix2.622 (5)
Sr1—O82.604 (8)O4—Cl21.422 (5)
Cl1—Sr13.931 (4)O5—Sr1viii2.650 (5)
Cl1—Sr1vii3.937 (4)O5—Cl11.436 (5)
Cl1—Sr1ii3.779 (4)O6—Sr1ii2.540 (5)
Cl1—Sr1viii3.669 (4)O6—Cl11.436 (5)
Cl1—O51.436 (5)O7—Sr1vii2.590 (5)
Cl1—O61.436 (5)O7—Cl11.426 (5)
Cl1—O71.426 (5)O8—Sr12.604 (8)
Cl1—O81.423 (5)O8—Cl11.423 (5)
O1v—Sr1—O2ii71.2 (3)O5iii—Sr1—O8136.8 (4)
O1v—Sr1—O3vi84.3 (4)O6ii—Sr1—O7i139.7 (3)
O1v—Sr1—O4iv138.9 (3)O6ii—Sr1—O874.3 (4)
O1v—Sr1—O5iii144.9 (3)O7i—Sr1—O878.3 (4)
O1v—Sr1—O6ii109.2 (4)O5—Cl1—O6109.7 (5)
O1v—Sr1—O7i88.9 (4)O5—Cl1—O7108.0 (7)
O1v—Sr1—O871.3 (4)O5—Cl1—O8113.5 (7)
O2ii—Sr1—O3vi77.6 (3)O6—Cl1—O7105.4 (7)
O2ii—Sr1—O4iv143.6 (3)O6—Cl1—O8110.3 (7)
O2ii—Sr1—O5iii75.5 (3)O7—Cl1—O8109.7 (7)
O2ii—Sr1—O6ii73.8 (3)O1—Cl2—O2110.0 (7)
O2ii—Sr1—O7i146.3 (3)O1—Cl2—O3111.1 (7)
O2ii—Sr1—O8118.0 (4)O1—Cl2—O4108.8 (6)
O3vi—Sr1—O4iv117.7 (3)O2—Cl2—O3110.7 (8)
O3vi—Sr1—O5iii77.9 (3)O2—Cl2—O4106.4 (6)
O3vi—Sr1—O6ii141.9 (4)O3—Cl2—O4109.6 (6)
O3vi—Sr1—O7i73.4 (3)Sr1x—O1—Cl2139.0 (5)
O3vi—Sr1—O8142.8 (4)Sr1ii—O2—Cl2139.4 (5)
O4iv—Sr1—O5iii76.1 (3)Sr1xi—O3—Cl2157.0 (6)
O4iv—Sr1—O6ii75.8 (3)Sr1ix—O4—Cl2153.3 (6)
O4iv—Sr1—O7i67.6 (3)Sr1viii—O5—Cl1125.1 (4)
O4iv—Sr1—O871.2 (4)Sr1ii—O6—Cl1142.2 (5)
O5iii—Sr1—O6ii70.9 (4)Sr1vii—O7—Cl1156.2 (7)
O5iii—Sr1—O7i114.04 (3)Sr1—O8—Cl1153.9 (7)
Symmetry codes: (i) x, y+1/2, z1/2; (ii) x+1, y+1, z+1; (iii) x+1, y+1/2, z+1/2; (iv) x+1/2, y+1, z1/2; (v) x+1/2, y+1/2, z+1; (vi) x+1/2, y, z+1/2; (vii) x, y+1/2, z+1/2; (viii) x+1, y1/2, z+1/2; (ix) x+1/2, y+1, z+1/2; (x) x1/2, y+1/2, z+1; (xi) x1/2, y, z+1/2.
 

Funding information

Funding for this research was provided by: the Creative Materials Discovery Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT and Future Planning (2015M3D1A1069707).

References

First citationBetteridge, P. W., Carruthers, J. R., Cooper, R. I., Prout, K. & Watkin, D. J. (2003). J. Appl. Cryst. 36, 1487.  Web of Science CrossRef IUCr Journals Google Scholar
First citationBrese, N. E. & O'Keeffe, M. (1991). Acta Cryst. B47, 192–197.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationBrown, I. D. (2002). In The Chemical Bond in Inorganic Chemistry. Oxford University Press.  Google Scholar
First citationBrown, I. D. & Altermatt, D. (1985). Acta Cryst. B41, 244–247.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationHennings, E., Schmidt, H. & Voigt, W. (2014). Acta Cryst. E70, 510–514.  CrossRef IUCr Journals Google Scholar
First citationHormillosa, C., Healy, S., Stephen, T. & Brown, I. D. (1993). Bond Valence Calculator. Version 2.0. McMaster University, Canada. https://www.CCP14.ac.uk/solution/bond_valence/Google Scholar
First citationLarson, A. C. & Von Dreele, R. B. (2000). General Structure Analysis System (GSAS). Report LAUR 86-748. Los Alamos National Laboratory, New Mexico, USA.  Google Scholar
First citationLee, D., Bu, H., Kim, D., Hyoung, J. & Hong, S.-T. (2018). Acta Cryst. E74, 514–517.  CrossRef IUCr Journals Google Scholar
First citationLee, J. H., Kang, J. H., Lim, S.-C. & Hong, S.-T. (2015). Acta Cryst. E71, 588–591.  Web of Science CrossRef IUCr Journals Google Scholar
First citationLim, H.-K., Choi, Y. S. & Hong, S.-T. (2011). Acta Cryst. C67, i36–i38.  Web of Science CrossRef IUCr Journals Google Scholar
First citationMomma, K. & Izumi, F. (2011). J. Appl. Cryst. 44, 1272–1276.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationPadigi, P., Goncher, G., Evans, D. & Solanki, R. (2015). J. Power Sources, 273, 460–464.  Web of Science CrossRef CAS Google Scholar
First citationPANalytical (2011). X'Pert Data Collector and X'Pert HighScore Plus. PANalytical BV, Almelo, The Netherlands.  Google Scholar
First citationRohlíček, J. & Hušák, M. (2007). J. Appl. Cryst. 40, 600–601.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSato, T., Sørby, M. H., Ikeda, K., Sato, S., Hauback, B. C. & Orimo, S. (2009). J. Alloys Compd. 487, 472–478.  Web of Science CrossRef CAS Google Scholar
First citationShannon, R. D. (1976). Acta Cryst. A32, 751–767.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationSheldrick, G. M. (2008). Acta Cryst. A64, 112–122.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationShirley, R. (2002). The Crysfire 2002 System for Automatic Powder Indexing: User's Manual. Guildford, UK: The Lattice Press.  Google Scholar
First citationTchitchekova, D. S., Monti, D., Johansson, P., Bardé, F., Randon-Vitanova, A., Palacín, M. R. & Pnrouch, A. (2017). J. Electrochem. Soc. 164, A1384–A1392.  CrossRef CAS Google Scholar
First citationWang, R. Y., Wessells, C. D., Huggins, R. A. & Cui, Y. (2013). Nano Lett. 13, 5748–5752.  Web of Science CrossRef CAS PubMed Google Scholar
First citationWerner, P. E. (1990). TREOR90. University of? Stockholm, Sweden.  Google Scholar
First citationWhittingham, M. S., Siu, C. & Ding, J. (2018). Acc. Chem. Res. 51, 258–264.  CrossRef CAS Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890
Follow Acta Cryst. E
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds