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

Journal logoCRYSTALLOGRAPHIC
COMMUNICATIONS
ISSN: 2056-9890

Ge0.57Ti0.43O2: a new high-pressure material with rutile-type crystal structure

CROSSMARK_Color_square_no_text.svg

aSandvik Hyperion, 6325 Huntley Road, Worthington, OH 43085, USA, bEyring Materials Center, Arizona State University, Tempe, AZ 85287-1604, USA, and cSchool of Molecular Sciences, PSD-102 MS-871604, Arizona State University, Tempe, AZ 85287-1604, USA
*Correspondence e-mail: emil.stoyanov@sandvik.com

Edited by M. Weil, Vienna University of Technology, Austria (Received 12 June 2018; accepted 19 June 2018; online 26 June 2018)

Single crystals of a GeO2–TiO2 solid solution with the corresponding composition Ge0.57Ti0.43O2 (germanium titanium tetra­oxide) were obtained by devitrification of germania-titania glass at high pressure and temperature. The new compound crystallizes in the rutile structure type (space group P42/mnm), where Ge and Ti share the same position M (site symmetry m.mm), with occupancy values of 0.57 (3) and 0.43 (3), respectively, and one O-atom position (m.2m). The M site is in a sixfold O-atom coordination and, as in the original TiO2 rutile structure, an elongation of the O—M—O bonds along the c-axis direction of the coordination polyhedron and deviation of the angles from 90° lead to a decrease in the coordination symmetry from octa­hedral to tetra­gonal. The Ge and Ti atoms are fully disordered in the structure, which indicates that the rutile structure is surprisingly pliant given the differing sizes of the two cations.

1. Chemical context

At ambient pressure, the GeO2–TiO2 phase diagram shows the formation of three phases: rutile-type GeO2, stable up to 1323 K, β-quartz-type GeO2, stable above 1323 K and TiO2 in the form of rutile. A metastable α-quartz-type structured GeO2 has also been reported as the result of the cooling of the β-quartz-type structure (Sarver, 1961[Sarver, J. F. (1961). Am. J. Sci. 259, 709-718.]). Additionally, at ambient pressure GeO2 and TiO2 exhibit only limited mutual solubility. The GeO2–TiO2 phase diagram at elevated pressures and temperatures has not been studied in great detail and the mutual solubility of Ge and Ti in the phases stable at these conditions is still largely unknown. GeO2 is dimorphous at ambient atmospheric conditions, represented by both rutile-type and α-quartz-structured phases depending on the temperature, but with increasing pressure the GeO2 rutile becomes more stable, and is the primary phase above two GPa (Micoulaut et al., 2006[Micoulaut, M., Cormier, L. & Henderson, G. S. (2006). J. Phys. Condens. Matter, 18, R753-R784.]). At pressures above 25 GPa, the tetra­gonal rutile-type phase transforms into an ortho­rhom­bic CaCl2-type phase (Haines et al., 2000[Haines, J., Léger, J. M., Chateau, C. & Pereira, A. S. (2000). Phys. Chem. Miner. 27, 575-582.]). TiO2 rutile undergoes two phase transitions under high pressure of up to 12 GPa: rutile-to-α-PbO2-type at around 7 GPa and α-PbO2-to-baddeleyite at 12 GPa (Gerward & Staun Olsen, 1997[Gerward, L. & Staun Olsen, J. (1997). J. Appl. Cryst. 30, 259-264.]). We synthesized the title compound while investigating the GeO2–TiO2 phase diagram at a pressure of 8 GPa at 2028 K by means of the multi-anvil high-pressure technique. Instead of forming Ge-bearing TiO2 and Ti-bearing GeO2, we discovered that the high pressure and temperature conditions led to the formation of a crystalline, single solid-solution material. At temperatures above 1873 K, crystal growth was significant and high-quality single crystals of the solid solution with a composition near TiGeO4 could be obtained.

2. Structural commentary

The crystal structure of Ge0.57Ti0.43O2 corresponds to the TiO2 rutile type (space group P42/mnm). The shared metal site M is in Wyckoff position 2a and is surrounded by six O atoms, thus forming a sixfold coordination polyhedron. 57% of the 2a positions are occupied by Ge and the remaining 43% are occupied by Ti. Each oxygen atom occupies a 4f position and is surrounded by three M sites, forming triangular MO3 groups in the (110) lattice plane (Fig. 1[link]). The structure is represented by chains of edge-sharing MO6 octa­hedra running parallel to the c-axis direction (Fig. 2[link]) and connected to each other by shared corners. Relevant bond lengths and angles are presented in Table 1[link]. In the MO6 coordination polyhedra, the M—O distances in the xy plane are 1.9080 (12) Å, while the M—O distances along the z axis increase to 1.9441 (19) Å. The MM distances are equal to 2.9121 (13) Å. The elongation of O—M—O bonds along the z direction of the MO6 coordination polyhedron and the deviation of the angles from 90° lead to a decrease in point group symmetry from octa­hedral Oh to tetra­gonal D4h. The unit-cell volume of Ge0.57Ti0.43O2 [58.79 (6) Å3] falls in between the rutile-type GeO2 [55.3424 (17) Å3] (Gullikson et al., 2015[Gullikson, A. L., Leinenweber, K., Stoyanov, E., Zhang, H. & Malik, A.-S. (2015). J. Am. Ceram. Soc. 98, 982-989.]) and TiO2 rutile [62.435 Å3] (Howard et al., 1991[Howard, C. J., Sabine, T. M. & Dickson, F. (1991). Acta Cryst. B47, 462-468.]) and indicates a linear relationship between the unit-cell volume and molar fraction of GeO2, adhering to Vegard's Law.

Table 1
Selected geometric parameters (Å, °)

Ge1—O1i 1.9080 (12) Ge1—Ti1ii 2.9121 (13)
Ge1—O1 1.9441 (19)    
       
O1i—Ge1—O1iii 99.48 (8) Ti1ii—Ge1—Ti1vi 180
O1i—Ge1—O1iv 80.52 (8) Ti1vii—O1—Ti1viii 99.48 (8)
O1v—Ge1—O1 180 Ge1vii—O1—Ge1 130.26 (4)
Symmetry codes: (i) [y-{\script{1\over 2}}, -x+{\script{1\over 2}}, -z+{\script{1\over 2}}]; (ii) x, y, z+1; (iii) [y-{\script{1\over 2}}, -x+{\script{1\over 2}}, -z-{\script{1\over 2}}]; (iv) [-y+{\script{1\over 2}}, x-{\script{1\over 2}}, z+{\script{1\over 2}}]; (v) -x, -y, -z; (vi) x, y, z-1; (vii) [-y+{\script{1\over 2}}, x+{\script{1\over 2}}, z+{\script{1\over 2}}]; (viii) [-y+{\script{1\over 2}}, x+{\script{1\over 2}}, z-{\script{1\over 2}}].
[Figure 1]
Figure 1
View of the structure of Ge0.57Ti0.43O2 looking down the c axis. The unit cell is outlined in white. The red ellipsoids represent the oxygen atoms and show the orientation of the displacement ellipsoids for 99% probability. The Ti atom is represented by light blue and the Ge atom by purple, with the percentage occupancy of the M site represented as a pie chart on the atom. MO6 octa­hedra are represented as transparent polyhedra.
[Figure 2]
Figure 2
View of the edge-sharing chain of MO6 octa­hedra in Ge0.57Ti0.43O2. The symbols are the same as in Fig. 1[link].

The somewhat large difference in the ionic radii of the sixfold coordinated Ge4+ and Ti4+ (0.53 and 0.605 Å, respectively; Shannon, 1976[Shannon, R. D. (1976). Acta Cryst. A32, 751-767.]) may be the reason for the limited mutual solubility of Ge and Ti in the rutile structured oxides at ambient pressure. This might explain why the single solid-solution phase is absent in the GeO2–TiO2 system, and why the synthesis of a material with composition near TiGeO4 requires high-pressure and high-temperature conditions. Disordering at high temperatures (significantly above the ambient-pressure melting point) could assist in the stability of the solid solution even with the two different sized cations.

3. Synthesis and crystallization

The title compound was synthesized by using an industrial multi-anvil high-pressure apparatus. The starting material was a GeO2–TiO2 glass produced from the corresponding oxide powders with a molar ratio of 60:40 (Sem-Com Company, Toledo, OH). A Pt foil capsule was loaded with the powdered glass and was subjected to high-pressure/high-temperature (HPHT) conditions of 8 GPa and 2028 K for 30 minutes, followed by cooling for 15 minutes to room temperature and releasing pressure non-isobarically to atmospheric pressure to recover the sample. The temperature was monitored with a W3%Re-W26%Re (C-type) thermocouple. The pressure was estimated by recovering and analyzing SiO2–GeO2 glass that was loaded in the Pt foil capsule and pressed in the same high-pressure cell. Thus, the pressure standard and the GeO2–TiO2 glass were treated at the same conditions. The details on the pressure calibration technique can be found elsewhere (Gullikson et al., 2015[Gullikson, A. L., Leinenweber, K., Stoyanov, E., Zhang, H. & Malik, A.-S. (2015). J. Am. Ceram. Soc. 98, 982-989.]; Leinenweber et al., 2015[Leinenweber, K., Gullikson, A. L., Stoyanov, E. & Malik, A. (2015). J. Solid State Chem. 229, 10-18.]). The applied temperature was sufficient to produce high-quality single crystals with uniform extinction in the optical microscope. A clear colourless tabular-like crystal from the recovered GeO2–TiO2 sample was used for the X-ray crystallographic analysis.

4. Refinement

Crystal data, data collection and structure refinement details are summarized in Table 2[link]. Structure data were standardized according to the STRUCTURE-TIDY program (Gelato & Parthé, 1987[Gelato, L. M. & Parthé, E. (1987). J. Appl. Cryst. 20, 139-143.]).

Table 2
Experimental details

Crystal data
Chemical formula Ge1.14Ti0.86O4
Mr 187.88
Crystal system, space group Tetragonal, P42/mnm
Temperature (K) 298
a, c (Å) 4.493 (2), 2.9121 (13)
V3) 58.79 (6)
Z 1
Radiation type Mo Kα
μ (mm−1) 17.23
Crystal size (mm) 0.08 × 0.08 × 0.07
 
Data collection
Diffractometer Bruker SMART APEX
Absorption correction Multi-scan (SADABS; Bruker, 2014[Bruker (2014). APEX2, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.])
Tmin, Tmax 0.31, 0.40
No. of measured, independent and observed [I > 2σ(I)] reflections 722, 77, 76
Rint 0.022
(sin θ/λ)max−1) 0.772
 
Refinement
R[F2 > 2σ(F2)], wR(F2), S 0.018, 0.050, 1.20
No. of reflections 77
No. of parameters 10
Δρmax, Δρmin (e Å−3) 0.60, −0.93
Computer programs: APEX2 and SAINT (Bruker, 2014[Bruker (2014). APEX2, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.]), SHELXT (Sheldrick, 2015a[Sheldrick, G. M. (2015a). Acta Cryst. A71, 3-8.]), SHELXL2014 (Sheldrick, 2015b[Sheldrick, G. M. (2015b). Acta Cryst. C71, 3-8.]), CrystalMaker (Palmer, 2015[Palmer, D. (2015). CrystalMaker. CrystalMaker Software Ltd, Bicester, Oxfordshire, England.]) and publCIF (Westrip, 2010[Westrip, S. P. (2010). J. Appl. Cryst. 43, 920-925.]).

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: SHELXL2014 (Sheldrick, 2015b); molecular graphics: CrystalMaker (Palmer, 2015); software used to prepare material for publication: publCIF (Westrip, 2010).

Germanium titanium tetraoxide top
Crystal data top
Ge1.14Ti0.86O4Dx = 5.307 Mg m3
Mr = 187.88Mo Kα radiation, λ = 0.71073 Å
Tetragonal, P42/mnmCell parameters from 538 reflections
a = 4.493 (2) Åθ = 4.5–33.3°
c = 2.9121 (13) ŵ = 17.23 mm1
V = 58.79 (6) Å3T = 298 K
Z = 1Tabular, clear colourless
F(000) = 870.08 × 0.08 × 0.07 mm
Data collection top
Bruker SMART APEX
diffractometer
77 independent reflections
Radiation source: fine-focus sealed tube, sealed tube76 reflections with I > 2σ(I)
Graphite monochromatorRint = 0.022
Detector resolution: 8.3330 pixels mm-1θmax = 33.3°, θmin = 6.4°
ω and φ scansh = 66
Absorption correction: multi-scan
(SADABS; Bruker, 2014)
k = 66
Tmin = 0.31, Tmax = 0.40l = 44
722 measured reflections
Refinement top
Refinement on F2Primary atom site location: iterative
Least-squares matrix: fullSecondary atom site location: notdet
R[F2 > 2σ(F2)] = 0.018 w = 1/[σ2(Fo2) + (0.0347P)2 + 0.0217P]
where P = (Fo2 + 2Fc2)/3
wR(F2) = 0.050(Δ/σ)max < 0.001
S = 1.20Δρmax = 0.60 e Å3
77 reflectionsΔρmin = 0.93 e Å3
10 parametersExtinction correction: SHELXL2014 (Sheldrick, 2015b)
0 restraintsExtinction coefficient: 0.58 (7)
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*/UeqOcc. (<1)
Ge10000.0049 (3)0.57 (3)
Ti10000.0049 (3)0.43 (3)
O10.3059 (3)0.3059 (3)00.0077 (6)
Atomic displacement parameters (Å2) top
U11U22U33U12U13U23
Ge10.0059 (4)0.0059 (4)0.0029 (4)0.00019 (9)00
Ti10.0059 (4)0.0059 (4)0.0029 (4)0.00019 (9)00
O10.0084 (7)0.0084 (7)0.0065 (8)0.0014 (5)00
Geometric parameters (Å, º) top
Ge1—O1i1.9080 (12)Ge1—Ti1vii2.9121 (13)
Ge1—O1ii1.9080 (12)Ge1—Ge1vi2.9121 (13)
Ge1—O1iii1.9080 (12)Ge1—Ge1vii2.9121 (13)
Ge1—O1iv1.9080 (12)O1—Ti1viii1.9080 (12)
Ge1—O1v1.9440 (19)O1—Ge1viii1.9080 (12)
Ge1—O11.9441 (19)O1—Ti1ix1.9080 (12)
Ge1—Ti1vi2.9121 (13)O1—Ge1ix1.9080 (12)
O1i—Ge1—O1ii99.48 (8)O1i—Ge1—Ge1vi40.26 (4)
O1i—Ge1—O1iii80.52 (8)O1ii—Ge1—Ge1vi139.74 (4)
O1ii—Ge1—O1iii180.0O1iii—Ge1—Ge1vi40.26 (4)
O1i—Ge1—O1iv180.0O1iv—Ge1—Ge1vi139.74 (4)
O1ii—Ge1—O1iv80.52 (8)O1v—Ge1—Ge1vi90.0
O1iii—Ge1—O1iv99.48 (8)O1—Ge1—Ge1vi90.0
O1i—Ge1—O1v90.0Ti1vi—Ge1—Ge1vi0
O1ii—Ge1—O1v90.0Ti1vii—Ge1—Ge1vi180.0
O1iii—Ge1—O1v90.0O1i—Ge1—Ge1vii139.74 (4)
O1iv—Ge1—O1v90.0O1ii—Ge1—Ge1vii40.26 (4)
O1i—Ge1—O190.0O1iii—Ge1—Ge1vii139.74 (4)
O1ii—Ge1—O190.0O1iv—Ge1—Ge1vii40.26 (4)
O1iii—Ge1—O190.0O1v—Ge1—Ge1vii90.0
O1iv—Ge1—O190.0O1—Ge1—Ge1vii90.0
O1v—Ge1—O1180.0Ti1vi—Ge1—Ge1vii180.0
O1i—Ge1—Ti1vi40.26 (4)Ti1vii—Ge1—Ge1vii0
O1ii—Ge1—Ti1vi139.74 (4)Ge1vi—Ge1—Ge1vii180.0
O1iii—Ge1—Ti1vi40.26 (4)Ti1viii—O1—Ge1viii0
O1iv—Ge1—Ti1vi139.74 (4)Ti1viii—O1—Ti1ix99.48 (8)
O1v—Ge1—Ti1vi90.0Ge1viii—O1—Ti1ix99.48 (8)
O1—Ge1—Ti1vi90.0Ti1viii—O1—Ge1ix99.5
O1i—Ge1—Ti1vii139.74 (4)Ge1viii—O1—Ge1ix99.48 (8)
O1ii—Ge1—Ti1vii40.26 (4)Ti1ix—O1—Ge1ix0
O1iii—Ge1—Ti1vii139.74 (4)Ti1viii—O1—Ge1130.3
O1iv—Ge1—Ti1vii40.26 (4)Ge1viii—O1—Ge1130.26 (4)
O1v—Ge1—Ti1vii90.0Ti1ix—O1—Ge1130.3
O1—Ge1—Ti1vii90.0Ge1ix—O1—Ge1130.26 (4)
Ti1vi—Ge1—Ti1vii180.0
Symmetry codes: (i) y1/2, x+1/2, z+1/2; (ii) y1/2, x+1/2, z1/2; (iii) y+1/2, x1/2, z+1/2; (iv) y+1/2, x1/2, z1/2; (v) x, y, z; (vi) x, y, z+1; (vii) x, y, z1; (viii) y+1/2, x+1/2, z+1/2; (ix) y+1/2, x+1/2, z1/2.
Atomic coordinates and equivalent isotropic atomic displacement parameters (Å2) for TiGeO4. U(eq) is defined as one third of the trace of the orthogonalized Uij tensor. top
x/ay/bz/cU(eq)
Ge(1)0.00.00.00.0049 (3)
Ti(1)0.00.00.00.0049 (3)
O(1)0.3059 (3)0.3059 (3)0.00.0077 (6)
Anisotropic atomic displacement parameters (Å2) for TiGeO4. top
U11U22U33U23U13U12
Ge(1)0.0059 (4)0.0059 (4)0.0029 (4)00-0.00019 (9)
Ti(1)0.0059 (4)0.0059 (4)0.0029 (4)00-0.00019 (9)
O(1)0.0084 (7)0.0084 (7)0.0065 (8)00-0.0014 (5)
Selected interatomic distances (Å) and angles for TiGeO4. top
Ge(1)-O(1)1.9080 (12)Ge(1)-O(1)1.9441 (19)
Ge1-Ti12.9121 (13)O1-Ti11.9080 (12)
O1-Ge11.9080 (12)
O(1)-Ge(1)-O(1)99.48 (8)O(1)-Ge(1)-O(1)80.52 (8)
O(1)-Ge(1)-O(1)180.0O(1)-Ge(1)-O(1)90.0
O(1)-Ge(1)-Ti(1)139.74 (4)O(1)-Ge(1)-Ti(1)40.26 (4)
O(1)-Ge(1)-Ti(1)90.0Ti(1)-Ge(1)-Ti(1)180.0
Ge1-O1-Ti199.48 (8)Ti1-O1-Ge1130.26 (4)

Acknowledgements

We acknowledge the use of facilities within the Eyring Materials Center at Arizona State University. The GeO2–TiO2 crystals were grown in-house at Sandvik Hyperion.

Funding information

The authors thank Sandvik Hyperion (Worthington, OH, USA) for providing research funding for exploratory experiments and X-ray diffraction related to this project. Exploration of the TiO2–GeO2 system at ASU was funded by Sandvik Hyperion. Use of the COMPRES Cell Assembly Project was supported by COMPRES under NSF cooperative agreement EAR 1661511.

References

First citationBruker (2014). APEX2, SAINT and SADABS. Bruker AXS Inc., Madison, Wisconsin, USA.  Google Scholar
First citationGelato, L. M. & Parthé, E. (1987). J. Appl. Cryst. 20, 139–143.  CrossRef Web of Science IUCr Journals Google Scholar
First citationGerward, L. & Staun Olsen, J. (1997). J. Appl. Cryst. 30, 259–264.  CrossRef Web of Science IUCr Journals Google Scholar
First citationGullikson, A. L., Leinenweber, K., Stoyanov, E., Zhang, H. & Malik, A.-S. (2015). J. Am. Ceram. Soc. 98, 982–989.  Web of Science CrossRef Google Scholar
First citationHaines, J., Léger, J. M., Chateau, C. & Pereira, A. S. (2000). Phys. Chem. Miner. 27, 575–582.  Web of Science CrossRef Google Scholar
First citationHoward, C. J., Sabine, T. M. & Dickson, F. (1991). Acta Cryst. B47, 462–468.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationLeinenweber, K., Gullikson, A. L., Stoyanov, E. & Malik, A. (2015). J. Solid State Chem. 229, 10–18.  Web of Science CrossRef Google Scholar
First citationMicoulaut, M., Cormier, L. & Henderson, G. S. (2006). J. Phys. Condens. Matter, 18, R753–R784.  Web of Science CrossRef Google Scholar
First citationPalmer, D. (2015). CrystalMaker. CrystalMaker Software Ltd, Bicester, Oxfordshire, England.  Google Scholar
First citationSarver, J. F. (1961). Am. J. Sci. 259, 709–718.  CrossRef 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 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 citationWestrip, S. P. (2010). J. Appl. Cryst. 43, 920–925.  Web of Science CrossRef CAS IUCr Journals 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