リケラボ論文検索は、全国の大学リポジトリにある学位論文・教授論文を一括検索できる論文検索サービスです。

リケラボ 全国の大学リポジトリにある学位論文・教授論文を一括検索するならリケラボ論文検索大学・研究所にある論文を検索できる

リケラボ 全国の大学リポジトリにある学位論文・教授論文を一括検索するならリケラボ論文検索大学・研究所にある論文を検索できる

大学・研究所にある論文を検索できる 「Phase and Size Selective Crystal Growth of Nanoparticles under Supercritical Water」の論文概要。リケラボ論文検索は、全国の大学リポジトリにある学位論文・教授論文を一括検索できる論文検索サービスです。

コピーが完了しました

URLをコピーしました

論文の公開元へ論文の公開元へ
書き出し

Phase and Size Selective Crystal Growth of Nanoparticles under Supercritical Water

藤田, 知樹 筑波大学 DOI:10.15068/0002005507

2022.11.17

概要

Physical properties of materials have a close relation with their atomic arrangement. Size, surface, interface and lattice defects of materials are relating to properties of materials such as electric and/or thermal conductivity. Fine nanoparticles below 10 nm size show interesting physical properties such as a narrower band gap or a lower melting point than bulk materials.

 Supercritical hydrothermal synthesis has attracted a lot of attention as a method to synthesize fine metal oxide nanoparticles with uniform size and shape [1]. Nanoparticles are prepared in the water above its critical temperature and pressure of 673 K and 22.1 MPa. Supercritical water is placed between a liquid phase and a gas phase on the phase diagram. Density and dielectric constant of supercritical water are continuously varied with temperature and pressure. Highly functional nanoparticles with large specific surface area can be synthesized in supercritical water [2]. Complex nanoparticles can be synthesized at lower temperature than a conventional method [3]. Mass production of nanoparticles is possible due to its short reaction time. It typically takes a few seconds to one hour.

 Supercritical water has superior properties as a reaction field for preparing fine nanoparticles with controllable functionality [4]. Lower solubility can suppress crystal growth through Ostwald ripening in conventional hydrothermal synthesis in a liquid phase. A higher density than that of a gas phase can prevent the collision and aggregation between nucleuses of nanoparticles which are frequently observed in gas phase reaction like a chemical vapor decomposition (CVD) method. Organic solvent can be dispersed in supercritical water due to its low dielectric constants. This makes it possible to synthesize nanoparticles with a highly reactive surface modified by organic molecules. Controlling structure of fine functional nanoparticles is possible by changing properties of a reaction field continuously.

 Control of crystalline phase is one of the most important topics of supercritical hydrothermal synthesis of nanoparticle. Functional metal oxide nanoparticles have several crystalline phases. It strongly affects their physical or chemical properties [5], [6]. Phase stability depends on the size of nanoparticles and/or a precursor used in the synthesis [7], [8]. Various crystalline phases were observed for nanoparticles synthesized in supercritical water [3], [9]–[11]. Controllable parameters of synthesis are temperature, pressure, pH of solution, type of solute molecule, ionic composition of solution, and reaction time. It is difficult to carry out complete investigation for all combinations of controllable parameters.

 Physical properties of nanoparticles can be changed by local defects in atomic arrangement. Disorder at interface of core-shell structure of nanoparticles attracts attention in terms of controlling optical resonance of quantum dots [12]. Strain induced at interface by lattice mismatch causes upper offset of valence band and red shift of photoluminescence spectra. The improvement of thermoelectric performance was reported for composite nanoparticles [13]. The high density of twin faulting led to higher phonon scattering and lower thermal conductivity. A figure of merit was higher by 50 % than bulk material.

 Methods of structure determination in atomic-to-nanoscale are powerful tools, which can help with the development of a new area of materials science which focuses on local structure of nanoparticles. Simultaneous determination of size, shape and local defect of nanoparticles provides important clues to identify structural parameters relating to important properties of nanoparticles. Elucidating local structure in fine nanoparticles enables us to create a guideline for an improvement of nanomaterials properties which depends on local lattice defects.

 Monitoring of nanoparticles structure during synthesis can provide clues to phase stability, mechanism of growth or correlation between structure of nanoparticles and physical properties of supercritical water. The critical size between polymorphs of nanoparticles is reported from observation of size and crystalline phase of nanoparticles [8]. Surface stress of nanoparticles was investigated from anomalous lattice expansion or contraction of nanoparticles [14]. The measurement at multi-temperature and pressure conditions can provide an evidence of correlation between microscopic structure and macroscopic properties of a reaction field.

 This “in-situ” measurement of nanostructure during synthesis requires intense beam with an angstrom wavelength. An X-ray, a neutron and an electron beam are commonly used for evaluating structure of materials in atomic and nanometer scale [15]. Both X-ray and neutron beams can penetrate the reaction vessel, while electron beam is scattered by electrons on the wall of vessel. A neutron beam is a tool for evaluation of structure of materials composed with light atoms like lithium and oxygen [16]. A neutron experiment requires a sample size of the order of centimeter due to its low intensity per area. Electron diffraction and microscopy can be carried out in milliseconds due to strong interaction between electrons [17]. High vacuum should be kept around samples for preventing the electron from being scattered by air.

 Synchrotron radiation (SR) is an intense X-ray source. A highly intense and parallelized X-ray are emitted by accelerated relativistic electrons in a magnetic field with GeV order energy. The brilliance at the third generation SR facility is more than 1010 times higher than that of an X-ray tube. The high brilliance decreases the time for collecting powder X-ray diffraction (PXRD) data which can be used for structure analysis. A large space can be assured around samples in the setup of SR-PXRD measurement, since a highly parallelized X-ray does not require optics elements such as soller slit. The various measurement systems can be installed for conducting measurement at the high temperature, the high pressure or the vacuum conditions.

 SR is a powerful tool for evaluation of structure of nanoparticles. Diffraction data can be collected from samples with weight of microgram. It is possible to observe the weak scattering from nanoparticles, discriminating it with scattering from a pressure vessel. The composition of crystalline phase of nanoparticles can be identified from a set of positions of diffraction lines. The average diameter of nanoparticles can be estimated from width of diffraction profiles.

 In-situ SR-PXRD measurement has been carried out for observation of nucleation and crystal growth of nanoparticles during synthesis [18], [19]. A reaction vessel or equipment for synthesis was installed at a sample position. PXRD data were collected from nanoparticles during synthesis with a second time scale. Reaction processes of various synthesis were investigated. An anatase phase of TiO2 nanoparticle was monitored in reaction of a sol-gel method [20]. The formation of Ru nanoparticle with fcc and hcp structure was observed in solvothermal reaction with supercritical ethanol [21]. Phase transition during the mechanochemical reaction was reported for ball- milling process [22].

 High quality PXRD data are required for determination of structure of nanoparticles from in-situ experiment. Diffraction data should be collected simultaneously from a wide range of diffraction angles. In the case of 10 nm ZrO2, data with resolution d > 0.6 Å were required for determining lattice constants with 0.001 Å accuracy. The broadening of diffraction profiles due to energy dispersion should be minimized for estimating a size of nanoparticles. Crystal growth was investigated at the beamline with the 10-4 scale of the energy dispersion [23].

 Evaluation of size and lattice constants can be complicated due to local lattice defects. For example, a shift and broadening of diffraction lines can be caused by planar defects such as twin and deformation faulting in the stacking structure [24]. Lattice defects of nanoparticles cause a greater change to PXRD data than those of bulk material, since larger fraction of atoms is related to defect structures. We found the height of 200 reflection of Ag nanosphere was almost half of the calculated data of Rietveld refinement. Several reflections of ZrO2 nanoparticle prepared by supercritical hydrothermal synthesis were not fitted by Rietveld refinement.

 Issues about reproducibility of parameters obtained from analysis of diffraction data of in-situ SR-PXRD experiment were reported by Iversen et al. [25]. In work presented by this group, diffraction data were collected 10 times for in-situ experiments of hydrothermal synthesis in identical conditions. In Rietveld refinement, the absolute value of an intensity scaling factor of diffraction data were not reproduced. It was partly due to occasional movement of nanoparticles in initial heating process, since it made difficult to control amount of samples exposed by an X-ray. The diameter of nanoparticles estimated from diffraction line width showed uncertainty of about 3 nm. Iversen et al. found that diffraction lines with relatively narrow width corresponding to 15 nm diameter required a careful treatment than those corresponding to the diameter of less than 10 nm. The time dependence of an intensity scaling factor and that of the diameter were qualitatively reproduced for 10 times of experiments.

 The methods have been developed for analyzing powder diffraction data of nanoparticles. Leonardi et al. reported effect of dislocation to PXRD data for the rod-shaped nanoparticle of iridium and palladium [26]. Bertolotti et al. investigated atomic arrangement of colloidal nanoplatelet [27]. Parakh et al. reported high pressure study about gold nanoparticles structure [28]. Atomic arrangements were modeled by using a combination of diffraction data calculation and a molecular dynamics simulation. Some of these studies were performed on the large-scale parallel computing platform.

 In the present study, we carried out in-situ SR-PXRD experiment to investigate the crystal growth of nanoparticles in supercritical water. The measurement system was developed for in-situ SR-PXRD experiment at the beamline BL02B2 of the third generation SR facility SPring-8. Powder diffraction data were collected at thirteen combinations from the synthesis of ZrO2 nanoparticles. The analysis method was developed for analyzing powder diffraction data of nanoparticles in the reaction vessel. Debye Scattering Equation (DSE) was used for the calculation of powder profiles of nanoparticles. DSE was used for the analysis of a series of the powder diffraction data of nanoparticles by Ozawa [29]. The calculation was performed on the large-scale parallel computing platform with GPU.

 The crystal growth of ZrO2 nanoparticles in supercritical water was investigated by in-situ SR-PXRD experiment. The crystalline phases and the size of nanoparticles were determined from the analysis of the powder diffraction data. The relation between the crystalline phases, size of nanoparticles and synthesis conditions were investigated by the analysis.

この論文で使われている画像

参考文献

[1] T. Adschiri, K. Kanazawa, and K. Arai, “Rapid and Continuous Hydrothermal Crystallization of Metal Oxide Particles in Supercritical Water,” J. Am. Ceram. Soc., vol. 75, pp. 1019–1041, 1992, doi: 10.1111/j.1151-2916.1992.tb04179.x.

[2] Q. X. Zheng et al., “Synthesis of YVO4 and rare earth-doped YVO4 ultra-fine particles in supercritical water,” J. Supercrit. Fluids, vol. 46, pp. 123–128, 2008, doi: 10.1016/j.supflu.2008.04.014.

[3] H. Hayashi and Y. Hakuta, “Hydrothermal Synthesis of metal oxide nanoparticles in supercritical water,” Materials, vol. 3, no. 7, pp. 3794–3817, 2010, doi: 10.3390/ma3073794.

[4] T. Adschiri and A. Yoko, “Supercritical fluids for nanotechnology,” J. Supercrit. Fluids, vol. 134, pp. 167– 175, 2018, doi: 10.1016/j.supflu.2017.12.033.

[5] R. C. Garvie, R. H. Hannink, and R. T. Pascoe, “Ceramic steel?,” Nature, vol. 258, pp. 703–704, 1975, doi: 10.1038/258703a0.

[6] J. Panpranot, K. Kontapakdee, and P. Praserthdam, “Effect of TiO2 crystalline phase composition on the physicochemical and catalytic properties of Pd/TiO2 in selective acetylene hydrogenation,” J. Phys. Chem. B, vol. 110, pp. 8019–8024, 2006, doi: 10.1021/jp057395z.

[7] Z. Wang, D. Xia, G. Chen, T. Yang, and Y. Chen, “The effects of different acids on the preparation of TiO2 nanostructure in liquid media at low temperature,” Mater. Chem. Phys., vol. 111, pp. 313–316, 2008, doi: 10.1016/j.matchemphys.2008.04.015.

[8] S. Shukla and S. Seal, “Mechanisms of room temperature metastable tetragonal phase stabilisation in zirconia,” Int. Mater. Rev., vol. 50, pp. 45–64, 2005, doi: 10.1179/174328005X14267.

[9] Z. Fang, H. Assaaoudi, H. Lin, X. Wang, I. S. Butler, and J. A. Kozinski, “Synthesis of nanocrystalline SnO2 in supercritical water,” J. Nanopart. Res., vol. 9, pp. 683–687, 2007, doi: 10.1007/s11051-006-9171-9.

[10] N. Talebian and F. Jafarinezhad, “Morphology-controlled synthesis of SnO2 nanostructures using hydrothermal method and their photocatalytic applications,” Ceram. Inter., vol. 39, pp. 8311–8317, 2013, doi: 10.1016/j.ceramint.2013.03.101.

[11] K. Sue, K. Kimura, K. Murata, and K. Arai, “Effect of cations and anions on properties of zinc oxide particles synthesized in supercritical water,” J. Supercrit. Fluids, vol. 30, pp. 325–331, 2004, doi: 10.1016/j.supflu.2003.09.009.

[12] S. Mangel, L. Houben, and M. Bar Sadan, “The effect of atomic disorder at the core-shell interface on stacking fault formation in hybrid nanoparticles,” Nanoscale, vol. 8, pp. 17568–17572, 2016, doi: 10.1039/c6nr04867f.

[13] M. Ibáñez et al., “Crystallographic control at the nanoscale to enhance functionality: Polytypic Cu2GeSe3 nanoparticles as thermoelectric materials,” Chemistry of Materials, vol. 24, no. 23, pp. 4615–4622, Dec. 2012, doi: 10.1021/cm303252q.

[14] P. M. Diehm, P. Ágoston, and K. Albe, “Size-dependent lattice expansion in nanoparticles: Reality or anomaly?,” ChemPhysChem, vol. 13, pp. 2443–2454, 2012, doi: 10.1002/cphc.201200257.

[15] C. Weidenthaler, “Pitfalls in the characterization of nanoporous and nanosized materials,” Nanoscale, vol. 3, pp. 792–810, 2011, doi: 10.1039/c0nr00561d.

[16] H. Akiba et al., “Nanometer-Size Effect on Hydrogen Sites in Palladium Lattice,” J. Am. Chem. Soc., vol. 138, pp. 10238–10243, 2016, doi: 10.1021/jacs.6b04970.

[17] I. Lignos, S. Stavrakis, A. Kilaj, and A. J. deMello, “Millisecond-Timescale Monitoring of PbS Nanoparticle Nucleation and Growth Using Droplet-Based Microfluidics,” Small, vol. 11, pp. 4009–4017, 2015, doi: 10.1002/smll.201500119.

[18] R. I. Walton and D. O’Hare, “Watching solids crystallise using in situ powder diffraction,” Chem. Commun., pp. 2283–2291, 2000, doi: 10.1039/b007795j.

[19] K. M. Ø. Jensen, C. Tyrsted, M. Bremholm, and B. B. Iversen, “In situ studies of solvothermal synthesis of energy materials,” ChemSusChem, vol. 7, pp. 1594–1611, 2014, doi: 10.1002/cssc.201301042.

[20] H. Jensen et al., “In situ high-energy synchrotron radiation study of sol-gel nanoparticle formation in supercritical fluids,” Angew. Chem. Int. Ed., vol. 46, pp. 1113–1116, 2007, doi: 10.1002/anie.200603386.

[21] J. L. Mi, Y. Shen, J. Becker, M. Bremholm, and B. B. Iversen, “Controlling allotropism in ruthenium nanoparticles: A pulsed-flow supercritical synthesis and in situ synchrotron x-ray diffraction study,” J. Phys. Chem. C, vol. 118, pp. 11104–11110, 2014, doi: 10.1021/jp501229p.

[22] T. Friščić et al., “Real-time and in situ monitoring of mechanochemical milling reactions,” Nat. Chem., vol. 5, pp. 66–73, 2013, doi: 10.1038/nchem.1505.

[23] O. Aalling-Frederiksen, M. Juelsholt, A. S. Anker, and K. M. Ø. Jensen, “Formation and growth mechanism for niobium oxide nanoparticles: Atomistic insight from: In situ X-ray total scattering,” Nanoscale, vol. 13, pp. 8087–8097, 2021, doi: 10.1039/d0nr08299f.

[24] B. E. Warren and E. P. Warekois, “Stacking faults in cold worked alpha-brass,” Acta Metall. Mater., vol. 3, pp. 473–479, 1955, doi: 10.1016/0001-6160(55)90138-3.

[25] H. L. Andersen, E. D. Bøjesen, S. Birgisson, M. Christensen, and B. B. Iversen, “Pitfalls and reproducibility of in situ synchrotron powder X-ray diffraction studies of solvothermal nanoparticle formation,” J. Appl. Cryst., vol. 51, pp. 526–540, 2018, doi: 10.1107/S1600576718003552.

[26] A. Leonardi and P. Scardi, “Dislocation Effects on the Diffraction Line Profiles from Nanocrystalline Domains,” Metall. Mater. Trans. A, vol. 47, pp. 5722–5732, 2016, doi: 10.1007/s11661-015-2863-y.

[27] F. Bertolotti et al., “Crystal Structure, Morphology, and Surface Termination of Cyan-Emissive, Six- Monolayers-Thick CsPbBr3 Nanoplatelets from X-ray Total Scattering,” ACS Nano, vol. 13, pp. 14294– 14307, 2019, doi: 10.1021/acsnano.9b07626.

[28] A. Parakh et al., “Nucleation of Dislocations in 3.9 nm Nanocrystals at High Pressure,” Phys. Rev. Lett., vol. 124, p. 106104, 2020, doi: 10.1103/PhysRevLett.124.106104.

[29] H. Ozawa, “X線散乱に基づく粒子の原子配列解析法の開発,” 2018.

[30] W. Wagner et al., “The IAPWS Industrial Formulation 1997 for the Thermodynamic Properties of Water and Steam,” J. Eng. Gas Turb. Power, vol. 122, pp. 150–182, 2000.

[31] M. Uematsu and E. U. Franck, “Static Dielectric Constant of Water and Steam,” J. Phys. Chem. Ref. Data, vol. 9, pp. 1291–1306, 1980, doi: 10.1063/1.555632.

[32] W. L. Marshall and E. U. Franck, “Ion Product of Water Substance, 0-1000℃, 1-10,000 Bars New International Formulation and Its Background,” J. Phys. Chem. Ref. Data, vol. 10, pp. 295–304, 1981, doi: 10.1063/1.555643.

[33] M. Fisher and B. Widom, “Decay of Correlations in Linear Systems,” J. Chem. Phys., vol. 50, p. 3756, 1969, doi: 10.1063/1.1671624.

[34] T. Morita, K. Kusano, H. Ochiai, K.-I. Saitow, and K. Nishikawa, “Study of inhomogeneity of supercritical water by small-angle x-ray scattering,” J. Chem. Phys., vol. 112, pp. 4203–4211, 2000, doi: 10.1063/1.480965.

[35] N. Yoshida, M. Matsugami, Y. Harano, K. Nishikawa, and F. Hirata, “Structure and Properties of Supercritical Water: Experimental and Theoretical Characterizations,” J, vol. 4, pp. 698–726, 2021, doi: 10.3390/j4040049.

[36] Y. Hakuta, S. Onai, H. Terayama, T. Adschiri, and K. Arai, “Production of ultra - fine ceria particles by hydrothermal synthesis under supercritical conditions,” J. Mater. Sci. Lett., vol. 17, pp. 1211–1213, 1009, doi: 10.1023/A:1006597828280.

[37] J. Zhang, S. Ohara, M. Umetsu, T. Naka, Y. Hatakeyama, and T. Adschiri, “Colloidal ceria nanocrystals: A tailor-made crystal morphology in supercritical water,” Adv. Mater., vol. 19, pp. 203–206, 2007, doi: 10.1002/adma.200600964.

[38] N. Aoki et al., “Kinetics study to identify reaction-controlled conditions for supercritical hydrothermal nanoparticle synthesis with flow-type reactors,” J. Supercrit. Fluids, vol. 110, pp. 161–166, 2016, doi: 10.1016/j.supflu.2015.11.015.

[39] Y. Hakuta, T. Adschiri, H. Hirakoso, and K. Arai, “Chemical equilibria and particle morphology of boehmite(AlOOH) in sub and supercritical water,” Fluid Phase Equilibr., vol. 158–160, pp. 733–742, 1999, doi: 10.1016/S0378-3812(99)00118-1.

[40] J. Becker et al., “Experimental setup for in situ X-ray SAXS/WAXS/PDF studies of the formation and growth of nanoparticles in near-and supercritical fluids,” J. Appl. Cryst., vol. 43, pp. 729–736, 2010, doi: 10.1107/S0021889810014688.

[41] S. Kawasaki, Y. Xiuyi, K. Sue, Y. Hakuta, A. Suzuki, and K. Arai, “Continuous supercritical hydrothermal synthesis of controlled size and highly crystalline anatase TiO2 nanoparticles,” J. Supercrit. Fluids, vol. 50, pp. 276–282, 2009, doi: 10.1016/j.supflu.2009.06.009.

[42] F. H. Gjørup, J. v. Ahlburg, and M. Christensen, “Laboratory setup for rapid in situ powder X-ray diffraction elucidating Ni particle formation in supercritical methanol,” Rev. Sci. Instrum., vol. 90, p. 073902, 2019, doi: 10.1063/1.5089592.

[43] Y. Yang et al., “In Situ X-ray Absorption Spectroscopy of a Synergistic Co-Mn Oxide Catalyst for the Oxygen Reduction Reaction,” J. Am. Chem. Soc., vol. 141, pp. 1463–1466, 2019, doi: 10.1021/jacs.8b12243.

[44] M. Bremholm, H. Jensen, S. B. Iversen, and B. B. Iversen, “Reactor design for in situ X-ray scattering studies of nanoparticle formation in supercritical water syntheses,” J. Supercrit. Fluids, vol. 44, pp. 385– 390, 2008, doi: 10.1016/j.supflu.2007.09.029.

[45] E. Bøjesen and B. Iversen, “The chemistry of nucleation ,” CrystEngComm, vol. 18, pp. 8332–8353, 2016.

[46] S. J. Moorhouse, N. Vranje, A. Jupe, M. Drakopoulos, and D. O’Hare, “The oxford-diamond in situ cell for studying chemical reactions using time-resolved x-ray diffraction,” Rev. Sci. Instrum., vol. 83, p. 084101, 2012, doi: 10.1063/1.4746382.

[47] J. S. O. Evans et al., “An apparatus for the study of the kinetics and mechanism of hydrothermal reactions by in situ energy dispersive x-ray diffraction,” Rev. Sci. Instrum., vol. 66, pp. 2442–2445, 1995, doi: 10.1063/1.1146451.

[48] E. Nishibori et al., “The large Debye Scherrer camera installed at SPring-8 BL02B2 for charge density studies,” Nucl. Instr. Meth., vol. A467, p. 1045, 2001.

[49] S. Kawaguchi et al., “High-throughput powder diffraction measurement system consisting of multiple MYTHEN detectors at beamline BL02B2 of SPring-8,” Rev. Sci. Instrum. , vol. 88, p. 085111, 2017, doi: 10.1063/1.4999454.

[50] T. Fujita, H. Kasai, and E. Nishibori, “Hydrothermal reactor for in-situ synchrotron radiation powder diffraction at SPring-8 BL02B2 for quantitative design for nanoparticle,” J. Supercrit. Fluids, vol. 147, pp. 172–178, 2019, doi: 10.1016/j.supflu.2018.10.016.

[51] X. Guo, “Property degradation of tetragonal zirconia induced by low-temperature defect reaction with water molecules,” Chem. Mater., vol. 16, pp. 3988–3994, 2004, doi: 10.1021/cm040167h.

[52] C. Piconi and G. Maccauro, “Zirconia as a ceramic biomaterial,” Biomaterials, vol. 20, pp. 1–25, 1999, doi: 10.1016/s0142-9612(98)00010-6.

[53] V. Lughi and V. Sergo, “Low temperature degradation -aging- of zirconia: A critical review of the relevant aspects in dentistry,” Dent. Mater., vol. 26, pp. 807–820, 2010, doi: 10.1016/j.dental.2010.04.006.

[54] Y. Hakuta, T. Ohashi, H. Hayashi, and K. Arai, “Hydrothermal synthesis of zirconia nanocrystals in supercritical water,” J. Mater. Res., vol. 19, pp. 2230–2234, 2004, doi: 10.1557/JMR.2004.0314.

[55] J. Becker et al., “Critical size of crystalline ZrO2 nanoparticles synthesized in near- and supercritical water and supercritical isopropyl alcohol,” ACS Nano, vol. 2, pp. 1058–1068, 2008, doi: 10.1021/nn7002426.

[56] J. Qi and X. Zhou, “Formation of tetragonal and monoclinic-HfO2 nanoparticles in the oil/water interface,” Colloid. Surface A, vol. 487, pp. 26–34, 2015, doi: 10.1016/j.colsurfa.2015.09.037.

[57] A. Sahraneshin, S. Takami, D. Hojo, K. Minami, T. Arita, and T. Adschiri, “Synthesis of shape-controlled and organic-hybridized hafnium oxide nanoparticles under sub- and supercritical hydrothermal conditions,” J. Supercrit. Fluids, vol. 62, pp. 190–196, 2012, doi: 10.1016/j.supflu.2011.10.019.

[58] A. Sahraneshin et al., “Surfactant-assisted hydrothermal synthesis of water-dispersible hafnium oxide nanoparticles in highly alkaline media,” Cryst. Growth Des., vol. 12, pp. 5219–5226, 2012, doi: 10.1021/cg3005739.

[59] H. M. Rietveld, “A Profile Refinement Method for Nuclear and Magnetic Structures,” J. Appl. Cryst., vol. 2, pp. 65–71, 1969, doi: 10.1107/S0021889869006558.

[60] H. Toraya, “Array-Type Universal Profile Function for Powder Pattern Fitting,” J. Appl. Cryst, vol. 23, pp. 485–491, 1990, doi: 10.1107/S002188989000704X.

[61] E. Nishibori et al., “Accurate structure factors and experimental charge densities from synchrotron X-ray powder diffraction data at SPring-8,” Acta Cryst. , vol. A63, pp. 43–52, 2007, doi: 10.1107/S0108767306047210.

[62] P. Debye, “Scattering from non-crystalline substances,” Ann. Physik, pp. 809–823, 1915.

[63] F. Bertolotti, D. Moscheni, A. Guagliardi, and N. Masciocchi, “When Crystals Go Nano – The Role of Advanced X-ray Total Scattering Methods in Nanotechnology,” Eur. J. Inorg. Chem., vol. 2018, pp. 3789– 3803, 2018, doi: 10.1002/ejic.201800534.

[64] A. P. Thompson et al., “LAMMPS - a flexible simulation tool for particle-based materials modeling at the atomic, meso, and continuum scales,” Comput. Phys. Commun., vol. 271, p. 108171, 2022, doi: 10.1016/j.cpc.2021.108171.

[65] L. Martinez, R. Andrade, E. G. Birgin, and J. M. Martínez, “PACKMOL: A package for building initial configurations for molecular dynamics simulations,” J. Comput. Chem., vol. 30, pp. 2157–2164, 2009, doi: 10.1002/jcc.21224.

[66] A. Stukowski, “Visualization and analysis of atomistic simulation data with OVITO-the Open Visualization Tool,” Model. Simul. Mater. Sc., vol. 18, p. 015012, 2010, doi: 10.1088/0965-0393/18/1/015012.

[67] S. Narioka, 2004.

[68] P. Scherrer, “Bestimmung der Grösse und der inneren Struktur von Kolloidteilchen mittels Röntgenstrahlen,” Nachrichten von der Gesellschaft der Wissenschaften, pp. 98–100, 1918.

[69] M. Leoni, J. Martinez-Garcia, and P. Scardi, “Dislocation effects in powder diffraction,” J. Appl. Cryst., vol. 40, pp. 719–724, 2007, doi: 10.1107/S002188980702078X.

[70] F. Bertolotti et al., “A total scattering Debye function analysis study of faulted Pt nanocrystals embedded in a porous matrix,” Acta Cryst., vol. A72, pp. 632–644, 2016, doi: 10.1107/S205327331601487X.

[71] S. K. Huang, N. Li, Y. H. Wen, J. Teng, S. Ding, and Y. G. Xu, “Effect of Si and Cr on stacking fault probability and damping capacity of Fe-Mn alloy,” Mater. Sci. Eng. A, vol. 479, pp. 223–228, 2008, doi: 10.1016/j.msea.2007.06.063.

[72] K. R. Beyerlein, R. L. Snyder, and P. Scardi, “Faulting in finite face-centered-cubic crystallites,” Acta Cryst. , vol. A67, pp. 252–263, 2011, doi: 10.1107/S0108767311009482.

[73] G. Hura, J. M. Sorenson, R. M. Glaeser, and T. Head-Gordon, “High-quality X-ray scattering experiment on liquid water at ambient conditions,” J. Chem. Phys., vol. 113, pp. 9140–9148, 2000, doi: 10.1063/1.1319614.

[74] K. Nishikawa and N. Kitagawa, “X-Ray Diffraction Study of Liquid Water,” Bull. Chem. Soc. Jpn., vol. 53, pp. 2804–2808, 1980.

[75] M.-C. Bellissent-Funel, “Structure of Supercritical Water,” J. Mol. Liq., vol. 90, pp. 313–322, 2001, doi: 10.1016/S0167-7322(01)00135-0.

[76] C. J. Howard, R. J. Hill, and B. E. Reichert, “Structures of the ZrO2 Polymorphs at Room Temperature by High-Resolution Neutron Powder Diffraction,” Acta Cryst. , vol. B44, pp. 116–120, 1988, doi: 10.1107/S0108768187010279.

[77] K. R. Whittle, G. R. Lumpkin, and S. E. Ashbrook, “Neutron diffraction and MAS NMR of cesium tungstate defect pyrochlores,” J. Solid State Chem., vol. 179, pp. 512–521, 2006, doi: 10.1016/j.jssc.2005.11.011.

[78] T. Kiguchi, T. Shiraishi, T. Shimizu, H. Funakubo, and T. J. Konno, “Domain orientation relationship of orthorhombic and coexisting monoclinic phases of YO1.5-doped HfO2 epitaxial thin films,” Jpn. J. Appl. Phys., vol. 57, pp. 11–16, 2018, doi: 10.7567/JJAP.57.11UF16.

[79] D. A. Navarro, S. W. Depner, D. F. Watson, D. S. Aga, and S. Banerjee, “Partitioning behavior and stabilization of hydrophobically coated HfO2, ZrO2 and Hf xZr1-xO2 nanoparticles with natural organic matter reveal differences dependent on crystal structure,” J. Hazard. Mater., vol. 196, pp. 302–310, 2011, doi: 10.1016/j.jhazmat.2011.09.028.

[80] Q. Mahmood, A. Afzal, H. M. Siddiqi, and A. Habib, “Sol-gel synthesis of tetragonal ZrO2 nanoparticles stabilized by crystallite size and oxygen vacancies,” J. Sol-Gel Sci. Techn., vol. 67, pp. 670–674, 2013, doi: 10.1007/s10971-013-3112-8.

[81] A. S. Barnard, R. R. Yeredla, and H. Xu, “Modelling the effect of particle shape on the phase stability of ZrO 2 nanoparticles,” Nanotechnology, vol. 17, no. 12, pp. 3039–3047, 2006, doi: 10.1088/0957- 4484/17/12/038.

[82] R. C. Garvie, “Stabilization of the Tetragonal Structure in Zirconia Microcrystals,” J. Phys. Chem., vol. 82, pp. 218–223, 1978, doi: 10.1021/j100491a016.

[83] R. C. Garvie, “The occurrence of metastable tetragonal zirconia as a crystallite size effect,” J. Phys. Chem., vol. 69, pp. 1238–1243, 1965, doi: 10.1021/j100888a024.

[84] M. Taguchi et al., “One-pot synthesis of monoclinic ZrO2 nanocrystals under subcritical hydrothermal conditions,” J. Supercrit. Fluids, vol. 85, pp. 57–61, 2014, doi: 10.1016/j.supflu.2013.11.001.

[85] R. P. Denkewicz, K. S. TenHuisen, and J. H. Adair, “Hydrothermal Crystallization Kinetics of m-ZrO2 and t-ZrO2,” J. Mater. Res., vol. 5, pp. 2698–2705, 1990, doi: 10.1557/JMR.1990.2698.

[86] K. Sato, H. Abe, and S. Ohara, “Selective growth of monoclinic and tetragonal zirconia nanocrystals,” J. Am. Chem. Soc., vol. 132, pp. 2538–2539, 2010, doi: 10.1021/ja910712r.

[87] H. J. Noh, D. S. Seo, H. Kim, and J. K. Lee, “Synthesis and crystallization of anisotropic shaped ZrO2 nanocrystalline powders by hydrothermal process,” Mater. Lett., vol. 57, pp. 2425–2431, 2003, doi: 10.1016/S0167-577X(02)01248-X.

[88] M. Bucko, K. Haberko, and M. Faryna, “Crystallization of Zirconia under Hydrothermal Conditions,” J. Am. Ceram. Soc., vol. 78, pp. 3397–3400, 1999, doi: 10.1111/j.1151-2916.1995.tb07985.x.

[89] A. C. Dippel et al., “Towards atomistic understanding of polymorphism in the solvothermal synthesis of ZrO2 nanoparticles,” Acta Cryst., vol. A72, pp. 645–650, 2016, doi: 10.1107/S2053273316012675.

[90] M. Bremholm, J. Becker-Christensen, and B. B. Lversen, “High-pressure, high-temperature formation of phase-pure monoclinic zirconia nanocrystals studied by time-resolved in situ synchrotron X-ray diffraction,” Adv. Mater., vol. 21, pp. 3572–3575, 2009, doi: 10.1002/adma.200803431.

[91] Y. Murase and E. Kato, “Preparation of Zirconia Whiskers from Zirconium Hydroxide in Sulfuric Acid Solutions under Hydrothermal Conditions at 200°C,” J. Am. Ceram. Soc., vol. 84, pp. 2705–2706, 2001, doi: 10.1111/j.1151-2916.2001.tb01076.x.

[92] X. Guo, “On the degradation of zirconia ceramics during low-temperature annealing in water or water vapor,” J. Phys. Chem. Solids, vol. 60, pp. 539–546, 1999, doi: 10.1016/S0022-3697(98)00301-1.

[93] Y. S. Kim, C. H. Jung, and J. Y. Park, “Low temperature degradation of yttria-stabilized tetragonal zirconia polycrystals under aqueous solutions,” Journal of Nucl. Mater., vol. 209, pp. 326–331, 1994, doi: 10.1016/0022-3115(94)90271-2.

[94] J.-F. Li, R. Watanabe, B.-P. Zhang, K. Asami, and K. Hashimoto, “X-ray Photoelectron Spectroscopy Investigation on the Low-Temperature Degradation of 2 mol% Y2O3-ZrO2 Ceramics,” J. Am. Chem. Soc., vol. 79, pp. 3109–3112, 1996, doi: 10.1111/j.1151-2916.1996.tb08084.x.

[95] M. Yoshimura, T. Noma, K. Kawabata, and S. Somiya, “Role of H2O on the degradation process of Y-TZP,” J. Mater. Sci. Lett., vol. 6, pp. 465–467, 1987, doi: 10.1007/BF01756800.

[96] O. Kruse, H. Carstanjen, Kountouros. P., H. Schubert, and G. Petzow, “Characterization of H2O-Aged TZP by Elastic Recoil Detection Analysis (ERDA),” in Science and Technology of Zirconia V, S. Badwal, M. Bannister, and R. Hannink, Eds. Lancaster: Technomic Publishing Company, Inc., 1993, p. 163.

[97] T. Fujita, H. Kasai, and E. Nishibori, “Ion Product Scale for Phase and Size Selective Crystal Growth of Zirconia Nanoparticles,” Cryst. Growth Des., vol. 20, pp. 5589–5595, 2020, doi: 10.1021/acs.cgd.0c00765.

参考文献をもっと見る

全国の大学の
卒論・修論・学位論文

一発検索!

この論文の関連論文を見る