digital.csic.esdigital.csic.es/bitstream/10261/77596/1/Physics of the... · Web viewHigh-resolution...

33
Post-print of: Physics of the Earth and Planetary Interiors volumes 190–191, January 2012, pages 87–94 Aluminum incorporation in α-PbO2 type TiO2 at pressures up to 20 GPa Alberto Escudero, Falko Langenhorst Bayerisches Geoinstitut, Universität Bayreuth, D-95440 Bayreuth, Germany Abstract Aluminum incorporation into the high pressure polymorph of TiO2 with the structure of α-PbO2 has been studied from 10 to 20 GPa and 1300 °C by XRD, high-resolution 27Al MAS-NMR and TEM. Al- doped α-PbO2 type TiO2 can be recovered at atmospheric pressure. Al2O3 solubility in α-PbO2 type TiO2 increases with increasing the synthesis pressure. The α-PbO2 type TiO2 polymorph is able to incorporate up to 35 wt.% Al2O3 at 13.6 GPa and 1300 °C, being the substitution of Ti4+ by Al3+ on normal octahedral sites the mechanism of solubility. The transition to the higher pressure TiO2 polymorph with the ZrO2 baddeleyite structure, Akaogiite, has not been observed in the quenched samples at room pressure. The microstructure of the recovered sample synthesized at 16 GPa and 1300 °C points to the existence of a non- quenchable aluminum titanium oxide phase at these conditions. Keywords TiO2-II; Alumina; High-pressure; TEM; 27Al MAS-NMR; Non- quenchable phases; Solid solutions; Stacking faults 1. Introduction Titanium dioxide TiO2 has been intensively studied for a few decades due to both basic and applied interests in geology and material science. Rutile is a common minor mineral in metamorphic and plutonic igneous rocks. Rutile transforms to a high pressure TiO2 polymorph with the structure of α-PbO2 at about 9 GPa and 1300 °C, despite experiments on the phase 1

Transcript of digital.csic.esdigital.csic.es/bitstream/10261/77596/1/Physics of the... · Web viewHigh-resolution...

Post-print of: Physics of the Earth and Planetary Interiors volumes 190–191, January 2012, pages 87–94

Aluminum incorporation in α-PbO2 type TiO2 at pressures up to 20 GPa

Alberto Escudero, Falko Langenhorst

Bayerisches Geoinstitut, Universität Bayreuth, D-95440 Bayreuth, Germany

Abstract

Aluminum incorporation into the high pressure polymorph of TiO2 with the structure of α-PbO2 has been studied from 10 to 20 GPa and 1300 °C by XRD, high-resolution 27Al MAS-NMR and TEM. Al-doped α-PbO2 type TiO2 can be recovered at atmospheric pressure. Al2O3 solubility in α-PbO2 type TiO2 increases with increasing the synthesis pressure. The α-PbO2 type TiO2 polymorph is able to incorporate up to 35 wt.% Al2O3 at 13.6 GPa and 1300 °C, being the substitution of Ti4+ by Al3+ on normal octahedral sites the mechanism of solubility. The transition to the higher pressure TiO2 polymorph with the ZrO2 baddeleyite structure, Akaogiite, has not been observed in the quenched samples at room pressure. The microstructure of the recovered sample synthesized at 16 GPa and 1300 °C points to the existence of a non-quenchable aluminum titanium oxide phase at these conditions.

Keywords

TiO2-II; Alumina; High-pressure; TEM; 27Al MAS-NMR; Non-quenchable phases; Solid solutions; Stacking faults

1. Introduction

Titanium dioxide TiO2 has been intensively studied for a few decades due to both basic and applied interests in geology and material science. Rutile is a common minor mineral in metamorphic and plutonic igneous rocks. Rutile transforms to a high pressure TiO2 polymorph with the structure of α-PbO2 at about 9 GPa and 1300 °C, despite experiments on the phase boundary between rutile and the α-PbO2 type TiO2 polymorph yielded quite controversial results (Akaogi et al., 1992, Olsen et al., 1999 and Withers et al., 2003). This last polymorph has been used as a geobarometer to indicate ultra high pressure metamorphism (UHPM) (Hwang et al., 2000 and Wu et al., 2005), despite the suitableness of using nano-structured α-PbO2 type TiO2 as an indicator of ultra high pressure is still under discussion (Chen and Fu, 2006, Escudero et al., 2012 and Wu et al., 2006). At about 17 GPa and 1300 °C TiO2 transforms into a polymorph with the structure of ZrO2 baddeleyite (Sato et al., 1991). This mineral has been recently encountered in heavily shocked garnet-cordierite-sillimanite gneiss in the Suevite breccia of the Ries meteorite impact crater in Germany and has been named as Akaogiite (El Goresy et al., 2010). Higher pressure TiO2 polymorphs have also been reported in the literature (Dubrovinskaia et al., 2001).

1

Aluminum is one of the trace elements in rutile that has been suggested to provide information on the P-T conditions of the rutile-bearing metamorphic rocks. In fact, exsolved corundum lamellae have been reported in diamondiferous eclogites from South Africa (Sobolev and Yefimova, 2000). Recent studies have shown that Al solubility in TiO2 rutile increases drastically with pressure, with two different mechanism of solubility. Enhanced aluminum concentration in TiO2 rutile as well as (1 1 0) twinned CaCl2-type structure TiO2 grains are thus a clear indication of high-pressure conditions (Escudero et al., 2011b). However, in order to evaluate the possibility of using the Al incorporation into TiO2 phases as a possible indicator of high pressure conditions, the effect of the nature of the different TiO2 polymorphs at higher pressure on the Al incorporation into TiO2 phases should be determined. We present in this paper a chemical and microstructural study of the Al incorporation into TiO2 phases at pressures between 7 and 20 GPa, where the high pressure polymorphs of TiO2 are expected to be stable.

2. Experiments and methods

2.1. Synthesis

High-pressure experiments were performed from commercial Al2TiO5 (Sigma–Aldrich nanopowder, <25 nm particle size (BET), 98.5% trace metals basis) in the range from 7 to 20 GPa were carried out in a 1200 ton MA8 Kawai-type multi-anvil press, using 18, 14, and 10 mm Cr2O3-doped MgO octahedra with a stepped LaCrO3 furnace, depending on the pressure. The high pressure octahedral assembly was compressed using 8 tungsten carbide anvils with corner truncations of 11, 8, and 5 mm lengths, also depending on the pressure. Samples were placed in a Pt metal foil capsule and inserted into the center of the furnace inside an MgO sleeve. All ceramic parts of the high pressure cell were fired at 1000 °C for 30 min prior to assembling. A W97Re3–W75Re25 thermocouple was inserted to measure the temperature at the top surface of the capsule. Samples were first pressurized to the target pressure and then heated at 1300 °C for 3 h. A summary of the synthesized samples is shown in Table 1.

An extra sample with a 85% TiO2 + 15% Al2O3 (wt.%) composition was also prepared in order to carry out 27Al MAS-NMR measurements without the interference of the signal corresponding to corundum, which overlaps with the signals corresponding to Al in TiO2. TiO2 (rutile nanopowder, 99.5% sigma) and Al2O3 (corundum nanopowder, 99.97%, Chempur) were mixed in a ball mill at 600 rpm during 60 min. This precursor material was later submitted to high pressure treatments at 14 GPa and 1400 °C in a 5000 ton MA8 Kawai-type multi-anvil press during 1 h. This multi anvil press allows synthesizing sample volumes 10 times larger than the obtained in conventional systems.

2.2. Characterization techniques

Powder diffraction patterns of the quenched samples were obtained at atmospheric pressure with a Stoe STADI-P diffractometer operating in transmission mode, using Co Kα1 radiation selected with a focusing germanium monochromator and a linear position-sensitive detector. High quality XRD patterns (recorded from 2θ 20° to 120° with 0.016° steps and 1000 s counting time) were analyzed using the Le Bail algorithm of the program package GSAS (Larson and Von

2

Dreele, 1994) to determine the unit cell parameters. Refined parameters were background coefficients, lattice constants, line widths, asymmetry parameters, and zero error.

High-resolution 27Al magic angle spinning nuclear magnetic resonance (MAS NMR) spectroscopy was carried out in a Bruker DRX800 (18.8 T) spectrometer installed at the University of Lille, France. The spectrometer is equipped with a multinuclear probe operating at 208.49 MHz for 27Al. Powdered samples were packed in 2.5 mm zirconia rotors and spun at 30 kHz. A single-pulse sequence was used with an observation frequency for 27Al of 208.49 MHz, and a pulse width of 0.5 μs (π/2 pulse length of 5 μs). The delay time was optimized to 500 ms. The chemical shifts are reported in parts per million from 0.1 M AlCl3 solution. The spectrum was simulated using the DMFit program (Massiot et al., 2002) handling finite spinning speeds and mixed quadrupolar/CSA interaction in MAS experiments. 1H MAS-NMR spectroscopy was used to determine the possible presence of H species in the samples.

An analytical transmission electron microscope (ATEM, Philips CM-20FEG) operating at 200 kV and equipped with an energy-dispersive X-ray spectrometer (EDXS, NORAN Ge detector) was used to observe the microstructures of the TiO2 grains. Samples were thinned in an Ar ion milling machine at 4.5 kV and 1 mA until electron transparency was reached. ELNES spectra were collected with a Gatan PEELS 666 parallel electron spectrometer attached to the Philips CM20 FEG TEM. For the calibration of energy scale, the C K-edge of amorphous carbon was measured simultaneously. Aluminum content in TiO2 has been quantified by EDX analysis. The quantification included an absorption correction and the calibration of Cliff–Lorimer factors for aluminum and titanium with respect to oxygen, using standard samples (Langenhorst et al., 1995). At least ten Al-doped α-PbO2 type TiO2 single crystals have been analyzed at each pressure to calculate mean compositions. A counting time from 100 to 300 s has been used in order to accumulate enough X-ray counts.

3. Results

3.1. X-ray Diffraction data

Representative XRD patterns of the samples synthesized at 7, 10, 13.6, 16, and 20 GPa and 1300 °C are shown in Fig. 1. The diagram of the sample synthesized at 7 GPa shows reflections corresponding to CaCl2 – structured rutile (Escudero et al., 2011b), and corundum (α-Al2O3, PDF 10-0173). The diagrams of the samples synthesized from 10 upto 20 GPa show the same reflections of corundum (α-Al2O3), and new peaks corresponding to the TiO2 high pressure polymorph with the structure of α-PbO2 (PDF 21-1236). The further transformation of TiO2 into the baddeleyite-type TiO2 is not observed in the quenched samples. Increasing the synthesis pressure produces very similar α-PbO2 TiO2 patterns. There are, however, small variations in the peak positions, which shift slightly towards larger angles until 16 GPa. However, the signals corresponding to the α-PbO2 type TiO2 observed in the sample synthesized at 20 GPa shift to lower values of 2θ, respect to the signals of the 16 GPa sample (see peaks under the vertical guide lines). The position of the corundum signals does not change appreciably with the synthesis pressure. Two differences can be observed in the XRD diagram corresponding to the sample synthesized at 16 GPa. On one hand, some low intensity reflections situated at 2θ = 28.3°, 38.5° and 46.5° appear on the XRD diagram. The reflections have been marked with asteriks on the Figure. On the other hand, the intensity of the peaks

3

corresponding to α-PbO2 type TiO2, especially those situated at high 2θ angles is much higher than those corresponding to the corundum phase, if the intensity radio is compared to the lower pressure samples.

To refine the lattice parameters of α-PbO2 type TiO2 recovered from different pressures, the XRD patterns have been analyzed with the Le Bail method using the GSAS software (Larson and Von Dreele, 1994), as described in the experimental section. The starting parameters have been taken from Grey, Li, Madsen and Braunshausen for α-PbO2 type TiO2 (Grey et al., 1988), and from Toebbens, Stuesser, Knorr, Mayer and Lampert for corundum (Tobbens et al., 2001). All the reflections corresponding to the TiO2 phase of the samples synthesized from 10 to 20 GPa could be fitted on the basis of an orthorhombic unit cell with space group Pbcn, which is the crystal structure of α-PbO2 type TiO2.

Fig. 2 shows the variation of the unit cell parameters of for α-PbO2 type TiO2 as a function of synthesis pressure. All the lattice parameters (a, b, c as well as the unit cell volume) decrease linearly when increasing the synthesis pressure up to 16 GPa. The lattice parameters of the recovered α-PbO2 type TiO2 phase after the treatment at 20 GPa do not follow this trend. These parameters are higher than it would be expected, according to the behavior of the samples synthesized at lower pressure, and are similar to the ones of an Al-doped α-PbO2 type TiO2 phase synthesized at about 13 GPa. Lattice parameters of the recovered samples are shown in Table 1.

3.2. High-resolution 27Al MAS-NMR and 1H MAS-NMR

Fig. 3 shows the 27Al MAS-NMR spectrum of the 85% TiO2 + 15% Al2O3 sample synthesized at 14 GPa and 1400 °C. The spectrum consists of two different signals centered at 15 and 5.5 ppm, and correspond to hexacoordinated Al environments. No signals corresponding to other octahedral or tetrahedral Al environments are observed. The XRD pattern of this recovered sample (not shown) indicates the presence of α-PbO2 type TiO2 as a very major phase with some very small reflections of α-Al2O3 corundum. Thus, the first signal can be attributed to remaining α-Al2O3 corundum, and the second signal must correspond to an octahedral Al environment in α-PbO2 type TiO2. This corresponds to Al3+ replacing a Ti4+ on the normal octahedral Ti sites of the α-PbO2 type TiO2 polymorph. This signal can be readily fitted with CQ = 6.3 MHz, quadrupolar asymmetry parameter η = 0.5, and isotropic chemical shift δiso = 10.3 ppm. Single-pulse MAS NMR carried out at high external magnetic field has demonstrated to be very sensitive to the different Al environment in Al-doped TiO2 rutile synthesized at high pressure (Escudero et al., 2011a). This indicates that despite other Al species have been observed in Al-doped rutile, only one octahedral Al environment exists in α-PbO2 type TiO2, at least at relative low pressures (c.a. 10–14 GPa). The 1H MAS-NMR spectrum (not shown) did not show significant signals corresponding to H and thus the samples can be considered to be anhydrous.

3.3. Electron microscopic observations

Fig. 4 shows two micrographs of the sample synthesized at 10 GPa. The microstructure consists of α-PbO2 type TiO2 grains decorated with lamellae which present different thickness. Relative wide lamellae of this secondary phase can be observed in Fig. 4A, as well as the corresponding

4

electron diffraction pattern. The observed extra reflections on the SAED pattern not corresponding to the α-PbO2 type TiO2 are compatible with CaCl2-type TiO2, which is an orthorhombic distortion of Al-doped TiO2 rutile produced when quenching from high pressure (Escudero et al., 2011b). The high resolution TEM image of this sample as well as its SAED shown in Fig. 4B reveal that the lamellae also exist in a nanometric level, and are parallel to the (0 0 1) α-PbO2 type TiO2 plane.

Samples synthesized at 12 and 13.6 GPa (not shown) exhibit normal α-PbO2 type TiO2 grains with no extra phases or defects, and corundum grains. Fig. 5 shows a TEM micrograph of an α-PbO2 type TiO2 grain of the sample synthesized at 16 GPa. A microstructure consisting of ordered and homogeneous lamellae can be observed. The HRTEM image shown in Fig. 5B indicates that the lamellae are parallel to the (3 View the MathML source1¯ 1) plane of the α-PbO2 type TiO2 polymorph. The lamellae are about 5 or 6 lattice planes wide, being the d spacing around 6.7 Å.

This observation suggests the presence of another phase, almost no detectable by XRD, and that could be responsible for the low intensity reflections observed in its XRD diagram. To clarify the nature of this phase, many SAED were collected in order to detect extra diffraction spots which could correspond to this secondary phases. Only along three different zone axes some extra and generally weak spots not corresponding to the α-PbO2 type TiO2 polymorph could be observed (Fig. 5C, D, and E). The extra spots are compatible with a material possessing an orthorhombic symmetry, as it will be discussed in the next section. The ELNES spectrum of the Al-doped sample synthesized at 16 GPa is shown in Fig. 6. Spectra corresponding to both Al-doped and Si-doped TiO2 synthesized at 12 GPa have been added to the Figure for comparative reasons. ELNES spectra have demonstrated to be sensitive to the valence state of Ti (Stoyanov et al., 2007). There are no appreciable changes between the different spectra, which consist of four bands centered at 458.4, 460.5, 463.7 and 465.7 eV. The signals correspond to the typical Ti L3 and L2 edges, whose positions are the expected for Ti4+ (Stoyanov et al., 2007). This indicates that no changes in the valence state of Ti have been produced during the high pressure experiments.

Fig. 7 shows a TEM micrograph of an α-PbO2 type TiO2 grain of the sample synthesized at 20 GPa. The microstructure of the TiO2 is different to all previous samples and consists of stacking faults.

Fig. 8 shows the change in the Al2O3 solubility in TiO2 with pressure. In general terms the effect of high pressure is to enhance the Al2O3 solubility in TiO2, but two drops at 10 and 20 GPa can be observed. The TiO2 grains are able to accommodate up to at least 35% Al2O3 (wt.%) at 13.6 GPa and 1300 °C. EDX data from the sample synthesized at 16 GPa have not been taken into account for reasons that will be explained below. Al2O3 contents observed in the TiO2 grains are shown in Table 1.

4. Discussion

4.1. Phase stability of the Al-doped α-PbO2 type TiO2. Absence of Akaogiite at higher pressures in the quenched samples

5

XRD data indicate that CaCl2-structured TiO2 is the stable Al-doped TiO2 phase in the quenched samples up to at least 7 GPa, transforming into the α-PbO2 type TiO2 between this pressure and 10 GPa. This last is the stable phase above 10 GPa until at least 13.6 GPa. These results are in good agreement with the previous studies of Withers (Withers et al., 2003) and Akaogi (Akaogi et al., 1992) for pure TiO2. Thus, the Al-doped α-PbO2 type TiO2 can be recovered at atmospheric pressure and does not transform back into rutile or CaCl2-type TiO2 when releasing the pressure. The transformation of pure α-PbO2 type TiO2 to the baddeleyite phase, Akaogiite, has been reported at about 17 GPa at 1300 °C (Olsen et al., 1999). However, this last phase is not observed in the quenched sample synthesized at 20 GPa and 1300 °C, being the pure α-PbO2 type TiO2 the detected polymorph by XRD. Two explanations can be given to this observation. On one hand, the presence of Al3+ in the TiO2 structure may stabilize the α-PbO2 type TiO2 phase, thus shifting the phase boundary with the baddeleyite phase toward higher pressures. On the other hand, the baddeleyite phase may be the stable one at higher pressures but cannot be observed at atmospheric pressure, transforming back to the α-PbO2 type TiO2 when releasing the pressure. Based on the behavior of the lattice parameters with the synthesis pressure, the second hypothesis seems to be more reasonable. All the lattice parameters, a, b, c and the unit cell volume decrease linearly with increasing pressure up to 16 GPa, suggesting the formation of a solid solution between TiO2 and Al2O3 in which Al3+ replaces Ti4+ on normal α-PbO2 type TiO2 octahedral sites. This mechanism of solubility is consistent with the NMR data, which do not show the presence of interstitial Al3+, at least at relatively low pressures (up to 14 GPa). A similar behavior has been observed in the lattice parameters of Al-doped rutile at pressures up to 2 GPa at 1300 °C. This was associated with a substitutional solid solution in which oxygen vacancies compensate the charge difference between Al3+ and Ti4+ ( Escudero et al., 2011b and Slepetys and Vaughan, 1969). Due to the similar ionic radii (0.61 Å for Ti4+ and 0.53 Å for Al3+ in octahedral coordination) (Shannon, 1976) aluminum can occupy regular cation positions giving a substitutional solid solution. The slightly different ionic radii of the cations explains the decrease of the unit cell parameters when this substitution takes place. However, the change of the Al-doped α-PbO2 type TiO2 lattice parameters at increasing pressures is larger than the observed in the rutile structure at pressures in which the subsitutional solid solution is the main solubility mechanism. This indicates that the substitution is more effective at higher pressures. Taking into account that interstitial Al was found in the TiO2 rutile structure at much lower pressures and above 3.2 wt.% of Al2O3 (Escudero et al., 2011b), that no other Al environments have been detected by high-resolution 27Al MAS-NMR, that the 1H MAS-NMR data strongly suggest that the samples are anhydrous and that the ELNES spectra indicate the absence of other Ti species such as interstitial Ti3+, it seems that the main mechanism of Al incorporation into α-PbO2 type TiO2 with increasing pressure consists of the substitution of Ti4+ by Al3+, with the formation of oxygen vacancies to compensate the charge, at least until about 14 GPa. Although the formation of oxygen vacancies would not be favored at high pressure, the α-PbO2 type TiO2 seems to be able to accommodate them up to a certain level. Above 14 GPa, the structure is not able to accommodate more oxygen vacancies and thus a phase transition takes place, as it would be discussed in the 4.3 section. The substitution of Ti4+ by Al3+ would thus energetically compensate the instability produced by the presence of oxygen vacancies at

6

relatively lower pressures (10–14 GPa). The same substitutional mechanism has been reported for Fe3+ doped α-PbO2 type TiO2 at high pressure (Bromiley et al., 2004), and only hexacoordinated Si has been observed in Si-doped α-PbO2 type TiO2 at the same pressures (Mosenfelder et al., 2010).

As it was commented above, the trend in the behavior of the Al-doped α-PbO2 type TiO2 lattice parameters with increasing the synthesis pressure is not observed in the sample synthesized at 20 GPa, which shows larger lattice parameters than expected. This indicates a minor Al incorporation into the TiO2 structure and suggests that another TiO2 polymorph, which presents a different affinity for Al, is the stable one at 20 GPa and 1300 °C. In fact, according to the TiO2 phase diagram (Olsen et al., 1999), the baddeleyite polymorph is the expected for pure TiO2 at this conditions and probably it is the one formed in the conditions of the experiment. However, TiO2 baddeleyite type could not be observed experimentally at room pressure after decompression (Dubrovinskaia et al., 2001, Sato et al., 1991 and Tang and Endo, 1993). Thus, this polymorph reverts to α-PbO2 type TiO2 when releasing the pressure, being this last one the observed at ambient conditions and keeping the Al content existing in the baddeleyite phase. The Al2O3 content determined from the recovered TiO2 grains should reflect the solubility of Al2O3 in TiO2 baddeleyite, Akaogiite, at 20 GPa. The increase of the recovered α-PbO2 type TiO2 lattice parameters in the sample synthesized at 20 GPa compared to the ones determined at 16 GPa indicates that the solubility of Al2O3 in TiO2 baddeleyite is lower than in the α-PbO2 type TiO2 phase. The lattice parameters of the of the recovered α-PbO2 type TiO2 of the sample synthesized at 20 GPa are close to those corresponding to an Al-doped sample synthesized between 12 and 14 GPa and 1300 °C. This comparison suggests that the solubility of Al2O3 in TiO2 baddeleyite at 20 GPa is similar than the one shown by Al2O3 in α-PbO2 type TiO2 at about 12–14 GPa.

4.2. Solubility of Al2O3 in α-PbO2 type TiO2

In general terms the solubility of Al2O3 in TiO2 phases increases with increasing pressure. This behavior has been observed in Al-doped rutile (Escudero et al., 2011b). Al2O3 solubility in α-PbO2 type TiO2 increases drastically with the synthesis pressure. The α-PbO2 type TiO2 polymorph is able to accommodate up to at least 35% Al2O3 (wt.%) at 13.6 GPa and 1300 °C, compared to 1.22% at atmospheric pressure, in the rutile polymorph (Slepetys and Vaughan, 1969). However, two drops in the solubility of Al2O3 in TiO2 can be observed in Fig. 8. Both decreases in the Al incorporation in TiO2 phases can be associated with phase transitions. On one hand, TiO2 incorporates less Al at 12 GPa than at 10 GPa. This last sample showed inhomogeneous nanometer-size rutile lamellae in the α-PbO2 type TiO2 grains that can give rise to higher global Al2O3 content in TiO2. In fact, the uncertainty of the measurements is higher for this sample. This indicates that the solubility of Al2O3 in rutile in this range of pressure is higher than in the α-PbO2 type TiO2. This is consistent with the absence of interstitial Al3+ in the α-PbO2 type TiO2, which does not present a second mechanism of solubility to enhance the global Al2O3 solubility in TiO2.

On the other hand, the Al2O3 content of the TiO2 grains of the sample synthesized at 20 GPa is much lower than the expected, suggesting again a TiO2 phase transition between 16 and 20 GPa. As it was indicated by the behavior of the lattice parameters, the baddeleyite phase

7

seems to be the stable polymorph at 20 GPa, transforming into α-PbO2 type TiO2 when releasing the pressure but keeping its original Al content. The Al2O3 measured in this sample can be compared to the one shown by an Al-doped α-PbO2 type TiO2 synthesized between 12 and 14 GPa and 1300 °C. All these observations support the non-quenchable character of the TiO2 baddeleyite phase, and are consistent with the XRD data. The different microstructure shown by the recovered 20 GPa sample must be thus connected with this reversible phase transition from the TiO2 baddeleyite polymorph to the α-PbO2 type TiO2.

Some crystal-chemical explanations can be given to explain the different affinity of Al for the different TiO2 polymorphs. Except around the phase boundary between rutile and α-PbO2 type TiO2, Al solubility in α-PbO2 type TiO2 is markedly higher than in TiO2 rutile. Taking into account the geometry of the Ti4+ octahedral site in both rutile and α-PbO2 type TiO2, the octahedral site in the α-PbO2 type TiO2 structure is more distorted that the one of the rutile structure. Three different Ti–O bond distances of 1.87, 1.96, and 2.06 Å are found in α-PbO2 type TiO2 (Grey et al., 1988), whereas two almost identical distances of 1.96, and 1.97 Å are reported for TiO2 rutile (Bokhimi et al., 2002). This suggests that the more distorted octahedral Ti site in the α-PbO2 type TiO2 polymorph might facilitate the accommodation of an atom of different nature in this structure, making thus the incorporation of dopants into this high pressure polymorph of TiO2 more effective. The coordination number of Ti increases from six to seven across the α-PbO2 type TiO2 to baddeleyite transition (Sato et al., 1991). Given the radius radio criterion (Müller, 2007), the incorporation of a relatively small cation such as Al3+ into heptacoordinated sites would be less favored than into hexacoordinated sites, even at very high pressures. This explains the observed lower affinity of Al for the Akaogiite polymorph.

4.3. Possible non-quenchable phase at 16 GPa

The microstructure of the Al-doped TiO2 grains in the sample synthesized at 16 GPa suggests the existence of a non-quenchable phase at this pressure and temperature, which reverts to α-PbO2 type TiO2 when decreasing temperature and pressure. However, some rest of this phase are observed in the quenched sample. EDX analyses of the different TiO2 grains indicate a AlTi0.64O2.8 composition. This composition reflects very likely the original composition of a possible non-quenchable aluminum titanium oxide phase, and it is close to a ABO3 stoichiometry. Two important mineral structures are well-known for this stoichiometry: ilmenite and perovskite. In fact, the extra diffractions observed in the XRD pattern of this sample (Fig. 1d) are compatible with both the structures of FeTiO3 ilmenite (space group RView the MathML source3¯, PDF 01-071-1140) and with a family of orthorhombic rare earth titanates perovskites (NdTiO3, PDF 29-922; GdTiO3, PDF 29-0614, space group Pnma). Despite the cation size difference between the rare earths and Al, this perovskite-type structure is also known for FeTiO3 at high pressure ( Leinenweber et al., 1991, Ming et al., 2006 and Wilson et al., 2005). However, experimental studies have shown that it reverts to lower pressure polymorphs when decompressing, being thus not observable at atmospheric conditions (Ming et al., 2006). The extra weak electron diffraction spots observed in Fig. 5C, D, and E have been tried to be indexed considering both possible structures. The planes could not be indexed considering a hexagonal symmetry but are compatible with the orthorhombic structure shown by the rare earth titanate perovskites. Taking into account that Fe3+ and Al3+ show a similar

8

size (Shannon, 1976), the non-quenchable character of the perovskite-type high pressure phase of FeTiO3 is consistent with the similar behavior of the supposed perosvkite-type structure of Al-doped TiO2 at 16 GPa. However, the formation of this supposed non-quenchable aluminum titanium oxide phase with an orthorhombic perovskite structure would imply changes in the valence state of one of its elements in order to keep the electroneutrality of the structure. ELNES spectra have shown that Ti remains as Ti4+. This indicate that the supposed perovskite structure should be cation-deficient. In fact, cation-deficient rare earth titanates with the perovskite structure have been reported in the literature. These compounds are A-site deficient perovskite-type and exhibit an orthorhombic or tetragonal symmetry with an orthorhombic distortion observed at room temperature due to octahedral tilting ( Howard and Zhang, 2004, Jung, 2005 and Yoshii, 2000). They show a RE2/3TiO3 stoichiometry. The rare earth cation is the bigger one and occupies the A perovskite site. In case of the aluminum titanium oxide phase, the bigger cation should occupy the A site, giving thus a Ti2/3AlO3 compound, which is exactly the composition experimentally determined by EDX analyses on the TEM. Indeed, both the extra and low intensity reflections observed on the XRD of the sample synthesized at 16 GPa are also compatible with this structure (Nd0.66TiO3, PDF 00-049-0244) and the extra weak electron diffraction spots appearing on the electron diffraction patterns can also be indexed taking into account the orthorhombic structures shown by the cation deficient rare earth titanates perovskites.

The observed microstructure of the Al-doped TiO2 grains recovered at room temperature and pressure, the extra spots appearing on the SAED patterns, the composition determined by EDX and the studies existing in the literature for similar compounds, suggest that a new non-quenchable phase with a cation – deficient perovskite structure is formed in the Al2O3–TiO2 system at 16 GPa and 1300 °C. However, in situ measurements at high pressure and high temperature are demanded in order to clarify the nature of this possible new phase.

5. Conclusions

Aluminum incorporation in α-PbO2 type TiO2 increases with increasing the synthesis pressure. Al-doped CaCl2-type TiO2 transforms into the α-PbO2 type TiO2 polymorph between 7 and 10 GPa in the quenched samples. The further transition into the baddeleyite type polymorph, Akaogiite, has not been directly observed. The microstructure of the TiO2 grains with coexisting rutile and α-PbO2 type TiO2 consists of α-PbO2 type TiO2 grains decorated with CaCl2-type TiO2 lamellae which are parallel to the [0 0 1] α-PbO2 type TiO2 plane. The transition from rutile to the α-PbO2 type TiO2 polymorph produces a decrease of the Al2O3 solubility in TiO2, given the lower affinity of Al for the α-PbO2 type TiO2. A second decrease in the Al2O3 incorporation in α-PbO2 type TiO2 exists between 16 and 20 GPa. This has been ascribed to the non-quenchable character of the Akaogiite polymorph, which reverts to α-PbO2 type TiO2 when decompressing. The α-PbO2 type TiO2 grains recovered from the sample synthesized at 20 GPa are decorated with stacking faults, which are produced during this reversible phase transition. The affinity of Al for the Akaogiite polymorph is lower than for α-PbO2 type TiO2. The microstructure of the Al-doped sample recovered from 16 GPa and 1300 °C points to the existence of a non-quenchable aluminum titanium oxide compound at this conditions. This phase might show a cation-deficient orthorhombic perosvskite structure. However, new in situ experiments are demanded in order to clarify the nature of this phase.

9

Acknowledgments

Uwe Dittmann, Laurent Delevoye, and Nobuyoshi Miyajima are gratefully acknowledged for help with thin section samples preparation, MAS-NMR, and ELNES measurements, respectively. This work was supported by the European Union VI Framework Programme as an HRM Activity (Contract number MRTN-CT-2006-035957), the Spanish Ministerio de Ciencia e Innovación (Postdoctoral Fellowship MICINN-FECYT), the Visitors Programme of the Bayerisches Geoinstitut, and the Leibniz program of the Deutsche Forschungsgemeinschaft (LA 830/14-1 to FL).

10

References

Akaogi et al., 1992; M. Akaogi, K. Kusaba, J.I. Susaki, T. Yagi, M. Matsui, T. Kikegawa, H. Yusa, E. Ito

High-pressure high-temperature stability of α-PbO2-type TiO2 and MgSiO3 majorite: calorimetric and in situ X-ray diffraction studies

High-Pressure Research Application to Earth and Planetary Sciences, American Geophysical Union (1992)

Amow and Greedan, 1996; G. Amow, J.E. Greedan

The structural and magnetic properties of Nd1−xTiO3 for x = 0, 0.05, and 0.10

J. Solid State Chem., 121 (1996), pp. 443–450

Bokhimi et al., 2002; X. Bokhimi, A. Morales, F. Pedraza

Crystallography and crystallite morphology of rutile synthesized at low temperature

J. Solid State Chem., 169 (2002), pp. 176–181

Bromiley et al., 2004; Bromiley, G., Hilaret, N., McCammon, C., 2004. Solubility of hydrogen and ferric iron in rutile and TiO2 (II): implications for phase assemblages during ultrahigh-pressure metamorphism and for the stability of silica polymorphs in the lower mantle. Geophysical Research Letters 31.

Chen and Fu, 2006; J. Chen, Z.D. Fu

Alpha-PbO2-type nanophase of TiO2 from coesite-bearing eclogite in the Dabie Mountains, China – Comment

Am. Mineral., 91 (2006), pp. 1699–1700

Dubrovinskaia et al., 2001; N.A. Dubrovinskaia, L.S. Dubrovinsky, R. Ahuja, V.B. Prokopenko, V. Dmitriev, H.P. Weber, J.M. Osorio-Guillen, B. Johansson

Experimental and theoretical identification of a new high-pressure TiO2 polymorph

Phys. Rev. Lett., 87 (2001), p. 275501

11

El Goresy et al., 2010; A. El Goresy, L. Dubrovinsky, P. Gillet, G. Graup, M. Chen

Akaogiite: an ultra-dense polymorph of TiO2 with the baddeleyite-type structure, in shocked garnet gneiss from the Ries Crater, Germany

Am. Mineral., 95 (2010), pp. 892–895

Escudero et al., 2011a; A. Escudero, L. Delevoye, F. Langenhorst

Aluminium Incorporation in TiO2 Rutile at High Pressure: An XRD and High-Resolution 27Al NMR Study

J. Phys. Chem. C, 115 (2011), pp. 12196–12201

Escudero and Langenhorst, 2011; Escudero, A., Langenhorst, F., 2011. Incorporation of Si into TiO2 phases at high pressure. American Mineralogist, in press.

Escudero et al., 2011b; A. Escudero, F. Langenhorst, W.F. Müller

Aluminum solubility in rutile at high pressure and experimental evidence for a CaCl2 structured polymorph

American Mineralogist, submitted (2011)

Escudero et al., 2012; Escudero, A., Miyajima, N., Langenhorst, F., 2012. Microstructure, composition and P-T conditions of rutile from diamondiferous gneiss of the Saxonian Erzgebirge, Germany. Chemie der Erde, doi: 10.1016/j.chemer.2011.11.001.

Grey et al., 1988; I.E. Grey, C. Li, I.C. Madsen, G. Braunshausen

TiO2-II. Ambient pressure preparation and structure refinement

Mater. Res. Bull., 23 (1988), pp. 743–753

Howard and Zhang, 2004; C.J. Howard, Z.M. Zhang

Structure for perovskites with layered ordering of A-site cations

Acta Crystallogr. Sect. B-Struct. Sci., 60 (2004), pp. 249–251

12

Hwang et al., 2000; S.L. Hwang, P.Y. Shen, H.T. Chu, T.F. Yui

Nanometer-size alpha-PbO2-type TiO2 in garnet: A thermobarometer for ultrahigh-pressure metamorphism

Science, 288 (2000), pp. 321–324

Jung, 2005; W.H. Jung

Synthesis and dc conductivity of Nd2/3TiO2.988

Mater. Lett., 59 (2005), pp. 2408–2411

Langenhorst et al., 1995; F. Langenhorst, P. Joreau, J.C. Doukhan

Thermal and shock metamorphism of the Tenham chondrite – a TEM examination

Geochim. Cosmochim. Acta, 59 (1995), pp. 1835–1845

Larson and Von Dreele, 1994; A.C. Larson, R.B. Von Dreele

GSAS: General Structural Analysis system

The Regents of the University of California, Los Alamos National Laboratory. Los Alamos, NM (1994)

Leinenweber et al., 1991; K. Leinenweber, W. Utsumi, Y. Tsuchida, T. Yagi, K. Kurita

Unquencheable high-pressure perovskite polymorphs of MnSnO3 and FeTiO3

Phys. Chem. Miner., 18 (1991), pp. 244–250

Massiot et al., 2002; D. Massiot, F. Fayon, M. Capron, I. King, S. Le Calve, B. Alonso, J.O. Durand, B. Bujoli, Z.H. Gan, G. Hoatson

Modelling one- and two-dimensional solid-state NMR spectra

Magn. Reson. Chem., 40 (2002), pp. 70–76

Ming et al., 2006; L.C. Ming, Y.H. Kim, T. Uchida, Y. Wang, M. Rivers

13

In situ X-ray diffraction study of phase transitions of FeTiO3 at high pressures and temperatures using a large-volume press and synchrotron radiation

Am. Mineral., 91 (2006), pp. 120–126

Mosenfelder et al., 2010; J.L. Mosenfelder, N. Kim, J.F. Stebbins

Silicon coordination in rutile and TiO2-II at ambient and high pressures: Si-29 NMR

Am. Mineral., 95 (2010), pp. 968–973

Müller, 2007; Müller, U., 2007. Inorganic structural Chemistry, 2nd ed. Wiley.

Olsen et al., 1999; J.S. Olsen, L. Gerward, J.Z. Jiang

On the rutile/alpha-PbO2-type phase boundary of TiO2

J. Phys. Chem. Solids, 60 (1999), pp. 229–233

Sato et al., 1991; H. Sato, S. Endo, M. Sugiyama, T. Kikegawa, O. Shimomura, K. Kusaba

Baddeleyite-type high-pressure phase of TiO2

Science, 251 (1991), pp. 786–788

Shannon, 1976; R.D. Shannon

Revised effective ionic-radii and systematic studies of interatomic distances in halides and chalcogenides

Acta Crystallographica Section A, 32 (1976), pp. 751–767

Slepetys and Vaughan, 1969; R.A. Slepetys, P.A. Vaughan

Solid solution of aluminum oxide in rutile titanium oxide

J. Phys. Chem., 73 (1969), pp. 2157–2162

Sobolev and Yefimova, 2000; N.V. Sobolev, E.S. Yefimova

14

Composition and petrogenesis of Ti-oxides associated with diamonds

Int. Geology Rev., 42 (2000), pp. 758–767

Stoyanov et al., 2007; E. Stoyanov, F. Langenhorst, G. Steinle-Neumann

The effect of valence state and site geometry on Ti L-3, L-2 and OK electron energy-loss spectra of TixOy phases

Am. Mineral., 92 (2007), pp. 577–586

Tang and Endo, 1993; J. Tang, S. Endo

P-T bounday of alpha-PbO2 type and baddeleyite type high-pressure phases of titanium-dioxide

J. Am. Ceram. Soc., 76 (1993), pp. 796–798

Tobbens et al., 2001; D.M. Tobbens, N. Stusser, K. Knorr, H.M. Mayer, G. Lampert

E9: the new high-resolution neutron powder diffractometer at the Berlin neutron scattering center

R. Delhez, E.J. Mittemeijer (Eds.), Epdic 7: European Powder Diffraction, Pts 1 and 2, Trans Tech Publications Ltd, Zurich-Uetikon (2001), pp. 288–293

Wilson et al., 2005; Wilson, N.C., Russo, S.P., Muscat, J., Harrison, N.M., 2005. High-pressure phases of FeTiO3 from first principles. Phys. Rev. B 72.

Withers et al., 2003; A.C. Withers, E.J. Essene, Y.X. Zhang

Rutile/TiO2II phase equilibria

Contrib. Mineral. Petrol., 145 (2003), pp. 199–204

Wu et al., 2005; X.L. Wu, D.W. Meng, Y.J. Han

Alpha-PbO2-type nanophase of TiO2 from coesite-bearing eclogite in the Dabie Mountains, China

15

Am. Mineral., 90 (2005), pp. 1458–1461

Wu et al., 2006; X.L. Wu, D.W. Meng, Y.J. Han

Alpha-PbO2-type nanophase of TiO2 from coesite-bearing eclogite in the Dabie Mountains, China – Reply

Am. Mineral., 91 (2006), pp. 1701–1702

Yoshii, 2000; K. Yoshii

Synthesis and magnetic properties of Ln2/3TiO3 (Ln = Pr and Nd)

J. Solid State Chem., 149 (2000), pp. 354–359

16

Figure captions

Figure 1. Selected portions of the X-ray diffraction patterns of Al2TiO5 calcined at 1300 °C at (a) 7 GPa, (b) 10 GPa, (c) 13.6 GPa, (d) 16 GPa, and (e) 20 GPa. R = TiO2 rutile, C = α-Al2O3 corundum, α = α-PbO2 type TiO2. The vertical guide lines are a help for the eye.

Figure 2. Al-doped α-PbO2 type TiO2 unit cell parameters (a, b, c and volume) plotted as a function of the synthesis pressure. The error bars are approximately the size of the symbols.

Figure 3. High-resolution 27Al MAS-NMR spectrum of the 85% TiO2 + 15% Al2O3 sample synthesized at 14 GPa and 1400 °C (black crosses). The fit of the spectrum (red line) as well as the individual contributions (gray lines) are also included. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

Figure 4. (A) TEM micrograph of an Al-doped α-PbO2 type TiO2 grain synthesized at 10 GPa and 1300 °C. Zone axes: [0 View the MathML source1¯ 2] α-PbO2 type TiO2; [0 View the MathML source1¯ 1] CaCl2-type TiO2. (B) High resolution TEM image. Zone axis [View the MathML source1¯ 1 0] α-PbO2 type TiO2.

Figure 5. (A) TEM micrograph of an Al-doped α-PbO2 type TiO2 grain of the sample synthesized at 16 GPa and 1300 °C. Zone axis [1 2 5]. (B) High resolution image. (C, D, and E) SAED patterns of the grains taken along different zone axes. α = α-PbO2 type TiO2; Pv = Perovskite phase. Zone axes α = α-PbO2 type TiO2 [3 2 View the MathML source7¯], [2 View the MathML source5¯ 1], and [0 View the MathML source1¯ 2], respectively. The extra and weak diffraction spots not corresponding to α-PbO2 type TiO2 have been indexed taking into account the RETiO3 perovskite structure (NdTiO3, orthorhombic symmetry, space group P nma (Amow and Greedan, 1996)), zone axes [View the MathML source2¯ 1 0], [1 View the MathML source1¯ 0], and [1 0 1], respectively, but are also compatible with the cation-deficient structures (Nd0.66TiO3, orthorhombic symmetry, space group P mmm (Yoshii, 2000)), zone axes [View the MathML source2¯ 2 1], [View the MathML source1¯ 0 1], and [0 1 0], respectively.

Figure 6. ELNES spectra of Al-doped α-PbO2 type TiO2 grains synthesized at 16 GPa and 1300 °C (a), and 12 GPa and 1300 °C. The spectrum shown in (c) corresponds to a Si-doped α-PbO2 type TiO2 grain synthesized at 12 GPa which shows 1.6 wt.% SiO2 (Escudero and Langenhorst, 2011).

Figure 7. TEM micrograph of an recovered Al-doped α-PbO2 type TiO2 grain synthesized at 20 GPa and 1300 °C. Zone axis [1 2 1¯].

Figure 8. Solubility of Al2O3 in TiO2 versus pressure at 1300 °C. Data have been obtained from EDX analyses on single crystals carried out on the TEM. The standard deviation of the measurements is also shown. The straight lines are only a help for the eye. The dotted line between 14 and 16 GPa indicates that the composition shown by the TiO2 grains synthesized at 16 GPa may not correspond to Al2O3 in TiO2 but to another aluminum titanium phase. Vertical guides indicate the reported transitions with pressure between the different TiO2 polymorphs. The α-PbO2-type TiO2 polymorph is the unique TiO2 phase observed in this study at synthesis pressures above 12 GPa.

17

Table 1

Table 1. Summary of the Al-doped α-PbO2 type TiO2 samples synthesized at high pressure at 1300 °C. The lattice parameters obtained from the Le Bail analysis and the Al2O3 compositions in TiO2 determined by TEM-EDX are also included. Numbers in brackets indicate standard deviations.

GPa a (Å) b (Å) c (Å) Unit cell volume (Å3) wt.% Al2O3

10 4.4662 5.5527 4.9325 122.3234 18(2)12 4.4454 5.5313 4.9102 120.7361 15.5(6)13.6 4.4394 5.5170 4.9004 120.0224 35(1)16 4.4128 5.5001 4.8914 118.7184 50(1)20 4.4408 5.5254 4.9080 120.4286 26.2(7)

18

Figure 1

19

Figure 2

20

Figure 3

21

Figure 4

22

Figure 5

23

Figure 6

24

Figure 7

25

Figure 8

26