Next Article in Journal
Applications of Gene Expression Programming and Regression Techniques for Estimating Compressive Strength of Bagasse Ash based Concrete
Next Article in Special Issue
Comparative Ab Initio Calculations of ReO3, SrZrO3, BaZrO3, PbZrO3 and CaZrO3 (001) Surfaces
Previous Article in Journal
The Role of Substitution in the Apex Position of the Bent-Core on Mesomorphic Properties of New Series of Liquid Crystalline Materials
Previous Article in Special Issue
The Electronic Properties of Extended Defects in SrTiO3—A Case Study of a Real Bicrystal Boundary
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electron-Beam-Induced Current and Cathodoluminescence Study of Dislocations in SrTiO3

1
National Institute for Materials Science, Tsukuba 305-0044, Japan
2
Faculty of Pure and Applied Sciences, University of Tsukuba, Tsukuba 305-8577, Japan
*
Author to whom correspondence should be addressed.
Crystals 2020, 10(9), 736; https://doi.org/10.3390/cryst10090736
Submission received: 30 July 2020 / Revised: 18 August 2020 / Accepted: 20 August 2020 / Published: 21 August 2020
(This article belongs to the Special Issue Electronic Phenomena of Transition Metal Oxides)

Abstract

:
Electron-beam-induced current (EBIC) and cathodoluminescence (CL) have been applied to investigate the electrical and optical behaviors of dislocations in SrTiO3. The electrical recombination activity and defect energy levels of dislocations have been deduced from the temperature-dependent EBIC measurement. Dislocations contributed to resistive switching were clarified by bias-dependent EBIC. The distribution of oxygen vacancies around dislocations has been obtained by CL mapping. The correlation between switching, dislocation and oxygen vacancies was discussed.
Keywords:
SrTiO3; dislocation; EBIC; CL

Graphical Abstract

1. Introduction

The perovskite oxides, especially SrTiO3 and its related oxides, have attracted much attention because of their dielectric and ferroelectric properties. The oxygen vacancies (Vo) directly govern the bulk and interfacial properties of perovskites, including Vo related dielectric relaxation and electrical conduction [1], n-type doping in SrTiO3 [2], as well as the introduction of two-dimensional electron gas at the interface of SrTiO3/LaAlO3 [3,4]. With the rising interest in resistance switchable oxides, the behavior of Vo has again become the highlight of the study on switching mechanisms [5,6,7]. The switching mechanism has been interpreted either by the interface model due to charge trapping and detrapping or the filament model based on Vo migration [8,9,10,11]. The distribution and diffusion kinetics of Vo in single-crystal SrTiO3 has been investigated by the combination of isotope exchange and time-of-flight secondary ion mass spectrometry, and it is indicated that the Gibbs formation energy of Vo at the interface is lower than in the bulk [12].
On the other hand, the presence of crystallographic defects such as dislocations would also affect the distribution of Vo. In present SrTiO3 crystals, there are still high-density dislocations in the order of 106~108 cm−2. Hence, the interaction between Vo and dislocations are of principal importance to the performances of resistive switching devices. The role of dislocations on the defect chemistry and oxide ion transport properties has been assessed by atomistic simulations with lower formation energies Vo at dislocation cores [13]. To investigate the switching activities of dislocations in SrTiO3, electron-beam-induced current (EBIC) has been applied to metal/SrTiO3 Schottky contacts. For instance, in non-doped (100) SrTiO3 crystal, a transition of EBIC contrast of dislocation arrays from dark to bright was discovered when subjected to electrical stress, and it was explained that the change of EBIC contrast is due to Vo diffusion along dislocations [14]. Our group has performed systematic EBIC studies of differently Nb-doped (111) and (100) SrTiO3 crystals [15,16], and we found that dislocations can act as either recombination centers or conduction paths depending on their character, bias condition and Nb doping concentrations. These studies also suggested that not all the dislocations contribute to the switching phenomenon.
However, there is still a lack of an experiment method to directly identify the local distribution of Vo around dislocations. It has been confirmed that Vo in SrTiO3 would emit light under beam or light injection [17,18,19,20], the emission is attributed to the radiative recombination of carriers via defect levels introduced by Vo. As many studies suggested, dislocations can modulate the distribution of oxygen vacancies; recently, we have applied cathodoluminescence (CL) to non-doped SrTiO3, and enhanced Vo-related luminescence from certain dislocations was found [21].
In the study, we attempt to clarify the electrical activities of dislocations in Nb-doped SrTiO3 by temperature-dependent EBIC measurement. According to the temperature-dependence of EBIC contrast, the energy levels of defects would be known. Generally, shallow level defects become EBIC active at lower temperatures while deep level defects are active at room temperature. This dependence is explained based on Shockley–Read–Hall (SRH) statistics [22], and thus temperature-dependent EBIC has been extensively applied in the study of extended defects in Si-based materials [23,24,25,26] as well as wide-gap semiconductor materials [27]. However, for dislocations in SrTiO3, so far as we know there is no temperature-dependent EBIC performed. Furthermore, a combinational EBIC and CL study of the electrical/optical behaviors of dislocations in SrTiO3 will be carried out. The correlation between dislocation defect level, resistive switching and Vo distribution will be investigated.

2. Materials and Methods

Commercially available n-type Nb-doped SrTiO3 single crystals grown by Verneuil process (SHINKOSYA, Yokohama, Japan) were used in this study. The doping concentration of Nb was about 0.01% by weight, corresponding to a carrier concentration of 2~4 × 1018 cm−3. For the EBIC experiment, in order to obtain an atomically flat surface for the fabrication of better Schottky contact, the specimens were first annealed at 1000 °C in air for 2 h. Schottky contacts were prepared by e-beam deposition of Pt with a thickness of 20 nm on the top surface, and Ohmic contacts were prepared by e-beam deposition of Al with a thickness of 30 nm on the back. EBIC measurements were carried out by a JEOL JSM-7600F field emission scanning electron microscope (FE-SEM) (JEOL, Tokyo, Japan) with the accelerating voltage of 7 kV. The beam current is about 250 pA. Kleindiek micromanipulators installed inside the SEM chamber were used for the electrical connections for EBIC measurement and bias voltage applying. The temperature was varied from 120 K to room temperature by a cooling stage. To obtain an EBIC contrast-temperature profile, images were taken every 10 K steps. The EBIC contrast C is defined by:
C = 100% × (IB − Id)/Ib
where Ib and Id are the EBIC currents collected at the background and dislocation, respectively. For CL observation, bare SrTiO3 specimens which had undergone same annealing procedure were used. CL measurement was conducted by a HORIBA MP32 CL system (HORIBA, Kyoto, Japan) attached to a Hitachi SU6600 FE-SEM (Hitachi High-Tech, Tokyo, Japan). The accelerating voltage was 7 kV same as that of EBIC observation. The temperature was set at 80 K and 300 K, respectively.

3. Results and Discussion

EBIC images were taken in the temperature range of 120~300 K, and in this temperature range the EBIC contrast of dislocations increased with rising temperature. Representative EBIC images taken at low (120 K), medium (200 K) and room temperature (300 K) were demonstrated here. Figure 1a–d shows the secondary electron (SE) and temperature-dependent EBIC images of dislocations in (111) SrTiO3. The in-plane orientations were indicated by the Thompson tetrahedron in Figure 1a. In the EBIC image at 120 K, there were straight and curved/tangled lines with gray contrast. The straight lines were along <112> directions. The straight lines are related to slip dislocations, and the curved/tangled lines are dislocations interacted with each other or in cluster. Bright EBIC contrast was observed around these curved and tangled dislocations when temperature below 200 K. When temperature increased to 200 K, the EBIC contrast of most dislocations showed no obvious change, except for the tangled dislocations in the upright showed stronger dark contrast than the others. The bright EBIC contrast was still visible around the curved/tangled dislocations at this temperature. When temperature further increased to 300 K, the dislocations became more significant with dark contrast. The EBIC contrast varied greatly among different dislocations, i.e., the curved/tangled dislocations exhibited strong dark contrast than the slip lines. The bright EBIC contrast around the curved/tangled dislocations became invisible at 300 K.
To describe the recombination activity and defect levels, we plotted the variation of EBIC contrast of typical dislocations (denoted as sites 1~8 in Figure 1d) with respect to temperature as shown in Figure 2. According to the EBIC contrast statistics, it is clear to see that the EBIC contrast of all dislocations increased monotonically with temperature. At 120 K, the EBIC contrast of all dislocations was weak about 5~10%. The EBIC contrast of tangled dislocations (site 8) quickly increased from 150 K, and became extremely strong (~55%) at 300 K. The EBIC contrast of the curved dislocations (sites 5~7) started to increase from 200 K and was about 30% at 300 K. The EBIC contrast of the straight slip lines (sites 1~4) gradually increased with temperature and was 10~20% at 300 K.
EBIC contrast reflects the recombination activity of minority carries via a defect level, and the temperature-dependence of EBIC contrast is strongly dependent on the position of defect level [26]. Shallow-level defects show weak contrast at room temperature and strong contrast at low temperature, which is due to the change in the occupation of shallow level caused by the shift in the Fermi level with temperature. On the contrary, deep-level defects show stronger contrast at room temperature. This is because that deep level is far from the Fermi level and the occupation of deep level is not greatly affected by the shifts in Fermi level with temperature. Hence, the EBIC contrast of deep-level defect shows a temperature dependence of the form C ∝ T1/2 (temperature dependence of the thermal velocity of carriers).
Based on temperature-dependent EBIC results, we can discuss the recombination activity of dislocations in SrTiO3. The EBIC contrast of all dislocations at room temperature tends to be stronger than that at low temperatures, which suggests that the dislocations in SrTiO3 act as strong recombination centers for minority carriers. The contrast-temperature characteristics indicate that dislocations are accompanied with deep levels, especially the tangled dislocations with more deeper levels. The bright EBIC contrast around the tangled dislocations is due to the formation of denuded zones by impurity gettering. Dislocation clusters could act as strong gettering sites of impurities and/or point defects, and thus to form a region with comparatively longer diffusion length of minority carrier around the defects. A similar phenomenon has been found in our previous EBIC study of grain boundaries in contaminated multicrystalline Si [25]. The appearance of denuded zones at low temperatures is due to the smaller carrier diffusion length at low temperatures. The presence of denuded zones around dislocation clusters could be explained in terms of gettering effect. On the other hand, the dislocation clusters may also act as a sink of oxygen vacancies and result in the formation of denuded zones. So far, there are very few reports on the background impurities in SrTiO3. An early report by Chan et al. [28] suggested that there are over 18 unintentional impurities (Si, Fe, Mg, Cl, C, etc.) in SrTiO3 crystals. At present, it is difficult to figure out the level of impurities decorated around dislocations due to the limitation of analysis techniques.
Next, EBIC studies under bias voltages were conducted to find out the active dislocations for resistive switching. Figure 3a,b shows the room-temperature EBIC images of dislocations taken under zero bias (0 V) and negative bias voltage (−2 V) in (111) SrTiO3. At 0 V, all the dislocations were observed with dark EBIC contrast. At −2 V, bright contrast became visible around and/or at the straight lines related to slip dislocations along <112> directions. However, no bright contrast was found near the curved/tangled dislocations. The narrowing of line contrast under the negative bias indicated that the band bending had suppressed the carrier diffusion and recombination.
The origin of EBIC contrast changing from dark to bright under negative bias voltage is related to the change of depletion layer width under biasing compared to the depth of carrier generation by e-beam injection. The depletion layer width of the Pt/SrTiO3 Schottky was about 130 nm at zero bias and 280 nm at −2 V. The depth of carrier generation by e-beam injection is about 200 nm at the accelerating voltage of 7 kV. When carrier generation is mainly in the bulk region (below the depletion layer), dislocations can act as active recombination centers and show dark EBIC contrast. When biased at −2 V, the width of depletion layer increases and carrier generation is mainly located inside the depletion layer, the recombination process is greatly suppressed and the enhanced transportation of minority carriers via dislocations could be observed.
The bias-dependent EBIC results suggest that some dislocations could act as conduction paths for resistive switching. As for the origin of bright contrast, there are several explanations, such as the migration of Vo along dislocations under bias voltage or carrier tunneling through empty levels associated with these dislocations. It is quite interesting to find that bright EBIC contrast appeared either at low temperatures or under bias voltage. However, the origins of these two bright contrasts are different. The bright contrast around tangled dislocations is related to impurity gettering, while the bright contrast which appeared under negative bias voltages could be explained by the enhanced carrier transport via dislocations inside the depletion region.
To confirm the distribution of Vo around dislocations, CL image and spectrum were taken at 80 K and 300 K. Vo related emission (~2.8 eV) was detected at both temperatures. Hence, we performed CL mapping of dislocations in SrTiO3 at an emission energy of 2.8 eV. Figure 4a,b shows the monochromatic CL images of dislocations taken at 80 K and 300 K in (110) SrTiO3, respectively. For this sample, the emission from background region is not homogenous, suggesting an inhomogeneous distribution of Vo. Dislocations were observed with dark contrast in the room-temperature CL image, and became faint at 80 K. The CL appearance of dislocations at low- and room-temperature is consistent with the temperature-dependent EBIC study. The recombination via dislocations is non-radiative and associated with deep levels. However, it should be noted that bright CL contrast appeared at some of the straight slip lines.
Figure 5 shows the room temperature CL point spectra from three different sites marked in Figure 4b. Site A is close to a straight line with bright CL contrast, Site B is the background, and Site C is a curved dislocation with dark CL contrast. The CL intensity was normalized to the background intensity. It was found that CL spectrum from the different regions showed similar a broad peak centered at 2.8 eV. The peak intensity of Site A is about 1.3 times higher than that of the background, while the peak intensity at the curved dislocation (Site C) slightly decreased.
The bight CL contrast detected from slip lines suggests the enrichment of Vo at these dislocations. Correspondingly, the bright EBIC contrast of dislocations under bias voltage could be related to the migration of Vo along dislocations. The difference between individual and tangled dislocations may be explained from the viewpoint of diffusion path for Vo migration. An individual dislocation, when it is smooth and perpendicular to the surface, could act as a fast diffusion path for Vo. On the contrary, the diffusion Vo would be difficult through tangled dislocations. CL results have indicated that some dislocations in SrTiO3 would modulate the distribution of Vo. However, bright CL contrast was also observed in the regions without dislocations. It is speculated that the distribution of Vo would be affected by a variety of factors. In addition to dislocations, there may be other factors such as strain and impurities. Future studies are needed to clarify this.

4. Conclusions

This paper reported temperature-dependent EBIC and CL studies of dislocations in Nb-doped SrTiO3. The major findings include the following: (1) EBIC indicates that dislocations in SrTiO3 are electrically active at room temperature and act as recombination centers of minority carriers. In particular, the tangled dislocations are strong recombination centers associated with deep levels and denuded zones are formed around them; (2) dislocations arrayed along slip lines are active for resistive switching, while curved and tangled dislocations are not; (3) CL suggests that nonradiative recombination via dislocations is strong at room temperature, which is in good correlation with EBIC. An enrichment of Vo along dislocation slips has been found.

Author Contributions

Conceptualization, J.C. and T.S.; methodology, W.Y. and J.C.; writing—original draft preparation, W.Y. and J.C.; writing—review and editing, J.C. and T.S.; funding acquisition, J.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by JSPS KAKENHI, grant number 18K04248.

Acknowledgments

We are grateful to HORIBA Ltd. for the technique support in cathodoluminescence.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ang, C.; Yu, Z.; Cross, L.E. Oxygen-vacancy-related low-frequency dielectric relaxation and electrical conduction in Bi:SrTiO3. Phys. Rev. B 2000, 62, 228–236. [Google Scholar] [CrossRef] [Green Version]
  2. Ohtomo, A.; Hwang, H.Y. Growth mode control of the free carrier density in SrTiO3−δ films. J. Appl. Phys. 2007, 102, 083704. [Google Scholar] [CrossRef] [Green Version]
  3. Kalabukhov, A.; Gunnarsson, R.; Börjesson, J.; Olsson, E.; Claeson, T.; Winkler, D. Effect of oxygen vacancies in the SrTiO3 substrate on the electrical properties of the LaAlO3/SrTiO3 interface. Phys. Rev. B 2007, 75, 121404. [Google Scholar] [CrossRef] [Green Version]
  4. Siemons, W.; Koster, G.; Yamamoto, H.; Harrison, W.A.; Lucovsky, G.; Geballe, T.H.; Blank, D.H.; Beasley, M.R. Origin of charge density at LaAlO3 on SrTiO3 heterointerfaces: Possibility of intrinsic doping. Phys. Rev. Lett. 2007, 98, 196802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Nian, Y.B.; Strozier, J.; Wu, N.J.; Chen, X.; Ignatiev, A. Evidence for an oxygen diffusion model for the electric pulse induced resistance change effect in transition-metal oxides. Phys. Rev. Lett. 2007, 98, 146403. [Google Scholar] [CrossRef] [Green Version]
  6. Janousch, M.; Meijer, G.I.; Staub, U.; Delley, B.; Karg, S.F.; Andreasson, B.P. Role of oxygen vacancies in Cr-doped SrTiO3 for resistance-change memory. Adv. Mater. 2007, 19, 2232–2235. [Google Scholar] [CrossRef] [Green Version]
  7. Yang, J.J.; Miao, F.; Pickett, M.D.; Ohlberg, D.A.A.; Stewart, D.R.; Lau, C.N.; Williams, R.S. The mechanism of electroforming of metal oxide memristive switches. Nanotechnology 2009, 20, 215201. [Google Scholar] [CrossRef] [Green Version]
  8. Fujii, T.; Kawasaki, M.; Sawa, A.; Kawazoe, Y.; Akoh, H.; Tokura, Y. Electrical properties and colossal electroresistance of heteroepitaxial SrRuO3/SrTi1−xNbxO3 (0.0002 ≤ x ≤ 0.02) Schottky junctions. Phys. Rev. B 2007, 75, 165101. [Google Scholar] [CrossRef]
  9. Li, J.Y.; Ohashi, N.; Okushi, H.; Haneda, H. Temperature dependence of carrier transport and resistance switching in Pt/SrTi1−xNbxO3 Schottky junctions. Phys. Rev. B 2011, 83, 125317. [Google Scholar] [CrossRef]
  10. Szot, K.; Speier, W.; Bihlmayer, G.; Waser, R. Switching the electrical resistance of individual dislocations in single-crystalline SrTiO3. Nat. Mater. 2006, 5, 312–320. [Google Scholar] [CrossRef]
  11. Sawa, A. Resistive switching in transition metal oxides. Mater. Today 2008, 11, 28–36. [Google Scholar] [CrossRef]
  12. De Souza, R.A.; Metlenko, V.; Park, D.; Weirich, T.E. Behavior of oxygen vacancies in single-crystal SrTiO3: Equilibrium distribution and diffusion kinetics. Phys. Rev. B 2012, 85, 174109. [Google Scholar] [CrossRef]
  13. Marrocchelli, D.; Sun, L.; Yildiz, B. Dislocations in SrTiO3: Easy to reduce but not so fast for oxygen transport. J. Am. Chem. Soc. 2015, 137, 4735–4748. [Google Scholar] [CrossRef] [PubMed]
  14. Jiang, W.; Evans, D.; Bain, J.A.; Skowronski, M.; Salvador, P.A. Electron beam induced current investigations of Pt/SrTiO3−x interface exposed to chemical and electrical stresses. Appl. Phys. Lett. 2010, 96, 092102. [Google Scholar] [CrossRef]
  15. Chen, J.; Sekiguchi, T.; Li, J.Y.; Ito, S.; Yi, W.; Ogura, A. Investigation of dislocations in Nb-doped SrTiO3 by electron-beam-induced current and transmission electron microscopy. Appl. Phys. Lett. 2015, 106, 102109. [Google Scholar] [CrossRef]
  16. Chen, J.; Sekiguchi, T.; Li, J.Y.; Ito, S. Investigation of dislocations in Nb-doped (100) SrTiO3 single crystals and their impacts on resistive switching. Superlattices Microstruct. 2016, 99, 182–185. [Google Scholar] [CrossRef]
  17. Kan, D.; Terashima, T.; Kanda, R.; Masuno, A.; Tanaka, K.; Chu, S.; Kan, H.; Ishizumi, A.; Kanemitsu, Y.; Shimakawa, Y.; et al. Blue-light emission at room temperature from Ar+-irradiated SrTiO3. Nat. Mater. 2005, 4, 816–819. [Google Scholar] [CrossRef]
  18. Bruno, F.Y.; Tornos, J.; Olmo, M.G.; Santolino, G.S.; Nemes, N.M.; Garcia-Hernandez, M.; Mendez, B.; Piqueras, J.; Antorrena, G.; Morellon, L.; et al. Anisotropic magnetotransport in SrTiO3 surface electron gases generated by Ar+ irradiation. Phys. Rev. B 2011, 83, 245120. [Google Scholar] [CrossRef] [Green Version]
  19. Lim, J.; Lee, Y.S.; Bu, S.D. Surface-direction dependence of the oxygen vacancy formation in SrTiO3 single crystals. Ceram. Int. 2018, 44, S93–S95. [Google Scholar] [CrossRef]
  20. Lee, D.; Wang, H.W.; Noesges, B.A.; Asel, T.J.; Pan, J.B.; Lee, J.W.; Yan, Q.; Brillson, L.J.; Wu, X.; Eom, C.B. Identification of a functional point defect in SrTiO3. Phys. Rev. Mater. 2018, 2, 060403(R). [Google Scholar] [CrossRef]
  21. Wang, P.; Yi, W.; Chen, J.; Ito, S.; Cui, C.; Sekiguchi, T. Oxygen vacancy migration along dislocations in SrTiO3 by cathodoluminescence. J. Phys. D Appl. Phys. 2019, 52, 475103. [Google Scholar] [CrossRef]
  22. Shockley, W.; Read, W.T. Statistics of the recombination of holes and electrons. Phys. Rev. 1952, 87, 835–842. [Google Scholar] [CrossRef]
  23. Kusanagi, S.; Sekiguchi, T.; Shen, B.; Sumino, K. Electrical activity of extended defects and gettering of metallic impurities in silicon. Mater. Sci. Technol. 1995, 11, 685–690. [Google Scholar] [CrossRef]
  24. Kveder, V.; Kittler, M.; Schröter, W. Recombination activity of contaminated dislocations in silicon: A model describing electron-beam-induced current contrast behavior. Phys. Rev. B 2001, 63, 115208. [Google Scholar] [CrossRef]
  25. Chen, J.; Sekiguchi, T.; Yang, D.; Yin, F.; Kido, K.; Tsurekawa, S. Electron-beam-induced current study of grain boundaries in multicrystalline silicon. J. Appl. Phys. 2004, 96, 5490–5495. [Google Scholar] [CrossRef]
  26. Kittler, M.; Seifert, W. Two types of electron-beam-induced current behaviour of misfit dislocations in Si(Ge): Experimental observations and modelling. Mater. Sci. Eng. B 1994, 24, 78–81. [Google Scholar] [CrossRef]
  27. Maximenko, S.I.; Freitas, J.A., Jr.; Myers-Ward, R.L.; Lew, K.-K.; VanMil, B.L.; Eddy, C.R., Jr.; Gaskill, D.K.; Muzykov, P.G.; Sudarshan, T.S. Effect of threading screw and edge dislocations on transport properties of 4H–SiC homoepitaxial layers. J. Appl. Phys. 2010, 108, 013708. [Google Scholar] [CrossRef] [Green Version]
  28. Chan, N.-H.; Sharma, R.K.; Smyth, D.M. Nonstoichiometry in SrTiO3. J. Electrochem. Soc. 1981, 128, 1762–1769. [Google Scholar] [CrossRef]
Figure 1. SE and temperature-dependent electron-beam-induced current (EBIC) images of dislocations in (111) SrTiO3. (a) SE; (b) EBIC 120 K; (c) EBIC 200 K; (d) EBIC 300 K. Site 1–4 are dislocations slips and site 5–8 are tangled dislocations.
Figure 1. SE and temperature-dependent electron-beam-induced current (EBIC) images of dislocations in (111) SrTiO3. (a) SE; (b) EBIC 120 K; (c) EBIC 200 K; (d) EBIC 300 K. Site 1–4 are dislocations slips and site 5–8 are tangled dislocations.
Crystals 10 00736 g001
Figure 2. Variation of EBIC contrast of dislocation related defects with respect to temperature. The different sites were denoted by 1~8 in Figure 1d.
Figure 2. Variation of EBIC contrast of dislocation related defects with respect to temperature. The different sites were denoted by 1~8 in Figure 1d.
Crystals 10 00736 g002
Figure 3. Room-temperature EBIC images of dislocations in (111) SrTiO3 taken at different bias voltages. (a) 0 V; (b) −2 V.
Figure 3. Room-temperature EBIC images of dislocations in (111) SrTiO3 taken at different bias voltages. (a) 0 V; (b) −2 V.
Crystals 10 00736 g003
Figure 4. Monochromatic cathodoluminescence (CL) images of dislocations in (110) oriented SrTiO3 taken at the emission energy of 2.8 eV. (a) 80 K and (b) 300 K. Site A–C: A—straight line with bright contrast; B—background; C—curved dislocation with dark contrast.
Figure 4. Monochromatic cathodoluminescence (CL) images of dislocations in (110) oriented SrTiO3 taken at the emission energy of 2.8 eV. (a) 80 K and (b) 300 K. Site A–C: A—straight line with bright contrast; B—background; C—curved dislocation with dark contrast.
Crystals 10 00736 g004
Figure 5. Room temperature CL spectra of different sites denoted as A–C in Figure 4b.
Figure 5. Room temperature CL spectra of different sites denoted as A–C in Figure 4b.
Crystals 10 00736 g005

Share and Cite

MDPI and ACS Style

Yi, W.; Chen, J.; Sekiguchi, T. Electron-Beam-Induced Current and Cathodoluminescence Study of Dislocations in SrTiO3. Crystals 2020, 10, 736. https://doi.org/10.3390/cryst10090736

AMA Style

Yi W, Chen J, Sekiguchi T. Electron-Beam-Induced Current and Cathodoluminescence Study of Dislocations in SrTiO3. Crystals. 2020; 10(9):736. https://doi.org/10.3390/cryst10090736

Chicago/Turabian Style

Yi, Wei, Jun Chen, and Takashi Sekiguchi. 2020. "Electron-Beam-Induced Current and Cathodoluminescence Study of Dislocations in SrTiO3" Crystals 10, no. 9: 736. https://doi.org/10.3390/cryst10090736

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