The Electrodegradation Process in PZT Ceramics under Exposure to Cosmic Environmental Conditions
Abstract
:1. Introduction
2. Results
2.1. Electrically-Stressed PZT Ceramics under Vacuum Conditions at an Elevated Temperature (T > TC)
- Stage I—A steep step-like increase of the current flow (green curve in Figure 1c) due to displacement and Ohmic currents after a high voltage is switched on;
- Stage II—Decreasing the electric current due to discharging of the capacitor; only the Ohmic contribution of the so-called leakage current contributing to the total current remains;
- Stage IV—A dramatic increase of the kinetics of the electrodegradation; positive feedback from a Joule heating (J-H peak in Figure 1c) is presented. In this way, in a short time, the resistance of the sample is strongly reduced. The sample temperature can increase in this phase by about 10–20 °C [42];
- Stage V—In this final stage, the electrical decomposition process stagnates, and the resistance change is minimal because, in the current flow, the transport of electrons dominates, and the ionic transport, which is responsible for reducing the oxygen stoichiometry in the ceramic, does not play an important role.
2.2. Oxidation of Electro-Degraded Ceramics. Reversibility of I/M Transition
2.3. Inhomogeneous Electrodegradation. Raman Scattering Analysis of the Electro-Degraded Ceramics
2.4. Resistance and Potential Distribution during Electrodegradation: The Deoxidation and Reoxidation Processes
2.5. Electrical Conductivity in the Nano-Scale of PZT Polycrystalline Thin Films
2.6. DFT Calculations and Source of the Charge Carriers in Electro-Degraded Ceramic
3. Discussion
- Electrodegradation progression increases under reduced oxygen activity, as is the case under cosmic conditions;
- Resistance change during electrodegradation under DC action is comparable to that observed in ternary oxides in the paraelectric phase, as it is in SrTiO3 or KTaO3 crystals [20]. A difference between them centres on the type of ions moving towards the cathode in the PIC ceramics, which creates a virtual anode. Based on the EDS mapping and XPS data (Figure S1e, we found that the concentration of Pb in the grain boundary did not increase. However, a reduction in PbO already occurs at 200 °C, and hence the lead in a metallic state was observed at 400 °C on the surface. This kind of electrodegradation is at a relatively low electric field of 100 V/cm. Hence, we state that these are the Pb cations that move into the cathode, creating the aforementioned virtual anode;
- Through in operando measurements of the electric potential distribution, we showed that in a high electric field, the oxygen ions’ movement towards the anode predominates. In this process, the part of the sample lying between the cathode’s geometric and virtual positions is more conductive than the rest of the ceramic, and the maximal potential drop is localised at the virtual cathode. In contrast to the single crystal, the cathode front’s position is not sharp in the ceramic, and when the virtual cathode is close to the anode, maximum oxygen effusion occurs. The oxygen escape allows the electrodegradation process to be classified as electro-induced deoxidation or solid-state electrolysis [20];
- The determination of the transfer number showed that the electric charge transport is a mixture of electrons and ions, but in all phases of the electrodegradation, the electrons predominate. Ionic transport disappears when metallic conductivity, in the last stage IV of electrodegradation, occurs;
- The electrodegradation is an activated process described by the following current power law [5] I(t) = I0 × t−n, where t is the time and the n value is the function of the temperature. With increasing temperature, the exponent n becomes smaller, and so the idea of Sidebottom [56]—stating that the exponent of the power law decreases with decreasing dimensionality of the ionic conducting paths—can be applied to a percolation network of conducting dislocations in electro-degraded PZT (see Supplemental Materials S1). The LCAFM measurement and etching studies have shown the existence of the network of dislocations (conducting filaments) in the grain boundary of thin films. In particular, the dendrite-like fractal structure was proposed by Scott, et al. [15] to analyse the origin of the fatigue effect. An additional argument for the involvement of the mentioned network derives from the discolouration of the ceramics sample with the bar-like geometry. Due to a higher concentration of dislocations in the surface region generated during polishing, this region was mostly electro-degraded;
- The extremely low oxygen outflow, leading to the insulator-metal transition, suggests that the regions (sources) of escaping oxygen are small and galvanically connected. Despite this, these regions are inhomogeneously distributed in the ceramic sample, which is reflected by non-uniform conductivity at the nano-scale;
- The network of conducting filaments is created at the grain boundaries. The core of dislocations constitutes semiconducting nanowires with an additional d1 state of Ti and Zr, which follow the invariance of Burger’s vector and constitute a galvanic short circuit of the insulating grains. During the electrodegradation, the current flows through such a filament network. Hence, the deoxidation process is selective and preferentially reduced to the core of dislocations. Incorporating additional oxygen vacancies in such filaments can generate Ti and Zr states close to the bottom of the conduction band. Then, the filaments can switch into metallic nanowires, and this behaviour is responsible for the current power law;
- Although we did not perform measurements of the local electric field near the virtual anode (e.g., through nano-potentiometric studies), we submit that these fields are extremely high. This agrees with the electric field-driven cold emission of electrons reported for PLZT ceramics above TC, which may even lead to PZT ceramics damage, as was observed [57]. The strength of the electric field locally reaches a value higher than 106 V/cm and transforms the dislocation network into metallic filaments through destruction by hot electrons of the TiO2 or ZrO2 bonding in the vicinity of the virtual anode. The valence of Ti or Zr atoms is thereby reduced, and a new metallic state is created on the core of the electro-degraded network. Therefore, the metallic component of the network is extended.
4. Materials and Methods
5. Conclusions
Supplementary Materials
Author Contributions
Funding
Institutional Review Board Statement
Informed Consent Statement
Data Availability Statement
Conflicts of Interest
Sample Availability
References
- Wang, Y.; Fu, N.; Fu, Z.; Lu, X.; Liao, X.; Wang, H.; Qin, S. A Semi-Automatic Coupling Geophone for Tunnel Seismic Detection. Sensors 2019, 19, 3734. [Google Scholar] [CrossRef] [PubMed]
- Wang, H.; Yang, H.; Jiang, H.; Chen, Z.; Feng, P.X.-L.; Xie, H. A Multi-Frequency PMUT Array Based on Ceramic PZT for Endoscopic Photoacoustic Imaging. In Proceedings of the 2021 21st International Conference on Solid-State Sensors, Actuators and Microsystems (Transducers), Orlando, FL, USA, 20–24 June 2021; pp. 30–33. [Google Scholar]
- Ramanathan, S.; Varadan, V.V.; Varadan, V.K. Deicing of helicopter blades using piezoelectric actuators. In Proceedings of the Smart Structures and Materials 2000: Smart Electronics and MEMS, Newport Beach, CA, USA, 6–8 March 2000; pp. 281–292. [Google Scholar]
- Dahl-Hansen, R.P.; Polfus, J.M.; Vøllestad, E.; Akkopru-Akgun, B.; Denis, L.; Coleman, K.; Tyholdt, F.; Trolier-McKinstry, S.; Tybell, T. Electrochemically driven degradation of chemical solution deposited ferroelectric thin-films in humid ambient. J. Appl. Phys. 2020, 127, 244101. [Google Scholar] [CrossRef]
- Donnelly, N.J.; Randall, C.A. Refined model of electromigration of Ag/Pd electrodes in multilayer PZT ceramics under extreme humidity. J. Am. Ceram. Soc. 2009, 92, 405–410. [Google Scholar] [CrossRef]
- Le Letty, R.; Barillot, F.; Fabbro, H.; Claeyssen, F.; Guay, P.; Cadiergues, L. Miniature Piezo Mechanisms for Optical and Space Applications. In Proceedings of the 9th International Conference on New Actuators, Bremen, Germany, 14–16 June 2004. [Google Scholar]
- Janker, P.; Claeyssen, F.; Grohmann, B.; Christmann, M.; Lorkowski, T.; LeLetty, R.; Sosniki, O.; Pages, A. New Actuators for Aircraft and Space Applications. In Proceedings of the 11th International Conference on New Actuators, Bremen, Germany, 9–11 June 2008. [Google Scholar]
- Allegranza, C.; Gaillard, L.; Le Letty, R.; Patti, S.; Scolamiero, L.; Toso, M. Actuators for Space Applications: State of the Art and New Technologies. In Proceedings of the 14th International Conference on New Actuators, Bremen, Germany, 23–25 June 2014. [Google Scholar]
- Aplin, K.; Middleton, K. A Michelson interferometer system for testing the stability of a piezo-electric actuator intended for use in space. J. Phys. Conf. Ser. 2007, 85, 012013. [Google Scholar] [CrossRef]
- Scott, J.; Dawber, M. Oxygen-vacancy ordering as a fatigue mechanism in perovskite ferroelectrics. Appl. Phys. Lett. 2000, 76, 3801–3803. [Google Scholar] [CrossRef]
- Lou, X.; Hu, X.; Zhang, M.; Morrison, F.; Redfern, S.; Scott, J. Phase separation in lead zirconate titanate and bismuth titanate during electrical shorting and fatigue. J. Appl. Phys. 2006, 99, 044101. [Google Scholar] [CrossRef]
- Lou, X.J.; Zhang, M.; Redfern, S.A.T.; Scott, J.F. Fatigue as a local phase decomposition: A switching-induced charge-injection model. Phys. Rev. B 2007, 75, 224104. [Google Scholar] [CrossRef]
- Bouregba, R.; Sama, N.; Soyer, C.; Poullain, G.; Remiens, D. Interface depolarization field as common denominator of fatigue and size effect in Pb(Zr0.54Ti0.46)O3 ferroelectric thin film capacitors. J. Appl. Phys. 2010, 107, 104102. [Google Scholar] [CrossRef]
- Dausch, D.E. Ferroelectric polarization fatigue in PZT-based RAINBOWs and bulk ceramics. J. Am. Ceram. Soc. 1997, 80, 2355–2360. [Google Scholar] [CrossRef]
- Scott, J.F.; Araujo, C.A.; Melnick, B.M.; Mcmillan, L.D.; Zuleeg, R. Quantitative Measurement of Space-Charge Effects in Lead Zirconate-Titanate Memories. J. Appl. Phys. 1991, 70, 382–388. [Google Scholar] [CrossRef]
- Grossmann, M.; Bolten, D.; Lohse, O.; Boettger, U.; Waser, R.; Tiedke, S. Correlation between switching and fatigue in PbZr0.3Ti0.7O3 thin films. Appl. Phys. Lett. 2000, 77, 1894–1896. [Google Scholar] [CrossRef]
- Lynch, C.; Yang, W.; Collier, L.; Suo, Z.; McMeeking, R. Electric field induced cracking in ferroelectric ceramics. Ferroelectrics 1995, 166, 11–30. [Google Scholar] [CrossRef]
- Pan, M.J.; Park, S.E.; Park, C.W.; Markowski, K.A.; Yoshikawa, S.; Randall, C.A. Superoxidation and Electrochemical Reactions during Switching in Pb (Zr2Ti)O3 Ceramics. J. Am. Ceram. Soc. 1996, 79, 2971–2974. [Google Scholar] [CrossRef]
- Lin, S.; Beom, H.; Tao, D.; Kim, Y. Dielectric breakdown of an unpoled piezoelectric material with a conductive channel. Fatigue Fract. Eng. Mater. Struct. 2009, 32, 580–586. [Google Scholar] [CrossRef]
- Szot, K.; Bihlmayer, G.; Speier, W. Nature of the Resistive Switching Phenomena in TiO2 and SrTiO3: Origin of the Reversible Insulator–Metal Transition. In Solid State Physics; Elsevier: Amsterdam, The Netherlands, 2014; Volume 65, pp. 353–559. [Google Scholar]
- Wojtyniak, M.; Szot, K.; Wrzalik, R.; Rodenbücher, C.; Roth, G.; Waser, R. Electro-degradation and resistive switching of Fe-doped SrTiO3 single crystal. J. Appl. Phys. 2013, 113, 083713. [Google Scholar] [CrossRef]
- Jankowska-Sumara, I.; Szot, K.; Majchrowski, A.; Roleder, K. Effect of resistive switching and electrically driven insulator-conductor transition in PbZrO3 single crystals. Phys. Status Solidi 2013, 210, 507–512. [Google Scholar] [CrossRef]
- Wouters, D.J.; Willems, G.J.; Maes, H.E. Electrical conductivity in ferroelectric thin films. Microelectron. Eng. 1995, 29, 249–256. [Google Scholar] [CrossRef]
- Ossmer, H.; Slouka, C.; Andrejs, L.; Blaha, P.; Friedbacher, G.; Fleig, J. Electrocoloration of donor-doped lead zirconate titanate under DC field stress. Solid State Ion. 2015, 281, 49–59. [Google Scholar] [CrossRef]
- Holzlechner, G.; Kastner, D.; Slouka, C.; Hutter, H.; Fleig, J. Oxygen vacancy redistribution in PbZrxTi1−xO3 (PZT) under the influence of an electric field. Solid State Ion. 2014, 262, 625–629. [Google Scholar] [CrossRef]
- Andrejs, L.; Fleig, J. Resistance degradation in donor-doped PZT ceramic stacks with Ag/Pd electrodes: I. Phenomenology of processes. J. Eur. Ceram. Soc. 2013, 33, 779–794. [Google Scholar] [CrossRef]
- Andrejs, L.; Oßmer, H.; Friedbacher, G.; Fleig, J. Resistance degradation in donor-doped PZT ceramic stacks with Ag/Pd electrodes: II. Distribution of conduction paths. J. Eur. Ceram. Soc. 2013, 33, 1165–1176. [Google Scholar] [CrossRef]
- Wang, R.-V.; McIntyre, P.C. 18O tracer diffusion in Pb (Zr, Ti) O3 thin films: A probe of local oxygen vacancy concentration. J. Appl. Phys. 2005, 97, 023508. [Google Scholar] [CrossRef]
- Gottschalk, S.; Hahn, H.; Flege, S.; Balogh, A. Oxygen vacancy kinetics in ferroelectric PbZr0.4Ti0.6O3. J. Appl. Phys. 2008, 104, 114106. [Google Scholar] [CrossRef]
- Slouka, C.; Holzlechner, G.; Andrejs, L.; Navickas, E.; Hutter, H.; Fleig, J. Oxygen ion conduction in bulk and grain boundaries of nominally donor-doped lead zirconate titanate (PZT): A combined impedance and tracer diffusion study. J. Am. Ceram. Soc. 2015, 98, 3259–3269. [Google Scholar] [CrossRef]
- Härdtl, K.; Rau, H. PbO vapour pressure in the Pb (Ti1−x)O3 system. Solid State Commun. 1969, 7, 41–45. [Google Scholar] [CrossRef]
- Holman, R.L.; Fulrath, R.M. Intrinsic nonstoichiometry in the lead zirconate-lead titanate system determined by Knudsen effusion. J. Appl. Phys. 1973, 44, 5227–5236. [Google Scholar] [CrossRef]
- Randall, C.; Yousefian, P. Fundamentals and practical dielectric implications of stoichiometry and chemical design in a high-performance ferroelectric oxide: BaTiO3. J. Eur. Ceram. Soc. 2022, 42, 1445–1473. [Google Scholar] [CrossRef]
- Valdes, L. Characteristics of M-1768 Transistor. In Proceedings of the Institute of Radio Engineers, San Francisco, CA, USA, July 1953; p. 412. [Google Scholar]
- Sakalauskas, S.; Sodeika, A. Automated measuring instrument of the surface electric potential and potential distribution. Rev. Sci. Instrum. 1998, 69, 466–468. [Google Scholar] [CrossRef]
- Szot, K.; Keppels, J.; Speier, W.; Besocke, K.; Teske, M.; Eberhardt, W. Surface chemistry and molecular reactions on KNbO3 single crystal surfaces. Surf. Sci. 1993, 280, 179–184. [Google Scholar] [CrossRef]
- Liu, Q.; Turhan, A.; Zawodzinski, T.A.; Mench, M.M. In situ potential distribution measurement in an all-vanadium flow battery. Chem. Commun. 2013, 49, 6292–6294. [Google Scholar] [CrossRef]
- Bannani, A.; Bobisch, C.; Möller, R. Local potentiometry using a multiprobe scanning tunneling microscope. Rev. Sci. Instrum. 2008, 79, 083704. [Google Scholar] [CrossRef] [PubMed]
- Smits, E.C.; Mathijssen, S.G.; Cölle, M.; Mank, A.J.; Bobbert, P.A.; Blom, P.W.; de Boer, B.; de Leeuw, D.M. Unified description of potential profiles and electrical transport in unipolar and ambipolar organic field-effect transistors. Phys. Rev. B 2007, 76, 125202. [Google Scholar] [CrossRef]
- Celano, U. Electrical Atomic Force Microscopy for Nanoelectronics; Springer: Berlin/Heidelberg, Germany, 2019. [Google Scholar]
- Rodenbücher, C.; Bittkau, K.; Bihlmayer, G.; Wrana, D.; Gensch, T.; Korte, C.; Krok, F.; Szot, K. Mapping the conducting channels formed along extended defects in SrTiO3 by means of scanning near-field optical microscopy. Sci. Rep. 2020, 10, 17763. [Google Scholar] [CrossRef]
- Rodenbücher, C.; Menzel, S.; Wrana, D.; Gensch, T.; Korte, C.; Krok, F.; Szot, K. Current channeling along extended defects during electroreduction of SrTiO3. Sci. Rep. 2019, 9, 2502. [Google Scholar] [CrossRef]
- Andrejs, L.; Oßmer, H.; Friedbacher, G.; Bernardi, J.; Limbeck, A.; Fleig, J. Conductive AFM and chemical analysis of highly conductive paths in DC degraded PZT with Ag/Pd electrodes. Solid State Ion. 2013, 244, 5–16. [Google Scholar] [CrossRef]
- Zhang, J.; Pan, P.; Jiang, P.; Qin, J.; Hu, J. Electric degradation in PZT piezoelectric ceramics under a DC bias. Sci. Eng. Compos. Mater. 2020, 27, 464–468. [Google Scholar] [CrossRef]
- Zhang, Y.; Zheng, D.Y.; Zhang, C.; Luo, S. Investigation on the piezo-ceramic discoloration during the DC electrical degradation process. AIP Adv. 2021, 11, 065121. [Google Scholar] [CrossRef]
- Kröger, F.; Vink, H. Relations Between the Concentrations of Imperfections in Crystalline Solids. In Solid State Physics; Elsevier: Amsterdam, The Netherlands, 1956; pp. 307–435. [Google Scholar]
- Raymond, M.; Smyth, D. Defects and charge transport in perovskite ferroelectrics. J. Phys. Chem. Solids 1996, 57, 1507–1511. [Google Scholar] [CrossRef]
- Bock, J.A.; Lee, S.; Trolier-McKinstry, S.; Randall, C.A. Metallic-like to nonmetallic transitions in a variety of heavily oxygen deficient ferroelectrics. Appl. Phys. Lett. 2015, 107, 092902. [Google Scholar] [CrossRef]
- Kolodiazhnyi, T. Insulator-metal transition and anomalous sign reversal of the dominant charge carriers in perovskite BaTiO3−δ. Phys. Rev. B 2008, 78, 045107. [Google Scholar] [CrossRef]
- Lee, S.; Bock, J.A.; Trolier-McKinstry, S.; Randall, C.A. Ferroelectric-thermoelectricity and Mott transition of ferroelectric oxides with high electronic conductivity. J. Eur. Ceram. Soc. 2012, 32, 3971–3988. [Google Scholar] [CrossRef]
- Pereira, M.; Peixoto, A.; Gomes, M. Effect of Nb doping on the microstructural and electrical properties of the PZT ceramics. J. Eur. Ceram. Soc. 2001, 21, 1353–1356. [Google Scholar] [CrossRef]
- Kozielski, L.; Buixaderas, E.; Clemens, F.; Bujakiewicz-Korońska, R. PZT Microfibre defect structure studied by Raman spectroscopy. J. Phys. D Appl. Phys. 2010, 43, 415401. [Google Scholar] [CrossRef]
- Buixaderas, E.; Gregora, I.; Kamba, S.; Petzelt, J.; Kosec, M. Raman spectroscopy and effective dielectric function in PLZT x/40/60. J. Phys. Condens. Matter 2008, 20, 345229. [Google Scholar] [CrossRef]
- Delimova, L.; Guschina, E.; Seregin, D.; Vorotilov, K.; Sigov, A. Unexpected behavior of transient current in thin PZT films caused by grain-boundary conduction. J. Appl. Phys. 2017, 121, 224104. [Google Scholar] [CrossRef]
- Lazar, I.; Oboz, M.; Kubacki, J.; Majchrowski, A.; Piecha, J.; Kajewski, D.; Roleder, K. Weak ferromagnetic response in PbZr1−xTixO3 single crystals. J. Mater. Chem. C 2019, 7, 11085–11089. [Google Scholar] [CrossRef]
- Sidebottom, D.L. Dimensionality dependence of the conductivity dispersion in ionic materials. Phys. Rev. Lett. 1999, 83, 983. [Google Scholar] [CrossRef]
- Gundel, H.; Hańderek, J.; Riege, H. Time-dependent electron emission from ferroelectrics by external pulsed electric fields. J. Appl. Phys. 1991, 69, 975–982. [Google Scholar] [CrossRef]
- Zhou, D.; Kamlah, M.; Munz, D. Effects of uniaxial prestress on the ferroelectric hysteretic response of soft PZT. J. Eur. Ceram. Soc. 2005, 25, 425–432. [Google Scholar] [CrossRef]
- Wu, Z.; Cohen, R.E. More accurate generalized gradient approximation for solids. Phys. Rev. B 2006, 73, 235116. [Google Scholar] [CrossRef]
- Rahmanizadeh, K.; Wortmann, D.; Bihlmayer, G.; Blügel, S. Charge and orbital order at head-to-head domain walls in PbTiO3. Phys. Rev. B 2014, 90, 115104. [Google Scholar] [CrossRef]
- Tan, X.; He, H.; Shang, J.K. In situ transmission electron microscopy studies of electric-field-induced phenomena in ferroelectrics. J. Mat. Res. 2005, 20, 1641–1653. [Google Scholar] [CrossRef]
- Weaver, P.; Cain, M.; Stewart, M.; Anson, A.; Franks, J.; Lipscomb, I.; McBride, J.; Zheng, D.; Swingler, J. The effects of porosity, electrode and barrier materials on the conductivity of piezoelectric ceramics in high humidity and dc electric field. Smart Mat. and Struct. 2012, 21, 045012. [Google Scholar] [CrossRef]
- Szot, K.; Rodenbücher, C.; Bihlmayer, G.; Speier, W.; Ishikawa, R.; Shibata, N.; Ikuhara, Y. Influence of dislocations in transition metal oxides on selected physical and chemical properties. Crystals 2018, 8, 241. [Google Scholar] [CrossRef]
- Paladino, A.; Rubin, L.; Waugh, J. Oxygen ion diffusion in single crystal SrTiO3. J. Phy. and Chem. of Solids 1965, 26, 391–397. [Google Scholar] [CrossRef]
- Szot, K.; Speier, W.; Carius, R.; Zastrow, U.; Beyer, W. Localized metallic conductivity and self-healing during thermal reduction of SrTiO 3. Phys. Rev. Lett. 2002, 88, 075508. [Google Scholar] [CrossRef]
- Waugh, J.; Paladino, A.; DiBenedetto, B.; Wantman, R. Effect of Dislocations on Oxidation and Reduction of Single-Crystal SrTiO3. J. Am. Ceram. Soc. 1963, 46, 60. [Google Scholar] [CrossRef]
- Rhodes, W. Etching and Chemical Polishing of Single-Crystal SrTiO3. J. Am. Ceram. Soc. 1966, 49, 110–112. [Google Scholar] [CrossRef]
- Spalding, G.; Murphy, W.; Davidsmeier, T.; Elenewski, J. Faceting of single-crystal SrTiO3 during wet chemical etching. MRS Online Proc. Libr. (OPL) 1999, 587. [Google Scholar] [CrossRef]
- Hensling, F.V.; Du, H.; Raab, N.; Jia, C.-L.; Mayer, J.; Dittmann, R. Engineering antiphase boundaries in epitaxial SrTiO3 2019. APL Mater. 2019, 7, 101127. [Google Scholar] [CrossRef]
- Marrocchelli, D.; Sun, L.; Yildiz, B. Dislocations in SrTiO3: Easy to reduce but not so fast for oxygen transport. J. Am. Ceram. Soc. 2015, 137, 4735–4748. [Google Scholar] [CrossRef] [PubMed]
- Navickas, E.; Chen, Y.; Lu, Q.; Wallisch, W.; Huber, T.M.; Bernardi, J.; Stöger-Pollach, M.; Friedbacher, G.; Hutter, H.; Yildiz, B. Dislocations accelerate oxygen ion diffusion in La0. 8Sr0. 2MnO3 epitaxial thin films. ACS nano 2017, 11, 11475–11487. [Google Scholar] [CrossRef] [PubMed]
- Lamoreaux, R.; Hildenbrand, D.; Brewer, L. High-Temperature Vaporization Behavior of Oxides II. Oxides of Be, Mg, Ca, Sr, Ba, B, Al, Ga, In, Tl, Si, Ge, Sn, Pb, Zn, Cd, and Hg. J. Phys. Chem. Ref. Data 1987, 16, 419–443. [Google Scholar] [CrossRef]
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. |
© 2023 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).
Share and Cite
Lazar, I.; Rodenbücher, C.; Bihlmayer, G.; Randall, C.A.; Koperski, J.; Nielen, L.; Roleder, K.; Szot, K. The Electrodegradation Process in PZT Ceramics under Exposure to Cosmic Environmental Conditions. Molecules 2023, 28, 3652. https://doi.org/10.3390/molecules28093652
Lazar I, Rodenbücher C, Bihlmayer G, Randall CA, Koperski J, Nielen L, Roleder K, Szot K. The Electrodegradation Process in PZT Ceramics under Exposure to Cosmic Environmental Conditions. Molecules. 2023; 28(9):3652. https://doi.org/10.3390/molecules28093652
Chicago/Turabian StyleLazar, Iwona, Christian Rodenbücher, Gustav Bihlmayer, Clive A. Randall, Janusz Koperski, Lutz Nielen, Krystian Roleder, and Krzysztof Szot. 2023. "The Electrodegradation Process in PZT Ceramics under Exposure to Cosmic Environmental Conditions" Molecules 28, no. 9: 3652. https://doi.org/10.3390/molecules28093652
APA StyleLazar, I., Rodenbücher, C., Bihlmayer, G., Randall, C. A., Koperski, J., Nielen, L., Roleder, K., & Szot, K. (2023). The Electrodegradation Process in PZT Ceramics under Exposure to Cosmic Environmental Conditions. Molecules, 28(9), 3652. https://doi.org/10.3390/molecules28093652