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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890
Volume 71| Part 6| June 2015| Pages 592-596

Crystal structures of spinel-type Na2MoO4 and Na2WO4 revisited using neutron powder diffraction

aISIS Facility, Rutherford Appleton Laboratory, Harwell Science and Innovation Campus, Didcot, Oxfordshire OX11 0QX, England, bDepartment of Earth Sciences, University College London, Gower Street, London WC1E 6BT, England, and cDepartment of Earth and Planetary Sciences, Birkbeck, University of London, Malet Street, London WC1E 7HX, England
*Correspondence e-mail: andrew.fortes@ucl.ac.uk

Edited by M. Weil, Vienna University of Technology, Austria (Received 24 April 2015; accepted 5 May 2015; online 9 May 2015)

Time-of-flight neutron powder diffraction data have been collected from Na2MoO4 and Na2WO4 to a resolution of sin (θ)/λ = 1.25 Å−1, which is substanti­ally better than the previous analyses using Mo Kα X-rays, providing roughly triple the number of measured reflections with respect to the previous studies [Okada et al. (1974[Okada, K., Morikawa, H., Marumo, F. & Iwai, S. (1974). Acta Cryst. B30, 1872-1873.]). Acta Cryst. B30, 1872–1873; Bramnik & Ehrenberg (2004[Bramnik, K. G. & Ehrenberg, H. (2004). Z. Anorg. Allg. Chem. 630, 1336-1341.]). Z. Anorg. Allg. Chem. 630, 1336–1341]. The unit-cell parameters are in excellent agreement with literature data [Swanson et al. (1962[Swanson, H. E., Morris, M. C., Stinchfield, R. P. & Evans, E. H. (1962). NBS Monograph No. 25, sect. 1, pp. 46-47.]). NBS Monograph No. 25, sect. 1, pp. 46–47] and the structural parameters for the molybdate agree very well with those of Bramnik & Ehrenberg (2004[Bramnik, K. G. & Ehrenberg, H. (2004). Z. Anorg. Allg. Chem. 630, 1336-1341.]). However, the tungstate structure refinement of Okada et al. (1974[Okada, K., Morikawa, H., Marumo, F. & Iwai, S. (1974). Acta Cryst. B30, 1872-1873.]) stands apart as being conspicuously inaccurate, giving significantly longer W—O distances, 1.819 (8) Å, and shorter Na—O distances, 2.378 (8) Å, than are reported here or in other simple tungstates. As such, this work represents an order-of-magnitude improvement in precision for sodium molybdate and an equally substantial improvement in both accuracy and precision for sodium tungstate. Both compounds adopt the spinel structure type. The Na+ ions have site symmetry .-3m and are in octa­hedral coordination while the transition metal atoms have site symmetry -43m and are in tetra­hedral coordination.

1. Chemical context

Both Na2MoO4 and Na2WO4 have rich phase diagrams in pressure and temperature space (Pistorius, 1966[Pistorius, C. W. F. T. (1966). J. Chem. Phys. 44, 4532-4537.]). The stable form at room temperature is the β-Ag2MoO4 cubic spinel structure type, space group Fd[\overline{3}]m, which has been known for almost a century (Wyckoff, 1922[Wyckoff, R. W. G. (1922). J. Am. Chem. Soc. 44, 1994-1998.]). Among the alkali metal sulfates, chromates, molybdates and tungstates, only Na2MoO4 and Na2WO4 adopt the normal spinel structure at ambient pressure. Li2MoO4 forms a cubic spinel structure at high pressure (Liebertz & Rooymans, 1967[Liebertz, J. & Rooymans, C. J. M. (1967). Solid State Commun. 5, 405-409.]). Li2WO4 forms a `spinel-like' phase at high pressure (Pistorius, 1975[Pistorius, C. W. F. T. (1975). J. Solid State Chem. 13, 325-329.]; Horiuchi et al., 1979[Horiuchi, H., Morimoto, N. & Yamaoka, S. (1979). J. Solid State Chem. 30, 129-135.]). Cubic sodium molybdate and sodium tungstate have been examined inter­mittently over subsequent decades using a variety of crystallographic techniques (Lindqvist, 1950[Lindqvist, I. (1950). Acta Chem. Scand. 4, 1066-1074.]; Becka & Poljak, 1958[Becka, L. N. & Poljak, R. J. (1958). Anales Asoc. Quim. Arg. 46, 204-209.]; Swanson et al., 1957[Swanson, H. E., Gilfrich, N. T. & Cook, M. I. (1957). Natl. Bur. Stand. (US) Circ. 539, Vol. 7, p. 45.], 1962[Swanson, H. E., Morris, M. C., Stinchfield, R. P. & Evans, E. H. (1962). NBS Monograph No. 25, sect. 1, pp. 46-47.]; Singh Mudher et al., 2005[Singh Mudher, K. D., Keskar, M., Krishnan, K. & Venugopal, V. (2005). J. Alloys Compd. 396, 275-279.]) and vibrational spectroscopic methods (Busey & Keller, 1964[Busey, R. H. & Keller, O. L. Jr (1964). J. Chem. Phys. 41, 215-225.]; Preudhomme & Tarte, 1972[Preudhomme, J. & Tarte, P. (1972). Spectrochim. Acta Part A, 28, 69-79.]; Breitinger et al., 1981[Breitinger, D. K., Emmert, L. & Kress, W. (1981). Ber. Bunsenges. Phys. Chem. 85, 504-505.]; Luz Lima et al., 2010[Luz Lima, C., Saraiva, G. D., Souza Filho, A. G., Paraguassu, W., Freire, P. T. C. & Mendes Filho, J. (2010). J. Raman Spectrosc. 41, 576-581.], 2011[Luz Lima, C., Saraiva, G. D., Freire, P. T. C., Maczka, M., Paraguassu, W., de Sousa, F. F. & Mendes Filho, J. (2011). J. Raman Spectrosc. 42, 799-802.]), or by nuclear magnetic resonance and quadrupole coupling (Lynch & Segel, 1972[Lynch, G. F. & Segel, S. L. (1972). Can. J. Phys. 50, 567-572.]). However, the extant structural information on both phases is derived from X-ray diffraction data of low to modest precision. The first published structure refinement of Na2MoO4 was only reported recently (Bramnik & Ehrenberg, 2004[Bramnik, K. G. & Ehrenberg, H. (2004). Z. Anorg. Allg. Chem. 630, 1336-1341.]) from X-ray powder diffraction data measured to sin (θ)/λ = 0.71 Å−1; the last structure refinement of Na2WO4 was reported by Okada et al. (1974[Okada, K., Morikawa, H., Marumo, F. & Iwai, S. (1974). Acta Cryst. B30, 1872-1873.]) from X-ray single-crystal diffraction data to sin (θ)/λ = 0.81 Å−1. Both compounds are highly soluble in water, crystallizing at room temperature as ortho­rhom­bic dihydrates (space group Pbca, Atovmyan & D'yachenko, 1969[Atovmyan, L. O. & D'yachenko, O. A. (1969). J. Struct. Chem. 10, 416-418.]; Farrugia, 2007[Farrugia, L. J. (2007). Acta Cryst. E63, i142.]). Below 283.5 K for the molybdate and 279.2 K for the tungstate, crystals grow with ten water mol­ecules per formula unit (Funk, 1900[Funk, R. (1900). Ber. Dtsch. Chem. Ges. 33, 3696-3703.]; Cadbury, 1955[Cadbury, W. E. Jr (1955). J. Phys. Chem. 59, 257-260.]; Zhilova et al., 2008[Zhilova, S. B., Karov, Z. G. & El'mesova, R. M. (2008). Russ. J. Inorg. Chem. 53, 628-635.]). The high solubility in water and propensity towards forming hydrogen-bonded hydrates (unlike the heavier alkali metal molybdates and tungstates) suggests that both compounds would be excellent candidates for formation of hydrogen-bonded complexes with water soluble organics, such as amino acids, producing metal–organic crystals with potentially useful optical properties (cf., glycine lithium molybdate; Fleck et al., 2006[Fleck, M., Schwendtner, K. & Hensler, A. (2006). Acta Cryst. C62, m122-m125.]).

In the course of preparing deuterated specimens of the dihydrated and deca­hydrated forms of Na2MoO4 and Na2WO4 for neutron diffraction analysis, the anhydrous phases were synthesised and an opportunity arose to acquire neutron powder diffraction data. The advantage of using a neutron radiation probe is that the scattering lengths of the atoms concerned are fairly similar, coherent scattering lengths being 6.715 fm for Mo, 4.86 fm for W, 3.63 fm for Na and 5.803 fm for O (Sears, 2006[Sears, V. F. (2006). Neutron News, 3, 26-37.]). Secondly, with the time-of-flight method, particularly with a very long primary flight path and high-angle backscattering detectors, one can acquire unparalleled resolution at very short flight times (i.e., small d-spacings), ensuring an order of magnitude improvement in parameter precision over the previous studies. In this work, usable data were obtained at a resolution of sin (θ)/λ = 1.25 Å−1, roughly tripling the number of measured reflections with respect to Okada et al. (1974[Okada, K., Morikawa, H., Marumo, F. & Iwai, S. (1974). Acta Cryst. B30, 1872-1873.]) and Bramnik & Ehrenberg (2004[Bramnik, K. G. & Ehrenberg, H. (2004). Z. Anorg. Allg. Chem. 630, 1336-1341.]). This work provides the most accurate and precise foundation on which to build future discussion of the hydrated forms of Na2MoO4 and Na2WO4. Neutron powder diffraction data for Na2MoO4 and Na2WO4 are given in Figs. 1[link] and 2[link].

[Figure 1]
Figure 1
Neutron powder diffraction data for Na2MoO4; red points are the observations, the green line is the calculated profile and the pink line beneath the diffraction pattern represents Obs − Calc. Vertical black tick marks report the expected positions of the Bragg peaks. The inset shows the data measured at very short flight times (i.e., small d-spacing).
[Figure 2]
Figure 2
Neutron powder diffraction data for Na2WO4; red points are the observations, the green line is the calculated profile and the pink line beneath the diffraction pattern represents Obs − Calc. Vertical black tick marks report the expected positions of the Bragg peaks. The inset shows the data measured at very short flight times (i.e., small d-spacing).

2. Structural commentary

The structure of both compounds is the normal spinel type with Na+ ions on the 16c sites in octa­hedral coordination and Mo6+/W6+ ions on 8b sites in tetra­hedral coordination. The coordinating oxygen atoms occupy the 32e general positions, their location being defined by a single variable parameter u. For ideal cubic close packing, the u coordinate adopts a value of 0.25 although for various spinels is found in the range 0.24 to 0.275. In Na2MoO4 the u parameter has a value of 0.262710 (15) and in Na2WO4 it has a value of 0.262246 (15). The practical consequence of this compared with the `ideal' value of u = 0.25 is that the shared edges of the NaO6 octa­hedra are shorter than the unshared edges (Fig. 3b[link]). In the molybdate, these lengths are 3.2288 (2) and 3.5479 (2) Å, the ratio being 1.0988 (1); in the tungstate, the lengths of the two inequivalent octa­hedral edges are 3.2356 (2) Å and 3.5441 (2) Å, their ratio being 1.0953 (1). The MoO42− and WO42− tetra­hedra have perfect Td symmetry with Mo—O and W—O bond lengths of 1.7716 (3) and 1.7830 (2) Å, respectively. The unit-cell parameters for both compounds are in excellent agreement with those of Swanson et al. (1962[Swanson, H. E., Morris, M. C., Stinchfield, R. P. & Evans, E. H. (1962). NBS Monograph No. 25, sect. 1, pp. 46-47.]) and the structural parameters for the molybdate agree very well with those of Bramnik & Ehrenberg (2004[Bramnik, K. G. & Ehrenberg, H. (2004). Z. Anorg. Allg. Chem. 630, 1336-1341.]). However, the Na2WO4 structure refinement of Okada et al. (1974[Okada, K., Morikawa, H., Marumo, F. & Iwai, S. (1974). Acta Cryst. B30, 1872-1873.]) stands apart as being conspicuously inaccurate, giving significantly longer W—O distances, 1.819 (8) Å, and shorter Na—O distances, 2.378 (8) Å, than are reported here or in many other simple tungstates. Indeed the ionic radii of four-coordinated Mo6+ and W6+ obtained from analysis of a large range of crystal structures are nearly identical, being 0.41 and 0.42 Å, respectively (Shannon, 1976[Shannon, R. D. (1976). Acta Cryst. A32, 751-767.]). The values reported here agree very well with the majority of Mo—O and W—O bond lengths in isolated MoO42− and WO42− tetra­hedral oxyanions from a range of alkali metal and alkaline earth compounds tabulated in the literature (e.g., Zachariasen & Plettinger, 1961[Zachariasen, W. H. & Plettinger, H. A. (1961). Acta Cryst. 14, 229-230.]; Gatehouse & Leverett, 1969[Gatehouse, B. M. & Leverett, P. (1969). J. Chem. Soc. A, pp. 849.]; Koster et al., 1969[Koster, A. S., Kools, F. X. N. M. & Rieck, G. D. (1969). Acta Cryst. B25, 1704-1708.]; Gürmen et al., 1971[Gürmen, E. (1971). J. Chem. Phys. 55, 1093-1097.]; Wandahl & Christensen, 1987[Wandahl, G. & Christensen, A. N. (1987). Acta Chem. Scand. Ser. A, 41, 358-360.]; Farrugia, 2007[Farrugia, L. J. (2007). Acta Cryst. E63, i142.]; van den Berg & Juffermans, 1982[Berg, A. J. van den & Juffermans, C. A. H. (1982). J. Appl. Cryst. 15, 114-116.]). As such, this work represents an improvement in accuracy for sodium molybdate and an improvement in both accuracy and precision for sodium tungstate.

[Figure 3]
Figure 3
(a) Arrangement of molybdate ions in the unit cell of Na2MoO4; anisotropic displacement ellipsoids are drawn at the 75% probability level. (b) Connectivity of the NaO6 octa­hedra, with shorter shared edges and longer unshared edges, to the MoO4 tetra­hedra in Na2MoO4; as in (a), the ellipsoids are drawn at the 75% probability level.

3. Synthesis and crystallization

Na2MoO4·2H2O (Sigma Aldrich M1003, > 99.5%) and Na2WO4·2H2O (Sigma Aldrich 14304, > 99.0%) were heated to 673 K in ceramic crucibles for 24 hr. Loss of water was confirmed by Raman spectroscopy; X-ray powder diffraction confirmed the phase identity and purity of the two anhydrous products, Na2MoO4 and Na2WO4.

Raman spectra were acquired using a B&WTek i-Raman plus portable spectrometer; this device uses a 532 nm laser (37 mW power at the fiber-optic probe tip) to stimulate Raman scattering, which is measured in the range 170–4000 cm−1 with a spectral resolution of 3 cm−1. Data were collected in a series of 20 x 9 sec integrations for Na2MoO4 and 20 x 7 sec integrations for Na2WO4; after summation, the background was removed and peaks fitted using Pseudo-Voigt functions in OriginPro (OriginLab, Northampton, MA) (Fig. 4[link]). These data are provided as an electronic supplement in the form of an ASCII file.

[Figure 4]
Figure 4
Raman spectra of Na2MoO4 (left) and Na2WO4 (right) in the range 0–1200 cm−1 (the full range of data to 4000 cm−1 is given in the electronic supplement). Band positions and vibrational assignments are indicated. For the tungstate these agree very well with literature values (e.g., Busey & Keller, 1964[Busey, R. H. & Keller, O. L. Jr (1964). J. Chem. Phys. 41, 215-225.]) whereas for the molybdate, these data show a systematic shift to lower frequencies by 3–4 wavenumbers with respect to published values (Luz Lima et al., 2010[Luz Lima, C., Saraiva, G. D., Souza Filho, A. G., Paraguassu, W., Freire, P. T. C. & Mendes Filho, J. (2010). J. Raman Spectrosc. 41, 576-581.], 2011[Luz Lima, C., Saraiva, G. D., Freire, P. T. C., Maczka, M., Paraguassu, W., de Sousa, F. F. & Mendes Filho, J. (2011). J. Raman Spectrosc. 42, 799-802.]).

4. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 1[link]. For the neutron scattering experiments, each specimen was loaded into a vanadium tube of 11 mm internal diameter to a depth of approximately 25 mm. The exact sample volume and mass were measured in order to determine the number density for correction of the specimen self-shielding. The samples were mounted on the HRPD beamline (Ibberson, 2009[Ibberson, R. M. (2009). Nucl. Instrum. Methods Phys. Res. A, 600, 47-49.]) at the ISIS neutron spallation source and data were collected in the 10–110 ms time-of-flight window for 2.5 h (Na2MoO4) and 3.5 h (Na2WO4). Data were corrected for self-shielding, focussed to a common scattering angle and normalized to the incident spectrum by reference to a V:Nb null-scattering standard before being output in a format suitable for Rietveld refinement with GSAS/Expgui (Larsen & Von Dreele, 2000[Larsen, A. C. & Von Dreele, R. B. (2000). General Structure Analysis System (GSAS). Los Alamos National Laboratory Report LAUR 86-748, Los Alamos, New Mexico, USA. http://www.ncnr.NIST.gov/Xtal/software/GSAS.html .]: Toby, 2001[Toby, B. H. (2001). J. Appl. Cryst. 34, 210-213.]).

Table 1
Experimental details

  Na2MoO4 Na2WO4
Crystal data
Chemical formula Na2MoO4 Na2WO4
Mr 205.92 293.83
Crystal system, space group Cubic, Fd[\overline{3}]m Cubic, Fd[\overline{3}]m
Temperature (K) 298 298
a (Å) 9.10888 (3) 9.12974 (4)
V3) 755.78 (1) 760.98 (1)
Z 8 8
Radiation type Neutron Neutron
μ (mm−1) 0.014 + 0.0018 * λ 0.014 + 0.0097 * λ
Specimen shape, size (mm) Cylinder, 25 × 11 Cylinder, 27 × 11
 
Data collection
Diffractometer HRPD, high-resolution neutron powder HRPD, high-resolution neutron powder
Specimen mounting Vanadium tube Vanadium tube
Data collection mode Transmission Transmission
Scan method Time of flight Time of flight
Absorption correction Analytical Analytical
2θ values (°) 2θfixed = 168.329 2θfixed = 168.329
Distance from source to specimen (mm) 95000 95000
Distance from specimen to detector (mm) 965 965
 
Refinement
R factors and goodness of fit Rp = 0.037, Rwp = 0.043, Rexp = 0.022, R(F2) = 0.06364, χ2 = 3.842 Rp = 0.037, Rwp = 0.044, Rexp = 0.024, R(F2) = 0.06245, χ2 = 3.423
No. of data points 7716 7716
No. of parameters 24 24
Computer programs: HRPD control software, GSAS/Expgui (Larsen & Von Dreele, 2000[Larsen, A. C. & Von Dreele, R. B. (2000). General Structure Analysis System (GSAS). Los Alamos National Laboratory Report LAUR 86-748, Los Alamos, New Mexico, USA. http://www.ncnr.NIST.gov/Xtal/software/GSAS.html .]; Toby, 2001[Toby, B. H. (2001). J. Appl. Cryst. 34, 210-213.]), MANTID (Arnold et al., 2014[Arnold, O., & 27 co-authors (2014). Nucl. Instrum. Methods Phys. Res. A, 764, 156-166.]; Mantid, 2013[Mantid (2013). Manipulation and Analysis Toolkit for Instrument Data; Mantid Project. http://dx.doi.org/10.5286/SOFTWARE/MANTID .]), DIAMOND (Putz & Brandenburg, 2006[Putz, H. & Brandenburg, K. (2006). DIAMOND. Crystal Impact GbR, Bonn, Germany. http://www.crystalimpact.com/diamond .]) and publCIF (Westrip, 2010[Westrip, S. P. (2010). J. Appl. Cryst. 43, 920-925.]).

Supporting information


Chemical context top

Both Na2MoO4 and Na2WO4 have rich phase diagrams in pressure and temperature space (Pistorius, 1966). The stable form at room temperature is the β-Ag2MoO4 cubic spinel structure type, space group Fd3m, which has been known for almost a century (Wyckoff, 1922). Among the alkali metal sulfates, chromates, molybdates and tungstates, only Na2MoO4 and Na2WO4 adopt the normal spinel structure at ambient pressure. Li2MoO4 forms a cubic spinel structure at high pressure (Liebertz & Rooymans, 1967). Li2WO4 forms a 'spinel-like' phase at high pressure (Pistorius, 1975; Horiuchi et al., 1979). Cubic sodium molybdate and sodium tungstate have been examined inter­mittently over subsequent decades using a variety of crystallographic techniques (Lindqvist, 1950; Becka & Poljak, 1958; Swanson et al., 1957, 1962; Singh Mudher et al., 2005) and vibrational spectroscopic methods (Busey & Keller, 1964; Preudhomme & Tarte, 1972; Breitinger et al., 1981; Luz Lima et al., 2010, 2011), or by nuclear magnetic resonance and quadrupole coupling (Lynch & Segel, 1972). However, the extant structural information on both phases is derived from X-ray diffraction data of low to modest precision. The first published structure refinement of Na2MoO4 was only reported recently (Bramnik & Ehrenberg, 2004) from X-ray powder diffraction data measured to sin(θ)/λ = 0.71 Å-1; the last structure refinement of Na2WO4 was reported by Okada et al. (1974) from X-ray single-crystal diffraction data to sin(θ)/λ = 0.81 Å-1. Both compounds are highly soluble in water, crystallizing at room temperature as orthorhombic dihydrates (space group Pbca, Atovmyan & D'yachenko, 1969; Farrugia, 2007). Below 283.5 K for the molybdate and 279.2 K for the tungstate, crystals grow with ten water molecules per formula unit (Funk, 1900; Cadbury, 1955; Zhilova et al., 2008). The high solubility in water and propensity towards forming hydrogen-bonded hydrates (unlike the heavier alkali metal molybdates and tungstates) suggests that both compounds would be excellent candidates for formation of hydrogen-bonded complexes with water soluble organics, such as amino acids, producing metal–organic crystals with potentially useful optical properties (cf., glycine lithium molybdate; Fleck et al., 2006).

In the course of preparing deuterated specimens of the dihydrated and decahydrated forms of Na2MoO4 and Na2WO4 for neutron diffraction analysis, the anhydrous phases were synthesised and an opportunity arose to acquire neutron powder diffraction data. The advantage of using a neutron radiation probe is that the scattering lengths of the atoms concerned are fairly similar, coherent scattering lengths being 6.715 fm for Mo, 4.86 fm for W, 3.63 fm for Na and 5.803 fm for O (Sears, 206). Secondly, with the time-of-flight method, particularly with a very long primary flight path and high-angle backscattering detectors, one can acquire unparalleled resolution at very short flight times (i.e., small d-spacings), ensuring an order of magnitude improvement in parameter precision over the previous studies. In this work, usable data were obtained at a resolution of sin(θ)/λ = 1.25 Å-1, roughly tripling the number of measured reflections with respect to Okada et al. (1974) and Bramnik & Ehrenberg (2004). This work provides the most accurate and precise foundation on which to build future discussion of the hydrated forms of Na2MoO4 and Na2WO4. Neutron powder diffraction data for Na2MoO4 and Na2WO4 are given in Figs. 1 and 2.

Structural commentary top

The structure of both compounds is the normal spinel type with Na+ ions on the 16c sites in o­cta­hedral coordination and Mo6+/W6+ ions on 8b sites in tetra­hedral coordination. The coordinating oxygen atoms occupy the 32e general positions, their location being defined by a single variable parameter u. For ideal cubic close packing, the u coordinate adopts a value of 0.25 although for various spinels is found in the range 0.24 to 0.275. In Na2MoO4 the u parameter has a value of 0.262710 (15) and in Na2WO4 it has a value of 0.262246 (15). The practical consequence of this compared with the `ideal' value of u = 0.25 is that the shared edges of the NaO6 o­cta­hedra are shorter than the unshared edges (Fig. 3). In the molybdate, these lengths are 3.2288 (2) and 3.5479 (2) Å, the ratio being 1.0988 (1); in the tungstate, the lengths of the two inequivalent o­cta­hedral edges are 3.2356 (2) Å and 3.5441 (2) Å, their ratio being 1.0953 (1). The MoO42- and WO42- tetra­hedra have perfect Td symmetry with Mo—O and W—O bond lengths of 1.7716 (3) and 1.7830 (2) Å, respectively. The unit-cell parameters for both compounds are in excellent agreement with those of Swanson et al. (1962) and the structural parameters for the molybdate agree very well with those of Bramnik & Ehrenberg (2004). However, the Na2WO4 structure refinement of Okada et al. (1974) stands apart as being conspicuously inaccurate, giving significantly longer W—O distances, 1.819 (8) Å, and shorter Na—O distances, 2.378 (8) Å, than are reported here or in many other simple tungstates. Indeed the ionic radii of four-coordinated Mo6+ and W6+ obtained from analysis of a large range of crystal structures are nearly identical, being 0.41 and 0.42 Å, respectively (Shannon, 1976). The values reported here agree very well with the majority of Mo—O and W—O bond lengths in isolated MoO42- and WO42- tetra­hedral oxyanions from a range of alkali metal and alkaline earth compounds tabulated in the literature (e.g., Zachariasen & Plettinger, 1961; Gatehouse & Leverett, 1969; Koster et al., 1969; Gürmen et al., 1971; Wandahl & Christensen, 1987; Farrugia, 2007; van den Berg & Juffermans, 1982). As such, this work represents an improvement in accuracy for sodium molybdate and an improvement in both accuracy and precision for sodium tungstate.

Synthesis and crystallization top

Na2MoO4·2H2O (Sigma Aldrich M1003, > 99.5 %) and Na2WO4·2H2O (Sigma Aldrich 14304, > 99.0 %) were heated to 673 K in ceramic crucibles for 24 hr. Loss of water was confirmed by Raman spectroscopy; X-ray powder diffraction confirmed the phase identity and purity of the two anhydrous products, Na2MoO4 and Na2WO4.

Raman spectra were acquired using a B&WTek i-Raman plus portable spectrometer; this device uses a 532 nm laser (37 mW power at the fiber-optic probe tip) to stimulate Raman scattering, which is measured in the range 170–4000 cm-1 with a spectral resolution of 3 cm-1. Data were collected in a series of 20 x 9 sec integrations for Na2MoO4 and 20 x 7 sec integrations for Na2WO4; after summation, the background was removed and peaks fitted using Pseudo-Voigt functions in OriginPro (Fig. 4). These data are provided as an electronic supplement in the form of an ASCII file.

Refinement top

Crystal data, data collection and structure refinement details are summarized in Table 1. For the neutron scattering experiments, each specimen was loaded into a vanadium tube of 11 mm diameter to a depth of approximately 25 mm. The exact sample height and mass were measured in order to determine the number density for correction of the specimen self-shielding. The samples were mounted on the HRPD beamline (Ibberson, 2009) at the ISIS neutron spallation source and data were collected in the 10–110 ms time-of-flight window for 2.5 h (Na2MoO4) and 3.5 h (Na2WO4). Data were corrected for self-shielding, focussed to a common scattering angle and normalized to the incident spectrum by reference to a V:Nb null-scattering standard before being output in a format suitable for Rietveld refinement with GSAS/Expgui (Larsen & Von Dreele, 2000: Toby, 2001).

Related literature top

For related literature, see: Atovmyan & D'yachenko (1969); Becka & Poljak (1958); Bramnik & Ehrenberg (2004); Breitinger et al. (1981); Busey & Keller (1964); Cadbury (1955); Farrugia (2007); Fleck et al. (2006); Funk (1900); Gürmen et al. (1971); Gatehouse & Leverett (1969); Horiuchi et al. (1979); Ibberson (2009); Koster et al. (1969); Larsen & Von Dreele (2000); Liebertz & Rooymans (1967); Lindqvist (1950); Luz Lima, Saraiva, Freire, Maczka, Paraguassu, de Sousa & Mendes Filho (2011); Luz Lima, Saraiva, Souza Filho, Paraguassu, Freire & Mendes Filho (2010); Okada et al. (1974); Pistorius (1966, 1975); Preudhomme & Tarte (1972); Sears (2006); Shannon (1976); Singh Mudher, Keskar, Krishnan & Venugopal (2005); Swanson et al. (1957, 1962); Toby (2001); van den Berg & Juffermans (1982); Wandahl & Christensen (1987); Wyckoff (1922); Zachariasen & Plettinger (1961); Zhilova et al. (2008).

Computing details top

For both compounds, data collection: HRPD control software; cell refinement: GSAS/Expgui (Larsen & Von Dreele, 2000; Toby, 2001); data reduction: MANTID (Arnold et al., 2014; Mantid, 2013); program(s) used to solve structure: coordinates taken from a previous refinement. Program(s) used to refine structure: GSAS/Expgui (Larsen & Von Dreele, 2000, Toby, 2001) for Na2MoO4; GSAS/Expgui (Larsen & Von Dreele, 2000; Toby, 2001) for Na2WO4. For both compounds, molecular graphics: DIAMOND (Putz & Brandenburg, 2006); software used to prepare material for publication: publCIF (Westrip, 2010).

Figures top
[Figure 1] Fig. 1. Neutron powder diffraction data for Na2MoO4; red points are the observations, the green line is the calculated profile and the pink line beneath the diffraction pattern represents Obs-Calc. Vertical black tick marks report the expected positions of the Bragg peaks. The inset shows the data measured at very short flight times (i.e., small d-spacing).
[Figure 2] Fig. 2. Neutron powder diffraction data for Na2WO4; red points are the observations, the green line is the calculated profile and the pink line beneath the diffraction pattern represents Obs-Calc. Vertical black tick marks report the expected positions of the Bragg peaks. The inset shows the data measured at very short flight times (i.e., small d-spacing).
[Figure 3] Fig. 3. (a) Arrangement of molybdate ions in the unit cell of Na2MoO4; anisotropic displacement ellipsoids are drawn at the 75% probability level. (b) Connectivity of the NaO6 octahedra, with shorter shared edges and longer unshared edges, to the MoO4 tetrahedra in Na2MoO4; as in (a), the ellipsoids are drawn at the 75% probability level.
[Figure 4] Fig. 4. Raman spectra of Na2MoO4 (left) and Na2WO4 (right) in the range 0–1200 cm-1 (the full range of data to 4000 cm-1 is given in the electronic supplement). Band positions and vibrational assignments are indicated. For the tungstate these agree very well with literature values (e.g., Busey & Keller, 1964) whereas for the molybdate, these data show a systematic shift to lower frequencies by 3–4 wavenumbers with respect to published values (Luz Lima et al., 2010, 2011).
(Na2MoO4) Disodium molybdenum(VI) oxide top
Crystal data top
Na2MoO4Melting point: 961 K
Mr = 205.92Neutron radiation
Cubic, Fd3mµ = 0.01+ 0.0018 * λ mm1
Hall symbol: -F 4vw 2vw 3T = 298 K
a = 9.10888 (3) Åwhite
V = 755.78 (1) Å3cylinder, 25 × 11 mm
Z = 8Specimen preparation: Prepared at 673 K and 100 kPa
Dx = 3.619 Mg m3
Data collection top
HRPD, High resolution neutron powder
diffractometer
Absorption correction: analytical
Data were corrected for self shielding using σscatt = 29.198 barns and σab(λ) = 3.541 barns at 1.798 Å during the normalization procedure. The linear absorption coefficient is wavelength dependent and is calculated as: µ = 0.014 + 0.0018 * λ (mm-1).
Radiation source: ISIS Facility, Neutron spallation sourceTmin = 1.000, Tmax = 1.000
Specimen mounting: vanadium tube2θfixed = 168.329
Data collection mode: transmissionDistance from source to specimen: 95000 mm
Scan method: time of flightDistance from specimen to detector: 965 mm
Refinement top
Least-squares matrix: fullExcluded region(s): Data at d-spacings smaller than 0.4 Å were excluded since the counting statistics became progressively poorer at very short flight times due to the lower neutron flux at the shortest wavelengths.
Rp = 0.037Profile function: TOF profile function #3 (21 terms). Profile coefficients for exp pseudovoigt convolution [Von Dreele, 1990 (unpublished)] (α) = 0.1919, (β0) = 0.025953, (β1) = 0.005213, (σ0) = 0, (σ1) = 196.3, (σ2) = 23.5, (γ0) = 0, (γ1) = 14.91, (γ2) = 0, (γ2s) = 0, (γ1e) = 0, (γ2e) = 0, (εi) = 0, (εa) = 0, (εA) = 0, (γ11) = 0, (γ22) = 0, (γ33) = 0, (γ12) = 0, (γ13) = 0, (γ23) = 0. Peak tails ignored where intensity <0.0005x peak. Aniso. broadening axis 0.0 0.0 1.0
Rwp = 0.04324 parameters
Rexp = 0.0220 restraints
R(F2) = 0.063640 constraints
χ2 = 3.842(Δ/σ)max = 0.03
7716 data pointsBackground function: GSAS Background function #1 (10 terms). Shifted Chebyshev function of 1st kind 1: 1.18715, 2: -7.466630x10-3, 3:8.117230x10-2, 4: -5.411800x10-2, 5: -1.714140x10-2, 6: -1.882400x10-2, 7: -1.930110x10-2, 8: -6.255180x10-3, 9: 6.598230x10-3, 10: 8.478560x10-3
Crystal data top
Na2MoO4Z = 8
Mr = 205.92Neutron radiation
Cubic, Fd3mµ = 0.01+ 0.0018 * λ mm1
a = 9.10888 (3) ÅT = 298 K
V = 755.78 (1) Å3cylinder, 25 × 11 mm
Data collection top
HRPD, High resolution neutron powder
diffractometer
Tmin = 1.000, Tmax = 1.000
Specimen mounting: vanadium tube2θfixed = 168.329
Data collection mode: transmissionDistance from source to specimen: 95000 mm
Scan method: time of flightDistance from specimen to detector: 965 mm
Absorption correction: analytical
Data were corrected for self shielding using σscatt = 29.198 barns and σab(λ) = 3.541 barns at 1.798 Å during the normalization procedure. The linear absorption coefficient is wavelength dependent and is calculated as: µ = 0.014 + 0.0018 * λ (mm-1).
Refinement top
Rp = 0.037χ2 = 3.842
Rwp = 0.0437716 data points
Rexp = 0.02224 parameters
R(F2) = 0.063640 restraints
Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
O0.262710 (16)0.262710 (16)0.262710 (16)0.01182
Mo0.3750.3750.3750.00740
Na0.00.00.00.01381
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
O0.01182 (6)0.01182 (6)0.01182 (6)0.00156 (6)0.00156 (6)0.00156 (6)
Mo0.00740 (8)0.00740 (8)0.00740 (8)0.00.00.0
Na0.01381 (11)0.01381 (11)0.01381 (11)0.00068 (12)0.00068 (12)0.00068 (12)
Geometric parameters (Å, º) top
Mo—O1.7716 (3)Naiv—Ovii2.3986 (2)
Mo—Oi1.7716 (3)Naiv—Oviii2.3986 (2)
Mo—Oii1.7716 (3)Naiv—Oix2.3986 (2)
Mo—Oiii1.7716 (3)Na—Naiv3.2205 (1)
Naiv—O2.3986 (2)Mo—Naiv3.7763 (1)
Naiv—Ov2.3986 (2)O—Ovi3.2288 (2)
Naiv—Ovi2.3986 (2)O—Ov3.5479 (4)
O—Mo—Oi109.4712 (1)O—Naiv—Ov95.393 (6)
O—Mo—Oii109.4712 (1)O—Naiv—Oix180.000 (1)
O—Mo—Oiii109.4712 (1)O—Naiv—Oviii84.607 (6)
Oi—Mo—Oii109.4712 (1)O—Naiv—Ovii95.393 (6)
Oi—Mo—Oiii109.4712 (1)Ov—Naiv—Ovi180.000 (1)
Oii—Mo—Oiii109.4712 (1)Mo—O—Naiv129.178 (5)
O—Naiv—Ovi84.607 (6)
Symmetry codes: (i) x+3/4, y, z+3/4; (ii) x+3/4, y+3/4, z; (iii) x, y+3/4, z+3/4; (iv) x+1/4, y+1/4, z; (v) x, y+1/4, z+1/4; (vi) y+1/2, x+1/4, z1/4; (vii) x+1/4, y, z+1/4; (viii) y+1/4, x+1/2, z1/4; (ix) y+1/2, x+1/2, z.
(Na2WO4) Disodium tungsten(VI) oxide top
Crystal data top
Na2WO4Melting point: 969 K
Mr = 293.83Neutron radiation
Cubic, Fd3mµ = 0.01+ 0.0097 * λ mm1
Hall symbol: -F 4vw 2vw 3T = 298 K
a = 9.12974 (4) Åwhite
V = 760.98 (1) Å3cylinder, 27 × 11 mm
Z = 8Specimen preparation: Prepared at 673 K and 100 kPa
Dx = 5.129 Mg m3
Data collection top
HRPD, High resolution neutron powder
diffractometer
Absorption correction: analytical
Data were corrected for self shielding using σscatt = 28.088 barns and σab(λ) = 19.361 barns at 1.798 Å during the normalisation procedure. The linear absorption coefficient is wavelength dependent and is calculated as: µ = 0.014 + 0.0097 * λ [mm-1]
Radiation source: ISIS Facility, Neutron spallation sourceTmin = 1.000, Tmax = 1.000
Specimen mounting: vanadium tube2θfixed = 168.329
Data collection mode: transmissionDistance from source to specimen: 95000 mm
Scan method: time of flightDistance from specimen to detector: 965 mm
Refinement top
Least-squares matrix: fullExcluded region(s): Data at d-spacings smaller than 0.4 Å were excluded since the counting statistics became progressively poorer at very short flight times due to the lower neutron flux at the shortest wavelengths.
Rp = 0.037Profile function: TOF profile function #3 (21 terms). Profile coefficients for exp pseudovoigt convolution [Von Dreele, 1990 (unpublished)] (α) = 0.1603, (β0) = 0.026115, (β1) = 0.004558, (σ0) = 0, (σ1) = 237.2, (σ2) = 45.0, (γ0) = 0, (γ1) = 14.21, (γ2) = 0, (γ2s) = 0, (γ1e) = 0, (γ2e) = 0, (εi) = 0, (εa) = 0, (εA) = 0, (γ11) = 0, (γ22) = 0, (γ33) = 0, (γ12) = 0, (γ13) = 0, (γ23) = 0. Peak tails ignored where intensity <0.0005x peak. Aniso. broadening axis 0.0 0.0 1.0
Rwp = 0.04424 parameters
Rexp = 0.0240 restraints
R(F2) = 0.062450 constraints
χ2 = 3.423(Δ/σ)max = 0.01
7716 data pointsBackground function: GSAS Background function # 1 (10 terms). Shifted Chebyshev function of 1st kind 1: 0.884779, 2: 4.212470x10-2, 3: 4.210950x10-2, 4: -4.489520x10-2, 5: -2.683690x10-2, 6: -1.892450x10-2, 7: -2.248710x10-2, 8: -2.821970x10-3, 9: 6.467340x10-3, 10: 6.167050x10-3
Crystal data top
Na2WO4Z = 8
Mr = 293.83Neutron radiation
Cubic, Fd3mµ = 0.01+ 0.0097 * λ mm1
a = 9.12974 (4) ÅT = 298 K
V = 760.98 (1) Å3cylinder, 27 × 11 mm
Data collection top
HRPD, High resolution neutron powder
diffractometer
Tmin = 1.000, Tmax = 1.000
Specimen mounting: vanadium tube2θfixed = 168.329
Data collection mode: transmissionDistance from source to specimen: 95000 mm
Scan method: time of flightDistance from specimen to detector: 965 mm
Absorption correction: analytical
Data were corrected for self shielding using σscatt = 28.088 barns and σab(λ) = 19.361 barns at 1.798 Å during the normalisation procedure. The linear absorption coefficient is wavelength dependent and is calculated as: µ = 0.014 + 0.0097 * λ [mm-1]
Refinement top
Rp = 0.037χ2 = 3.423
Rwp = 0.0447716 data points
Rexp = 0.02424 parameters
R(F2) = 0.062450 restraints
Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
O0.262246 (15)0.262246 (15)0.262246 (15)0.01312
W0.3750.3750.3750.00903
Na0.00.00.00.01538
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
O0.01312 (6)0.01312 (6)0.01312 (6)0.00161 (5)0.00161 (5)0.00161 (5)
W0.00903 (11)0.00903 (11)0.00903 (11)0.00.00.0
Na0.01538 (11)0.01538 (11)0.01538 (11)0.00045 (12)0.00045 (12)0.00045 (12)
Geometric parameters (Å, º) top
W—O1.7830 (2)Naiv—Ovii2.3995 (1)
W—Oi1.7830 (2)Naiv—Oviii2.3995 (1)
W—Oii1.7830 (2)Naiv—Oix2.3995 (1)
W—Oiii1.7830 (2)Na—Naiv3.2279 (1)
Naiv—O2.3995 (1)W—Naiv3.7850 (1)
Naiv—Ov2.3995 (1)O—Ovi3.2356 (2)
Naiv—Ovi2.3995 (1)O—Ov3.5441 (4)
O—W—Oi109.4712 (3)O—Naiv—Ov95.211 (6)
O—W—Oii109.4712 (3)O—Naiv—Oix180.000 (1)
O—W—Oiii109.4712 (3)O—Naiv—Oviii84.789 (6)
Oi—W—Oii109.4712 (3)O—Naiv—Ovii95.211 (6)
Oi—W—Oiii109.4712 (3)Ov—Naiv—Ovi180.000 (1)
Oii—W—Oiii109.4712 (3)W—O—Naiv129.043 (4)
O—Naiv—Ovi84.789 (6)
Symmetry codes: (i) x+3/4, y, z+3/4; (ii) x+3/4, y+3/4, z; (iii) x, y+3/4, z+3/4; (iv) x+1/4, y+1/4, z; (v) x, y+1/4, z+1/4; (vi) y+1/2, x+1/4, z1/4; (vii) x+1/4, y, z+1/4; (viii) y+1/4, x+1/2, z1/4; (ix) y+1/2, x+1/2, z.

Experimental details

(Na2MoO4)(Na2WO4)
Crystal data
Chemical formulaNa2MoO4Na2WO4
Mr205.92293.83
Crystal system, space groupCubic, Fd3mCubic, Fd3m
Temperature (K)298298
a (Å)9.10888 (3) 9.12974 (4)
V3)755.78 (1)760.98 (1)
Z88
Radiation typeNeutronNeutron
µ (mm1)0.01+ 0.0018 * λ0.01+ 0.0097 * λ
Specimen shape, size (mm)Cylinder, 25 × 11Cylinder, 27 × 11
Data collection
DiffractometerHRPD, High resolution neutron powder
diffractometer
HRPD, High resolution neutron powder
diffractometer
Specimen mountingVanadium tubeVanadium tube
Data collection modeTransmissionTransmission
Scan methodTime of flightTime of flight
Absorption correctionAnalytical
Data were corrected for self shielding using σscatt = 29.198 barns and σab(λ) = 3.541 barns at 1.798 Å during the normalization procedure. The linear absorption coefficient is wavelength dependent and is calculated as: µ = 0.014 + 0.0018 * λ (mm-1).
Tmin, Tmax1.000, 1.000
2θ values (°)2θfixed = 168.3292θfixed = 168.329
Distance from source to specimen (mm)9500095000
Distance from specimen to detector (mm)965965
Refinement
R factors and goodness of fitRp = 0.037, Rwp = 0.043, Rexp = 0.022, R(F2) = 0.06364, χ2 = 3.842Rp = 0.037, Rwp = 0.044, Rexp = 0.024, R(F2) = 0.06245, χ2 = 3.423
No. of data points77167716
No. of parameters2424

Computer programs: HRPD control software, GSAS/Expgui (Larsen & Von Dreele, 2000; Toby, 2001), MANTID (Arnold et al., 2014; Mantid, 2013), coordinates taken from a previous refinement, GSAS/Expgui (Larsen & Von Dreele, 2000, Toby, 2001), DIAMOND (Putz & Brandenburg, 2006), publCIF (Westrip, 2010).

 

Acknowledgements

The author thanks the STFC ISIS facility for beam-time access and acknowledges financial support from the STFC, grant No. ST/K000934/1.

References

First citationArnold, O., & 27 co-authors (2014). Nucl. Instrum. Methods Phys. Res. A, 764, 156–166.  Google Scholar
First citationAtovmyan, L. O. & D'yachenko, O. A. (1969). J. Struct. Chem. 10, 416–418.  CrossRef Google Scholar
First citationBecka, L. N. & Poljak, R. J. (1958). Anales Asoc. Quim. Arg. 46, 204–209.  CAS Google Scholar
First citationBerg, A. J. van den & Juffermans, C. A. H. (1982). J. Appl. Cryst. 15, 114–116.  CrossRef Web of Science IUCr Journals Google Scholar
First citationBramnik, K. G. & Ehrenberg, H. (2004). Z. Anorg. Allg. Chem. 630, 1336–1341.  Web of Science CrossRef CAS Google Scholar
First citationBreitinger, D. K., Emmert, L. & Kress, W. (1981). Ber. Bunsenges. Phys. Chem. 85, 504–505.  CrossRef CAS Google Scholar
First citationBusey, R. H. & Keller, O. L. Jr (1964). J. Chem. Phys. 41, 215–225.  CrossRef CAS Web of Science Google Scholar
First citationCadbury, W. E. Jr (1955). J. Phys. Chem. 59, 257–260.  CrossRef CAS Web of Science Google Scholar
First citationFarrugia, L. J. (2007). Acta Cryst. E63, i142.  Web of Science CrossRef IUCr Journals Google Scholar
First citationFleck, M., Schwendtner, K. & Hensler, A. (2006). Acta Cryst. C62, m122–m125.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationFunk, R. (1900). Ber. Dtsch. Chem. Ges. 33, 3696–3703.  CrossRef CAS Google Scholar
First citationGatehouse, B. M. & Leverett, P. (1969). J. Chem. Soc. A, pp. 849.  Google Scholar
First citationGürmen, E. (1971). J. Chem. Phys. 55, 1093–1097.  Google Scholar
First citationHoriuchi, H., Morimoto, N. & Yamaoka, S. (1979). J. Solid State Chem. 30, 129–135.  CrossRef CAS Web of Science Google Scholar
First citationIbberson, R. M. (2009). Nucl. Instrum. Methods Phys. Res. A, 600, 47–49.  Web of Science CrossRef CAS Google Scholar
First citationKoster, A. S., Kools, F. X. N. M. & Rieck, G. D. (1969). Acta Cryst. B25, 1704–1708.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationLarsen, A. C. & Von Dreele, R. B. (2000). General Structure Analysis System (GSAS). Los Alamos National Laboratory Report LAUR 86-748, Los Alamos, New Mexico, USA. http://www.ncnr.NIST.gov/Xtal/software/GSAS.htmlGoogle Scholar
First citationLiebertz, J. & Rooymans, C. J. M. (1967). Solid State Commun. 5, 405–409.  CrossRef CAS Web of Science Google Scholar
First citationLindqvist, I. (1950). Acta Chem. Scand. 4, 1066–1074.  CrossRef CAS Web of Science Google Scholar
First citationLuz Lima, C., Saraiva, G. D., Freire, P. T. C., Maczka, M., Paraguassu, W., de Sousa, F. F. & Mendes Filho, J. (2011). J. Raman Spectrosc. 42, 799–802.  Google Scholar
First citationLuz Lima, C., Saraiva, G. D., Souza Filho, A. G., Paraguassu, W., Freire, P. T. C. & Mendes Filho, J. (2010). J. Raman Spectrosc. 41, 576–581.  CAS Google Scholar
First citationLynch, G. F. & Segel, S. L. (1972). Can. J. Phys. 50, 567–572.  CrossRef CAS Google Scholar
First citationMantid (2013). Manipulation and Analysis Toolkit for Instrument Data; Mantid Project. http://dx.doi.org/10.5286/SOFTWARE/MANTIDGoogle Scholar
First citationOkada, K., Morikawa, H., Marumo, F. & Iwai, S. (1974). Acta Cryst. B30, 1872–1873.  CrossRef IUCr Journals Web of Science Google Scholar
First citationPistorius, C. W. F. T. (1966). J. Chem. Phys. 44, 4532–4537.  CrossRef CAS Web of Science Google Scholar
First citationPistorius, C. W. F. T. (1975). J. Solid State Chem. 13, 325–329.  CrossRef CAS Web of Science Google Scholar
First citationPreudhomme, J. & Tarte, P. (1972). Spectrochim. Acta Part A, 28, 69–79.  CrossRef CAS Web of Science Google Scholar
First citationPutz, H. & Brandenburg, K. (2006). DIAMOND. Crystal Impact GbR, Bonn, Germany. http://www.crystalimpact.com/diamondGoogle Scholar
First citationSears, V. F. (2006). Neutron News, 3, 26–37.  CrossRef Google Scholar
First citationShannon, R. D. (1976). Acta Cryst. A32, 751–767.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationSingh Mudher, K. D., Keskar, M., Krishnan, K. & Venugopal, V. (2005). J. Alloys Compd. 396, 275–279.  Web of Science CrossRef Google Scholar
First citationSwanson, H. E., Gilfrich, N. T. & Cook, M. I. (1957). Natl. Bur. Stand. (US) Circ. 539, Vol. 7, p. 45.  Google Scholar
First citationSwanson, H. E., Morris, M. C., Stinchfield, R. P. & Evans, E. H. (1962). NBS Monograph No. 25, sect. 1, pp. 46–47.  Google Scholar
First citationToby, B. H. (2001). J. Appl. Cryst. 34, 210–213.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationWandahl, G. & Christensen, A. N. (1987). Acta Chem. Scand. Ser. A, 41, 358–360.  CrossRef Web of Science Google Scholar
First citationWestrip, S. P. (2010). J. Appl. Cryst. 43, 920–925.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationWyckoff, R. W. G. (1922). J. Am. Chem. Soc. 44, 1994–1998.  CrossRef CAS Google Scholar
First citationZachariasen, W. H. & Plettinger, H. A. (1961). Acta Cryst. 14, 229–230.  CrossRef IUCr Journals Web of Science Google Scholar
First citationZhilova, S. B., Karov, Z. G. & El'mesova, R. M. (2008). Russ. J. Inorg. Chem. 53, 628–635.  Web of Science CrossRef 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
Volume 71| Part 6| June 2015| Pages 592-596
Follow Acta Cryst. E
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds