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

Journal logoSTRUCTURAL
CHEMISTRY
ISSN: 2053-2296

Sodium sulfite hepta­hydrate and its relation to sodium carbonate hepta­hydrate

CROSSMARK_Color_square_no_text.svg

aInstitute for Chemical Technologies and Analytics, Division of Structural Chemistry, TU Wien, Getreidemarkt 9/164-SC, A-1060 Vienna, Austria
*Correspondence e-mail: matthias.weil@tuwien.ac.at

Edited by M. Gardiner, Australian National University, Australia (Received 4 February 2020; accepted 31 March 2020; online 20 April 2020)

The monoclinic crystal structure of Na2SO3(H2O)7 is characterized by an alternating stacking of (100) cationic sodium–water layers and anionic sulfite layers along [100]. The cationic layers are made up from two types of [Na(H2O)6] octa­hedra that form linear 1[Na(H2O)4/2(H2O)2/1] chains linked by dimeric [Na(H2O)2/2(H2O)4/1]2 units on both sides of the chains. The isolated trigonal–pyramidal sulfite anions are connected to the cationic layers through an intricate network of O—H⋯O hydrogen bonds, together with a remarkable O—H⋯S hydrogen bond, with an O⋯S donor–acceptor distance of 3.2582 (6) Å, which is about 0.05 Å shorter than the average for O—H⋯S hydrogen bonds in thio­salt hydrates and organic sulfur com­pounds of the type Y—S—Z (Y/Z = C, N, O or S). Structural relationships between monoclinic Na2SO3(H2O)7 and ortho­rhom­bic Na2CO3(H2O)7 are discussed in detail.

1. Introduction

Sodium sulfite is used extensively in industrial processes, for example, as an anti­oxidant and preservative in food industries (E number for food additives E221), as a corrosion inhibitor in aqueous media, as a bleaching agent, as a solubilizing agent for cellulose, straw and wood in the pulp and paper industry, or as an additive in dying processes. In the USA alone, the production of sodium sulfite reached 150 000 tons in 2002 (Weil et al., 2007[Weil, E. D., Sandler, S. R. & Gernon, M. (2007). Sulfur Com­pounds, 13.1 Sodium Sulfite. In Kirk-Othmer Concise Encyclopedia of Chemical Technology, 5th ed., Vol. 23, edited by A. Seidel, pp. 669-672. New York: John Wiley & Sons Inc.]). Solid sodium sulfite is stable in its anhydrous form and as the hepta­hydrate. Despite its use at industrial scales, structural details are known only for anhydrous Na2SO3 that crystallizes with two formula units in the trigonal system in the space group P[\overline{3}] (Larsson & Kierkegaard, 1969[Larsson, L. O. & Kierkegaard, P. (1969). Acta Chem. Scand. 23, 2253-2260.]). Bond lengths and near-neighbour distances of sodium sulfite in aqueous solution have been calculated by ab initio quantum mechanical charge field mol­ecular dynamics (QMCF MD) studies and determined experimentally by large-angle X-ray scattering (LAXS) by Eklund et al. (2012[Eklund, L., Hofer, T. S., Pribil, A. B., Rode, B. M. & Persson, I. (2012). Dalton Trans. 41, 5209-5216.]). For crystalline Na2SO3(H2O)7, lattice parameters and the space group (P21/n) have previously been determined from Weissenberg photographs without providing further structural details, except for a close metrical resemblance with ortho­rhom­bic Na2CO3(H2O)7 (Dunsmore & Speakman, 1963[Dunsmore, H. S. & Speakman, J. C. (1963). Acta Cryst. 16, 573-574.]). To obtain a more detailed picture of the relationship between the hepta­hydrates of Na2SO3 and Na2CO3, we grew single crystals of Na2SO3(H2O)7 and determined its crystal structure. Indeed, the two hepta­hydrates show not only a close metrical relationship (Table 1[link]), but also structural similarities, though they belong to different crystal systems and contain differently shaped divalent anions, viz. trigonal–pyramidal SO32− and trigonal–planar CO32−.

Table 1
Com­parison of lattice parameters (Å, °) for Na2SO3(H2O)7 (this work) and Na2CO3(H2O)7 (Betzel et al., 1982[Betzel, C., Saenger, W. & Loewus, D. (1982). Acta Cryst. B38, 2802-2804.])

  Na2SO3(H2O)7 Na2CO3(H2O)7
a 14.6563 (8) 14.492 (7)
b 19.7180 (9) 19.490 (5)
c 7.2197 (5) 7.017 (3)
α 90 90
β 90 90
γ 94.0997 (17) 90
V3) 2081.1 (2) 1981.95
T (K) 100 RT
Space group C1121/a Pbca

2. Experimental

2.1. Crystallization

Colourless prismatic crystals of Na2SO3(H2O)7 were grown by recrystallization of a commercial anhydrous sample (Merck, p.A. grade) from an aqueous solution at room tem­per­ature by slow evaporation over the course of several days. In order to remove adherent mother liquor, the crystals were placed on filter paper and subsequently immersed in Paratone oil. The crystal under investigation was cleaved from a larger specimen.

2.2. Crystallography and refinement

Crystal data, data collection and structure refinement details are summarized in Table 2[link]. The crystal structure of Na2SO3(H2O)7 was originally solved and refined in the space group P121/n1 (No. 14), with lattice parameters a = 11.8576 (8), b = 7.2197 (5), c = 12.6965 (9) Å and β = 106.7938 (17)° at 100 K (full crystal data in the setting P121/n1 are available in CIF format as supporting information). The values for the lattice parameters at 100 K are in good agreement with the previous study, with values of a = 11.922, b = 7.260, c = 12.765 Å and β = 107.22° obtained at room temperature from Weissenberg film data (note that a and c are inter­changed in the original description; Dunsmore & Speakman, 1963[Dunsmore, H. S. & Speakman, J. C. (1963). Acta Cryst. 16, 573-574.]). For a better com­parison with the reported crystal structure of β-Na2CO3(H2O)7 (Betzel et al., 1982[Betzel, C., Saenger, W. & Loewus, D. (1982). Acta Cryst. B38, 2802-2804.]), the nonconventional setting C1121/a was chosen for the final structural description of Na2SO3(H2O)7, using the matrix (101, 10[\overline{1}], 010) for transformation of the primitive cell to the C-centred cell with c as the unique axis; moreover, the atomic coordinates and the origin of the unit cell were chosen to ensure a similar packing of structural building blocks in the two unit cells of Na2SO3(H2O)7 and β-Na2CO3(H2O)7. All H atoms present in the crystal structure of Na2SO3(H2O)7 were located in a difference Fourier map and were refined freely.

Table 2
Experimental details

Crystal data
Chemical formula Na2SO3(H2O)7
Mr 252.15
Crystal system, space group Monoclinic, C1121/a
Temperature (K) 100
a, b, c (Å) 14.6563 (8), 19.7180 (9), 7.2197 (5)
γ (°) 94.0997 (17)
V3) 2081.1 (2)
Z 8
Radiation type Mo Kα
μ (mm−1) 0.42
Crystal size (mm) 0.15 × 0.13 × 0.12
 
Data collection
Diffractometer Bruker APEXII CCD
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 0.675, 0.747
No. of measured, independent and observed [I > 2σ(I)] reflections 16909, 4845, 4222
Rint 0.021
(sin θ/λ)max−1) 0.827
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.023, 0.063, 1.06
No. of reflections 4845
No. of parameters 174
H-atom treatment All H-atom parameters refined
Δρmax, Δρmin (e Å−3) 0.89, −0.33
Com­puter programs: APEX2 (Bruker, 2014[Bruker (2014). APEX2 and SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.]), SAINT (Bruker, 2014[Bruker (2014). APEX2 and SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.]), SHELXT (Sheldrick, 2015a[Sheldrick, G. M. (2015a). Acta Cryst. A71, 3-8.]), SHELXL2018 (Sheldrick, 2015b[Sheldrick, G. M. (2015b). Acta Cryst. C71, 3-8.]), ATOMS (Dowty, 2006[Dowty, E. (2006). ATOMS. Shape Software, Kingsport, Tennessee, USA.]), 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.]) and publCIF (Westrip, 2010[Westrip, S. P. (2010). J. Appl. Cryst. 43, 920-925.]).

3. Results and discussion

3.1. Crystal structure

In the crystal structure of Na2SO3(H2O)7, all atoms (2 Na, 1 S, 10 O and 14 H) are located on general sites. The two sodium cations are surrounded by six water mol­ecules, defining a distorted octa­hedral coordination polyhedron in each case. The Na—O distances range from 2.3690 (6) to 2.4952 (6) Å (Table 3[link]), with mean values of 2.42 (4) Å for Na1 and 2.43 (6) Å for Na2, in fairly good agreement with the mean value for Na[6]—O of 2.44 (11) Å calculated for 5520 individual bonds (Gagné & Hawthorne, 2016[Gagné, O. C. & Hawthorne, F. C. (2016). Acta Cryst. B72, 602-625.]). The bond valence sums (Brown, 2002[Brown, I. D. (2002). In The Chemical Bond in Inorganic Chemistry: The Bond Valence Model. Oxford University Press.]) for the sodium cations, as calculated with parameters provided by Brese & O'Keeffe (1991[Brese, N. E. & O'Keeffe, M. (1991). Acta Cryst. B47, 192-197.]), are 1.15 valence units (v.u.) for Na1 and 1.13 v.u. for Na2, and thus in the expected range for monovalent Na+. The O—Na—O angles deviate clearly from ideal values, with values for trans O atoms in the range 172.149 (16)–176.42 (2)° for Na1 and 165.81 (2)–174.23 (2)° for Na2, and for cis O atoms in the range 81.464 (19)–101.74 (2)° for Na1 and 81.51 (2)–103.23 (2)° for Na2. The two types of [Na(H2O)6] octa­hedra show a different linkage pattern. Octa­hedra centred by Na1 share common edges (O8/O10 and O8ii/O10i; see Table 3[link] for symmetry codes) to form infinite linear 1[Na1(H2O)4/2(H2O)2/1] chains running parallel to [001], whereas octa­hedra centred by Na2 make up dimeric [Na2(H2O)2/2(H2O)4/1]2 units by sharing an edge (O5 and O5iii). In both cases, the corresponding Na—O bonds to the shared O atoms at the edges are the shortest in the respective octa­hedron. The dimeric units connect adjacent chains by sharing the terminal water mol­ecules (O9 and O7) on both sides of the chains (corner-sharing links). This way, the sodium–water octa­hedra are assembled by edge- and corner-sharing into an infinite layer extending parallel to (100) (Fig. 1[link]a).

Table 3
Selected geometric parameters (Å, °)

Na1—O10 2.3690 (6) Na2—O4 2.3939 (6)
Na1—O8 2.3785 (6) Na2—O7 2.4093 (6)
Na1—O10i 2.4199 (6) Na2—O6 2.4928 (6)
Na1—O7 2.4199 (6) Na2—O9i 2.4952 (6)
Na1—O8ii 2.4436 (6) S1—O3 1.5224 (5)
Na1—O9 2.4599 (6) S1—O1 1.5234 (5)
Na2—O5iii 2.3787 (6) S1—O2 1.5338 (5)
Na2—O5 2.3805 (6)    
       
O3—S1—O1 105.85 (3) O1—S1—O2 105.87 (3)
O3—S1—O2 106.07 (3)    
Symmetry codes: (i) [-x+1, -y+{\script{1\over 2}}, z+{\script{1\over 2}}]; (ii) [-x+1, -y+{\script{1\over 2}}, z-{\script{1\over 2}}]; (iii) [-x+1], [-y, -z+3].
[Figure 1]
Figure 1
(a) View along [[\overline{1}]00] onto the cationic water–sodium (100) layer in the crystal structure of Na2SO3(H2O)7 made up from edge- and corner-sharing [Na(H2O)6] octa­hedra (turquoise). Anisotropic displacement ellipsoids are drawn at the 90% probability level and H atoms are shown as grey spheres of arbitrary radii. Symmetry codes refer to Table 3[link]. (b) The same type of layer in the crystal structure of Na2CO3(H2O)7, with atoms as spheres of arbitrary radii.

The sulfite anion has the characteristic trigonal–pyramidal configuration, with the SIV atom occupying the pyramidal position. Atom S1 is 0.5912 (4) Å above the basal plane formed by atoms O1, O2 and O3. The S—O bond lengths are in a narrow range 1.5224 (5)–1.5338 (5) Å [mean 1.527 (6) Å], just like the O—S—O angles [105.85 (3)–106.07 (3)°; mean 105.93 (16)°]. Again, these values are in good agreement with the grand mean SIV—O bond length of 1.529 (15) Å calculated for 90 bonds and with the O—SIV—O angles in the range ∼99–107° with a mean value of ∼104° (Gagné & Hawthorne, 2018[Gagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79-96.]). The bond valence sum for atom S1 is 4.12 v.u., using the parameters of Brese & O'Keeffe (1991[Brese, N. E. & O'Keeffe, M. (1991). Acta Cryst. B47, 192-197.]) for calculation. The sulfite anions are isolated from the sodium–water layer, lying alternatingly on both sides outside of an individual layer. In this way, cationic sodium–water layers at x ≃ 0, [1 \over 2] are sandwiched by sulfite layers at x[1 \over 4], [3 \over 4] and stacked along [100], with the sulfite anions situated approximately at the height in y where the [Na2O2/2O4/1]2 dimers are linked to the 1[Na1(H2O)4/2(H2O)2/1] chains (Fig. 2[link]a).

[Figure 2]
Figure 2
(a) The crystal structure of Na2SO3(H2O)7 in a projection along [001], showing the layered character with cationic water–sodium layers at x ≃ 0, [1 \over 2] alternating with sulfite layers (red polyhedra) at x[1 \over 4], [3 \over 4]. (b) The crystal structure of Na2CO3(H2O)7 in a projection along [00[\overline{1}]], showing the same type of layer stacking but a different orientation of the dimeric groups and adjacent carbonate anions at y[1 \over 2]. For clarity, hydrogen bonds are displayed only in the lower half of the figures, with moderate O—H⋯O hydrogen bonds shown as green lines, weak O—H⋯O bonds as yellow lines and O—H⋯S hydrogen bonds as orange lines.

The seven independent water mol­ecules possess approximately tetra­hedral coordination arrangements (including hydrogen bonds), except for O9, and five of them each bridge two sodium cations (O5, O7, O8, O9 and O10), whereas two are each bonded to only one sodium cation (O4 and O6). An intricate network of O—H⋯O hydrogen bonds between the water mol­ecules and the sulfite O atoms link the anionic layers to adjacent cationic layers, thus establishing a three-dimensional hydrogen-bonded network structure (Fig. 2[link]a). Based on the donor–acceptor distances between 2.7204 (7) and 2.9110 (8) Å (Table 4[link]), the hydrogen-bonding strength is moderate according to the classification of Jeffrey (1997[Jeffrey, G. A. (1997). In An Introduction to Hydrogen Bonding. New York: Oxford University Press Inc.]). Most of these hydrogen bonds are donated to sulfite atoms O1, O2 and O3 (Fig. 3[link]a). Thereby, atom O1 is the acceptor of three, O2 of four and O3 of three hydrogen bonds. It is worth noting that the S—O bond lengths reflect this situation nicely, with S1—O2 = 1.5338 (5) Å being about 0.01 Å longer than the remaining two. The O9 water mol­ecule, bonded to Na1, Na2 and via H9A to O2, lacks a clearcut hydrogen bond for its second H atom (H9B), which points to H6B of the O6—H6B⋯O3 hydrogen bond [H9B(x, y − [{1\over 2}], z + [{5\over 2}])⋯H6B = 2.49 Å], while distances from O9(x, y − [{1\over 2}], z + [{5\over 2}]) to O6 and O3 exceed 3.3 Å.

Table 4
Hydrogen-bond geometry (Å, °)

D—H⋯A D—H H⋯A DA D—H⋯A
O4—H4A⋯O3iv 0.800 (15) 2.031 (16) 2.8216 (8) 169.7 (15)
O4—H4B⋯O1v 0.821 (16) 1.983 (16) 2.7904 (7) 167.8 (15)
O5—H5A⋯O1 0.804 (13) 1.947 (13) 2.7503 (7) 175.9 (13)
O5—H5B⋯O2iv 0.798 (15) 1.994 (15) 2.7694 (7) 163.9 (15)
O6—H6A⋯O2v 0.777 (14) 2.072 (14) 2.8206 (7) 161.8 (14)
O6—H6B⋯O3 0.774 (15) 1.962 (15) 2.7204 (7) 166.5 (15)
O7—H7A⋯O2v 0.810 (13) 1.976 (14) 2.7761 (7) 169.3 (13)
O7—H7B⋯O6vi 0.773 (14) 2.171 (14) 2.9110 (8) 160.7 (15)
O8—H8A⋯O1iii 0.792 (15) 2.009 (15) 2.7900 (7) 168.9 (13)
O8—H8B⋯S1vii 0.825 (14) 2.455 (14) 3.2582 (6) 164.5 (13)
O9—H9A⋯O2vii 0.807 (14) 2.106 (14) 2.9096 (7) 174.0 (14)
O10—H10A⋯O4ii 0.839 (15) 1.962 (15) 2.7908 (8) 169.5 (14)
O10—H10B⋯O3vi 0.762 (14) 2.069 (14) 2.8210 (7) 169.1 (14)
Symmetry codes: (ii) [-x+1, -y+{\script{1\over 2}}, z-{\script{1\over 2}}]; (iii) [-x+1, -y, -z+3]; (iv) [-x+{\script{3\over 2}}, -y, z+{\script{1\over 2}}]; (v) [-x+{\script{3\over 2}}, -y, z-{\script{1\over 2}}]; (vi) [-x+1, -y, -z+2]; (vii) [x-{\script{1\over 2}}, y+{\script{1\over 2}}, z].
[Figure 3]
Figure 3
Com­parison of the hydrogen bonding to the anion in (a) Na2SO3(H2O)7, with displacement ellipsoids drawn at the 50% probability level, and (b) Na2CO3(H2O)7, with atoms as arbitrary spheres; hydrogen bonds are shown as thin solid lines. Atoms O1 and O3 in the sulfite structure accept three hydrogen bonds each, whereas O8 and O10 in the carbonate structure accept four each. Likewise, O2 in the sulfite structure accepts four hydrogen bonds, whereas the corresponding O9 atom in the carbonate accepts three. Note that the arrangement of the hydrogen-bonded water mol­ecules around SO32− is approximately mirror-symmetric (e.g. O5i and O6i), whereas it is less symmetric for the carbonate. Symmetry codes for atoms in Na2SO3(H2O)7 involved in hydrogen bonding with the SO32– anion are: (i) x, y, z; (ii) x + [{1\over 2}], y − [{1\over 2}], z; (iii) −x + [{3\over 2}], −y, z + [{1\over 2}]; (iv) −x + [{3\over 2}], −y, z − [{1\over 2}]; (v) −x + 1, −y, −z + 2; (vi) −x + 1, −y, −z + 3.

In addition to the inter­actions between water mol­ecules and sulfite O atoms, there are two hydrogen bonds between water mol­ecules only (O7⋯O6vi and O10⋯O4ii; symmetry codes refer to Table 4[link]), and, as a pecularity, an O—H⋯S hydrogen bond between O8 and S1vii. In general, S⋯H inter­actions involving divalent S atoms are considered as weak hydrogen bonds (Allen et al., 1997[Allen, F. H., Bird, C. M., Rowland, R. S. & Raithby, P. R. (1997). Acta Cryst. B53, 696-701.]). The H⋯S hydrogen-bonding strength becomes even weaker for H⋯SO3 contacts because the S atom is positively polarized in an SO32− anion with partial double-bond character for the S—O bonds (Nyberg & Larsson, 1973[Nyberg, B. & Larsson, R. (1973). Acta Chem. Scand. 27, 63-70.]). The hydrogen-bond acceptor ability of divalent sulfur was evaluated some time ago from 1811 substructures of mostly organic com­pounds, i.e. Y—S—Z systems (Y/Z = C, N, O or S) as acceptor groups retrieved from the Cambridge Structural Database, giving a mean inter­molecular >S⋯H distance of 2.67 (5) Å for O—H donors and a mean S⋯O distance of 3.39 (4) Å (Allen et al., 1997[Allen, F. H., Bird, C. M., Rowland, R. S. & Raithby, P. R. (1997). Acta Cryst. B53, 696-701.]; Groom et al., 2016[Groom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171-179.]). In com­parison, the first ever reported crystal structure determination of an inorganic com­pound with an O—H⋯S hydrogen bond and a clear location of the H atoms, viz. BaS2O3(H2O) from neutron single-crystal diffraction data (Manojlović-Muir, 1969[Manojlović-Muir, L. (1969). Nature, 224, 686-687.]), revealed a considerably shorter S⋯H distance of 2.367 (4) Å and a likewise shorter S⋯O distance of 3.298 (4) Å. The O—H⋯S angle in BaS2O3·H2O was determined as 163 (3)°. Corresponding values of the O—H⋯S hydrogen bond in the crystal structure of Na2SO3(H2O)7 are somewhat larger at 2.455 (14) Å for H8B⋯S1vii (X-ray data), slightly shorter at 3.2582 (6) Å for O8⋯S1vii and similar at 164.5 (13)° for the O8—H8B⋯S1vii angle. A com­parable O⋯S distance of 3.326 Å was found as the mean value for 86 hydrogen-bonding inter­actions between water mol­ecules and S atoms in a variety of thio­salt hydrates, such as Schlippes salt, Na3SbS4(H2O)9 (Mikenda et al., 1989[Mikenda, W., Mereiter, K. & Preisinger, A. (1989). Inorg. Chim. Acta, 161, 21-28.]). A literature search indicated that the O—H⋯S hydrogen bond in the title com­pound appears to be unprecedented thus far among hydrated sulfites. This suggests that in sulfite hydrates, O—H⋯O hydrogen bonding is clearly preferred over O—H⋯S hydrogen bonding and that a certain structural motif is needed to induce O—H⋯S hydrogen bonding like in the title com­pound. Invoking the results of an electron deformation density study of MgSO3(H2O)6 (Bats et al., 1986[Bats, J. W., Fuess, H. & Elerman, Y. (1986). Acta Cryst. B42, 552-557.]), the coordination capability of the sulfite S atom via its electron lone-pair lobe at the apex of the SO3 pyramid is not unexpected, but this capability seems to be weak in the context of hydrogen bonding because otherwise more examples with features com­parable to the title com­pound would have been encountered already. As soon as covalent bonding comes into play, the coordination capability of the sulfite S atom is well documented by transition-metal com­plexes like K2[Pd(SO3)2]·H2O (Messer et al., 1979[Messer, D., Breitinger, D. & Haegler, W. (1979). Acta Cryst. B35, 815-818.]) or K2[Hg(SO3)2]·2.25H2O (Weil et al., 2010[Weil, M., Baumann, S. O. & Breitinger, D. K. (2010). Acta Cryst. C66, i89-i91.]), with metal–sulfur bonds, or by hydrogen sulfites like CsHSO3 (Johansson et al., 1980[Johansson, L.-G., Lindqvist, O. & Vannerberg, N.-G. (1980). Acta Cryst. B36, 2523-2526.]) or K5(HSO3)(S2O5) (Magnusson et al., 1983[Magnusson, A., Johansson, L.-G. & Lindqvist, O. (1983). Acta Cryst. C39, 819-822.]) that contain HSO3 anions with hydrogen covalently bonded to sulfur.

The numerical values of the atomic distances for crystalline Na2SO3(H2O)7 (Tables 3[link] and 4[link]) are in good agreement with those of aqueous Na2SO3 solutions determined from LAXS studies, with S—O = 1.53 Å for the sulfite group and Na—O = 2.41 Å for the sodium—water distances (Eklund et al., 2012[Eklund, L., Hofer, T. S., Pribil, A. B., Rode, B. M. & Persson, I. (2012). Dalton Trans. 41, 5209-5216.]). In the latter study, the S⋯Owater distance in solution was determined as 3.68 Å, which is considerably longer than in the solid state, giving further evidence for a weak but existing O—H⋯S hydrogen bond in the crystalline material.

3.2. Com­parison with Na2CO3(H2O)7

The close structural relationship between monoclinic Na2SO3(H2O)7 and ortho­rhom­bic Na2CO3(H2O)7 (Table 1[link]) becomes evident from the similar arrangement of the principal building units in the crystal structures. The same type of cationic sodium–water layers made up from edge- and corner-sharing [Na(H2O)6] octa­hedra [mean Na—O distance = 2.43 (4) Å and O—Na—O angles = 81–102 and 164–180°; Fig. 1[link]b] is present in the carbonate, likewise situated at x ≃ 0, [1 \over 2] in the unit cell (Fig. 2[link]b). The carbonate groups do not show pyramidalization (Zemann, 1981[Zemann, J. (1981). Fortschr. Mineral. 59, 95-116.]) and occupy the same space as the sulfite groups between adjacent layers close to the [Na(H2O)2/2(H2O)4/1]2 dimers.

The main difference between the two structures is related to the orientation of the [Na(H2O)2/2(H2O)4/1]2 dimers in the layers. Whereas in the sulfite structure, the dimers at y ≃ 0 and [1 \over 2] in one layer and also the accom­panying anions close to them have the same orientation relative to (100), in the carbonate structure, the orientation of every second dimer (at y[1 \over 2]) and the accom­panying anions in a layer is reversed due to the presence of the c-glide plane (Fig. 2[link]).

The hydrogen-bonding schemes in the two hepta­hydrates are similar (Fig. 3[link]). In the carbonate structure, the anions are likewise hydrogen bonded to water mol­ecules through medium–strong hydrogen bonds [O⋯O = 2.690 (5)–3.060 (4) Å, with an additional weak inter­action of 3.223 (5) Å]. In analogy, two water–water O—H⋯O inter­actions with donor–acceptor distances of 2.827 (5) and 2.766 (5) Å are also observed. However, in contrast to the central sulfite S atom with its free electron lone pair, the central C atom of the carbonate anion cannot act as a hydrogen-bond acceptor, and thus this inter­action is missing in the carbonate structure.

3.3. Com­parison with related com­pounds

Crystal structures with sulfite groups anchored exclusively by hydrogen bonds are at present restricted to the title com­pound Na2SO3(H2O)7, to NH4SO3(H2O) (Battelle & Trueblood, 1965[Battelle, L. F. & Trueblood, K. N. (1965). Acta Cryst. 19, 531-535.]; Durand et al., 1977[Durand, J., Galigné, J. L. & Cot, L. (1977). Acta Cryst. B33, 1414-1417.]) and to MgSO3(H2O)6 (Andersen & Lindqvist, 1984[Andersen, L. & Lindqvist, O. (1984). Acta Cryst. C40, 584-586.]; Bats et al., 1986[Bats, J. W., Fuess, H. & Elerman, Y. (1986). Acta Cryst. B42, 552-557.]). In MgSO3(H2O)6, which is built up from [Mg(H2O)6] octa­hedra and isolated SO3 pyramids within a lattice of the space group type R3, and with Mg and S atoms both located on threefold rotation axes, there are two independent water mol­ecules that donate, apart from one water–water hydrogen bond, three water–Osulfite hydrogen bonds to each sulfite O atom, com­parable to O1 and O3 in Na2SO3(H2O)7, but with shorter O⋯O distances [2.687 (3), 2.701 (3) and 2.726 (3) Å] than in the latter. An electron deformation density study of MgSO3(H2O)6 (Bats et al., 1986[Bats, J. W., Fuess, H. & Elerman, Y. (1986). Acta Cryst. B42, 552-557.]) proved the presence of an extended lone-pair lobe at the apex of the SO3 pyramid, but neither MgSO3(H2O)6 nor NH4SO3(H2O) contain O—H⋯S or N—H⋯S hydrogen bonds.

A further com­parison with other hydrated sodium com­pounds com­prised of related oxo anions shows no close structural relationship to the title hepta­hydrate. For example, Na2SO4(H2O)7 (Oswald et al., 2008[Oswald, I. D. H., Hamilton, A., Hall, C., Marshall, W. G., Prior, T. J. & Pulham, C. R. (2008). J. Am. Chem. Soc. 130, 17795-17800.]) (I41/amd, Z = 4) has a com­pletely different arrangement of the principal building units. Its crystal structure is com­prised of [Na(H2O)]6 octa­hedra concatenated by edge- and corner-sharing into a three-dimensional network with isolated tetra­hedral sulfate anions hydrogen bonded to the chains. Also, for sodium com­pounds with analogous trigonal–pyramidal oxoanions and the same charge, i.e. XO32−, with X = Se and Te, no phases related structurally or com­positionally to Na2SO3(H2O)7 are known. For Na2SeO3, the anhydrous form (P21/c, Z = 4) is made up from [NaO6] octa­hedra and trigonal–pyramidal SeO32− anions (Wickleder, 2002[Wickleder, M. S. (2002). Acta Cryst. E58, i103-i104.]), and is isotypic with Na2TeO3 (Masse et al., 1980[Masse, R., Guitel, J. C. & Tordjman, I. (1980). Mater. Res. Bull. 15, 431-436.]). Hydrated forms are known only for the penta­hydrates Na2SeO3(H2O)5 (Mereiter, 2013[Mereiter, K. (2013). Acta Cryst. E69, i77-i78.]) and Na2TeO3(H2O)5 (Philippot et al., 1979[Philippot, E., Maurin, M. & Moret, J. (1979). Acta Cryst. B35, 1337-1340.]) that are, surprisingly, not isotypic (Pbcm, with Z = 8, and C2/c, with Z = 8, respectively). These structures are based on two- or three-dimensional assemblies of [NaO5] polyhedra (Se) and [NaO6] octa­hedra (Se and Te), to which SeO3/TeO3 groups are bonded via two (Se) or one (Te) O atom. The [NaO6] octa­hedra in these two salts share common faces and edges but no vertices. As pointed out by Philippot et al. (1979[Philippot, E., Maurin, M. & Moret, J. (1979). Acta Cryst. B35, 1337-1340.]) for Na2TeO3(H2O)5 and confirmed also for Na2SeO3(H2O)5 (Mereiter, 2013[Mereiter, K. (2013). Acta Cryst. E69, i77-i78.]), the electron lone pair of Se and Te in these structures shows no attracting inter­actions with neighbouring H atoms. This might be one reason why hydrates of Na2SeO3 and Na2TeO3 do not crystallize in the Na2SO3(H2O)7 structure and vice versa.

Supporting information


Computing details top

Data collection: APEX2 (Bruker, 2014); cell refinement: SAINT (Bruker, 2014); data reduction: SAINT (Bruker, 2014); program(s) used to solve structure: SHELXT (Sheldrick, 2015a); program(s) used to refine structure: SHELXL2018 (Sheldrick, 2015b); molecular graphics: ATOMS (Dowty, 2006) and Mercury (Macrae et al., 2020); software used to prepare material for publication: publCIF (Westrip, 2010).

Sodium sulfite heptahydrate top
Crystal data top
H14Na2O10SF(000) = 1056
Mr = 252.15Dx = 1.610 Mg m3
Monoclinic, C1121/aMo Kα radiation, λ = 0.71073 Å
Hall symbol: -C_2acCell parameters from 8292 reflections
a = 14.6563 (8) Åθ = 2.8–36.0°
b = 19.7180 (9) ŵ = 0.42 mm1
c = 7.2197 (5) ÅT = 100 K
β = 90°Fragment, colourless
V = 2081.1 (2) Å30.15 × 0.13 × 0.12 mm
Z = 8
Data collection top
Bruker APEXII CCD
diffractometer
4222 reflections with I > 2σ(I)
Radiation source: fine-focus sealed tubeRint = 0.021
ω– and φ–scanθmax = 36.0°, θmin = 2.8°
Absorption correction: multi-scan
(SADABS; Krause et al., 2015)
h = 2423
Tmin = 0.675, Tmax = 0.747k = 3230
16909 measured reflectionsl = 1111
4845 independent reflections
Refinement top
Refinement on F20 restraints
Least-squares matrix: fullHydrogen site location: difference Fourier map
R[F2 > 2σ(F2)] = 0.023All H-atom parameters refined
wR(F2) = 0.063 w = 1/[σ2(Fo2) + (0.0304P)2 + 0.806P]
where P = (Fo2 + 2Fc2)/3
S = 1.06(Δ/σ)max = 0.003
4845 reflectionsΔρmax = 0.89 e Å3
174 parametersΔρmin = 0.33 e Å3
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
Na10.49508 (2)0.24227 (2)1.10868 (4)0.00970 (6)
Na20.55111 (2)0.06011 (2)1.35665 (4)0.01015 (6)
S10.76676 (2)0.13751 (2)1.37190 (2)0.00782 (4)
O10.71232 (3)0.11448 (2)1.53732 (7)0.01115 (8)
O20.85230 (3)0.08810 (3)1.36082 (6)0.01132 (8)
O30.70978 (4)0.12350 (3)1.20165 (7)0.01248 (9)
O40.69179 (4)0.12961 (3)1.36691 (7)0.01302 (9)
O50.60016 (4)0.01074 (3)1.59875 (7)0.01119 (8)
O60.60771 (4)0.01777 (3)1.11731 (7)0.01330 (9)
O70.50388 (4)0.12044 (3)1.08692 (7)0.01129 (8)
O80.39237 (4)0.23269 (3)1.36393 (7)0.01119 (8)
O90.49463 (4)0.36695 (3)1.11835 (7)0.01244 (9)
O100.39308 (4)0.24819 (3)0.85559 (7)0.01182 (9)
H4A0.7253 (10)0.1272 (7)1.454 (2)0.040 (4)*
H4B0.7270 (10)0.1258 (7)1.280 (2)0.043 (4)*
H5A0.6351 (9)0.0397 (7)1.5804 (18)0.026 (3)*
H5B0.6230 (10)0.0138 (7)1.6763 (19)0.036 (4)*
H6A0.6299 (9)0.0083 (7)1.047 (2)0.035 (4)*
H6B0.6429 (10)0.0432 (8)1.1470 (18)0.033 (4)*
H7A0.5492 (9)0.1157 (6)1.0255 (19)0.030 (3)*
H7B0.4660 (10)0.0941 (7)1.053 (2)0.034 (4)*
H8A0.3667 (10)0.1966 (7)1.3816 (17)0.032 (4)*
H8B0.3519 (10)0.2597 (7)1.3636 (16)0.026 (3)*
H9A0.4574 (10)0.3785 (7)1.192 (2)0.030 (3)*
H9B0.5427 (11)0.3788 (7)1.1579 (19)0.034 (4)*
H10A0.3612 (10)0.2819 (8)0.8574 (17)0.032 (4)*
H10B0.3596 (10)0.2171 (7)0.8447 (17)0.026 (3)*
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Na10.01001 (12)0.01018 (12)0.00894 (11)0.00089 (9)0.00002 (8)0.00018 (8)
Na20.01087 (13)0.01032 (12)0.00923 (11)0.00053 (9)0.00029 (8)0.00100 (8)
S10.00777 (6)0.00762 (6)0.00812 (6)0.00102 (4)0.00010 (4)0.00042 (4)
O10.0113 (2)0.0120 (2)0.01032 (18)0.00238 (16)0.00291 (14)0.00012 (14)
O20.0090 (2)0.0124 (2)0.01216 (19)0.00153 (16)0.00037 (14)0.00118 (15)
O30.0144 (2)0.0131 (2)0.01010 (19)0.00150 (17)0.00474 (15)0.00016 (15)
O40.0119 (2)0.0158 (2)0.0114 (2)0.00138 (17)0.00043 (16)0.00018 (16)
O50.0113 (2)0.0101 (2)0.01232 (19)0.00204 (16)0.00185 (15)0.00181 (15)
O60.0132 (2)0.0122 (2)0.0147 (2)0.00227 (17)0.00056 (16)0.00217 (16)
O70.0112 (2)0.0112 (2)0.01143 (19)0.00058 (16)0.00061 (15)0.00022 (15)
O80.0099 (2)0.0104 (2)0.01337 (19)0.00126 (16)0.00133 (15)0.00157 (15)
O90.0139 (2)0.0117 (2)0.01171 (19)0.00068 (17)0.00033 (16)0.00053 (15)
O100.0101 (2)0.0117 (2)0.0136 (2)0.00013 (17)0.00087 (15)0.00148 (15)
Geometric parameters (Å, º) top
Na1—O102.3690 (6)Na2—O42.3939 (6)
Na1—O82.3785 (6)Na2—O72.4093 (6)
Na1—O10i2.4199 (6)Na2—O62.4928 (6)
Na1—O72.4199 (6)Na2—O9i2.4952 (6)
Na1—O8ii2.4436 (6)S1—O31.5224 (5)
Na1—O92.4599 (6)S1—O11.5234 (5)
Na2—O5iii2.3787 (6)S1—O21.5338 (5)
Na2—O52.3805 (6)
O5iii—Na2—O588.41 (2)H4A—O4—H4B101.7 (15)
O5iii—Na2—O4165.81 (2)Na2iii—O5—Na291.59 (2)
O5—Na2—O491.68 (2)Na2iii—O5—H5A110.7 (9)
O5iii—Na2—O791.10 (2)Na2—O5—H5A122.0 (9)
O5—Na2—O7173.08 (2)Na2iii—O5—H5B119.4 (10)
O4—Na2—O790.48 (2)Na2—O5—H5B106.9 (10)
O5iii—Na2—O6100.56 (2)H5A—O5—H5B106.6 (14)
O5—Na2—O691.13 (2)Na2—O6—H6A100.9 (10)
O4—Na2—O693.62 (2)Na2—O6—H6B118.2 (10)
O7—Na2—O682.177 (19)H6A—O6—H6B110.0 (14)
O5iii—Na2—O9i81.51 (2)Na2—O7—Na2118.38 (2)
O5—Na2—O9i83.523 (19)Na2—O7—H7A96.9 (9)
O4—Na2—O9i84.40 (2)Na1—O7—H7A104.6 (9)
O7—Na2—O9i103.23 (2)Na2—O7—H7B98.3 (10)
O6—Na2—O9i174.23 (2)Na1—O7—H7B127.0 (10)
O10—Na1—O8101.74 (2)H7A—O7—H7B107.6 (13)
O10—Na1—O10i172.170 (15)Na1—O8—Na1i97.465 (19)
O8—Na1—O10i81.77 (2)Na1—O8—H8A117.0 (9)
O10—Na1—O794.44 (2)Na1i—O8—H8A109.6 (9)
O8—Na1—O792.86 (2)Na1—O8—H8B115.3 (8)
O10i—Na1—O792.36 (2)Na1i—O8—H8B112.2 (9)
O10—Na1—O8ii81.464 (19)H8A—O8—H8B105.3 (14)
O8—Na1—O8ii172.149 (16)Na1—O9—Na2ii125.08 (2)
O10i—Na1—O8ii94.20 (2)Na1—O9—H9A110.5 (10)
O7—Na1—O8ii94.04 (2)Na2ii—O9—H9A97.1 (10)
O10—Na1—O985.75 (2)Na1—O9—H9B104.1 (11)
O8—Na1—O990.60 (2)Na2ii—O9—H9B112.2 (11)
O10i—Na1—O987.22 (2)H9A—O9—H9B106.7 (14)
O7—Na1—O9176.42 (2)Na1—O10—Na1ii98.38 (2)
O8ii—Na1—O982.451 (19)Na1—O10—H10A114.6 (9)
O3—S1—O1105.85 (3)Na1ii—O10—H10A111.2 (9)
O3—S1—O2106.07 (3)Na1—O10—H10B114.4 (10)
O1—S1—O2105.87 (3)Na1ii—O10—H10B112.4 (9)
Na2—O4—H4A119.9 (11)H10A—O10—H10B106.0 (14)
Na2—O4—H4B116.4 (11)
Symmetry codes: (i) x+1, y+1/2, z+1/2; (ii) x+1, y+1/2, z1/2; (iii) x+1, y, z+3.
Hydrogen-bond geometry (Å, º) top
D—H···AD—HH···AD···AD—H···A
O4—H4A···O3iv0.800 (15)2.031 (16)2.8216 (8)169.7 (15)
O4—H4B···O1v0.821 (16)1.983 (16)2.7904 (7)167.8 (15)
O5—H5A···O10.804 (13)1.947 (13)2.7503 (7)175.9 (13)
O5—H5B···O2iv0.798 (15)1.994 (15)2.7694 (7)163.9 (15)
O6—H6A···O2v0.777 (14)2.072 (14)2.8206 (7)161.8 (14)
O6—H6B···O30.774 (15)1.962 (15)2.7204 (7)166.5 (15)
O7—H7A···O2v0.810 (13)1.976 (14)2.7761 (7)169.3 (13)
O7—H7B···O6vi0.773 (14)2.171 (14)2.9110 (8)160.7 (15)
O8—H8A···O1iii0.792 (15)2.009 (15)2.7900 (7)168.9 (13)
O8—H8B···S1vii0.825 (14)2.455 (14)3.2582 (6)164.5 (13)
O9—H9A···O2vii0.807 (14)2.106 (14)2.9096 (7)174.0 (14)
O10—H10A···O4ii0.839 (15)1.962 (15)2.7908 (8)169.5 (14)
O10—H10B···O3vi0.762 (14)2.069 (14)2.8210 (7)169.1 (14)
Symmetry codes: (ii) x+1, y+1/2, z1/2; (iii) x+1, y, z+3; (iv) x+3/2, y, z+1/2; (v) x+3/2, y, z1/2; (vi) x+1, y, z+2; (vii) x1/2, y+1/2, z.
Comparison of lattice parameters (Å, °) for Na2SO3(H2O)7 (this work) and Na2CO3(H2O)7 (Betzel et al., 1982) top
Na2SO3(H2O)7Na2CO3(H2O)7
a14.6563 (8)14.492 (7)
b19.7180 (9)19.490 (5)
c7.2197 (5)7.017 (3)
α9090
β9090
γ94.0997 (17)90
V3)2081.1 (2)1981.95
T (K)100RT
Space groupC1121/aPbca
 

Acknowledgements

The X-ray centre of TU Wien is acknowledged for providing access to the single-crystal X-ray diffractometer.

References

First citationAllen, F. H., Bird, C. M., Rowland, R. S. & Raithby, P. R. (1997). Acta Cryst. B53, 696–701.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationAndersen, L. & Lindqvist, O. (1984). Acta Cryst. C40, 584–586.  CrossRef ICSD CAS Web of Science IUCr Journals Google Scholar
First citationBattelle, L. F. & Trueblood, K. N. (1965). Acta Cryst. 19, 531–535.  CrossRef ICSD IUCr Journals Web of Science Google Scholar
First citationBats, J. W., Fuess, H. & Elerman, Y. (1986). Acta Cryst. B42, 552–557.  CrossRef ICSD CAS Web of Science IUCr Journals Google Scholar
First citationBetzel, C., Saenger, W. & Loewus, D. (1982). Acta Cryst. B38, 2802–2804.  CrossRef ICSD CAS Web of Science 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: The Bond Valence Model. Oxford University Press.  Google Scholar
First citationBruker (2014). APEX2 and SAINT. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationDowty, E. (2006). ATOMS. Shape Software, Kingsport, Tennessee, USA.  Google Scholar
First citationDunsmore, H. S. & Speakman, J. C. (1963). Acta Cryst. 16, 573–574.  CrossRef IUCr Journals Google Scholar
First citationDurand, J., Galigné, J. L. & Cot, L. (1977). Acta Cryst. B33, 1414–1417.  CrossRef ICSD CAS IUCr Journals Google Scholar
First citationEklund, L., Hofer, T. S., Pribil, A. B., Rode, B. M. & Persson, I. (2012). Dalton Trans. 41, 5209–5216.  CrossRef CAS PubMed Google Scholar
First citationGagné, O. C. & Hawthorne, F. C. (2016). Acta Cryst. B72, 602–625.  Web of Science CrossRef IUCr Journals Google Scholar
First citationGagné, O. C. & Hawthorne, F. C. (2018). Acta Cryst. B74, 79–96.  Web of Science CrossRef IUCr Journals Google Scholar
First citationGroom, C. R., Bruno, I. J., Lightfoot, M. P. & Ward, S. C. (2016). Acta Cryst. B72, 171–179.  Web of Science CrossRef IUCr Journals Google Scholar
First citationJeffrey, G. A. (1997). In An Introduction to Hydrogen Bonding. New York: Oxford University Press Inc.  Google Scholar
First citationJohansson, L.-G., Lindqvist, O. & Vannerberg, N.-G. (1980). Acta Cryst. B36, 2523–2526.  CrossRef ICSD CAS IUCr Journals 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 citationLarsson, L. O. & Kierkegaard, P. (1969). Acta Chem. Scand. 23, 2253–2260.  CrossRef ICSD CAS Web of Science 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 citationMagnusson, A., Johansson, L.-G. & Lindqvist, O. (1983). Acta Cryst. C39, 819–822.  CrossRef ICSD CAS Web of Science IUCr Journals Google Scholar
First citationManojlović-Muir, L. (1969). Nature, 224, 686–687.  Google Scholar
First citationMasse, R., Guitel, J. C. & Tordjman, I. (1980). Mater. Res. Bull. 15, 431–436.  CrossRef ICSD CAS Web of Science Google Scholar
First citationMereiter, K. (2013). Acta Cryst. E69, i77–i78.  CSD CrossRef IUCr Journals Google Scholar
First citationMesser, D., Breitinger, D. & Haegler, W. (1979). Acta Cryst. B35, 815–818.  CrossRef ICSD CAS IUCr Journals Google Scholar
First citationMikenda, W., Mereiter, K. & Preisinger, A. (1989). Inorg. Chim. Acta, 161, 21–28.  CrossRef CAS Web of Science Google Scholar
First citationNyberg, B. & Larsson, R. (1973). Acta Chem. Scand. 27, 63–70.  CrossRef CAS Google Scholar
First citationOswald, I. D. H., Hamilton, A., Hall, C., Marshall, W. G., Prior, T. J. & Pulham, C. R. (2008). J. Am. Chem. Soc. 130, 17795–17800.  Web of Science CrossRef ICSD PubMed CAS Google Scholar
First citationPhilippot, E., Maurin, M. & Moret, J. (1979). Acta Cryst. B35, 1337–1340.  CrossRef ICSD CAS IUCr Journals Web of Science 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 citationWeil, E. D., Sandler, S. R. & Gernon, M. (2007). Sulfur Com­pounds, 13.1 Sodium Sulfite. In Kirk-Othmer Concise Encyclopedia of Chemical Technology, 5th ed., Vol. 23, edited by A. Seidel, pp. 669–672. New York: John Wiley & Sons Inc.  Google Scholar
First citationWeil, M., Baumann, S. O. & Breitinger, D. K. (2010). Acta Cryst. C66, i89–i91.  CrossRef ICSD IUCr Journals Google Scholar
First citationWestrip, S. P. (2010). J. Appl. Cryst. 43, 920–925.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationWickleder, M. S. (2002). Acta Cryst. E58, i103–i104.  Web of Science CrossRef ICSD CAS IUCr Journals Google Scholar
First citationZemann, J. (1981). Fortschr. Mineral. 59, 95–116.  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 logoSTRUCTURAL
CHEMISTRY
ISSN: 2053-2296
Follow Acta Cryst. C
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds