Vous êtes sur la page 1sur 7

journal

J. Am. Ceram. Soc., 83 [3] 60511 (2000)

Crystal Structure and Related Properties of Manganese-Doped Barium Titanate Ceramics


Hans Theo Langhammer, Thomas Muller, Karl-Heinz Felgner, and Hans-Peter Abicht
Departments of Physics and Chemistry, Martin-Luther-Universitat Halle-Wittenberg, D-06099 Halle, Germany The influence of manganese on the crystallographic phase and the microstructure of BaTi1 xMnxO3 ceramics (0 < x < 0.05) is investigated. The crystal structure at room temperature changes from tetragonal to hexagonal between x 0.005 and 0.017, which causes a drastic change in the microstructure. 3 The JahnTeller distortion caused by the MnTi ions is proposed as the driving force for the phase transition. Annealing of the as-fired samples in both a reducing and oxidizing atmosphere restores the tetragonal phase, which is accompanied by a change in the microstructure based on the percentage of tetragonal phase. I. Introduction sufficient to stabilize h-BT at room temperature. Recnik and Kolar14 16 also investigated the microstructure of h-BT sintered at 1360C in a H2/Ar atmosphere and found strongly exaggerated grain growth, which leads to large platelike hexagonal grains. Doping with 3d transition elements, such as manganese, iron, or nickel, is the second way of stabilizing h-BT at room temperature.8,1722 Because of their small ionic radii, they substitute for Ti4 , which is commonly accepted in the literature. Their valence state when they are incorporated into the lattice has been discussed in numerous experimental and theoretical papers. In the case of manganese,2,24 27 Glaister and Kay8 found that it is the most effective dopant for stabilizing h-BT. Dickson et al.,17 and recently Takeuchi et al.,12 proposed a chemical bonding between the two d orbitals of titanium and manganese as a reason for stabilization of the face-sharing TiO6/MnO6 octahedra in h-BT. Ren et al.19 stated that the small size of the Mn4 ion, which substitutes Ti4 , is a dominant factor favoring the formation of h-BT. A critical discussion of this assumption was published by the authors of the present paper.20 Similar to h-BT produced by sintering in reducing atmosphere, manganese-doped h-BT also exhibits a pronounced exaggerated growth of platelike grains.20,21 Comparing the published results, the whole picture of manganese-doped BaTiO3 is not quite clear. This is true for both the microscopic models describing the phase transition from cubic/tetragonal to hexagonal and also macroscopic properties, such as quantitative phase composition, depending on the doping level and the parameters of the sintering process or the microstructure formed. One reason for the inconsistency of the macroscopic results seems to be the strong influence of impurities of the raw materials on the phase composition and microstructure. As an example, the influence of strontium as an impurity is discussed in a subsequent paper.28 Hence, the aim of this paper is to clarify the influence of manganese on the macroscopic properties of BaTiO3 ceramics mentioned above (crystallographic phase, microstructure) as well as discuss the microscopic mechanism of the phase transition from cubic to hexagonal and the role of manganese in the promotion of this process. For this purpose, we performed detailed investigations of manganese-doped BaTiO3 at a doping level of between 0 and 5 mol% with regard to the following properties: X-ray diffraction (XRD), electron diffraction, microstructure, and the solubility of manganese in the BaTiO3 lattice. Both as-sintered specimens and samples annealed in oxidizing and reducing atmospheres were investigated. II. Experimental Procedure

ANGANESE-DOPED BaTiO3 ceramics are used for several purposes. The main applications are capacitors and devices using the positive temperature coefficient of resistance (PTCR), where, as an effective acceptor dopant, manganese influences both the electrical properties and the microstructure, because it is an effective acceptor dopant. In capacitor materials, manganese improves resistivity to reducing atmospheres during sintering. In donor-doped PTCR ceramics, manganese depresses the grain size and enhances the rise in electrical resistivity near the Curie temperature of BaTiO3. Much work has been done to elucidate the role of manganese regarding these effects.17 However, little is known about the influence of manganese on the crystallographic structure and the microstructure of otherwise undoped BaTiO3. Glaister and Kay8 first reported in 1960 that certain amounts of manganese are sufficient to stabilize the hexagonal phase of BaTiO3 (h-BT) at room temperature, whereas the hexagonal polymorph of undoped material is stable in air only at temperatures 1460C.9,10 Two possibilities are known to stabilize h-BT at lower temperatures. The first one is firing in reducing atmospheres.8,1116 In pure hydrogen, a temperature of 1330C is sufficient to stabilize the hexagonal phase.8 The formation of oxygen vacancies and the resulting reduction of Ti4 to Ti3 according to the reaction

OO

2TiTi VO

2TiTi

1 2

O21

(1)

seems to be the generally accepted reason for this stabilization effect, which is explained by the preferred formation of facesharing octahedra (Ti2O9 groups), because of the reduced electrostatic repulsion between Ti3 and Ti4 /Ti3 ions compared with the interaction between Ti4 ions. According to Wakamatsu, Takeuchi, and co-workers,11,12 a minimum of 0.3 mol% of Ti3 is

B. M. Kulwickicontributing editor

Manuscript No. 190362. Received February 20, 1998; approved July 23, 1999. Supported by the Kultusministerium des Landes Sachsen-Anhalt, Germany, and the Deutsche Forschungsgemeinschaft under Contract No. FKZ: 1434A/0083. Department of Physics. Department of Chemistry.

Ceramic powder with a nominal composition of BaTi1-xMnxO3 (0 x 0.05) was prepared by the conventional mixed-oxide powder technique. After mixing (agate balls, water) and calcining (1100C, 2 h) appropriate amounts of BaCO3 ( 0.1 mol% strontium; No. 3018, Leuchtstoffwerk Breitungen GmbH, Germany), TiO2, (No. 808, Merck, Darmstadt, Germany), and MnCO3 (Riedel de Haen, Germany), we fine-milled (agate balls, water) and densified the powder to disks with a diameter of 12 mm and a height of nearly 3 mm. Because of the thermodynamic instability of BaTiO3 in water, the Ba/Ti ratio of the green bodies amounted 605

606

Journal of the American Ceramic SocietyLanghammer et al.

Vol. 83, No. 3

to 1/1.015.29 The samples were sintered in air at Ts 1400C for 1 h and subsequently characterized with respect to their microstructure and phase composition (see below). Then, a portion of each sample was annealed in air and another portion was annealed in a gas stream mixture of H2 and argon (1/1), and characterized again. In both cases, annealing was accomplished at 1200C in 2 h. All heat treatments took place with heating and cooling rates of 10 K/min. To avoid interfering contamination, the samples were contained in ZrO2-covered Al2O3 dishes. The microstructure of polished and chemically etched specimens was examined by optical microscopy and by scanning electron microscopy (SEM). To determine the distribution of manganese in the grains and intergranular regions, wavelengthdispersive X-ray electron probe microanalysis (WDX-EPMA) (Model CAMEBAX, Cameca Instruments, Corbevoie, France) was performed. For the phase investigations at room temperature by XRD, the sintered samples were crushed again and mixed with silicon powder for the purpose of calibration. The phase composition was determined quantitatively by analyzing the intensity ratios (111)tetragonal/(103)hexagonal and (200)tetragonal/(103)hexagonal using a diffractometer (Model D5000, Siemens Aktiengesellschaft, Karlsruhe, Germany). Samples for transmission electron microscopy (TEM) and electron diffractometry (Model CM20Twin, Philips, Eindhoven, The Netherlands) were thinned by consecutive mechanical grinding, dimpling, and ion milling up to perforation of their central region.

most remarkable finding is the nearly complete restoration of the tetragonal phase after annealing in reducing atmosphere at a manganese concentration of 1.6 mol%. (2) Microstructure (A) As-Fired Samples: Figures 2(a)(e) show microstructures of samples with a nominal manganese content of between 0.5 and 5.0 mol%. The average grain sizes of all samples investigated are shown in Fig. 3. Up to a manganese content of 0.5 mol%, the microstructure looks like the well-known microstructure of undoped liquid-phase-sintered BaTiO3 ceramics. Simultaneously with the appearance of the hexagonal phase at 1 mol% manganese, the microstructure becomes bimodal. A new grain type with a platelike shape and a grain size of 400 m develops, while the grains of the other fraction, which exhibit the globular shape of the manganese-free ceramics, become smaller, and their percentage decreases. At 1.6 mol% manganese, the smaller globular grains have vanished, corresponding approximately to the disappearance of the tetragonal phase. For nominal manganese concentrations of 2.5 mol%, once again, the microstructure becomes bimodal. With increasing manganese content, the globular grains (diameter: 510 m) reappear, the percentage of which exceeds the percentage of the platelike fraction at a nominal manganese content of 5 mol%. Only for manganese concentrations 3 mol%, when the platelike grains are surrounded by a sufficient amount of small globular grains, is their platelike shape clearly developed. At lower manganese concentrations, these grains have grown into each other during the final stage of sintering and their platelike shape is deformed. (B) Annealed Samples: Whereas the samples annealed in air show no remarkable change in microstructure compared with the as-sintered specimens, the samples annealed in a highly reducing atmosphere exhibit a significant reduction in grain size and a change in the grain shape, which is especially pronounced for samples with manganese concentrations between 1.6 and 2.0 mol%, according to their considerable percentage of tetragonal phase (see Fig. 1). Corresponding to Figs. 2(a)(e), Figs. 2(f)(j) show the microstructures of the hydrogen-annealed samples, the average grain sizes of which are also shown in Fig. 3. (3) Electron Diffraction Samples were characterized by electron diffraction, because it is not easy to distinguish between the hexagonal and cubic phase of BaTiO3 (c-BT) by XRD measurements. An advantage of the method is also the possibility of local measurements. Kolar et al.16 reported on small cubic grains in undoped h-BT sintered in a reducing atmosphere. Because of the similarities to the microstructure of the manganese-doped samples, mainly small grains were investigated by electron diffraction. None of the numerous grains investigated exhibited the cubic phase. Figure 4 shows the TEM image of some small grains of a sample with 5 mol% manganese together with a typical hexagonal electron diffraction pattern of one of these grains. Hence, probably all the samples investigated with a manganese content of 1.7 mol% seem to be completely in the hexagonal phase. (4) Solubility of Manganese (A) As-Fired Samples: The results of our investigation into the solubility of manganese in BaTiO3 are shown in Fig. 5. The small deviation of the integrally measured values from the nominal manganese content seems to be due to an uncertain correction of atomic number effect, absorption, and fluorescence excitation (ZAF) at higher manganese concentrations, because we used metallic manganese as a standard sample. While the nominal amount of manganese is completely incorporated up to 1.5 mol%, for higher manganese concentrations, an increasing number of manganese ions are not soluble in the BaTiO3 lattice of the platelike grains but segregate at grain boundaries or form intergranular manganese-rich phases that are clearly visible in the EPMA-measured distribution images of manganese. Identification of these phases, the manganese content of which partially exceeds 7.5 at.%, has not been successful thus far. Two different phases

III.

Results

(1) X-Ray Diffraction The results of the XRD investigations both for the as-fired and annealed samples are shown in Fig. 1. Only the percentage of tetragonal phase as a function of the doping level of manganese is shown. If no other statement is given, the remaining percentage corresponds to the hexagonal phase. (A) As-Fired Samples: The region of coexistence of the tetragonal and hexagonal phases in the as-fired samples lies between 0.5 and 1.7 mol% manganese. Below and above this region, the samples are single-phase within the accuracy of the XRD method, which is not better than 25 mol%. (B) Annealed Samples: The annealed samples show a significant effect of restoration of the tetragonal phase. Although this holds for both types of annealing atmospheres, the effect is clearly more pronounced for reducing conditions. During annealing in a highly reducing atmosphere, some manganese is expelled from the specimens. Therefore, corrected data giving the remaining manganese concentration (see below) are also shown in Fig. 1. The

Fig. 1. XRD-determined percentage of tetragonal phase (20C) as a function of the nominal manganese content (x) of BaTi1-xMnxO3 before and after annealing under both oxidizing and reducing conditions.

March 2000

Crystal Structure and Related Properties of Manganese-Doped Barium Titanate Ceramics

607

Fig. 2. Optical micrographs of the microstructure of BaTi1-xMnxO3 ceramics of both (a)(e) as-fired and (f)(j) H2-annealed samples with (a), (f) x 0.005; (b), (g) 0.010; (c), (h) 0.016; (d), (i) 0.030; and (e), (j) 0.050. Note different magnifications.

608

Journal of the American Ceramic SocietyLanghammer et al.

Vol. 83, No. 3

Fig. 3. Average grain size as a function of the nominal manganese content (x) of BaTi1-xMnxO3 before and after annealing under oxidizing and reducing conditions. Percentage values at the data symbols denote roughly estimated area portions of the different grain fractions. Grain sizes presented are roughly estimated values, especially in the case of platelike grains that are obtained by random cuts in the two-dimensional polished plane.

Fig. 4. Bright-field TEM image (U 200 kV) of a thinned sample with a nominal composition of BaTi0.95Mn0.05O3 in the region of small globular grains. In the upper left-hand corner, a hexagonal electron diffraction pattern of the grain in the center is shown. Electron beam coincides with the [100] direction, and the vertical and horizontal direction of the pattern image belong to the (001) and the (010) plane, respectively.

Fig. 5. EPMA-measured manganese concentration both integrally in a square and inside the grown grains as a function of the nominal manganese content of manganese-doped BaTiO3.

seem to occurthe first with a Ti/Ba ratio of nearly one, and a second with a Ti/Ba ratio of nearly two. Contrary to the platelike grains, at least 5 mol% manganese seems to be soluble in the small

globular grains of the manganese-rich samples with a bimodal microstructure. But, indeed, a precise determination of the manganese content of the small grains by EPMA in compact specimens is hardly possible because of the rather large excitation volume of the X-rays. (B) Annealed Samples: Whereas the samples did not change their manganese content during annealing in air, some manganese

March 2000

Crystal Structure and Related Properties of Manganese-Doped Barium Titanate Ceramics

609

Fig. 6. Illustration of the reciprocal shift of the adjacent BaO lattice planes A and C, causing the transformation from the cubic (ABCABC) to the hexagonal (ABACBC) stacking. Only the layers BCA and BAC, respectively, are depicted.

is expelled from the specimens during annealing in a highly reducing atmosphere.30 The remaining manganese concentrations are measured by EPMA. The corrected values shown in Fig. 1 are related to mean values averaged over an area that covers various grains in the central region of the specimens. The amount of manganese incorporated inside the grains is even somewhat lower than the integrally measured quantities because of manganese segregation in triple points. IV. Discussion

(1) As-Fired Samples Taking into account all previous work on the factors influencing the cubic/hexagonal phase transformation of BaTiO3 (temperature, oxygen partial pressure, concentration of some acceptor dopants) (see, e.g., Refs. 8, 12, 14), it seems to be reasonable to assume that one common mechanism is responsible for the stability of h-BT. Two conditions must be accounted for by that mechanism. First, a sufficiently high density of oxygen vacancies is necessary to enable the microscopic steps of the transformation from the cubic to the hexagonal stacking of the BaO layers. The transition from the cubic ABCABC to the hexagonal ABACBC stacking can be summarized as a reciprocal gliding of the adjacent lattice planes A and C, which is shown schematically in Fig. 6. Second, a minimum 3 concentration of BTi ions (B titanium, manganese, iron, nickel, 3 . . .) is necessary, which was shown in the case of B Ti ([TiTi ] 0.3 mol%) by Wakamatsu and co-workers.11,12 With respect to all three factors mentioned above, these two conditions can be fulfilled. By means of the known defect models of BaTiO3 (see, e.g., Refs. 31, 32) it can be easily derived that, at a fixed oxygen partial pressure, an increasing temperature causes both an increasing concentration of oxygen vacancies and an increasing concen3 tration of TiTi ions. At a sufficiently high temperature, the 3 concentration [TiTi ] even exceeds the concentration of the metal vacancies. On the other hand, a decrease of the oxygen partial pressure decreases the temperature for the necessary minimum concentration of both the oxygen vacancies and the trivalent titanium site ions (see Eq. (1)), which is in accordance with the experiments. According to the equation 2BaO B2O3 2BaBa 2BTi VO 5OO
3

spin-coupling energy. In this case, only the electron configurations d1, d2, d 4, d 6, d7, and d 9 cause a JahnTeller effect (JTE) because of their partial filled sublevels. Moreover, the d4 and the d9 configurations cause a stronger JTE compared to the others, because their partial filling occurs in the higher sublevel. For that reason, the different effectivenesses of the dopants manganese, iron, and nickel in stabilizing the hexagonal polymorph of BaTiO38 could be explained easily. Whereas the Mn3 (d 4) ions cause a complete transformation into h-BT at 1400C for a rather small concentration of dopants (1.7 mol%), it is necessary to add 10 mol% FeO or NiO to reach 50% h-BT at the same temperature.8 Of course, the concentration of Fe2 (d6) and Ni3 (d7), exhibiting a weak JTE, is rather small at high temperatures in air as the ambient atmosphere.34 But concentrations on the order of 0.3 mol% should 3 be sufficient, such as in the case of TiTi . To prove this assumption with another electron configuration showing a strong JTE, we also investigated the effect of copper because of the d9 configuration of Cu2 . Samples with a nominal composition of Ba1.02Ti0.98Cu0.02O3 and sintered at 1400C for 1 h exhibit 31% hexagonal phase and a similar microstructure with large platelike grains similar to the case of manganese. Probably because of the rather high effective ionic radii of Cu2 and Cu at 73 and 77 pm, respectively,35 a definite excess of barium is necessary to incorporate a significant amount but, of course, far from all of the copper ions at titanium sites. A considerable amount of copper is a component of the secondary phase, which is necessary for liquid-phase-assisted grain growth. Furthermore, because of the well-known valence change from Cu2 to Cu at 1000C,36 the real concentration of 2 CuTi is even smaller. Hence, these results corroborate the high efficiency of the incorporation of Cu2 compared to the effect of iron and nickel. Because the d1 electron configuration of Ti3 also causes a JTE, this effect could be the driving force for the cubic 3 hexagonal phase transition in the case of undoped material as well. Looking at the microstructural development, three different grain growth mechanisms occur that alternate continuously for increasing manganese content. At low manganese concentrations of 0 x 0.015, the well-known exaggerated grain growth of undoped BaTiO3 occurs, which is progressively hindered by the increasing manganese content. For x 0.01, a second growth mechanism appears that results in large platelike grains of h-BT. Obviously, their growth mechanism is of the same type that is described by Kolar et al. for reduced h-BT as a liquid-phaseassisted one.14 16 The pronounced anisotropic grain growth results from the surface energy anisotropy of the hexagonal polymorph.15 At manganese concentrations 0.025, this exaggerated grain growth is progressively hindered again, probably because of the solubility limit of the manganese ions. A third growth mechanism gradually dominates, leading to hexagonal globular grains with a limited size. (2) Annealed Samples The partial restoration of the tetragonal phase of BaTiO3 (t-BT) during annealing in air can be understood easily if we assume that the results of the as-fired samples are not equilibrium values but frozen states. In Table I, the dependence of the percentage of t-BT on sintering temperature is shown. A 1.6 mol% manganese containing sample sintered at 1200C for 1 h in air exhibits 86% t-BT compared to only 3 mol% at a sintering temperature of 1400C. Thus, during annealing at 1200C (see Fig. 1), the sample tends to reach its equilibrium state by transformation of some h-BT into c/t-BT. The specimens annealed in a highly reducing atmosphere also exhibit a restoration of t-BT. Consequently, the same explanation should be valid. But, two questions arise that cannot be answered easily. First, following the discussion above, the ambient reducing atmosphere should originally favor a stabilization of the hexagonal phase and not vice versa. Second, annealing in a reducing

(2)

the incorporation of trivalent acceptor ions B at titanium sites also decreases the minimum temperature necessary to form h-BT in an analogous manner, as in the case of decreased ambient oxygen activity. A similar relation also holds for divalent acceptors. Besides the necessary conditions, the driving force for the phase transition must be elucidated. We proposeat least in the case of manganese-doped BaTiO3the influence of the Jahn Teller distortion (see, e.g., Ref. 33) as such a driving force, which is explained as follows. Considering the BO6 complex (B 3d transition elements) in terms of the crystal field approach, the electron configuration exhibits the high-spin ground state because of weak crystal field splitting of the oxygen ions compared to the

In the case of manganese and chromium, the high spin ground state was proved by ab initio calculations.25

610

Journal of the American Ceramic SocietyLanghammer et al. Table I. Percentage of Tetragonal Phase of BaTi1 xMnxO3 Ceramics as a Function of the Sintering Temperature, Ts (air), for Several Manganese Concentrations, x
Ts (C) 0.016 x 0.020 0.050

Vol. 83, No. 3

1000 1100 1200 1300 1400

92 90 86 84 3

88 88 85 74 0

74 69 43 12 0

explanation of the restoration of the cubic/tetragonal phase by annealing in a reducing atmosphere is more speculative and has to be checked by further investigations. Three different grain growth mechanisms control the microstructure in the whole doping range between 0 and 5 mol% manganese. One of them produces exaggerated platelike grains with hexagonal structure that corresponds to the analogous mechanism in undoped h-BT sintered in reducing atmosphere described by Kolar et al.15 The surprising grain size reduction and the rearrangement of the grains during annealing in reducing atmosphere at 1200C is not satisfactorily understood. Acknowledgment:

atmosphere is much more efficient than annealing in air. The 1.6 mol% manganese-containing sample even undergoes a nearly complete transformation from h-BT into c/t-BT. A possible explanation of this phenomenon could be as follows. The crystal structure of h-BT is not as symmetric as the perovskite structure of c-BT. Two types of O2 ions exist that occupy different lattice sites with a slightly different bonding distance to the neighboring ions. The O2 ions that occupy the face-sharing planes of the TiO6 octahedra (Ti2O9 groups) of h-BT are more weakly bound because of their greater bonding length. Hence, the activation energy for the creation of oxygen vacancies at those lattice sites is slightly lowered. Therefore, the oxygen vacancies that are produced during annealing under reducing conditions are created preferentially at the lattice sites of the face-sharing planes. This process destroys the Ti2O9 groups that make up two-thirds of the TiO6 octahedra of the hexagonal structure of BaTiO3. Consequently, the cubic/ tetragonal structure reappears. This mechanism is in competition 3 with the influence of the MnTi ions that should stabilize the hexagonal phase because of their JahnTeller distortion. Hence, the retransformed percentages of tetragonal phase decrease with increasing manganese content. The total restoration of the tetragonal phase at manganese concentrations lower than 2.0 mol% can be explained by a third factor controlling the stability of crystal2 lographic phases. The effective ionic radius (ri) of MnTi , which is 34 likely to be present in highly reducing conditions, amounts to 83 pm (high spin value).35 Thus, the Goldschmidt tolerance factor, 3 4 which is somewhat 1 for TiTi (ri 60.5 pm) and MnTi (ri 64.5 pm), decreases to 1, thus stabilizing the perovskite phase. Of course, these explanations are hypothetical in nature and have to be proved by further investigations. The microstructural development of samples annealed in reducing atmosphere exactly reflects the change of their phase composition related to the situation in the case of the as-sintered samples. But, the drastically changed microstructure of samples with a nominal manganese content of 1.0 and 1.6 mol% is difficult to explain. Because we can assume that at 1200C no liquid phase occurs, a totally diffusion-controlled material transport mechanism has to be taken into account to describe that size reduction and total rearrangement of the grains. In our opinion, the high density of oxygen vacancies during annealing has to be responsible for these extensive solid-state grain-size-reduction and grain growth processes. V. Summary

The authors thank Dr. Christian Eisenschmidt from the Department of Physics of the Martin-Luther-Universitat Halle for performing the quantitative XRD investiga tions and Dr. Stephan Senz from the Max-Planck-Institut fur Mikrostrukturphysik Halle for the careful TEM and electron diffraction investigations.

References
1 I. Burn and G. H. Maher, High Resistivity BaTiO3 Ceramics Sintered in COCO2 Atmospheres, J. Mater. Sci., 10, 633 (1975). 2 S. B. Desu and E. C. Subbarao, Mn-Doped BaTiO3; pp. 189 206 in Advances in Ceramics, Vol. 1, Grain Boundary Phenomena in Electronic Ceramics. Edited by L. M. Levinson. American Ceramic Society, Columbus, OH, 1981. 3 J. Rodel and G. Tomandl, Degradation of Mn-Doped BaTiO3 Ceramic under a High d.c. Electric Field, J. Mater. Sci., 19, 351523 (1984). 4 F. Batllo, E. Duverger, J.-C. Jules, J.-C. Niepce, B. Jannot, and M. Maglione, Dielectric and EPR Studies of Manganese-Doped Barium Titanate, Ferroelectrics, 109, 11318 (1990). 5 J. Illingsworth, H. M. Al-Allak, A. W. Brinkmann, and J. Woods, The Influence of Manganese on the Grain-Boundary Potential Barrier Characteristics of DonorDoped Barium Titanate Ceramics, J. Appl. Phys., 67, 2088 92 (1990). 6 D. Y. Wang and K. Umeya, Spontaneous Polarization Screening Effect and Trap-State Density at Grain Boundaries of Semiconducting Barium Titanate Ceramics, J. Am. Ceram. Soc., 74, 280 86 (1991). 7 Y.-C. Chen, G.-M. Lo, C.-R. Shih, L. Wu, M.-H. Chen, and K.-C. Huang, Influence of Manganese on Lanthanum-Doped BaTiO3, Jpn. J. Appl. Phys., 33, 141216 (1994). 8 R. M. Glaister and H. F. Kay, An Investigation of the CubicHexagonal Transition in Barium Titanate, Proc. Phys. Soc., London, 76, 763 (1960). 9 R. D. Burbank and H. T. Evans Jr., The Crystal Structure of Hexagonal Barium Titanate, Acta Crystallogr., 1, 330 37 (1948). 10 D. E. Rase and R. Roy, Phase Equilibria in the System BaOTiO2, J. Am. Ceram. Soc., 38, 10213 (1955). 11 M. N. Wakamatsu, N. Takeuchi, G. C. Lai, and S. Ishida, Effect of the Firing Atmosphere on the CubicHexagonal Transition and the Chemical State of Titanium in Barium Titanate, Yogyo Kyokaishi (Engl. abstract), 95, 1181 85 (1987). 12 N. Takeuchi, S. Ishida, and M. Wakamatsu, Mechanistic Study of Reactions between Ceramics and Gases during Firing under Various Atmospheres; pp. 51 66 in Memoirs of the Faculty of Engineering and Design, Vol. 43. Kyoto Institute of Technology, 1995. 13 O. Eibl, P. Pongratz, P. Skalicky, and H. Schmelz, Extended Defects in Hexagonal BaTiO3, Philos. Mag. A, 60, 60112 (1989). 14 A. Recnik and D. Kolar, Exaggerated Growth of Hexagonal Barium Titanate under Reducing Sintering Conditions, J. Am. Ceram. Soc., 79, 101518 (1996). 15 D. Kolar, U. Kunaver, and A. Recnik, Exaggerated Anisotropic Grain Growth in Hexagonal Barium Titanate Ceramics, Phys. Status Solidi A, 166, 219 30 (1998). 16 D. Kolar, A. Recnik, and M. Ceh, The Origin and Growth Kinetics of Plate-Like Abnormal Grains in Liquid Phase Sintered Barium Titanate; pp. 33138 in Ceramic Microstructure: Control at the Atomic Level, Plenum, New York, 1998. 17 J. G. Dickson, L. Katz, and R. Ward, Compounds with the Hexagonal Barium Titanate Structure, J. Am. Chem. Soc., 83, 3026 29 (1961). 18 L. Katz and R. Ward, Structure Relations in Mixed Metal Oxides, Inorg. Chem., 3, 20511 (1964). 19 F. Ren, S. Ishida, and S. Mineta, Effect of Manganese Addition on Phase Stability of Hexagonal BaTiO3, J. Ceram. Soc. Jpn. (Yogyo Kyokaishi), 102, 106 108 (1994). 20 H. T. Langhammer, T. Muller, A. Polity, K.-H. Felgner, and H.-P. Abicht, On the Crystal and Defect Structure of Manganese-Doped Barium Titanate Ceramics, Mater. Lett., 26, 20510 (1996). 21 J. Weiss and G. Rosenstein, Addition of Ba2TiSi2O8 to Manganese-Doped Barium Titanate: Effect on Oxygen Diffusion and Grain-Boundary Composition, J. Mater. Sci., 23, 326371 (1988). 22 I. Burn, Mn-Doped Polycrystalline BaTiO3, J. Mater. Sci., 14, 245358 (1979). 23 H.-J. Hagemann and H. Ihrig, Valence Change and Phase Stability of 3d-Doped BaTiO3 Annealed in Oxygen and Hydrogen, Phys. Rev. B: Condens. Matter, 20, 387178 (1979). 24 S. B. Desu and E. C. Subbarao, Effect of Oxidation States of Mn on the Phase Stability of Mn-Doped BaTiO3, Ferroelectrics, 37, 665 68 (1981). 25 P. Moretti and F. M. Michel-Calendini, Impurity Energy Levels and Stability of Cr and Mn Ions in Cubic BaTiO3, Phys. Rev. B: Condensed Matter, 36, 352227 (1987).

Manganese-doped BaTiO3 ceramics sintered at 1400C in air change their room-temperature crystallographic structure from tetragonal to hexagonal between 0.5 and 1.7 mol% manganese. This hexagonal phase is a nonequilibrium one and can be partially retransformed into the cubic/tetragonal one by annealing both in air and in highly reducing atmosphere. As a driving force of the transformation from the cubic to the hexagonal crystal structure, the influence of the JahnTeller distortion is proposed. Thus, it is possible to explain the known experimental data of the phase transformation cubic 7 hexagonal, induced or stabilized by 3d transition elements qualitatively in a satisfying manner. The

March 2000

Crystal Structure and Related Properties of Manganese-Doped Barium Titanate Ceramics

611

26 B. Milsch, Evaluation of Lattice Site and Valence of Manganese in Polycrystalline BaTiO3 and n-BaTiO3 by Electron Paramagnetic Resonance, Phys. Status Solidi A, 133, 455 64 (1992). 27 J. Y. Kim, C. R. Song, and H. I. Yoo, Mn-Doped BaTiO3: Electrical Transport Properties in Equilibrium State, J. Electroceram., 1, 2739 (1997). 28 H. T. Langhammer, T. Muller, K.-H. Felgner, and H.-P. Abicht, Influence of Strontium on Manganese-doped Barium Titanate Ceramics, Mater. Lett., 42, 2124 (2000). 29 H.-P. Abicht, D. Voltzke, A. Roder, R. Schneider, and J. Woltersdorf, The Influence of the Milling Liquid on the Properties of Barium Titanate Powders and Ceramics, J. Mater. Chem., 7, 48792 (1997). 30 V. Bheemineni, E. K. Chang, M. Lal, M. P. Harmer, and D. M. Smyth, Suppression of Acceptor Solubilities in BaTiO3 Densified in Highly Reducing Atmospheres, J. Am. Ceram. Soc., 77, 317376 (1994).

31 J. Daniels and K. H. Hardtl, Electrical Conductivity at High Temperatures of Donor-Doped Barium Titanate Ceramics, Philips Res. Rep., 31, 489 504 (1976). 32 N.-H. Chan, R. K. Sharma, and D. M. Smyth, Nonstoichiometry in Undoped BaTiO3, J. Am. Ceram. Soc., 64, 556 62 (1981). 33 S. F. A. Kettle, Physical Inorganic Chemistry; p. 166. Spectrum Academic Publishers, Oxford, U.K., 1996. 34 H.-J. Hagemann, Acceptor Ions in BaTiO3 and SrTiO3 and Their Consequences on the Properties of Titanate Ceramics; p. 44 in Ph.D. Thesis. RheinischWestfalische Technische Hochschule Aachen, Germany, 1980. 35 R. D. Shannon, Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides, Acta Crystallogr., Sect. A: Cryst. Phys., Diffr. Theor. Gen. Crystallogr., 32, 751 67 (1976). 36 A. F. Hollemann and E. Wiberg, Lehrbuch der Anorganischen Chemie; p. 792. W. de Gruyter, New York, 1976.

Vous aimerez peut-être aussi