Next Article in Journal
Performance and Characterization of Additively Manufactured BST Varactor Enhanced by Photonic Thermal Processing
Next Article in Special Issue
Telecom O-Band Quantum Dots Fabricated by Droplet Etching
Previous Article in Journal
Investigating Exchange Efficiencies of Sodium and Magnesium to Access Lithium from β-Spodumene and Li-Stuffed β-Quartz (γ-Spodumene)
Previous Article in Special Issue
Structural, Morphological, and Optical Properties of Nano- and Micro-Structures of ZnO Obtained by the Vapor–Solid Method at Atmospheric Pressure and Photocatalytic Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Growth and Characterization of High-Quality YTiO3 Single Crystals: Minimizing Ti4+ Containing Impurities and TiN Formation

Crystal Growth Facility, Institute of Physics, École Polytechnique Fédérale de Lausanne, CH-1015 Lausanne, Switzerland
*
Author to whom correspondence should be addressed.
Crystals 2024, 14(11), 989; https://doi.org/10.3390/cryst14110989
Submission received: 10 September 2024 / Revised: 11 November 2024 / Accepted: 11 November 2024 / Published: 16 November 2024

Abstract

:
We report the growth of YTiO3 single crystals using different starting materials with the nominal compositions, (1) stoichiometric YTiO3; (2) oxygen deficient YTiO2.925; (3) oxygen deficient YTiO2.85, and different atmospheres, (1) 97%Ar/3%H2; (2) Ar; (3) forming gas 95%N2/5%H2, using the laser floating zone growth technique. The oxygen-deficient starting materials were prepared by mixing Y2O3, Ti2O3, and Ti powder according to the YTiO3-δ stoichiometry. The addition of Ti powder to the starting materials effectively reacts with the oxygen in the floating zone furnace chamber, reducing Ti4+ ion-containing impurities. High-quality YTiO3 single crystals with (2 0 0) facet were grown from the starting materials corresponding to the nominal composition YTiO2.925. YTiO3 single crystals grown from different starting materials are characteristic of oxygen content of 3 in both pure crystals and crystals containing impurities, revealed by the same oxygen occupancy in single crystal X-ray diffraction measurements. When forming gas was used, a golden TiN coating formed on the surface of rod.

1. Introduction

RETiO3 (RE = rare earth elements) perovskites had been intensively studied, in particular since the discovery of high Tc superconductivity in rare earth cuprate perovskites [1,2,3,4]. In all RETiO3 compounds, except EuTiO3, Ti ions are trivalent (Ti3+), with 3d1 electronic configuration. It is well known that in high Tc cuprates, Cu2+ ions with 3d9 configuration have one unpaired spin, and the parent compounds exhibit antiferromagnetic interactions among Cu2+ ions. The two adjacent Cu2+ sites are bridged by the oxygen hole, and the hopping of these oxygen holes from site to site is responsible for the conductivity in high Tc cuprates [5]. A similar modification in electronic structure, governed by strong electron correlations in narrow d band, was anticipated in Mott-Hubbard insulators like RETiO3. The magnetic interactions and conductivity in this system can be tuned by substituting RE cations with varying ionic radius, from the largest La3+ cations to the smallest Gd3+ and Y3+ cations. Variation in ionic radii lead to uneven tiltings of the TiO6 octahedra, which alters the ratio between the electron bandwidth and Coulomb interaction [2,3,4,6]. For instance, LaTiO3, with the average Ti–O–Ti bond angle θ~157° (along the ac plane and the b axis), is a G-type antiferromagnet with a Néel temperature TN~140 K. In contrast, YTiO3 with a more bent Ti–O–Ti bond angle θ~142° is a ferromagnet with a Curie temperature of about Tc~30 K [1,2]. The evolution of magnetic transition had been well studied in the LaxY1-xTiO3 solid solution as a function of La/Y ratio [7,8]. Neutron scattering measurements have thoroughly investigated the magnetic fluctuations near the antiferromagnetic (AFM) to ferromagnetic (FM) phase transition boundary at a critical doping of x = 0.7 in LaxY1-xTiO3 [9,10]. Notably, the one-electron bandwidth and band filling in RETiO3 compounds can be tuned by both RE3+-cation-induced lattice distortions and the substitution of divalent Ca2+ or Sr2+ cations at RE sites [3]. By doping with Ca2+ or Sr2+ cations, a metal–insulator transition is introduced in La1-xSrxTiO3 and Y1-xCaxTiO3 systems [11,12,13,14,15,16,17,18,19,20]. In the end members SrTiO3 and CaTiO3, Ti4+ ions, instead of Ti3+ ions, balance the divalent Ca2+ or Sr2+ cations. Therefore, both Ti3+ and Ti4+ ions coexist in the doped compounds. Recently, by using femtosecond laser to excite the three phonon modes with center frequencies near 4, 9 and 17 THz in YTiO3 single crystals (Tc = 27 K), transient ferromagnetism up to 80 K was realized for the 9 THz oxygen rotation mode through the time-resolved magneto-optic Kerr effect (MOKE) measurements [21]. This finding strongly suggests that the FM phase transition can potentially be tuned up to 80 K by exploiting the strong spin-lattice coupling in YTiO3 [22,23].
The high-quality RETiO3 single crystals, as well as Ca2+- or Sr2+-doped samples, are highly desirable for studies of their physical property. Initially, RETiO3 single crystals were grown with Bridgman and Czochralski methods under an argon atmosphere [24,25]. However, thermalgravimetric analysis (TGA) indicated that the crystals were slightly oxidized, with the composition of LaTiO3.011 and CeTiO3.013 [25]. With the development of floating zone growth techniques [26], RETiO3 single crystals, including LaxY1-xTiO3, La1-xSrxTiO3 and Y1-xCaxTiO3, were grown [8,10,11,12,13,14,15,16,17,18,19,21,22,23,27,28]. It was realized that growing stoichiometric RETiO3, La1-xSrxTiO3 and Y1-xCaxTiO3 single crystals without excess oxygen is challenging. It was observed that magnetic transition temperatures decrease with increasing oxygen content in the samples. For instance, as nominal hole doping δ increases, LaTiO3+δ/2 undergoes a transition from an insulating phase to a metallic phase at δ~0.05, with the AFM transition disappearing at δ~0.08 [28]. In fact, with further oxidation, LaTiO3+δ/2 can transform into pyrochlore LaTiO3.5 (La2Ti2O7) phase with increasing excess oxygen δ [29] because the unpaired 3d1 electrons in Ti3+ ions are easily lost, forming Ti4+ with 3d0 configuration.
In this study, we thoroughly investigate the effect of growth conditions on the excess oxygen in YTiO3 single crystals by using the laser floating zone technique. Our results show that adding Ti powder to the starting materials is critical for growing YTiO3 single crystals with Tc~30 K, free from Ti4+-containing impurities. The Results and Discussion section is separated into the following three subsections: (1) growth from the starting materials with nominal composition YTiO3; (2) growth from the oxygen-deficient starting materials with nominal compositions YTiO2.925 and YTiO2.85; (3) the formation of TiN coating.

2. Materials and Methods

The laser-heated floating zone furnace (LFZ–Compact200EX, Model: TFL–1–P, TECHNO SEARCH CORPORATION, Tokyo, Japan) consists of 5 laser heads positioned circumferentially in a horizontal plane. The laser operates at a wavelength of 940 nm (invisible). The pressure in the chamber can be increased up to 10 bar when a thick quartz tube (thickness = 5 mm) is used.
Y2O3, Ti2O3, and Ti powder were used as starting materials. The rods were prepared with the following three nominal compositions: (1) YTiO3; (2) YTiO2.925; (3) YTiO2.85. The Y2O3, Ti2O3, and Ti powder were ground and mixed in an agate mortar. The mixture was pressed into rods with a diameter of ~6 mm and a length of ~100 mm using a pressing die, or loaded into a stripe balloon, and pressed under a pressure of 80 KN with a hydrostatic pressing system. The rods were sealed in quartz tubes, and sintered at 1100 °C for 24 h, as shown in Figure 1a,b. The floating zone growth conditions are summarized in Table 1.
The phase purity of the obtained crystals was checked by the powder X-ray diffraction (XRD) measurements using a PANalytical Empyrean X-ray diffractometer (Almelo, Overijssel, The Netherlands) equipped with a long-focused sealed X-ray tube of Cu-Kα radiation (λ = 1.5418 Å) and PIXcel 1D detector. Raw pattern reduction, profile fitting, phase identification and then the Le Bail fitting for both samples were performed with HighScore plus and ICDD PDF5+ [30].
The crystallinity of the crystals was confirmed by using a Laue X-ray diffractometer in backscattering geometry (MWL120 Real-Time Back-Reflection Laue Camera System, Multiwire Laboratories, Ithaca, NY, USA). High quality single crystals with suitable size were mounted onto the goniometer head of a Rigaku Synergy-I single crystal diffractometer (Rigaku Europe SE, Neu-Isenburg, Germany), equipped with Mo micro-focused source and Bantam HyPix X-Ray detector. Unit cell refinement, data collection & reduction, and absorption correction for all data sets were performed by CrysAlipro [31]. Structures were solved using ShelXT [32] within Olex2 [33] first, and then refined using ShelXL [34] within both Olex2 and ShelXle [35]. The final CIFs were edited with PublCIF [36] and passed IUCr checkCIF before depositing to CSD. The oxygen content in single crystals was checked through thermogravimetric analysis (TGA), which was conducted with a Setaram Labsys TGA-DSC thermal analyzer (KEP Technologies, Plan-les-Ouates, Switzerland).
Magnetization measurements were carried out using a Physical Property Measurement System (PPMS) DynaCool (Quantum Design, Inc., San Diego, CA, USA) equipped with a Vibrating Sample Magnetometer (VSM). Data were collected in both zero-field-cooling (ZFC) and field-cooling (FC) modes. For the ZFC mode, the sample was cooled down to the lowest temperature without applying a magnetic field, after which the magnetic field was applied and data were collected while warming up the sample. In the FC mode, data were collected while the sample was cooled down under an applied magnetic field.

3. Results and Discussion

3.1. Growth from the Starting Materials with the Nominal Composition YTiO3

We initiated the growth of single crystals using rods prepared with the stoichiometric YTiO3 composition. Figure 1c shows the XRD pattern obtained with the rods after sintering, prior to the single crystal growth. It reveals three phases, namely YTiO3, Y2O3 and Ti2O3. No Ti4+-containing phases are present in the rods. In the literature, a single-phase powder of YTiO3 was obtained, for instance by arc melting under argon atmosphere [20]. Polycrystalline LaxY1-xTiO3 (x = 0.15 and 0.3) samples were synthesized by ball milling followed by spark plasma sintering (SPS) at 1600 °C for 10 min [9]. It is well documented that the rods are at a risk of being partially oxidized when they are sintered at 1400 °C in an airtight chamber furnace under argon atmosphere [37].
After mounting the seed and feed rods in the floating zone furnace, the air in the chamber was purged by flowing argon gas over 6 h or by pumping for one hour followed by filling the chamber with a gas of mixture of 97% Ar + 3% H2. As the laser power increases, the rods melted and joined at 1800 °C, as indicated by the pyrometer. Three growth rates were applied to grow the crystals, as summarized in Table 1. Figure 2a shows an image of YTiO3 single crystals grown at a rate of 10 mm/h. The rod exhibits a shiny black appearance. When lower growth rates are applied (i.e., 5 and 2 mm/h), rods with similar appearance were obtained. However, they are fragile and tend to break into small pieces easily.
The Laue X-ray diffraction measurements were conducted to assess the crystallinity of the rod grown at 10 mm/h. Figure 2b,c show reflections taken from the surface of rod, perpendicular to its longitudinal direction indicating high crystallinity. Figure 2d,e present reflections from the rod cross-section. Figure 2f shows the tilted rod, with the Laue diffraction taken along this direction. The Laue spots form the two-fold symmetric pattern, as shown in Figure 2g, in agreement with the alignment of the crystal along one of the principal axes of orthorhombic YTiO3.
The result of the XRD measurement performed on a powder obtained by grinding part of the rod is shown in Figure 3. YTiO3 has a GdFeO3-type orthorhombic structure with the space group Pnma (62). The profile fitting of the XRD pattern yields lattice parameters a = 5.66971(9), b = 7.61722(9), and c = 5.33523(7) Å, respectively. In addition to the main YTiO3 phase, additional peaks are observed at 2θ = 15.16, 28.69, 29.61, 30.71, and 50.70°, respectively. Peaks at 2θ = 15.16, 29.61, and 30.71° are assigned to cubic Y2Ti2O7 and the peaks at 2θ = 28.69 and 50.70° to hexagonal Y2TiO5 phases, referring to the Y2O3-TiO2 phase diagram [38,39] and the structure data for hexagonal RE2TiO5 [40]. A comparison of orthorhombic YTiO3, cubic Y2Ti2O7 and hexagonal Y2TiO5 structures is shown in Figure 4. It should be pointed out that the α-Y2TiO5 with an orthorhombic structure is stable below 1330 °C, transforming to the hexagonal β-Y2TiO5 between 1330 °C and 1520 °C and to a fluorite-type phase at higher temperatures [38,39].
The presence of Y2Ti2O7 and Y2TiO5 in the rods after the floating growth process indicate Ti3+ is partially oxidized to Ti4+ even if hydrogen is present in the chamber. Previous reports on the growth of RETiO3 single crystals by the floating zone method [8,10,11,12,13,14,15,16,17,18,19,21,22,23,27,28] confirm the difficulty of avoiding Ti3+ oxidation. The presence of TiO2 in the Ti2O3 starting materials might be the cause of the formation of Ti4+ impurities in the crystals. It is also suggested that the residual oxygen in the chamber is responsible for the excess of oxygen in the crystals and the crystallization of Ti4+-containing impurities [27]. In conclusion, purging or evacuating the furnace is not efficient, as the chamber might be leaking and oxygen is inevitably introduced in the chamber during growth, which lasts for hours. A better solution is needed to impede Ti3+ oxidation and reduce the effect of the inevitable presence of a minute amount of TiO2 in Ti2O3.
The hydrogen gas was widely used to control the oxidation of Ti3+ ions during the crystal growth. However, our results clearly reveal that the hydrogen gas may not work as effectively as one assumed. In fact, TiO2 can be only slightly reduced to Ti2O3 under the hydrogen atmosphere at a high temperature. An effective way to reduce TiO2 to Ti metal was realized by using argon and hydrogen plasma [41]. Once Ti3+ ions are oxidized and transferred to Ti4+ ions, it is hard to reverse them to Ti3+ ions under the hydrogen atmosphere at 1800 °C.

3.2. Growth from the Oxygen-Deficient Starting Materials with Nominal Compositions YTiO2.925 and YTiO2.85

The idea behind the approach is to balance the excess oxygen by using the oxygen-deficient starting materials YTiO3-δ. In fact, this method had been successfully used in the previous work on the growth of RETiO3 single crystals. For example, to grow LaTiO3 and La0.95Sr0.05TiO3 single crystals, Ti, TiO2, La2O3, and SrTiO3 were weighed in prescribed ratios of LaTiO2.94 and La0.95Sr0.05TiO2.94 stoichiometry [13]. In a recent report on the growth of Y1-xLaxTiO3 ( x 0.25 ) and Y1-yCayTiO3 ( y 0.35 ) single crystals, the rods used for the crystal growth had nominal composition with oxygen deficiency Y1-xLa(Ca)xTiO3-δ [27]. In the case of YTiO3, several different values of δ in the range 0 δ 0.08 were tried with δ = 0.04 yielding the highest Tc~30 K [27].
In this study, rods with two nominal compositions (namely YTiO2.925 and YTiO2.825) were used for the crystal growth, as shown in Table 1. Crystal growth from the rod with nominal composition YTiO2.925 was performed at the growth rate of 4 and 10 mm/h. In both cases, the growth experiments were performed under an atmosphere of argon mixed with hydrogen. Figure 5a shows the rod obtained at 4 mm/h. Notably, the surface of the as-grown rod is coated by a grey crust, which can be mechanically removed using a razor blade. As observed, the rod is not completely round and two parallel flat surfaces run along the entire 70 mm length of the rod. The rod was cut into short sections and polished along the flat surfaces. XRD measurements revealed that the flat surface corresponds to the (2 0 0) plane of the YTiO3 crystal structure, allowing for the orientation of the a-axis in the rod, as shown in Figure 5b.
From the two-fold symmetric Laue pattern obtained from the flat surface of the rod, shown in Figure 5c,d, the crystallographic orientation of the rod can be determined. The b/c-axis is found tilted with an angle of approximately 20° from the longitudinal axis of the rod, as confirmed by Figure 5e,f. A similar tilt angle was obtained in the previous growth using rod with the nominal YTiO3 composition with 10mm/h growth rate.
One small disk cut from the middle part of the rod was crushed and ground into fine powder. Figure 6 shows the result of powder XRD measurements. The profile fitting of the XRD pattern confirms the absence of impurities. Pure YTiO3 phase is grown with the lattice parameters a = 5.69075(7), b = 7.61696(7), and c = 5.33880(8) Å, respectively. Compared to the lattice parameters obtained in Figure 3, the a-lattice parameter is substantially higher for the pure crystals.
Because the evaporation of Ti was observed when crystal was grown at a rate of 4 mm/h, an additional growth was performed with similar conditions but with a faster growth rate, i.e., 10 mm/h, and with a 2 bar pressure applied in the growth chamber. The objective was to reduce the evaporation of Ti and maintain the cationic stoichiometry of the rod in the crystal. Changing these parameters results in the absence of formation of the unwanted grey crust grey on the surface of the rod. The resulting crystal rod is almost perfectly circular, without the formation of a flat surface. The powder XRD pattern, similar to the one presented in Figure 6, confirm that a pure YTiO3 phase has been obtained.
In addition to powder X-ray diffraction measurements, YTiO3 single crystals grown under varying conditions were characterized using single-crystal X-ray diffraction at room temperature, with results summarized in Table 2. As shown, the lattice parameters obtained from powder and single-crystal X-ray diffraction measurements are comparable. The detailed structural information from single-crystal X-ray diffraction reveals minimal differences among single crystals grown under different conditions. Our findings rule out the possibility that excess oxygen atoms occupy interstitial sites (revealed by the low residual electron densities in structure refinement; See the Supplemental Materials). Moreover, our results do not support the presence of oxygen deficiency in single crystals grown from the nominal compositions YTiO2.925 and YTiO2.85. The TGA of the single crystal (same sample used for Figure 6) reveals that the excess oxygen is below 0.0126 in the formula YTiO3+0.0126; See the Supplemental Materials. Combining both powder and single crystal X-ray diffraction measurements, the slight variation among single crystals grown under different conditions is attributable to the presence of impurity phases rather than to the oxygen stoichiometry. We conclude that the oxygen content in YTiO3 single crystals is unambiguously stoichiometric, with an O content of 3.
Figure 7a shows the magnetic transition of YTiO3 single crystals grown under different conditions. For the crystals grown from the stoichiometric YTiO3 starting material, Tc varies between 25 K and 27 K. These crystals contain cubic Y2Ti2O7 and hexagonal Y2TiO5 impurity phases. For the crystals grown from the oxygen-deficient YTiO2.925 starting material, Tc increases to 30–31 K. The crystals grown from the starting material YTiO2.85 show the highest Tc~32 K; however, the floating zone growth with this starting material is unstable, leading to more defective and smaller crystals. The video clip available in the Supplemental Materials shows the formation of Ti bubbles in the molten zone, the bursting of which leads to the floating-zone instability during the growth using rods with the YTiO2.85 nominal composition. Figure 7b shows the magnetic field dependence of magnetization, M vs. H. The magnetization tends to saturate above H = 0.5 T. Crystals containing impurity phases, grown from the starting material YTiO3, exhibit a significantly lower magnetization, while pure crystals from oxygen-deficient materials have higher magnetization. Overall, single crystals grown from the starting materials YTiO2.925 show the best crystallinity and magnetic properties. The measurements of the heat capacity of the best crystals confirm the temperature of magnetic transition. Tc varies between 29.5 and 31 K, as shown in Figure 7c.

3.3. The Formation of TiN Coating

A TiN superconductor was accidentally formed during the growth of YTiO3 crystals using starting material with the YTiO2.85 nominal composition and hydrogen-containing Ar atmosphere. A leak in the reactor led to the formation of a rod with a golden coating, as can be seen in the Figure 8a. The powder XRD pattern recorded on the surface of the rod confirmed the presence of TiN, as shown in Figure 8b. The growth was repeated using the same starting material but using forming gas atmosphere. The obtained rod is homogeneously coated with TiN, as shown in Figure 8c and in the XRD pattern in Figure 8d.
The possible chemical reaction pathway for the formation of TiN from Ti2O3 could be [42]:
Ti2O3 + 1/10 N2 = 1/5 TiN + 3/5 Ti3O5
or Ti2O3 + 1/7N2 = 2/7 TiN + 3/7 Ti4O7
Based on these chemical reactions, the Ti2O3 compound is partially oxidized to form Ti3O5 or Ti4O7. As the valence of Ti increases, the formation of TiN also contributes to the formation of Ti4+-containing impurities.
The presence of TiN in the sample is further evidenced by magnetization measurements of the rod grown under forming gas atmosphere, as shown in Figure 9. Figure 9a shows that a superconducting transition is observed in the sample at ~4 K. Figure 9b shows the magnetic hysteresis loop measured at T = 3 K, which again confirms the superconducting state in the sample. For comparison, TiN single crystals exhibit a superconducting transition temperature of 6 K, while polycrystalline TiN shows a Tc of 1.7 K [43]. In the best TiN films, the superconducting transition temperature has been measured as high as 4.8 K [44]. Since TiN has a high melting point of 2947 °C, single crystals cannot be directly grown from polycrystalline TiN rods using the floating zone method. Our study demonstrates that TiN can be efficiently formed through the reaction of Ti2O3 with N2 by the floating zone method.

4. Conclusions

We have grown high-quality YTiO3 single crystals with visible (2 0 0) facets by using the laser floating zone technique. We observed that Tc decreases with the formation of Ti4+ ion-containing impurities, such as cubic Y2Ti2O7 and hexagonal Y2TiO5. Adding Ti powder to the starting materials to prepare the rods with the nominal compositions YTiO3-δ effectively minimizes the effect of residual air or the air leak on the formation of Ti4+-containing impurities. Ti addition allows a better control of the oxygen stoichiometry during the crystal growth, consequently maintaining the perovskite structure of YTiO3. Differences among crystals grown under varying conditions are attributable to impurity phases rather than crystal stoichiometry, which is confirmed as YTiO3. We recommend using YTiO2.925 as the optimal starting materials. The presence of larger quantity of Ti leads to unstable floating zone during the growth due to excessive Ti evaporation. Although slow growth may promote the formation of high-quality single crystals, it also increases the risk of oxidation during lengthy process. A growth rate of 4 mm/h is critical to obtain crystals with (2 0 0) facets. Growth under pressure could also mitigate the problem of air leaks. The presence of nitrogen in the chamber leads to the formation of superconductor TiN on the surface of the rods. Crystals grown in the best conditions exhibit a ferromagnetic phase transition between 29.5 and 31 K.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst14110989/s1, Video S1: Crystal grown from the starting materials YTiO2.85. Ti evaporation can be seen, leading to floating zone instability. Table S1. The details of structure refinement for YTiO3 single crystals at room temperature. The code 2392602, 2392601, and 2392600 are assigned to YTiO3 single crystals grown under different conditions: (A) from the nominal composition YTiO3 at a growth rate of 10 mm/h; (B) from the nominal composition YTiO2.925 at a growth rate of 4 mm/h; (C) from the nominal composition YTiO2.85 at a growth rate of 4 mm/h. Figure S1. Mass change of YTiO3 powder during heating under air up to 1100 °C. The Supplementary Materials S2 shows the results of single crystal XRD and thermal analysis measurements.

Author Contributions

Y.L. grew single crystals, performed X-ray diffraction, Laue, magnetization, thermal analysis and heat capacity measurements, and wrote the manuscript. D.W.B. performed X-ray diffraction and Laue measurements. A.M. and D.W.B. reviewed and edited the manuscript. A.M. supervised the project. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Single crystals and the data in this study are available from the corresponding authors upon request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. MacLean, D.A.; Ng, H.-N.; Greedan, J.E. Crystal structures and crystal chemistry of the RETiO3 perovskites: RE = La, Nd, Sm, Gd, Y. J. Solid State Chem 1979, 30, 35–44. [Google Scholar] [CrossRef]
  2. Greedan, J.E. The rare earth-titanium (III) perovskite oxides—An isostructural series with a remarkable variation in physical properties. J. Less-Common Met. 1985, 111, 335–345. [Google Scholar] [CrossRef]
  3. Tokura, Y. Fillingness dependence of electronic structures in strongly correlated electron systems Titanates and vanadates. J. Phys. Chm. Solids 1992, 53, 1619–1625. [Google Scholar] [CrossRef]
  4. Ishihara, S. Chapter 5. Titanates and Vanadates. In Physics of Transition Metal Oxides; Maekawa, S., Tohyama, T., Barnes, S.E., Ishihara, S., Koshibae, W., Khaliullin, G., Eds.; Springer Series in Solid-State Sciences 144; Springer: Berlin/Heidelberg, Germany, 2004; pp. 225–238. [Google Scholar]
  5. Guo, Y.; Langlois, J.-M.; Goddard, W.A. Electronic Structure and Valence-Bond Band Structure of Cuprate Superconducting Materials. Science 1988, 239, 896–899. [Google Scholar] [CrossRef]
  6. Mochizuki, M.; Imada, M. Orbital physics in the perovskite Ti oxides. New J. Phys. 2004, 6, 154. [Google Scholar] [CrossRef]
  7. Goral, J.P.; Greedan, J.E.; MacLean, D.A. Magnetic Behavior in the Series LaxY1-xTiO3. J. Solid State Chem. 1982, 43, 244. [Google Scholar] [CrossRef]
  8. Zhou, H.D.; Goodenough, J.B. Evidence for two electronic phases in Y1-xLaxTiO3 from thermoelectric and magnetic susceptibility measurements. Phys. Rev. B 2005, 71, 184431. [Google Scholar] [CrossRef]
  9. Li, B.; Louca, D.; Niedziela, J.; Li, Z.; Zhang, L.; Zhou, J.; Goodenough, J.B. Lattice and magnetic dynamics in perovskite Y1-xLaxTiO3. Phys. Rev. B 2016, 94, 224301. [Google Scholar] [CrossRef]
  10. Hameed, S.; El-Khatib, S.; Olson, K.P.; Yu, B.; Williams, T.J.; Hong, T.; Sheng, Q.; Yamakawa, K.; Zang, J.; Uemura, Y.J.; et al. Nature of the ferromagnetic-antiferromagnetic transition in Y1-xLaxTiO3. Phys. Rev. B 2021, 104, 024410. [Google Scholar] [CrossRef]
  11. Fujishima, Y.; Tokura, Y.; Arima, T.; Uchida, S. Optical-conductivity spectra of Sr1-xLaxTiO3: Filling-dependent effect of the electron correlation. Phys. Rev. B 1992, 46, 11167. [Google Scholar] [CrossRef]
  12. Tokura, Y.; Taguchi, Y.; Okada, Y.; Fujishima, Y.; Arima, T.; Kumagai, K.; Iye, Y. Filling dependence of electronic properties on the verge of metal–Mott-insulator transition in Sr1-xLaxTiO3. Phys. Rev. Lett. 1993, 70, 2126. [Google Scholar] [CrossRef]
  13. Okada, Y.; Arima, T.; Tokura, Y.; Murayama, C.; Môri, N. Doping- and pressure-induced change of electrical and magnetic properties in the Mott-Hubbard insulator LaTiO3 (La3+Ti3+O3). Phys. Rev. B 1993, 48, 9677. [Google Scholar] [CrossRef] [PubMed]
  14. Taguchi, Y.; Tokura, Y.; Arima, T.; Inaba, F. Change of electronic structures with carrier doping in the highly correlated electron system Y1-xCaxTiO3. Phys. Rev. B 1993, 48, 511. [Google Scholar] [CrossRef] [PubMed]
  15. Tokura, Y.; Taguchi, Y.; Moritomo, Y.; Kumagai, K.; Suzuki, T.; Iye, Y. Barely metallic states with enhanced carrier mass in Y1-xCaxTiO3. Phys. Rev. B 1993, 48, 14063. [Google Scholar] [CrossRef] [PubMed]
  16. Kumagai, K.; Suzuki, T.; Taguchi, Y.; Okada, Y.; Fujishima, Y.; Tokura, Y. Metal-insulator transition in La1-xSrxTiO3 and Y1-xCaxTiO3 investigated by specific-heat measurements. Phys. Rev. B 1993, 48, 7636. [Google Scholar] [CrossRef]
  17. Katsufuji, T.; Taguchi, Y.; Tokura, Y. Transport and magnetic properties of a Mott-Hubbard system whose bandwidth and band filling are both controllable: R1-xCaxTiO3+y/2. Phys. Rev. B 1997, 56, 10145. [Google Scholar] [CrossRef]
  18. Furukawa, Y.; Okamura, I.; Kumagai, K.; Goto, T.; Fukase, T.; Taguchi, Y.; Tokura, Y. La1-xSrxTiO3 by 47/49Ti and 139La nuclear magnetic resonance. Phys. Rev. B 1999, 59, 10550. [Google Scholar] [CrossRef]
  19. Tsubota, M.; Iga, F.; Nakano, T.; Uchihira, K.; Kura, S.; Takemura, M.; Bando, Y.; Umeo, K.; Takabatake, T.; Nishibori, E.; et al. Hole-doping and Pressure Effects on the Metal–Insulator Transition in Single Crystals of Y1-xCaxTiO3 (0.37<x<0.41). J. Phys. Soc. Jpn. 2003, 72, 3182–3188. [Google Scholar]
  20. Hays, C.C.; Zhou, J.-S.; Markert, J.T.; Goodenough, J.B. Electronic transition in La1-xSrxTiO3. Phys. Rev. B 1999, 60, 10367. [Google Scholar] [CrossRef]
  21. Disa, A.S.; Curtis, J.; Fechner, M.; Liu, A.; von Hoegen, A.; Först, M.; Nova, T.F.; Narang, P.; Maljuk, A.; Boris, A.V.; et al. Photo-induced high-temperature ferromagnetism in YTiO3. Nature 2023, 617, 73–78. [Google Scholar] [CrossRef]
  22. Komarek, A.C.; Roth, H.; Cwik, M.; Stein, W.-D.; Baier, J.; Kriener, M.; Bourée, F.; Lorenz, T.; Braden, M. Magnetoelastic coupling in RTiO3 (R = La, Nd, Sm, Gd, Y) investigated with diffraction techniques and thermal expansion measurements. Phys. Rev. B 2007, 75, 224402. [Google Scholar] [CrossRef]
  23. Knafo, W.; Meingast, C.; Boris, A.V.; Popovich, P.; Kovaleva, N.N.; Yordanov, P.; Maljuk, A.; Kremer, R.K.; v. Löhneysen, H.; Keimer, B. Ferromagnetism and lattice distortions in the perovskite YTiO3. Phys. Rev. B 2009, 79, 054431. [Google Scholar] [CrossRef]
  24. Garrett, J.D.; Greedan, J.E.; MacLean, D.A. Crystal growth and magnetic anisotropy of YTiO3. Mat. Res. Bull. 1981, 16, 145–148. [Google Scholar] [CrossRef]
  25. MacLean, D.A.; Greedan, J.E. Crystal growth, electrical resistivity, and magnetic properties of lanthanum titanate and cerium titanate Evidence for a metal-semiconductor transition. Inorg. Chem. 1981, 20, 1025–1029. [Google Scholar] [CrossRef]
  26. Kikugawa, N. Recent Progress of Floating-Zone Techniques for Bulk Single-Crystal Growth. Crystals 2024, 14, 552. [Google Scholar] [CrossRef]
  27. Hameed, S.; Joe, J.; Thoutam, L.R.; Garcia-Barriocanal, J.; Yu, B.; Yu, G.; Chi, S.; Hong, T.; Williams, T.J.; Freeland, J.W.; et al. Growth and characterization of large (Y, La)TiO3 and (Y, Ca)TiO3 single crystals. Phys. Rev. Mater. 2021, 5, 125003. [Google Scholar] [CrossRef]
  28. Taguchi, Y.; Okuda, T.; Ohashi, M.; Murayama, C.; Môri, N.; Iye, Y.; Tokura, Y. Critical behavior in LaTiO3+δ/2 in the vicinity of antiferromagnetic instability. Phys. Rev. B 1999, 59, 7917. [Google Scholar] [CrossRef]
  29. Lichtenberg, F.; Widmer, D.; Bednorz, J.G.; Williams, T.; Relier, A. Phase diagram of LaTiOx: From 2D layered ferroelectric insulator to 3D weak ferromagnetic semiconductor. Z. Phys. B-Condens. Matter 1991, 82, 211–216. [Google Scholar] [CrossRef]
  30. Degen, T.; Sadki, M.; Bron, E.; König, U.; Nénert, G. The HighScore suite. Powder Diffr. 2014, 29 (Suppl. S2), S13–S18. [Google Scholar] [CrossRef]
  31. Agilent. CrysAlis PRO; Agilent Technologies Ltd.: Yarnton, UK, 2014. [Google Scholar]
  32. Sheldrick, G.M. SHELXT—Integrated space-group and crystal. Acta Cryst. 2015, A71, 3–8. [Google Scholar]
  33. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Cryst. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  34. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, C71, 3–8. [Google Scholar]
  35. Hübschle, C.B.; Sheldrick, G.M.; Dittrich, B. ShelXle: A Qt graphical user interface for SHELXL. J. Appl. Cryst. 2011, 44, 1281–1284. [Google Scholar] [CrossRef] [PubMed]
  36. Westrip, S.P. publCIF: Software for editing, validating and formatting crystallographic information files. J. Appl. Cryst. 2010, 43, 920–925. [Google Scholar] [CrossRef]
  37. Roth, H. Single Crystal Growth and Electron Spectroscopy of d1-Systems. Ph.D. Thesis, The University of Cologne, Köln, Germany, 2008. [Google Scholar]
  38. Mizutani, N.; Tajima, Y.; Kato, M. Phase Relations in the System Y2O3-TiO2. J. Am. Ceram. Soc. 1976, 59, 168. [Google Scholar] [CrossRef]
  39. Villars, P. (Ed.) O-Ti-Y Vertical Section of Ternary Phase Diagram; Pauling File in: Inorganic Solid Phases, Springer Materials (online database); Springer: Berlin/Heidelberg, Germany, 2023; Available online: https://materials.springer.com/isp/phase-diagram/docs/c_0210439 (accessed on 1 January 2021).
  40. Shepelev, Y.F.; Petrova, M.A. Crystal Structures of Ln2TiO5 (Ln = Gd, Dy) Polymorphs. Inorg. Mater. 2008, 44, 1354–1361. [Google Scholar] [CrossRef]
  41. Palmer, R.A.; Doan, T.M.; Lloyd, P.G.; Jarvis, B.L.; Ahmed, N.U. Reduction of TiO2 with Hydrogen Plasma. Plasma Chem. Plasma Process. 2002, 22, 335–350. [Google Scholar] [CrossRef]
  42. Zheng, Q.; Li, Y.; Ma, C.; Sun, J.; Li, H.; Yue, X. Formation mechanism of the TiN containing coating on the Al2O3-Ti2O3 composite at 1400 °C nitrogen-blowing. Ceram. Int. 2023, 49, 13119–13124. [Google Scholar] [CrossRef]
  43. Spengler, W.; Kaiser, R.; Christensen, A.N.; Müller-Vogt, G. Raman scattering, superconductivity, and phonon density of states of stoichiometric and nonstoichiometric TiN. Phys. Rev. B 1978, 17, 1095–1101. [Google Scholar] [CrossRef]
  44. Torgovkin, A.; Chaudhuri, S.; Ruhtinas, A.; Lahtinen, M.; Sajavaara, T.; Maasilta, I.J. High quality superconducting titanium nitride thin film growth using infrared pulsed laser deposition. Supercond. Sci. Technol. 2018, 31, 055017. [Google Scholar] [CrossRef]
Figure 1. (a) Rods are wrapped in a tantalum foil and sealed in quartz ampoules. (b) The grey rods after sintering at 1100 °C for 24 h. (c) The XRD pattern shows three phases: YTiO3, Y2O3, and Ti2O3, in the sintered rod.
Figure 1. (a) Rods are wrapped in a tantalum foil and sealed in quartz ampoules. (b) The grey rods after sintering at 1100 °C for 24 h. (c) The XRD pattern shows three phases: YTiO3, Y2O3, and Ti2O3, in the sintered rod.
Crystals 14 00989 g001
Figure 2. (a) YTiO3 single crystals grown from the starting materials YTiO3 (growth rate 10 mm/h). (b) Laue X-ray diffraction pattern was performed on the surface of the rod after growth in the floating zone furnace. (c) The Laue pattern obtained from the surface of the rod. (d) The Laue X-ray diffraction measurement was performed on the cross-section of the rod. (e) The Laue pattern obtained from the cross-section of the rod. (f) The rod was titled until a symmetric Laue pattern was obtained. (g) The two-fold symmetric Laue pattern taken from the tilted rod.
Figure 2. (a) YTiO3 single crystals grown from the starting materials YTiO3 (growth rate 10 mm/h). (b) Laue X-ray diffraction pattern was performed on the surface of the rod after growth in the floating zone furnace. (c) The Laue pattern obtained from the surface of the rod. (d) The Laue X-ray diffraction measurement was performed on the cross-section of the rod. (e) The Laue pattern obtained from the cross-section of the rod. (f) The rod was titled until a symmetric Laue pattern was obtained. (g) The two-fold symmetric Laue pattern taken from the tilted rod.
Crystals 14 00989 g002
Figure 3. Profile fitting of powder XRD data from the crystal grown from the starting materials with nominal composition YTiO3 at a growth rate of 10 mm/h. The insets highlight the peaks assigned to the Y2Ti2O7 and Y2TiO5 impurities, as indicated by the arrows. The red dotted line, Yobs, presents the raw data, the black solid line, Ycalc by the Le Bail fitting using the HighScore plus program and ICDD PDF5+ data base, the red line Yobs-Ycalc, Rwp and GOOF (goodness of fit) values indicate the quality of fitting. The blue vertical bars indicate the Bragg peak position in the calculated pattern.
Figure 3. Profile fitting of powder XRD data from the crystal grown from the starting materials with nominal composition YTiO3 at a growth rate of 10 mm/h. The insets highlight the peaks assigned to the Y2Ti2O7 and Y2TiO5 impurities, as indicated by the arrows. The red dotted line, Yobs, presents the raw data, the black solid line, Ycalc by the Le Bail fitting using the HighScore plus program and ICDD PDF5+ data base, the red line Yobs-Ycalc, Rwp and GOOF (goodness of fit) values indicate the quality of fitting. The blue vertical bars indicate the Bragg peak position in the calculated pattern.
Crystals 14 00989 g003
Figure 4. (a) Crystal structure of orthorhombic YTiO3. (b) Pyrochlore Y2Ti2O7 structure with F d 3 ¯ m cubic space group. (c) Hexagonal Y2TiO5 with space group P63/mmc. The red, purple and grey spheres represent the O atoms, the Ti atoms, and the Y atoms, respectively.
Figure 4. (a) Crystal structure of orthorhombic YTiO3. (b) Pyrochlore Y2Ti2O7 structure with F d 3 ¯ m cubic space group. (c) Hexagonal Y2TiO5 with space group P63/mmc. The red, purple and grey spheres represent the O atoms, the Ti atoms, and the Y atoms, respectively.
Crystals 14 00989 g004
Figure 5. (a) YTiO3 single crystals grown from the starting materials with nominal composition YTiO2.925 at a growth rate of 4 mm/h. The rod has two parallel flat surfaces. The inset shows the flat surface after polishing. (b) The XRD measurement reveals that the flat surface corresponds to the (2 0 0) plane. (c) The Laue X-ray diffraction measurement performed on the flat surface. (d) The two-fold symmetric Laue pattern obtained from the flat surface. (e) View of the rod tilted to obtain a symmetric Laue pattern. (f) The two-fold symmetric Laue pattern taken from the tilted rod.
Figure 5. (a) YTiO3 single crystals grown from the starting materials with nominal composition YTiO2.925 at a growth rate of 4 mm/h. The rod has two parallel flat surfaces. The inset shows the flat surface after polishing. (b) The XRD measurement reveals that the flat surface corresponds to the (2 0 0) plane. (c) The Laue X-ray diffraction measurement performed on the flat surface. (d) The two-fold symmetric Laue pattern obtained from the flat surface. (e) View of the rod tilted to obtain a symmetric Laue pattern. (f) The two-fold symmetric Laue pattern taken from the tilted rod.
Crystals 14 00989 g005
Figure 6. Profile fitting of powder XRD data from the crystal grown from the starting materials with nominal composition YTiO2.925 at a growth rate of 4 mm/h. The red dotted line, Yobs, presents the raw data, the black solid line, Ycalc by the Le Bail fitting using the HighScore plus program and ICDD PDF5+ data base, the red line Yobs-Ycalc, Rwp and GOOF (goodness of fit) values indicate the quality of fitting.
Figure 6. Profile fitting of powder XRD data from the crystal grown from the starting materials with nominal composition YTiO2.925 at a growth rate of 4 mm/h. The red dotted line, Yobs, presents the raw data, the black solid line, Ycalc by the Le Bail fitting using the HighScore plus program and ICDD PDF5+ data base, the red line Yobs-Ycalc, Rwp and GOOF (goodness of fit) values indicate the quality of fitting.
Crystals 14 00989 g006
Figure 7. (a) The temperature dependence of the magnetic susceptibility of single-crystal crushed powder samples from the single crystals grown from the starting materials YTiO3, YTiO2.925 and YTiO2.85. (b) The magnetic field dependence of magnetization for the same powder samples as those in (a). (c) Heat capacity as a function of temperature for the crystals grown from rods with an oxygen-deficient YTiO2.925 composition.
Figure 7. (a) The temperature dependence of the magnetic susceptibility of single-crystal crushed powder samples from the single crystals grown from the starting materials YTiO3, YTiO2.925 and YTiO2.85. (b) The magnetic field dependence of magnetization for the same powder samples as those in (a). (c) Heat capacity as a function of temperature for the crystals grown from rods with an oxygen-deficient YTiO2.925 composition.
Crystals 14 00989 g007
Figure 8. (a) YTiO3 single crystals grown from the starting materials YTiO2.85 at 10 mm/h under Ar + H2 atmosphere. A thin golden layer formed on the surface of the rod due to the air leak. (b) The XRD pattern shows the presence of TiN phase in the sample shown in (a). (c) A homogeneous gold-like coating on the YTiO3 single crystal obtained when growth is performed under forming gas. (d) The XRD measurement was performed on the surface of the rod and TiN phase is identified.
Figure 8. (a) YTiO3 single crystals grown from the starting materials YTiO2.85 at 10 mm/h under Ar + H2 atmosphere. A thin golden layer formed on the surface of the rod due to the air leak. (b) The XRD pattern shows the presence of TiN phase in the sample shown in (a). (c) A homogeneous gold-like coating on the YTiO3 single crystal obtained when growth is performed under forming gas. (d) The XRD measurement was performed on the surface of the rod and TiN phase is identified.
Crystals 14 00989 g008
Figure 9. (a) The temperature dependence of magnetization reveals the superconducting transition at about 4 K in the powder obtained by grinding the rod grown under forming gas. (b) The magnetic hysteresis loop measured at T = 3 K also evidences the presence of the TiN superconducting phase in the powder.
Figure 9. (a) The temperature dependence of magnetization reveals the superconducting transition at about 4 K in the powder obtained by grinding the rod grown under forming gas. (b) The magnetic hysteresis loop measured at T = 3 K also evidences the presence of the TiN superconducting phase in the powder.
Crystals 14 00989 g009
Table 1. Summary of the rod preparation conditions and of the parameters during the crystal growth by laser floating zone technique.
Table 1. Summary of the rod preparation conditions and of the parameters during the crystal growth by laser floating zone technique.
Starting materialsStoichiometric ratio YTiO3Y2O3 + Ti2O3 → 2YTiO3
Oxygen deficient YTiO2.925Y2O3 + 0.95Ti2O3 + 0.1Ti → 2YTiO2.925
Oxygen deficient YTiO2.85Y2O3 + 0.9Ti2O3 + 0.2Ti → 2YTiO2.85
Gas atmosphereArgon
97% Ar + 3% H2
95% N2 + 5% H2
Growth rateSlow 2 mm/hThe growth rate indicates the travelling speed of seed rod. The feed rod moves 5–10% faster than the seed rod to maintain the stable molten zone and uniform shape of as-grown rod.
Middle 4–5 mm/h
Fast 10 mm/h
PressureAtmosphere pressure Gas flow rate 80–120 mL/min
2 bar97% Ar + 3% H2
Rotation speed15 rpmThe upper and lower shafts rotate in an opposite direction but at the same speed.
Table 2. Results of single crystal X-ray diffraction measurements of YTiO3 at room temperature. YTiO3 single crystals grown under different conditions: (A) from the nominal composition YTiO3 at a growth rate of 10 mm/h; (B) from the nominal composition YTiO2.925 at a growth rate of 4 mm/h; (C) from the nominal composition YTiO2.85 at a growth rate of 4 mm/h.
Table 2. Results of single crystal X-ray diffraction measurements of YTiO3 at room temperature. YTiO3 single crystals grown under different conditions: (A) from the nominal composition YTiO3 at a growth rate of 10 mm/h; (B) from the nominal composition YTiO2.925 at a growth rate of 4 mm/h; (C) from the nominal composition YTiO2.85 at a growth rate of 4 mm/h.
Batch ABC
a (Å)5.6807(3)5.6860(3)5.6995(1)
b (Å)7.6204(4)7.6108(3)7.6297(2)
c (Å)5.3391(2)5.3378(2)5.3455(1)
V3)231.13(2)230.99(2)232.45(1)
Bond angles (°)
Ti1—O1—Ti1
143.69(9)144.04(9)143.35(6)
Bond angles (°)
Ti1—O2—Ti1
140.97(13)140.73(13)140.09(10)
Atoms
Y(1)x0.42721(6)0.42719(7)0.42690(4)
y0.250.250.25
z0.02056(6)0.02060(6)0.02084(5)
Occ1 4c1 4c1 4c
Ti(1)x000
y000
z000
Occ1 4a1 4a1 4a
O(1)x0.04130(4)0.04140(4)0.04290(3)
y0.250.250.25
z−0.11860(5)−0.11930(5)−0.12120(3)
Occ1 4c1 4c1 4c
O(2)x0.19050(3)0.19160(3)0.19030(2)
y0.05800(2)0.05750(2)0.05883(16)
z0.30900(3)0.3090(3)0.30970(2)
Occ1 8d1 8d1 8d
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, Y.; Bi, D.W.; Magrez, A. Growth and Characterization of High-Quality YTiO3 Single Crystals: Minimizing Ti4+ Containing Impurities and TiN Formation. Crystals 2024, 14, 989. https://doi.org/10.3390/cryst14110989

AMA Style

Liu Y, Bi DW, Magrez A. Growth and Characterization of High-Quality YTiO3 Single Crystals: Minimizing Ti4+ Containing Impurities and TiN Formation. Crystals. 2024; 14(11):989. https://doi.org/10.3390/cryst14110989

Chicago/Turabian Style

Liu, Yong, David Wenhua Bi, and Arnaud Magrez. 2024. "Growth and Characterization of High-Quality YTiO3 Single Crystals: Minimizing Ti4+ Containing Impurities and TiN Formation" Crystals 14, no. 11: 989. https://doi.org/10.3390/cryst14110989

APA Style

Liu, Y., Bi, D. W., & Magrez, A. (2024). Growth and Characterization of High-Quality YTiO3 Single Crystals: Minimizing Ti4+ Containing Impurities and TiN Formation. Crystals, 14(11), 989. https://doi.org/10.3390/cryst14110989

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop