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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

A 1:1 solvate structure of succino­nitrile and lithium thio­cyanate

crossmark logo

aDepartment of Natural Sciences, Northeastern State University, 611 N. Grand Ave., Tahlequah, OK 74464, USA, and bDepartment of Chemistry and Biochemistry, University of Oklahoma, 101 Stephenson Parkway, Norman, OK 73019, USA
*Correspondence e-mail: burba@nsuok.edu

Edited by M. Zeller, Purdue University, USA (Received 5 October 2022; accepted 18 November 2022; online 30 November 2022)

A 1:1 solvate structure of succino­nitrile and lithium thio­cyanate, namely, catena-poly[lithium-di-μ-thio­cyanato-lithium-di-μ-butane­dinitrile], [Li(NCS)(C4H4N2)]n or LiSCN·NC(CH2)2CN, was isolated and its structure was solved. Lithium ions are tetra­hedrally coordinated by two nitrile groups from separate succino­nitrile mol­ecules, as well as S and N atoms of separate SCN anions. The succino­nitrile mol­ecules and Li+ ions form double-chain one-dimensional coordination polymers that are bridged by Li2(SCN)2 dimers. The coordination network extends along [[\overline{1}]01]. Weak hydrogen-bonding inter­actions are also noted among the constituent mol­ecules.

1. Chemical context

Most commercial lithium-ion batteries employ liquid-phase electrolyte solutions that are composed of organic solvents and lithium salts. It is now well documented, however, that these materials can present consumer safety risks (Chen et al., 2021[Chen, Y., Kang, Y., Zhao, Y., Wang, L., Liu, J., Li, Y., Liang, Z., He, X., Li, X., Tavajohi, N. & Li, B. (2021). J. Energy Chem., 59, 83-99.]). For example, inter­nal electrical shortages may lead to thermal runaway and solvent combustion. The battery community has responded to this problem by pursuing safer alternatives to `traditional' electrolyte solutions, such as all-solid-state polymer electrolytes (Armand, 1994[Armand, M. (1994). Solid State Ionics, 69, 309-319.]; Zhou et al., 2019[Zhou, D., Shanmukaraj, D., Tkacheva, A., Armand, M. & Wang, G. (2019). Chem, 5, 2326-2352.]). Despite the advantages provided by solid-state electrolytes, there are significant technological issues preventing their widespread commercialization (Zhou et al., 2019[Zhou, D., Shanmukaraj, D., Tkacheva, A., Armand, M. & Wang, G. (2019). Chem, 5, 2326-2352.]). For example, lithium-ion batteries require highly conductive electrolyte systems (>10−3 S cm−1) to support rapid charging and discharging rates. This requirement is probably the most formidable challenge currently facing polymer electrolytes since most candidate materials simply do not have high enough ionic conductivities to be commercially competitive.

Alternative solid-state ion conductors, such as those created from plastic crystalline materials (Zhou et al., 2019[Zhou, D., Shanmukaraj, D., Tkacheva, A., Armand, M. & Wang, G. (2019). Chem, 5, 2326-2352.]; Zhu et al., 2019[Zhu, H., MacFarlane, D. R., Pringle, J. M. & Forsyth, M. (2019). Trends Chem. 1, 126-140.]), are able to deliver high ionic conductivities without sacrificing mechanical integrity. Plastic crystalline materials possess long-range translational order with some degree of orientational or conformational disorder. Alarco and coworkers (2004[Alarco, P.-J., Abu-Lebdeh, Y., Abouimrane, A. & Armand, M. (2004). Nat. Mater. 3, 476-481.]) examined a family of plastic crystalline electrolytes based on succino­nitrile, a highly polar compound (dielectric constant ɛ = 66; Williams & Smyth, 1962[Williams, D. E. & Smyth, C. P. (1962). J. Am. Chem. Soc. 84, 1808-1812.]) capable of solvating lithium ions. These materials deliver good electrochemical performance and are promising candidates for lithium battery applications (Alarco et al., 2004[Alarco, P.-J., Abu-Lebdeh, Y., Abouimrane, A. & Armand, M. (2004). Nat. Mater. 3, 476-481.]).

Our contribution to this field is the isolation and analysis of a solvate structure formed between lithium thio­cyanate and succino­nitrile that we collected from a mixture of the two components. Additional details about the crystalline material and subsequent structure determination may be found in the Synthesis and Crystallization section. The crystallographic information reveals cation–solvent and cation–anion inter­action motifs that may guide the design of next generation plastic crystalline electrolytes for lithium-ion battery applications.

[Scheme 1]

2. Structural commentary

The succino­nitrile lithium thio­cyanate solvate, LiSCN·NC(CH2)2CN, belongs to the P21/n space group with Z = 4. A displacement ellipsoid plot constituting the asymmetric unit with the atom-labeling scheme is shown in Fig. 1[link]. Projections of the unit cell along the a, b, and c crystallographic axes are provided in Fig. 2[link], and the Li+ coordination environment and cation–succino­nitrile inter­actions are depicted in Fig. 3[link]. Lithium ions are tetra­hedrally coordinated in LiSCN·NC(CH2)2CN. Two of the ligating N atoms originate from two different succino­nitrile mol­ecules, yielding one-dimensional coordination polymer chains composed of cations and succino­nitrile. Each Li+ ion is also coordinated by S and N atoms from two different thio­cyanate anions to produce Li2(SCN)2 dimers that link adjacent lithium–succino­nitrile polymer chains. The resulting double-chain network is oriented along [[\overline{1}]01] (see Fig. 4[link]).

[Figure 1]
Figure 1
The asymmetric unit of LiSCN·NC(CH2)2CN solvate structure. Displacement ellipsoids are shown at the 50% probability level.
[Figure 2]
Figure 2
Projections of the unit cell for LiSCN·NC(CH2)2CN along the a (left), b (center), and c (right) axes. Hydrogen atoms are omitted for clarity.
[Figure 3]
Figure 3
Coordination environments for Li+, SCN, and succino­nitrile with atom labeling. Symmetry codes: (i) 1 − x, 1 − y, 2 − z; (ii) x − [{1\over 2}], [{3\over 2}] − y, z + [{1\over 2}]; (iii) x + [{1\over 2}], [{3\over 2}] − y, z − [{1\over 2}]; (vii) [{3\over 2}] − x, y − [{1\over 2}], [{3\over 2}] − z; (viii) [{3\over 2}] − x, [{1\over 2}] + y, [{3\over 2}] − z.
[Figure 4]
Figure 4
Double-chain structure of LiSCN·NC(CH2)2CN when viewed along the b axis. Hydrogen atoms are omitted for clarity.

The cation–anion Li1—N1 bond [1.985 (3) Å] is shorter than either Li–N bond formed between Li+ and the nitrile groups of succino­nitrile [Li1—N1A = 2.028 (3) Å and Li1—N2Aii = 2.093 (2) Å; symmetry code: (ii) x − [{1\over 2}], [{3\over 2}] − y, z + [{1\over 2}]]. By way of comparison, Li1—S1i is relatively longer at 2.572 (3) Å [symmetry code: (i) 1 − x, 1 − y, 2 − z]. The tetra­hedral coordination environment about Li+ is somewhat distorted with bond angles centered on the cation ranging from 125.75 (13) to 98.67 (9)°. The Li1—N1A—C1A bond angle approaches a linear geometry [170.60 (13)°], whereas Li1iii—N2A—C4A is noticeably bent [155.39 (13)°; symmetry code: (iii) x + [{1\over 2}], [{3\over 2}] − y, z − [{1\over 2}]]. By way of comparison, the Li1—N1—C1 and Li1—S1i—C1i bond angles are 159.26 (12) and 96.67 (6)°, respectively. The succino­nitrile mol­ecules adopt a gauche conformation with a 66.04 (13)° C1A—C2A—C3A—C4A torsion angle and comparable C≡N bond lengths. Both are similar to the low-temperature α phase of pure succino­nitrile (Hore et al., 2009[Hore, S., Dinnebier, R., Wen, W., Hanson, J. & Maier, J. (2009). Z. Anorg. Allg. Chem. 635, 88-93.]; Whitfield, et al., 2008[Whitfield, P. S., Le Page, Y., Abouimrane, A. & Davidson, I. J. (2008). Powder Diffr. 23, 292-299.]).

3. Supra­molecular features

Possible hydrogen bonds between C—H groups of succino­nitrile and N or S atoms are detected in LiSCN·NC(CH2)2CN (Table 1[link]). These inter­actions have relatively long distances and are of low directionality. Hence, we classify the hydrogen bonds as weak.

Table 1
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
C2A—H2AA⋯S1iv 0.99 2.93 3.5212 (15) 120
C2A—H2AB⋯N1v 0.99 2.63 3.4159 (17) 137
C3A—H3AA⋯N2Avi 0.99 2.61 3.4232 (17) 140
C3A—H3AB⋯N1vi 0.99 2.54 3.4257 (17) 150
Symmetry codes: (iv) [x-1, y, z-1]; (v) [x, y, z-1]; (vi) [x-{\script{1\over 2}}, -y+{\script{3\over 2}}, z-{\script{1\over 2}}].

4. Database survey

The structure of the SCN anion in our solvate structure may be compared against the LiSCN·xH2O (x = 0, 1, and 2) family of compounds. Joos et al. (2022[Joos, M., Conrad, M., Bette, S., Merkle, R., Dinnebier, R. E., Schleid, T. & Maier, J. (2022). J. Phys. Chem. Solids, 160, 110299.]) report crystallographic data for two phases of LiSCN·H2O: a room-temperature phase with space group C2/m and a high-temperature phase with space group Pnam. The structures for LiSCN and LiSCN·2H2O, both belonging to space group Pnma, are provided by Reckeweg et al. (2014[Reckeweg, O., Schulz, A., Blaschkowski, B., Schleid, T. & DiSalvo, F. J. (2014). Z. Naturforsch. 69, 17-24.]). Structural parameters for the anions in these compounds are similar to each other and to LiSCN·NC(CH2)2CN. The C1=S1 and C1=N1 bond lengths in LiSCN·NC(CH2)2CN [1.646 (1) and 1.167 (2) Å, respectively] are comparable to those found in LiSCN (1.643 and 1.162 Å, respectively). Furthermore, the C1=N1 bond length is only slightly longer than those reported for the LiSCN hydrates (e.g., the average length is 1.167 Å for LiSCN·xH2O). Finally, the 178.0 (1)° S1—C1—N1 bond angle in LiSCN·NC(CH2)2CN only deviates slightly from a linear geometry and is closest in value to that found in LiSCN·2H2O (177.55°).

5. Synthesis and crystallization

Succino­nitrile (NC(CH2)2CN, CAS# 110–61-2) and lithium thio­cyanate hydrate (LiSCN·xH2O, CAS# 123333–85-7) were both obtained from Sigma Aldrich. The LiSCN·xH2O was dehydrated in a vacuum oven. The two reagents were then stored inside a <1 p.p.m. H2O, argon-filled dry box. A total of 0.3005 g of LiSCN was mixed with 0.5550 g of succino­nitrile, and gentle heating on a hot plate promoted dissolution of the salt into the solvent. The resulting solution was stored under argon gas until crystals of sufficient size formed. A colorless, block-shaped crystal of dimensions 0.272 mm × 0.274 mm × 0.368 mm was selected for structural analysis.

6. Refinement

Crystal data, data collection, and structure refinement details are summarized in Table 2[link]. The positions of hydrogen atoms bonded to carbon atoms were initially determined by geometry and were refined using a riding model. Non-hydrogen atoms were refined with anisotropic displacement parameters. Hydrogen atom displacement parameters were set to 1.2 times the isotropic equivalent displacement parameters of the bonded atoms.

Table 2
Experimental details

Crystal data
Chemical formula [Li(NCS)(C4H4N2)]
Mr 145.11
Crystal system, space group Monoclinic, P21/n
Temperature (K) 100
a, b, c (Å) 7.5802 (13), 11.868 (3), 8.2724 (16)
β (°) 104.155 (7)
V3) 721.6 (3)
Z 4
Radiation type Mo Kα
μ (mm−1) 0.36
Crystal size (mm) 0.37 × 0.27 × 0.27
 
Data collection
Diffractometer Area detector κ–geometry diffractometer
Absorption correction Multi-scan (SADABS; Krause et al., 2015[Krause, L., Herbst-Irmer, R., Sheldrick, G. M. & Stalke, D. (2015). J. Appl. Cryst. 48, 3-10.])
Tmin, Tmax 0.403, 0.490
No. of measured, independent and observed [I > 2σ(I)] reflections 33422, 2400, 2210
Rint 0.097
(sin θ/λ)max−1) 0.735
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.049, 0.140, 1.04
No. of reflections 2400
No. of parameters 92
H-atom treatment H-atom parameters constrained
Δρmax, Δρmin (e Å−3) 0.41, −0.36
Computer programs: APEX3 (Bruker, 2018[Bruker (2018). APEX3. Bruker Nano Inc., Madison, Wisconsin, USA.]), SAINT (Bruker, 2016[Bruker (2016). SAINT. Bruker Nano Inc., Madison, Wisconsin, USA.]), SHELXT2014/5 (Sheldrick, 2015a[Sheldrick, G. M. (2015a). Acta Cryst. A71, 3-8.]), SHELXL2018/3 (Sheldrick, 2015b[Sheldrick, G. M. (2015b). Acta Cryst. C71, 3-8.]), and Mercury (Macrae et al., 2020[Macrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226-235.]).

Supporting information


Computing details top

Data collection: APEX3 (Bruker, 2018); cell refinement: SAINT (Bruker, 2016); data reduction: SAINT (Bruker, 2016); program(s) used to solve structure: SHELXT2014/5 (Sheldrick, 2015a); program(s) used to refine structure: SHELXL2018/3 (Sheldrick, 2015b); molecular graphics: Mercury (Macrae et al., 2020); software used to prepare material for publication: SHELXL2018/3 (Sheldrick, 2015b).

catena-Poly[lithium-di-µ-thiocyanato-lithium-di-µ-butanedinitrile] top
Crystal data top
[Li(NCS)(C4H4N2)]F(000) = 296
Mr = 145.11Dx = 1.336 Mg m3
Monoclinic, P21/nMo Kα radiation, λ = 0.71073 Å
a = 7.5802 (13) ÅCell parameters from 8809 reflections
b = 11.868 (3) Åθ = 2.8–33.7°
c = 8.2724 (16) ŵ = 0.36 mm1
β = 104.155 (7)°T = 100 K
V = 721.6 (3) Å3Block, colourless
Z = 40.37 × 0.27 × 0.27 mm
Data collection top
Area detector κ–geometry
diffractometer
2400 independent reflections
Radiation source: microfocus sealed tube, Incoatec IµS 3.02210 reflections with I > 2σ(I)
Multilayer mirror monochromatorRint = 0.097
ω and φ scansθmax = 31.5°, θmin = 3.1°
Absorption correction: multi-scan
(SADABS; Krause et al., 2015)
h = 1111
Tmin = 0.403, Tmax = 0.490k = 1717
33422 measured reflectionsl = 1212
Refinement top
Refinement on F2Secondary atom site location: difference Fourier map
Least-squares matrix: fullHydrogen site location: inferred from neighbouring sites
R[F2 > 2σ(F2)] = 0.049H-atom parameters constrained
wR(F2) = 0.140 w = 1/[σ2(Fo2) + (0.0936P)2 + 0.2195P]
where P = (Fo2 + 2Fc2)/3
S = 1.04(Δ/σ)max = 0.001
2400 reflectionsΔρmax = 0.41 e Å3
92 parametersΔρmin = 0.36 e Å3
0 restraintsExtinction correction: SHELXL2018/3 (Sheldrick 205b8), Fc*=kFc[1+0.001xFc2λ3/sin(2θ)]-1/4
Primary atom site location: shelxtExtinction coefficient: 0.044 (16)
Special details top

Geometry. All esds (except the esd in the dihedral angle between two l.s. planes) are estimated using the full covariance matrix. The cell esds are taken into account individually in the estimation of esds in distances, angles and torsion angles; correlations between esds in cell parameters are only used when they are defined by crystal symmetry. An approximate (isotropic) treatment of cell esds is used for estimating esds involving l.s. planes.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
S10.76963 (4)0.62537 (2)1.19303 (4)0.02384 (15)
N10.49689 (16)0.62469 (8)0.89339 (15)0.0233 (2)
C10.61234 (16)0.62517 (8)1.01617 (15)0.0184 (2)
Li10.2669 (3)0.5822 (2)0.7318 (3)0.0262 (4)
N1A0.27232 (17)0.59303 (10)0.48836 (14)0.0290 (3)
N2A0.50670 (16)0.86047 (9)0.23500 (16)0.0267 (2)
C1A0.25268 (17)0.60611 (10)0.34831 (16)0.0220 (2)
C2A0.22707 (18)0.62314 (9)0.16912 (15)0.0208 (2)
H2AA0.1167610.5820510.1091660.025*
H2AB0.3323800.5909280.1341430.025*
C3A0.20743 (15)0.74810 (10)0.11935 (14)0.0210 (2)
H3AA0.1746750.7539550.0037710.025*
H3AB0.1070910.7815630.1606450.025*
C4A0.37413 (15)0.81229 (9)0.18592 (14)0.0207 (2)
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
S10.0196 (2)0.0238 (2)0.0247 (2)0.00019 (8)0.00112 (13)0.00022 (9)
N10.0244 (5)0.0225 (5)0.0225 (5)0.0005 (3)0.0045 (4)0.0044 (3)
C10.0176 (5)0.0168 (5)0.0218 (5)0.0005 (3)0.0067 (4)0.0020 (3)
Li10.0232 (10)0.0341 (11)0.0204 (9)0.0029 (8)0.0034 (7)0.0021 (8)
N1A0.0357 (6)0.0275 (5)0.0231 (5)0.0024 (4)0.0061 (4)0.0007 (4)
N2A0.0249 (5)0.0210 (4)0.0318 (6)0.0011 (4)0.0021 (4)0.0008 (4)
C1A0.0227 (5)0.0210 (5)0.0220 (5)0.0022 (4)0.0049 (4)0.0018 (4)
C2A0.0231 (5)0.0229 (5)0.0174 (5)0.0043 (4)0.0068 (4)0.0040 (3)
C3A0.0180 (4)0.0260 (5)0.0180 (5)0.0006 (4)0.0022 (4)0.0000 (4)
C4A0.0219 (5)0.0191 (5)0.0204 (5)0.0022 (4)0.0043 (4)0.0010 (3)
Geometric parameters (Å, º) top
S1—C11.6459 (13)C1A—C2A1.4611 (17)
S1—Li1i2.572 (3)C2A—C3A1.5366 (17)
N1—C11.1674 (18)C2A—H2AA0.9900
N1—Li11.985 (3)C2A—H2AB0.9900
Li1—N1A2.028 (3)C3A—C4A1.4628 (16)
Li1—N2Aii2.093 (2)C3A—H3AA0.9900
N1A—C1A1.1420 (17)C3A—H3AB0.9900
N2A—C4A1.1411 (16)
C1—S1—Li1i96.67 (6)C1A—C2A—H2AA109.1
C1—N1—Li1159.26 (12)C3A—C2A—H2AA109.1
N1—C1—S1177.99 (11)C1A—C2A—H2AB109.1
N1—Li1—N1A115.14 (12)C3A—C2A—H2AB109.1
N1—Li1—N2Aii125.75 (13)H2AA—C2A—H2AB107.8
N1A—Li1—N2Aii103.89 (11)C4A—C3A—C2A112.56 (10)
N1—Li1—S1i102.07 (9)C4A—C3A—H3AA109.1
N1A—Li1—S1i109.28 (11)C2A—C3A—H3AA109.1
N2Aii—Li1—S1i98.67 (9)C4A—C3A—H3AB109.1
C1A—N1A—Li1170.60 (13)C2A—C3A—H3AB109.1
C4A—N2A—Li1iii155.39 (13)H3AA—C3A—H3AB107.8
N1A—C1A—C2A179.80 (16)N2A—C4A—C3A177.96 (13)
C1A—C2A—C3A112.68 (10)
C1A—C2A—C3A—C4A66.04 (13)
Symmetry codes: (i) x+1, y+1, z+2; (ii) x1/2, y+3/2, z+1/2; (iii) x+1/2, y+3/2, z1/2.
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
C2A—H2AA···S1iv0.992.933.5212 (15)120
C2A—H2AB···N1v0.992.633.4159 (17)137
C3A—H3AA···N2Avi0.992.613.4232 (17)140
C3A—H3AB···N1vi0.992.543.4257 (17)150
Symmetry codes: (iv) x1, y, z1; (v) x, y, z1; (vi) x1/2, y+3/2, z1/2.
 

Acknowledgements

The authors thank Northeastern State University for access to facilities needed to prepare the crystals. The authors declare no competing financial inter­ests.

Funding information

Funding for this research was provided by: National Science Foundation (grant No. CHE-1726630).

References

First citationAlarco, P.-J., Abu-Lebdeh, Y., Abouimrane, A. & Armand, M. (2004). Nat. Mater. 3, 476–481.  CrossRef PubMed CAS Google Scholar
First citationArmand, M. (1994). Solid State Ionics, 69, 309–319.  CrossRef CAS Google Scholar
First citationBruker (2016). SAINT. Bruker Nano Inc., Madison, Wisconsin, USA.  Google Scholar
First citationBruker (2018). APEX3. Bruker Nano Inc., Madison, Wisconsin, USA.  Google Scholar
First citationChen, Y., Kang, Y., Zhao, Y., Wang, L., Liu, J., Li, Y., Liang, Z., He, X., Li, X., Tavajohi, N. & Li, B. (2021). J. Energy Chem., 59, 83-99.  CrossRef CAS Google Scholar
First citationHore, S., Dinnebier, R., Wen, W., Hanson, J. & Maier, J. (2009). Z. Anorg. Allg. Chem. 635, 88–93.  CrossRef CAS Google Scholar
First citationJoos, M., Conrad, M., Bette, S., Merkle, R., Dinnebier, R. E., Schleid, T. & Maier, J. (2022). J. Phys. Chem. Solids, 160, 110299.  CrossRef Google Scholar
First citationKrause, L., Herbst-Irmer, R., Sheldrick, G. M. & Stalke, D. (2015). J. Appl. Cryst. 48, 3–10.  Web of Science CSD CrossRef ICSD CAS IUCr Journals Google Scholar
First citationMacrae, C. F., Sovago, I., Cottrell, S. J., Galek, P. T. A., McCabe, P., Pidcock, E., Platings, M., Shields, G. P., Stevens, J. S., Towler, M. & Wood, P. A. (2020). J. Appl. Cryst. 53, 226–235.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationReckeweg, O., Schulz, A., Blaschkowski, B., Schleid, T. & DiSalvo, F. J. (2014). Z. Naturforsch. 69, 17–24.  CrossRef CAS Google Scholar
First citationSheldrick, G. M. (2015a). Acta Cryst. A71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationSheldrick, G. M. (2015b). Acta Cryst. C71, 3–8.  Web of Science CrossRef IUCr Journals Google Scholar
First citationWhitfield, P. S., Le Page, Y., Abouimrane, A. & Davidson, I. J. (2008). Powder Diffr. 23, 292–299.  CrossRef CAS Google Scholar
First citationWilliams, D. E. & Smyth, C. P. (1962). J. Am. Chem. Soc. 84, 1808–1812.  CrossRef CAS Google Scholar
First citationZhou, D., Shanmukaraj, D., Tkacheva, A., Armand, M. & Wang, G. (2019). Chem, 5, 2326–2352.  CrossRef CAS Google Scholar
First citationZhu, H., MacFarlane, D. R., Pringle, J. M. & Forsyth, M. (2019). Trends Chem. 1, 126–140.  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