inorganic compounds\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Thortveitite-type Tm2Si2O7

aUniversity of Innsbruck, Institute of Mineralogy & Petrography, Innrain 52, A-6020 Innsbruck, Austria
*Correspondence e-mail: volker.kahlenberg@uibk.ac.at

(Received 24 May 2014; accepted 5 June 2014; online 11 June 2014)

Single crystals of dithulium disilicate, Tm2Si2O7, were obtained in flux synthesis experiments in the system SiO2–Tm2O3–LiF at ambient pressure. The compound belongs to the group of sorosilicates, i.e. it is based on [Si2O7]-units and crystallizes in the thortveitite (Sc2Si2O7) structure type. The Tm3+ cation (site symmetry .2.) occupies a distorted octa­hedral site, with Tm—O bond lengths in the range 2.217 (4)–2.289 (4) Å. Each of the octa­hedra shares three of its edges with adjacent [TmO6] groups, resulting in the formation of layers parallel to (001). The individual [SiO4] tetra­hedra are more regular, i.e. the differences between the bond lengths between Si and the bridging and non-bridging O atoms are not very pronounced. The layers containing the octa­hedra and the sheets containing the [Si2O7] groups (point group symmetry 2/m) form an alternating sequence. Linkage is provided by sharing common oxygen vertices.

Related literature

For applications of oxosilicates containing trivalent rare earth elements (REE), see: Kitai (2008[Kitai, A. (2008). Luminescent Materials and Applications, p. 298. London: John Wiley & Sons.]); Piccinelli et al. (2009[Piccinelli, F., Speghini, A., Mariotto, G., Bovo, L. & Bettinelli, M. (2009). J. Rare Earth, 27, 555-559.]); Qiao et al. (2014[Qiao, J., Zhang, J., Zhang, X., Hao, Z., Liu, Y. & Luo, Y. (2014). Dalton Trans. 43, 4146-4150.]); Luo et al. (2012[Luo, Y. Y., Jo, D. S., Senthil, L., Tezuka, S., Kakihana, M., Toda, K., Masaki, T. & Yoon, D. H. (2012). J. Solid State Chem. 189, 68-74.]); Streit et al. (2013[Streit, H. C., Kramer, J., Suta, M. & Wickleder, C. (2013). Materials, 6, 3079-3093.]); Han et al. (2006[Han, X., Lin, J., Li, Z., Qi, X., Li, M. & Wang, X. (2006). J. Rare Earth, 24, 108-110.]); Sun et al. (2012[Sun, Z., Wang, M., Song, X. & Jiang, Z. (2012). J. Rare Earth, 31, 957-961.]). For structures isotypic with that of the title compound, see: Zachariasen (1930[Zachariasen, W. H. (1930). Z. Kristallogr. 73, 1-6.]); Smolin et al. (1973[Smolin, Yu. I., Shepelev, Yu. F. & Titov, A. P. (1973). Sov. Phys. Crystallogr. 17, 749-750.]); Christensen (1994[Christensen, A. N. (1994). Z. Kristallogr. 209, 7-13.]); Redhammer & Roth (2003[Redhammer, G. J. & Roth, G. (2003). Acta Cryst. C59, i103-i106.]). For polymorphic forms of Tm2Si2O7 and other structure types adopted by (REE)2Si2O7 compounds, see: Bocquillon et al. (1977[Bocquillon, C., Chateau, C., Loriers, C. & Loriers, L. (1977). J. Solid State Chem. 20, 135-141.]); Hartenbach et al. (2003[Hartenbach, I., Lissner, F. & Schleid, T. (2003). Z. Naturforsch. Teil B, 58, 925-927.]); Felsche (1973[Felsche, J. (1973). Struct. Bond. 13, 99-197.]); Fleet & Liu (2005[Fleet, M. E. & Liu, X. (2005). J. Solid State Chem. 178, 3275-3283.]); Shannon & Prewitt (1970[Shannon, R. D. & Prewitt, C. T. (1970). J. Solid State Chem. 2, 199-202.]). For discussions of the [Si2O7]-unit with a linear bridging angle, see: Baur (1980[Baur, W. H. (1980). Acta Cryst. B36, 2198-2202.]); Bianchi et al. (1988[Bianchi, R., Pilati, T., Diella, V., Gramaccioli, C. M. & Mannucci, G. (1988). Am. Mineral. 73, 601-607.]); Cruickshank et al. (1962[Cruickshank, D. W. J., Lynton, H. & Barclay, G. A. (1962). Acta Cryst. 15, 491-498.]); Kimata et al. (1998[Kimata, M., Saito, S., Matsui, T., Shimizu, M. & Nishida, N. (1998). Neues Jarhb. Mineral. Monatsh. 1998, 361-372.]); Liebau (1961[Liebau, F. (1961). Acta Cryst. 14, 1103-1109.]). For general aspects on the crystal chemistry of silicates, see: Liebau (1985[Liebau, F. (1985). Structural Chemistry of Silicates, p. 347. Berlin, Heidelberg, New York, Tokyo: Springer.]). For definition of distortion parameters, see: Robinson et al. (1971[Robinson, K., Gibbs, G. V. & Ribbe, P. H. (1971). Science, 172, 567-570.]). For bond-valence analysis, see: Brown (2002[Brown, I. D. (2002). The Chemical Bond in Inorganic Chemistry: The Bond Valence Model, p. 292. Oxford University Press.]). For definition and calculation of similarity descriptors, see: Tasci et al. (2012[Tasci, E. S., de la Flor, G., Orobengoa, D., Capillas, C., Perez-Mato, J. M. & Aroyo, M. I. (2012). EPJ Web of Conferences, 22, 00009.]); Bergerhoff et al. (1999[Bergerhoff, G., Berndt, M., Brandenburg, K. & Degen, T. (1999). Acta Cryst. B55, 147-156.]). For ionic radii, see: Shannon (1976[Shannon, R. D. (1976). Acta Cryst. A32, 751-767.]). For the Inorganic Crystal Structure Database, see: ICSD (2014[ICSD (2014). Inorganic Crystal Structure Database. FIZ-Karlsruhe, Germany, and the National Institute of Standards and Technology (NIST), USA. http://www.fiz-karlsruhe.de/icsd_home.html]).

Experimental

Crystal data
  • Tm2Si2O7

  • Mr = 506.04

  • Monoclinic, C 2/m

  • a = 6.8205 (14) Å

  • b = 8.9062 (18) Å

  • c = 4.6937 (11) Å

  • β = 101.78 (2)°

  • V = 279.11 (10) Å3

  • Z = 2

  • Mo Kα radiation

  • μ = 31.99 mm−1

  • T = 293 K

  • 0.05 × 0.03 × 0.01 mm

Data collection
  • Agilent Xcalibur (Ruby, Gemini ultra) diffractometer

  • Absorption correction: multi-scan (CrysAlis PRO; Agilent, 2014[Agilent (2014). CrysAlis PRO. Agilent Technologies, Yarnton, England.]) Tmin = 0.231, Tmax = 1

  • 894 measured reflections

  • 340 independent reflections

  • 330 reflections with I > 2σ(I)

  • Rint = 0.020

Refinement
  • R[F2 > 2σ(F2)] = 0.018

  • wR(F2) = 0.045

  • S = 1.15

  • 340 reflections

  • 32 parameters

  • Δρmax = 1.62 e Å−3

  • Δρmin = −1.42 e Å−3

Data collection: CrysAlis PRO (Agilent, 2014[Agilent (2014). CrysAlis PRO. Agilent Technologies, Yarnton, England.]); cell refinement: CrysAlis PRO; data reduction: CrysAlis PRO; program(s) used to solve structure: SIR2002 (Burla et al., 2003[Burla, M. C., Camalli, M., Carrozzini, B., Cascarano, G. L., Giacovazzo, C., Polidori, G. & Spagna, R. (2003). J. Appl. Cryst. 36, 1103.]); program(s) used to refine structure: SHELXL97 (Sheldrick, 2008[Sheldrick, G. M. (2008). Acta Cryst. A64, 112-122.]); molecular graphics: ATOMS for Windows (Dowty, 2011[Dowty, E. (2011). ATOMS for Windows. Shape Software, Kingsport, USA.]); software used to prepare material for publication: publCIF (Westrip, 2010[Westrip, S. P. (2010). J. Appl. Cryst. 43, 920-925.]) and WinGX (Farrugia, 2012[Farrugia, L. J. (2012). J. Appl. Cryst. 45, 849-854.]).

Supporting information


Comment top

Oxosilicates that contain trivalent rare earth elements have been studied frequently because of their potential usage in the field of luminescense including applications in devices and circuits for electronic, optoelectronic as well as communication industries (Kitai, 2008; Piccinelli et al., 2009; Qiao et al., 2014; Luo et al., 2012; Streit et al., 2013; Han et al., 2006; Sun et al., 2012).

In the course of an ongoing project on the synthesis of alkali-REE-silicates (REE is a rare earth element), single-crystals of Tm2Si2O7 have been obtained and structurally characterized. Synthetic Tm2Si2O7 is isotypic with thortveitite (Sc2Si2O7), a rare scandium silicate mineral (Zachariasen, 1930; Smolin et al., 1973). The compound is a sorosilicate and contains isolated [Si2O7]-groups. The bridging oxygen atom of the dimer resides on a centre of symmetry, resulting in a linear Si—O—Si angle. The conformation of the group is staggered with a dihedral angle (or azimuth) of 60° (Fig. 1). In the past, the question whether or not Si—O—Si angles can exhibit a value of 180° has been discussed controversially and, actually, the thortveitite structure-type played an important role in this debate (Liebau, 1961; Cruickshank, et al., 1962). However, a critical analysis of published data performed by Baur (1980) revealed that linear Si—O—Si bridging angles do exist and cannot be attributed to incorrect space group assignments. To date, it is generally accepted that a description of the thortveitite structure-type in the centrosymmetric space group C2/m (implying a linear Si—O—Si angle) is correct (Bianchi et al., 1988; Kimata et al., 1998). The present structure determination of Tm2Si2O7 also confirms this model. The spread in the Si—O and O—Si—O angles is not very pronounced and the values are in the expected limits for silicates (Liebau, 1985). Numerically, the degree of distortion can be expressed by the quadratic elongation QE and the angle variance AV (Robinson et al., 1971). The values of these distortion parameters for a single [SiO4]-tetrahedron are very small: 1.001 (for QE) and 4.95 (for AV), respectively. The Tm3+ cations are octahedrally coordinated by O atoms (Fig. 2), with Tm—O bond lengths in the range 2.217 (4) – 2.289 (4) Å and an average of 2.247 Å. The mean value compares well with those observed for the thortveitite representatives of the directly neighbouring REE Yb (<Yb—O>=2.240 Å) and Er (<Er—O>=2.253 Å) (Christensen, 1994). The differences can be attributed to the increasing ionic radii of the trivalent cations in the series Yb3+ - Tm3+ - Er3+ (Shannon, 1976). The octahedra show a distortion with moderate QE-values (1.061) and very high values for the angle variance. The high AV value of 219.7 seems to be a characteristic feature of the thortveitite structure-type and has been also observed for other members of this family (Redhammer & Roth, 2003). Each of the octahedra shares three of its edges with adjacent [TmO6]-groups resulting in the formation of layers parallel to (001). These pseudo-hexagonal sheets (Fig. 3) are similar to the layers in dioctahedral micas. The above-mentioned pronounced angular distortions can be rationalized by a combination of (i) a shortening of the common edges of adjacent octahedra (in order to reduce the repulsive interactions between adjoining Tm3+ cations) and (ii) a widening of the corresponding opposite unshared O—O edges. Successive layers containing octahedra are linked by the [Si2O7]-groups in such a way that each of both tetrahedra shares two of its corners with two different octahedra from the same and one corner with an octahedron from the other surrounding layer.

Bond valence sum calculations using the parameter sets for the Tm—O and Si—O bonds given by Brown (2002) resulted in the following values (in v.u.) for the cation-anion interactions up to 3.4 Å: Tm: 3.09, Si: 4.01, O1: 2.12, O2: 1.99 and O3: 1.99.

As mentioned above, the present structure is isotypic with that of thortveitite. For the calculation of several quantitative descriptors for the characterization of the degree of similarity between the crystal structures of Tm2Si2O7 and Sc2Si2O7, the program COMPSTRU (Tasci et al., 2012) was employed. For the given two structures, the degree of lattice distortion (S), i.e. the spontaneous strain obtained from the eigenvalues of the finite Lagrangian strain tensor calculated in a Cartesian reference system, has a value of (S) = 0.0222. After application of an origin shift of p = (0, 0, 1/2) the structure of Tm2Si2O7 was transformed to the most similar configuration of Sc2Si2O7. The calculations revealed the following atomic displacements (in Å) between the corresponding atoms in Sc2Si2O7 (first entry) and Tm2Si2O7 (second entry): Sc—Tm: 0.025; Si—Si: 0.043; O1—O2: 0.000; O2—O1: 0.083; O3—O3: 0.070 i.e. the maximum displacement is lower than 0.10 Å. The measure of similarity (Δ) as defined by Bergerhoff et al. (1999) has a value of 0.059.

Since the beginning of the 1970ies a large number of different structure types have been described for rare earth element silicates with composition (REE)2Si2O7 (Felsche, 1973). To date, at least twelve different forms (A—I, K, L and X) have to be distinguished (Fleet & Liu, 2005). Tm2Si2O7, for example, exhibits a high degree of polymorphism where five different modifications can be realised. The synthesis of polycrystalline Tm2Si2O7 adopting the thortveitite- or C-type has been described by Bocquillon et al. (1977) in the temperature range between 1473 and 1673 K. However, the stability field of C-type Tm2Si2O7 extends to higher pressures as well: synthesis runs performed at 65 kbar/1773 K (Shannon & Prewitt, 1970) as well as 10 kbar/973 K and 18 kbar/973 K (Bocquillon et al., 1977) also resulted in the formation of the C-phase. Other high-pressure modifications of Tm2Si2O7 crystallize in the B–, D–, X– and L-types (Fleet & Liu, 2005; Shannon & Prewitt, 1970). The B-type, however, has been also prepared at ambient pressure and 1173 K (Hartenbach et al., 2003). In summary, one can say that more than forty years after the first systematic investigations to chart the p,T-behaviour of Tm2Si2O7, there are still open questions. The new flux synthesis route using lithium fluoride as a mineralizer offers the possibility to grow large single-crystals suited for in situ X-ray diffraction or Raman spectroscopic high-pressure studies in diamond anvil cells.

Related literature top

For applications of oxosilicates containing trivalent rare earth elements (REE), see: Kitai (2008); Piccinelli et al. (2009); Qiao et al. (2014); Luo et al. (2012); Streit et al. (2013); Han et al. (2006); Sun et al. (2012). For structures isotypic with that of the title compound, see: Zachariasen (1930); Smolin et al. (1973); Christensen (1994); Redhammer & Roth (2003). For polymorphic forms of Tm2Si2O7 and other structure types adopted by (REE)2Si2O7 compounds, see: Bocquillon et al. (1977); Hartenbach et al. (2003); Felsche (1973); Fleet & Liu (2005); Shannon & Prewitt (1970). For discussions of the [Si2O7]-unit with a linear bridging angle, see: Baur (1980); Bianchi et al. (1988); Cruickshank et al. (1962); Kimata et al. (1998); Liebau (1961). For general aspects on the crystal chemistry of silicates, see: Liebau (1985). For definition of distortion parameters, see: Robinson et al. (1971). For bond-valence analysis, see: Brown (2002). For definition and calculation of similarity descriptors, see: Tasci et al. (2012); Bergerhoff et al. (1999). For ionic radii, see: Shannon (1976). For the Inorganic Crystal Structure Database, see: ICSD (2014).

Experimental top

Starting materials for the flux growth experiments were dried reagent grade Tm2O3, SiO2 and LiF. Due to the pronounced hygroscopicity of the alkali fluoride, sample preparation was performed in a glove bag under nitrogen atmosphere. 0.1 g of the nutrient consisting of a mixture of Tm2O3:SiO2 in the molar ratio 1:4 was homogenized in an agate mortar with 0.1 g LiF. Subsequently, the educts were loaded into a platinum tube with an outer diameter of 3 mm and with 20 mm length. After sealing, the tube and its content were heated in a resistance furnace from 373 K to 1373 K with a rate of 50 K/h and isothermed for 2 h at the target temperature. The sample was cooled down to 1073 K with a rate of 5 K/h and, finally, the temperature was reduced to 373 K with a rate of 100 K/h. Removal of the flux with water left a residue of transparent, colorless, optically biaxial and highly birefringent crystals up to 500 µm in size. One of the optically biaxial crystals showing sharp extinction when observed between crossed polarizers was selected for further structural studies and mounted on the tip of a glass fiber using finger nail hardener as glue.

Refinement top

Similar sets of lattice parameters could be found in the recent WEB-based version of the Inorganic Crystal Structure Database (ICSD, 2014) for the chemically closely related thortveitite-type materials with composition (REE)2Si2O7 pointing to an isostructural relationship, which was confirmed by the subsequent structure analysis by direct methods. For structure determination a data set corresponding to a hemisphere of reciprocal space was collected.

Structure description top

Oxosilicates that contain trivalent rare earth elements have been studied frequently because of their potential usage in the field of luminescense including applications in devices and circuits for electronic, optoelectronic as well as communication industries (Kitai, 2008; Piccinelli et al., 2009; Qiao et al., 2014; Luo et al., 2012; Streit et al., 2013; Han et al., 2006; Sun et al., 2012).

In the course of an ongoing project on the synthesis of alkali-REE-silicates (REE is a rare earth element), single-crystals of Tm2Si2O7 have been obtained and structurally characterized. Synthetic Tm2Si2O7 is isotypic with thortveitite (Sc2Si2O7), a rare scandium silicate mineral (Zachariasen, 1930; Smolin et al., 1973). The compound is a sorosilicate and contains isolated [Si2O7]-groups. The bridging oxygen atom of the dimer resides on a centre of symmetry, resulting in a linear Si—O—Si angle. The conformation of the group is staggered with a dihedral angle (or azimuth) of 60° (Fig. 1). In the past, the question whether or not Si—O—Si angles can exhibit a value of 180° has been discussed controversially and, actually, the thortveitite structure-type played an important role in this debate (Liebau, 1961; Cruickshank, et al., 1962). However, a critical analysis of published data performed by Baur (1980) revealed that linear Si—O—Si bridging angles do exist and cannot be attributed to incorrect space group assignments. To date, it is generally accepted that a description of the thortveitite structure-type in the centrosymmetric space group C2/m (implying a linear Si—O—Si angle) is correct (Bianchi et al., 1988; Kimata et al., 1998). The present structure determination of Tm2Si2O7 also confirms this model. The spread in the Si—O and O—Si—O angles is not very pronounced and the values are in the expected limits for silicates (Liebau, 1985). Numerically, the degree of distortion can be expressed by the quadratic elongation QE and the angle variance AV (Robinson et al., 1971). The values of these distortion parameters for a single [SiO4]-tetrahedron are very small: 1.001 (for QE) and 4.95 (for AV), respectively. The Tm3+ cations are octahedrally coordinated by O atoms (Fig. 2), with Tm—O bond lengths in the range 2.217 (4) – 2.289 (4) Å and an average of 2.247 Å. The mean value compares well with those observed for the thortveitite representatives of the directly neighbouring REE Yb (<Yb—O>=2.240 Å) and Er (<Er—O>=2.253 Å) (Christensen, 1994). The differences can be attributed to the increasing ionic radii of the trivalent cations in the series Yb3+ - Tm3+ - Er3+ (Shannon, 1976). The octahedra show a distortion with moderate QE-values (1.061) and very high values for the angle variance. The high AV value of 219.7 seems to be a characteristic feature of the thortveitite structure-type and has been also observed for other members of this family (Redhammer & Roth, 2003). Each of the octahedra shares three of its edges with adjacent [TmO6]-groups resulting in the formation of layers parallel to (001). These pseudo-hexagonal sheets (Fig. 3) are similar to the layers in dioctahedral micas. The above-mentioned pronounced angular distortions can be rationalized by a combination of (i) a shortening of the common edges of adjacent octahedra (in order to reduce the repulsive interactions between adjoining Tm3+ cations) and (ii) a widening of the corresponding opposite unshared O—O edges. Successive layers containing octahedra are linked by the [Si2O7]-groups in such a way that each of both tetrahedra shares two of its corners with two different octahedra from the same and one corner with an octahedron from the other surrounding layer.

Bond valence sum calculations using the parameter sets for the Tm—O and Si—O bonds given by Brown (2002) resulted in the following values (in v.u.) for the cation-anion interactions up to 3.4 Å: Tm: 3.09, Si: 4.01, O1: 2.12, O2: 1.99 and O3: 1.99.

As mentioned above, the present structure is isotypic with that of thortveitite. For the calculation of several quantitative descriptors for the characterization of the degree of similarity between the crystal structures of Tm2Si2O7 and Sc2Si2O7, the program COMPSTRU (Tasci et al., 2012) was employed. For the given two structures, the degree of lattice distortion (S), i.e. the spontaneous strain obtained from the eigenvalues of the finite Lagrangian strain tensor calculated in a Cartesian reference system, has a value of (S) = 0.0222. After application of an origin shift of p = (0, 0, 1/2) the structure of Tm2Si2O7 was transformed to the most similar configuration of Sc2Si2O7. The calculations revealed the following atomic displacements (in Å) between the corresponding atoms in Sc2Si2O7 (first entry) and Tm2Si2O7 (second entry): Sc—Tm: 0.025; Si—Si: 0.043; O1—O2: 0.000; O2—O1: 0.083; O3—O3: 0.070 i.e. the maximum displacement is lower than 0.10 Å. The measure of similarity (Δ) as defined by Bergerhoff et al. (1999) has a value of 0.059.

Since the beginning of the 1970ies a large number of different structure types have been described for rare earth element silicates with composition (REE)2Si2O7 (Felsche, 1973). To date, at least twelve different forms (A—I, K, L and X) have to be distinguished (Fleet & Liu, 2005). Tm2Si2O7, for example, exhibits a high degree of polymorphism where five different modifications can be realised. The synthesis of polycrystalline Tm2Si2O7 adopting the thortveitite- or C-type has been described by Bocquillon et al. (1977) in the temperature range between 1473 and 1673 K. However, the stability field of C-type Tm2Si2O7 extends to higher pressures as well: synthesis runs performed at 65 kbar/1773 K (Shannon & Prewitt, 1970) as well as 10 kbar/973 K and 18 kbar/973 K (Bocquillon et al., 1977) also resulted in the formation of the C-phase. Other high-pressure modifications of Tm2Si2O7 crystallize in the B–, D–, X– and L-types (Fleet & Liu, 2005; Shannon & Prewitt, 1970). The B-type, however, has been also prepared at ambient pressure and 1173 K (Hartenbach et al., 2003). In summary, one can say that more than forty years after the first systematic investigations to chart the p,T-behaviour of Tm2Si2O7, there are still open questions. The new flux synthesis route using lithium fluoride as a mineralizer offers the possibility to grow large single-crystals suited for in situ X-ray diffraction or Raman spectroscopic high-pressure studies in diamond anvil cells.

For applications of oxosilicates containing trivalent rare earth elements (REE), see: Kitai (2008); Piccinelli et al. (2009); Qiao et al. (2014); Luo et al. (2012); Streit et al. (2013); Han et al. (2006); Sun et al. (2012). For structures isotypic with that of the title compound, see: Zachariasen (1930); Smolin et al. (1973); Christensen (1994); Redhammer & Roth (2003). For polymorphic forms of Tm2Si2O7 and other structure types adopted by (REE)2Si2O7 compounds, see: Bocquillon et al. (1977); Hartenbach et al. (2003); Felsche (1973); Fleet & Liu (2005); Shannon & Prewitt (1970). For discussions of the [Si2O7]-unit with a linear bridging angle, see: Baur (1980); Bianchi et al. (1988); Cruickshank et al. (1962); Kimata et al. (1998); Liebau (1961). For general aspects on the crystal chemistry of silicates, see: Liebau (1985). For definition of distortion parameters, see: Robinson et al. (1971). For bond-valence analysis, see: Brown (2002). For definition and calculation of similarity descriptors, see: Tasci et al. (2012); Bergerhoff et al. (1999). For ionic radii, see: Shannon (1976). For the Inorganic Crystal Structure Database, see: ICSD (2014).

Computing details top

Data collection: CrysAlis PRO (Agilent, 2014); cell refinement: CrysAlis PRO (Agilent, 2014); data reduction: CrysAlis PRO (Agilent, 2014); program(s) used to solve structure: SIR2002 (Burla et al., 2003); program(s) used to refine structure: SHELXL97 (Sheldrick, 2008); molecular graphics: ATOMS for Windows (Dowty, 2011); software used to prepare material for publication: publCIF (Westrip, 2010) and WinGX (Farrugia, 2012).

Figures top
[Figure 1] Fig. 1. Representation of a single [Si2O7]-unit. Ellipsoids are drawn at the 60% probability level. [Symmetry codes: (i) x, y, -1 + z (ii) -x, y, -z (iii) -x, y, 1 - z (iv) x, -y, -1 + z].
[Figure 2] Fig. 2. Representation of the coordination around the trivalent Tm ion. Ellipsoids are drawn at the 60% probability level. [Symmetry codes: (i) 1 - x, y, 1 - z (ii) 1/2 + x, 1/2 - y, z (iii) 1/2 - x, 1/2 - y, 1 - z].
[Figure 3] Fig. 3. Single layer of edge-sharing octahedra and one of the two adjacent sheets containing [Si2O7]-units in a projection parallel to [001]. Red, grey and blue spheres represent oxygen, silicon and thulium ions.
Dithulium disilicate top
Crystal data top
Tm2Si2O7F(000) = 444
Mr = 506.04Dx = 6.021 Mg m3
Monoclinic, C2/mMo Kα radiation, λ = 0.71073 Å
Hall symbol: -C 2yCell parameters from 762 reflections
a = 6.8205 (14) Åθ = 3.8–29.3°
b = 8.9062 (18) ŵ = 31.99 mm1
c = 4.6937 (11) ÅT = 293 K
β = 101.78 (2)°Platy fragment, colourless
V = 279.11 (10) Å30.05 × 0.03 × 0.01 mm
Z = 2
Data collection top
Agilent Xcalibur (Ruby, Gemini ultra)
diffractometer
340 independent reflections
Radiation source: Enhance (Mo) X-ray Source330 reflections with I > 2σ(I)
Graphite monochromatorRint = 0.020
Detector resolution: 10.3575 pixels mm-1θmax = 27.6°, θmin = 3.8°
ω scansh = 88
Absorption correction: multi-scan
(CrysAlis PRO; Agilent, 2014)
k = 1111
Tmin = 0.231, Tmax = 1l = 46
894 measured reflections
Refinement top
Refinement on F2Primary atom site location: structure-invariant direct methods
Least-squares matrix: fullSecondary atom site location: difference Fourier map
R[F2 > 2σ(F2)] = 0.018 w = 1/[σ2(Fo2) + (0.0265P)2]
where P = (Fo2 + 2Fc2)/3
wR(F2) = 0.045(Δ/σ)max < 0.001
S = 1.15Δρmax = 1.62 e Å3
340 reflectionsΔρmin = 1.42 e Å3
32 parametersExtinction correction: SHELXL97 (Sheldrick, 2008), Fc*=kFc[1+0.001xFc2λ3/sin(2θ)]-1/4
0 restraintsExtinction coefficient: 0.0072 (6)
Crystal data top
Tm2Si2O7V = 279.11 (10) Å3
Mr = 506.04Z = 2
Monoclinic, C2/mMo Kα radiation
a = 6.8205 (14) ŵ = 31.99 mm1
b = 8.9062 (18) ÅT = 293 K
c = 4.6937 (11) Å0.05 × 0.03 × 0.01 mm
β = 101.78 (2)°
Data collection top
Agilent Xcalibur (Ruby, Gemini ultra)
diffractometer
340 independent reflections
Absorption correction: multi-scan
(CrysAlis PRO; Agilent, 2014)
330 reflections with I > 2σ(I)
Tmin = 0.231, Tmax = 1Rint = 0.020
894 measured reflections
Refinement top
R[F2 > 2σ(F2)] = 0.01832 parameters
wR(F2) = 0.0450 restraints
S = 1.15Δρmax = 1.62 e Å3
340 reflectionsΔρmin = 1.42 e Å3
Special details top

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

Refinement. Refinement of F2 against ALL reflections. The weighted R-factor wR and goodness of fit S are based on F2, conventional R-factors R are based on F, with F set to zero for negative F2. The threshold expression of F2 > σ(F2) is used only for calculating R-factors(gt) etc. and is not relevant to the choice of reflections for refinement. R-factors based on F2 are statistically about twice as large as those based on F, and R-factors based on ALL data will be even larger.

Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzUiso*/Ueq
Tm0.50.19345 (4)0.50.0045 (2)
Si0.2186 (3)00.9130 (5)0.0044 (5)
O10.3804 (9)00.2130 (12)0.0069 (12)
O20000.0129 (19)
O30.2357 (6)0.1505 (5)0.7213 (9)0.0073 (8)
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Tm0.0032 (3)0.0043 (3)0.0058 (3)00.00056 (15)0
Si0.0049 (11)0.0043 (11)0.0042 (11)00.0013 (9)0
O10.006 (3)0.007 (3)0.005 (3)00.004 (2)0
O20.008 (4)0.019 (5)0.011 (5)00.001 (4)0
O30.006 (2)0.007 (2)0.009 (2)0.0042 (17)0.0026 (17)0.0045 (17)
Geometric parameters (Å, º) top
Tm—O3i2.217 (4)Si—O3vi1.632 (4)
Tm—O3ii2.217 (4)Si—O31.632 (4)
Tm—O12.236 (4)O1—Sivii1.602 (6)
Tm—O1iii2.236 (4)O1—Tmiii2.236 (4)
Tm—O32.289 (4)O2—Siviii1.624 (2)
Tm—O3iv2.289 (4)O2—Sivii1.624 (2)
Si—O1v1.602 (6)O3—Tmii2.217 (4)
Si—O2v1.624 (2)
O3i—Tm—O3ii102.3 (2)O3—Tm—O3iv160.7 (2)
O3i—Tm—O1154.9 (2)O1v—Si—O2v106.4 (2)
O3ii—Tm—O193.47 (17)O1v—Si—O3vi111.8 (2)
O3i—Tm—O1iii93.47 (17)O2v—Si—O3vi108.14 (18)
O3ii—Tm—O1iii154.9 (2)O1v—Si—O3111.8 (2)
O1—Tm—O1iii79.2 (2)O2v—Si—O3108.14 (18)
O3i—Tm—O3117.09 (17)O3vi—Si—O3110.4 (3)
O3ii—Tm—O375.78 (17)Sivii—O1—Tm129.10 (12)
O1—Tm—O385.4 (2)Sivii—O1—Tmiii129.10 (12)
O1iii—Tm—O379.74 (18)Tm—O1—Tmiii100.8 (2)
O3i—Tm—O3iv75.78 (17)Siviii—O2—Sivii180.00 (14)
O3ii—Tm—O3iv117.09 (17)Si—O3—Tmii130.3 (3)
O1—Tm—O3iv79.74 (18)Si—O3—Tm122.6 (2)
O1iii—Tm—O3iv85.43 (19)Tmii—O3—Tm104.22 (17)
Symmetry codes: (i) x+1/2, y+1/2, z; (ii) x+1/2, y+1/2, z+1; (iii) x+1, y, z+1; (iv) x+1, y, z+1; (v) x, y, z+1; (vi) x, y, z; (vii) x, y, z1; (viii) x, y, z+1.

Experimental details

Crystal data
Chemical formulaTm2Si2O7
Mr506.04
Crystal system, space groupMonoclinic, C2/m
Temperature (K)293
a, b, c (Å)6.8205 (14), 8.9062 (18), 4.6937 (11)
β (°) 101.78 (2)
V3)279.11 (10)
Z2
Radiation typeMo Kα
µ (mm1)31.99
Crystal size (mm)0.05 × 0.03 × 0.01
Data collection
DiffractometerAgilent Xcalibur (Ruby, Gemini ultra)
Absorption correctionMulti-scan
(CrysAlis PRO; Agilent, 2014)
Tmin, Tmax0.231, 1
No. of measured, independent and
observed [I > 2σ(I)] reflections
894, 340, 330
Rint0.020
(sin θ/λ)max1)0.653
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.018, 0.045, 1.15
No. of reflections340
No. of parameters32
Δρmax, Δρmin (e Å3)1.62, 1.42

Computer programs: CrysAlis PRO (Agilent, 2014), SIR2002 (Burla et al., 2003), SHELXL97 (Sheldrick, 2008), ATOMS for Windows (Dowty, 2011), publCIF (Westrip, 2010) and WinGX (Farrugia, 2012).

 

References

First citationAgilent (2014). CrysAlis PRO. Agilent Technologies, Yarnton, England.  Google Scholar
First citationBaur, W. H. (1980). Acta Cryst. B36, 2198–2202.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationBergerhoff, G., Berndt, M., Brandenburg, K. & Degen, T. (1999). Acta Cryst. B55, 147–156.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationBianchi, R., Pilati, T., Diella, V., Gramaccioli, C. M. & Mannucci, G. (1988). Am. Mineral. 73, 601–607.  CAS Google Scholar
First citationBocquillon, C., Chateau, C., Loriers, C. & Loriers, L. (1977). J. Solid State Chem. 20, 135–141.  CrossRef CAS Web of Science Google Scholar
First citationBrown, I. D. (2002). The Chemical Bond in Inorganic Chemistry: The Bond Valence Model, p. 292. Oxford University Press.  Google Scholar
First citationBurla, M. C., Camalli, M., Carrozzini, B., Cascarano, G. L., Giacovazzo, C., Polidori, G. & Spagna, R. (2003). J. Appl. Cryst. 36, 1103.  CrossRef IUCr Journals Google Scholar
First citationChristensen, A. N. (1994). Z. Kristallogr. 209, 7–13.  CrossRef CAS Web of Science Google Scholar
First citationCruickshank, D. W. J., Lynton, H. & Barclay, G. A. (1962). Acta Cryst. 15, 491–498.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationDowty, E. (2011). ATOMS for Windows. Shape Software, Kingsport, USA.  Google Scholar
First citationFarrugia, L. J. (2012). J. Appl. Cryst. 45, 849–854.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationFelsche, J. (1973). Struct. Bond. 13, 99–197.  CrossRef CAS Google Scholar
First citationFleet, M. E. & Liu, X. (2005). J. Solid State Chem. 178, 3275–3283.  Web of Science CrossRef CAS Google Scholar
First citationHan, X., Lin, J., Li, Z., Qi, X., Li, M. & Wang, X. (2006). J. Rare Earth, 24, 108–110.  Google Scholar
First citationHartenbach, I., Lissner, F. & Schleid, T. (2003). Z. Naturforsch. Teil B, 58, 925–927.  CAS Google Scholar
First citationICSD (2014). Inorganic Crystal Structure Database. FIZ-Karlsruhe, Germany, and the National Institute of Standards and Technology (NIST), USA. http://www.fiz-karlsruhe.de/icsd_home.html  Google Scholar
First citationKimata, M., Saito, S., Matsui, T., Shimizu, M. & Nishida, N. (1998). Neues Jarhb. Mineral. Monatsh. 1998, 361–372.  Google Scholar
First citationKitai, A. (2008). Luminescent Materials and Applications, p. 298. London: John Wiley & Sons.  Google Scholar
First citationLiebau, F. (1961). Acta Cryst. 14, 1103–1109.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationLiebau, F. (1985). Structural Chemistry of Silicates, p. 347. Berlin, Heidelberg, New York, Tokyo: Springer.  Google Scholar
First citationLuo, Y. Y., Jo, D. S., Senthil, L., Tezuka, S., Kakihana, M., Toda, K., Masaki, T. & Yoon, D. H. (2012). J. Solid State Chem. 189, 68–74.  Web of Science CrossRef CAS Google Scholar
First citationPiccinelli, F., Speghini, A., Mariotto, G., Bovo, L. & Bettinelli, M. (2009). J. Rare Earth, 27, 555–559.  Web of Science CrossRef Google Scholar
First citationQiao, J., Zhang, J., Zhang, X., Hao, Z., Liu, Y. & Luo, Y. (2014). Dalton Trans. 43, 4146–4150.  Web of Science CrossRef CAS PubMed Google Scholar
First citationRedhammer, G. J. & Roth, G. (2003). Acta Cryst. C59, i103–i106.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationRobinson, K., Gibbs, G. V. & Ribbe, P. H. (1971). Science, 172, 567–570.  CrossRef PubMed CAS Web of Science Google Scholar
First citationShannon, R. D. (1976). Acta Cryst. A32, 751–767.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationShannon, R. D. & Prewitt, C. T. (1970). J. Solid State Chem. 2, 199–202.  CrossRef CAS Google Scholar
First citationSheldrick, G. M. (2008). Acta Cryst. A64, 112–122.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationSmolin, Yu. I., Shepelev, Yu. F. & Titov, A. P. (1973). Sov. Phys. Crystallogr. 17, 749–750.  Google Scholar
First citationStreit, H. C., Kramer, J., Suta, M. & Wickleder, C. (2013). Materials, 6, 3079-3093.  Web of Science CrossRef CAS Google Scholar
First citationSun, Z., Wang, M., Song, X. & Jiang, Z. (2012). J. Rare Earth, 31, 957–961.  CrossRef Google Scholar
First citationTasci, E. S., de la Flor, G., Orobengoa, D., Capillas, C., Perez-Mato, J. M. & Aroyo, M. I. (2012). EPJ Web of Conferences, 22, 00009.  CrossRef Google Scholar
First citationWestrip, S. P. (2010). J. Appl. Cryst. 43, 920–925.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationZachariasen, W. H. (1930). Z. Kristallogr. 73, 1–6.  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