Next Article in Journal
Nematogenic Laterally Substituted Supramolecular H-Bonded Complexes Based on Flexible Core
Previous Article in Journal
Cementitious Behavior of Argon Oxygen Decarburization Stainless Steel Slag and Its Stabilization on Chromium
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Latent Tracks in Ion-Irradiated LiTaO3 Crystals: Damage Morphology Characterization and Thermal Spike Analysis

1
Institute of Frontier and Interdisciplinary Science and Key Laboratory of Particle Physics and Particle Irradiation (MOE), Shandong University, Qingdao 266237, China
2
Department of Materials Science & Engineering, University of Tennessee, Knoxville, TN 37996, USA
3
Materials Science & Technology Division, Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA
4
Shanghai Institute of Applied Physics, Chinese Academy of Sciences (CAS), Shanghai 201800, China
*
Author to whom correspondence should be addressed.
Crystals 2020, 10(10), 877; https://doi.org/10.3390/cryst10100877
Submission received: 2 September 2020 / Revised: 24 September 2020 / Accepted: 26 September 2020 / Published: 28 September 2020
(This article belongs to the Section Hybrid and Composite Crystalline Materials)

Abstract

:
Systematic research on the response of crystal materials to the deposition of irradiation energy to electrons and atomic nuclei has attracted considerable attention since it is fundamental to understanding the behavior of various materials in natural and manmade radiation environments. This work examines and compares track formation in LiTaO3 induced by separate and combined effects of electronic excitation and nuclear collision. Under 0.71–6.17 MeV/u ion irradiation with electronic energy loss ranging from 6.0 to 13.8 keV/nm, the track damage morphologies evolve from discontinuous to continuous cylindrical zone. Based on the irradiation energy deposited via electronic energy loss, the subsequently induced energy exchange and temperature evolution processes in electron and lattice subsystems are calculated through the inelastic thermal spike model, demonstrating the formation of track damage and relevant thresholds of lattice energy and temperature. Combined with a disorder accumulation model, the damage accumulation in LiTaO3 produced by nuclear energy loss is also experimentally determined. The damage characterizations and inelastic thermal spike calculations further demonstrate that compared to damage-free LiTaO3, nuclear-collision-damaged LiTaO3 presents a more intense thermal spike response to electronic energy loss owing to the decrease in thermal conductivity and increase in electron–phonon coupling, which further enhance track damage.

Graphical Abstract

1. Introduction

Recent research has demonstrated that perovskite-type LiTaO3 crystals, as representative functional oxides, have extensive potential applications in various radiation environments, such as tritium breeding in fusion reactors [1,2], nuclear waste immobilization [3], and radiation detection [4]. In the research field of ion–solid interactions and irradiation effects, systematic study of the response of crystal materials to the deposition of irradiation energy to electrons and atomic nuclei has always attracted considerable attention. As a research focus, understanding and clarifying the relevant physical mechanisms is not only fundamental to predicting and evaluating the damage behaviors of materials in natural and man-made radiation environments [5], but it is also key to the applications of ion and electron beam techniques in atomic-level defect manipulation, material modification, and micro/nanofabrication [6,7,8,9,10,11,12,13,14,15,16].
In this work, through appropriate choices of irradiating ion energies, velocities, and energy loss intensities, the role of irradiation energy deposited to electron and lattice subsystems and the induced lattice-defect production and structure evolution in LiTaO3 crystals are studied comprehensively. By combining the microstructure characterizations with inelastic thermal spike model calculations, the formation of latent tracks with different damage morphologies under the action of electronic energy loss and the corresponding threshold conditions are analyzed. Based on a disorder accumulation model [17,18], the damage accumulation in LiTaO3 produced by the nuclear energy loss process is also determined. Then, the coupled effect between the electronic excitation and nuclear collision on the thermal spike response and the induced enhanced track damage behavior are further experimentally characterized and theoretically demonstrated.

2. Materials and Methods

2.1. Energetic Ion Irradiation and Damage Characterization

Optically polished single-crystal LiTaO3 samples with a (006) surface normal zone axis direction were used in this work. Irradiation experiments with 1 MeV (0.005 MeV/u) 197Au+ and 20 MeV (0.71 MeV/u) 28Si3+ at 300 K were performed utilizing a 1.7 MV tandem acceleratior (NEC, Middleton, USA) and a 6 MV tandem accelerator (HVEE, Amersfoort, The Netherlands), respectively, within the State Key Laboratory of Nuclear Physics and Technology, Peking University. Irradiation experiments with 247 MeV (6.17 MeV/u) 40Ar12+ and 358 MeV (6.17 MeV/u) 58Ni19+ at 300 K were performed at the Heavy Ion Research Facility in Lanzhou (HIRFL), Institute of Modern Physics, Chinese Academy of Sciences. Relatively low ion-beam current density and ion flux (8.3 × 1010 cm−2·s−1 for Au+, 1.9 × 1010 cm−2·s−1 for Si3+, 6.5 × 1010 cm−2·s−1 for Ar12+, and 8.1 × 1010 cm−2·s−1 for Ni19+) were maintained during the irradiation process to avoid undesired heating and charge accumulation effects on the LiTaO3 samples. Rutherford backscattering (RBS)/channeling analysis was carried out with the abovementioned 1.7 MV tandem accelerator. A 2.0 MeV He+ beam was extracted, and a Si detector located at a scattering angle of 160° relative to the He+ beam was used to collect the backscattered He+ signal and assess the irradiation damage. Cross-sectional LiTaO3 samples for transmission electron microscopy (TEM) observations were prepared through polishing, dimpling, and Ar-ion milling with a Gatan 695 precision polishing system (Gatan Inc., Pleasanton, USA), and the corresponding irradiation damage regions were characterized utilizing an FEI Tecnai G2 F20 transmission electron microscope (FEI Co., Hillsboro, OR, USA).

2.2. Simulations and Calculations of Irradiation Damage Models

As indicated in Table 1, the electronic/nuclear energy losses (Eele/Enucl) and displacements per atom (dpa) in LiTaO3 induced by energetic Au+, Si3+, Ar12+, and Ni19+ irradiation were determined through the stopping and range of ions in matter (SRIM) 2013 simulation code with full-damage cascade mode [19,20], in which the density and default displacement energies of Li, Ta, and O atoms were set to 7.45 g·cm−3, 25 eV, 25 eV, and 28 eV, respectively [21,22]. Based on the irradiation energy deposited by electronic energy loss, the induced energy exchange and diffusion and temperature evolution processes in the electron and lattice subsystems of the LiTaO3 crystal were numerically calculated through two heat-conduction differential equations described in the inelastic thermal spike (iTS) model [23,24,25]. In the iTS model, the energy exchange between the electronic and lattice subsystems is described by a set of thermal diffusion equations, one for the electronic subsystem (Equation (1)) and one for the atomic subsystem (Equation (2))
C e ( T e ) T e t = 1 r r [ r K e ( T e ) T e r ] g ( T e T a ) + A ( r , t )
C a ( T a ) T a t = 1 r r [ r K a ( T a ) T a r ] + g ( T e T a )
where the thermal conductivity K a and specific heat coefficient C a of the lattice subsystem versus lattice temperature were extracted from [26]; the thermal conductivity K e and specific heat coefficient C e of the electron subsystem were fixed to 100 W m−1·K−1 and 1.0 J cm−3·K−1, respectively, owing to the constant thermodynamic parameters in the electron subsystem [23,24]; the electron-phonon mean free path λ was deduced from the band gap (4.7 eV) and set to 4.2 nm [27]; the electron-phonon coupling parameter g was set to 5.7 × 1018 W·m−3·K−1 [26]; the electron-phonon mean free time τ was 1.75 × 10−13 s owing to τ = C e / g [28]; and the latent heat of fusion was set to 150 J/g. The term A ( r , t ) describes the spatial and temporal energy deposition from the incident ion to the electrons within 10−17–10−15 s, and the electrons would deposit their energy close to the ion path within 10−15 s [24,29]. The damage accumulation in LiTaO3 produced by nuclear energy loss originating from low-energy ion irradiation was also fitted through a disorder accumulation model.

3. Results and Discussion

3.1. Latent Ion Tracks Induced by Electronic Energy Loss

As mentioned above, SRIM 2013 simulation code is primarily used to determine the profile of electronic energy loss as a function of ion penetration path, and could not provide the radial distribution of deposited irradiation energy via electronic energy loss process at a certain penetration depth, which actually is essential for the iTS calculations so as to obtain the radial evolutions of electron and lattice temperatures and further describe the lattice melting behavior. During the ion slowing process, the interaction cross section between the irradiating ion and target electrons and the subsequently induced electron cascade in the target directly depend on the ion velocity, which would further determine the radial distribution of the energy deposited in the electron subsystem via the electronic energy loss process. Considering a cylindrical volume in which 0.66 of the total electronic energy is deposited, the corresponding cylindrical radius (radial distance from the ion trajectory) is defined as the absorption radius α e to describe the energy distribution profile [24]. The value of α e would increase with increasing ion velocity (velocity effect), resulting in a large energy spread region and locally low energy deposition density [24]. In this work, as shown in Figure 1a, the α e of the LiTaO3 crystal used for electronic energy deposition increases from 2.1 to 4.7 nm as the ion velocity increases from 0.71 to 6.17 MeV/u. Under ion irradiation with velocities of 0.71, 1.15, 2.15, and 6.17 MeV/u, the fraction of irradiation energy deposited on the electron subsystem derived from the Katz model and MC calculations [24,29,30] and the corresponding normalized energy deposition distribution F ( r ) versus the radial distance from the ion trajectory are shown in Figure 1b. Then, considering the SRIM-simulated electronic energy loss (Eele) relative to the ion velocity and ensuring that the integral of the energy deposition density over the radial area equals Eele [24], the determined distribution profile of the energy deposition density in the electron subsystem is shown in Figure 1c, illustrating that during the ion irradiation process, for a similar Eele (6.0–6.2 keV/nm), the deposition density of electronic energy would accordingly increase with decreasing ion velocity from 6.17 to 0.71 MeV/u (orange and wine curves in Figure 1c). For the same ion velocity (6.17 MeV/u), the deposition density of electronic energy would accordingly increase with increasing Eele from 6.0 to 12.0 keV/nm (orange and blue curves in Figure 1c). For the ion (1.15 MeV/u) with an Eele of 8.4 keV/nm and the ion (6.17 MeV/u) with an Eele of 12.0 keV/nm, the deposition density of electronic energy in the cylindrical core area are similar (pink and blue curves in Figure 1c). For the ion (2.15 MeV/u) with an Eele of 13.8 keV/nm, the deposition density of electronic energy reaches the maximum (green curve in Figure 1c).
Under the action of electronic energy deposition, the induced energy exchange and diffusion and temperature evolution in electron and lattice subsystems are discussed utilizing the inelastic thermal spike (iTS) model [23,24,25]. Figure 1d illustrates the iTS-calculated energy evolution in a LiTaO3 crystal under 0.71 MeV/u ion irradiation with an Eele of 6.2 keV/nm. As shown, the kinetic energy of the incident ion is gradually deposited into the electronic subsystem via electronic energy loss, reaching a peak (6.2 keV/nm) at approximately 10−15 s. The radial distribution of the energy deposition density in the electron subsystem at this time is described by the wine curve in Figure 1c. Then, as shown in the inset of Figure 1d, owing to the much higher temperature in the electron subsystem than in the lattice subsystem, the deposited electronic energy is gradually transferred from the hot electronic subsystem to the cooler lattice through electron-phonon coupling, leading to significant evolution of the electronic and lattice temperatures; once the energy transferred and deposited to lattice atoms exceeds a certain threshold, local melting occurs, and lattice damage along the ion trajectory is formed via a subsequent rapid quenching process. In this energy exchange stage, the sum of irradiation energy deposited into the electron and lattice subsystems remains constant (6.2 keV/nm). Finally, due to the decrease in electron temperature and increase in lattice temperature, the energy exchange driven by the temperature difference ceases to occur since the temperatures between the electron and lattice subsystems reach equilibrium within approximately 10−10 s.
As shown in Figure 2, the temporal evolution profiles of the energy deposition to atoms (Figure 2a,d,g,j,m) and the temporal evolution of the atomic temperature (Figure 2b,e,h,k,n) are shown for the different irradiation conditions at different radii (distance from the ion path), obtained by the iTS model [23,24,25]. The related cross-sectional TEM images in LiTaO3 under ion irradiation with different ion velocities and electronic energy losses are also presented (Figure 2c,f,i,l,o). For 0.71 MeV/u Si3+ irradiation with an Eele of 6.2 keV/nm, the energy deposited to the lattice atoms within a 1 nm radius exceeds the criterion E m for melting phase formation, and the corresponding lattice temperature exceeds the melting temperature T m . It should be pointed out that, according to the distribution of specific heat C a as a function of the temperature T a in LiTaO3 crystal [26], the Em for melting phase formation is calculated to be 0.54 eV/atom (1113 J/g) via E m =   T 0 T m   C a ( T a ) d T a + E l , where T 0 is set to 300 K, and the latent heat of fusion E l is set to 150 J/g [31]. This results in the formation of discontinuous tracks with a radius of ~1.0 nm, as shown in Figure 2c. For 6.17 MeV/u Ar12+ irradiation with an Eele of 6.0 keV/nm, the energy deposited to the lattice atoms along the ion path reaches a peak at 0.47 eV/atom within ~10−13 s, and the corresponding atomic temperature increases to peak at 1860 K, which is still lower than the melting temperature T m (1923 K). Therefore, as shown in Figure 2f, no melting phase or lattice damage would occur [32]. For 1.15 MeV/u Ar12+ irradiation (Eele of 8.4 keV/nm) and 6.17 MeV/u Ni19+ irradiation (Eele of 12.0 keV/nm), accompanied by increasing energy deposition on the lattice atoms, the radii of the produced cylindrical melting zones increase to ~1.5 nm and ~2.0 nm, respectively. As shown in Figure 2i,l, this results in the formation of discontinuous tracks of larger size, with ~1.5 nm and ~2.0 nm diameter, respectively. In the case of 2.15 MeV/u Ni19+ irradiation with an Eele of 13.8 keV/nm, the energy deposited to the lattice atoms in the cylinder center reaches 1.20 eV/atom, significantly exceeding the melting criterion, and lattice melting occurs in a cylindrical region with a radius of ~3.0 nm. Owing to the much higher distributions of atomic energy and temperature, continuous track damage with a radius of ~2.5 nm shown in Figure 2o is formed during the subsequent rapid quenching process [26].

3.2. Damage Accumulation Induced by Nuclear Energy Loss

Utilizing 1 MeV Au+ irradiation, the damage production and accumulation of LiTaO3 crystals due to nuclear energy loss only, originating from the elastic collisions between irradiating ions and target atomic nuclei, were determined. The SRIM-simulated nuclear energy loss (Enucl) profile and displacements per atom (dpa) profile corresponding to the Au+ fluence of 1 × 1014 cm−2 are shown in Figure 3a, illustrating that the dpa peak at 0.71 is located at a depth of ~100 nm. The measured RBS/channeling spectra of Au+-irradiated LiTaO3 under different fluences are shown in Figure 3b, and then, based on an iterative procedure [33], the profiles of Ta sublattice disorder as a function of depth are calculated and shown in Figure 3c, indicating the long-range disorder in LiTaO3 under Au+ irradiation to the local dpa of 0.25. The Ta sublattice disorder in the damage peak region versus the local irradiation dose (dpa) is further plotted in Figure 3d, and the obtained damage accumulation curve is described by a disorder accumulation model [17,18], where the irradiation-induced disorder S (black curve) measured by the RBS/channeling technique would arise from the sum of two contributions: amorphous fraction f a (blue curve) and interstitials and small interstitial clusters in residual crystalline regions S d (red curve). The related fitting parameters—including the amorphization cross section σ a , the effective cross section σ s for defect-stimulated amorphization, the saturation value S d * for the defect-induced disorder, and the coefficient B proportional to an effective defect recombination volume—are also indicated.

3.3. Enhanced Track Damage Induced by Coupled Nuclear and Electronic Energy Losses

To study the coupled effects of electronic and nuclear energy losses, the Au+-irradiated LiTaO3 sample with a fluence of 2 × 1013 cm−2 was subsequently irradiated with 20 MeV (0.71 MeV/u) Si3+ to a fluence of 1 × 1013 cm−2. The measured RBS/channeling spectra, corresponding Ta disorder profiles, cross-sectional TEM images, and FFT patterns of LiTaO3 crystals under different irradiation conditions are shown in Figure 4, and they are used to discuss the Eele-induced track damage response in pristine and Enucl-damaged LiTaO3. According to the RBS/channeling spectra shown in Figure 4a, the determined disorder profiles of the Ta sublattice in ion-irradiated LiTaO3 samples shown in Figure 4b,e,h quantitatively indicated that in contrast to pristine LiTaO3, the same electronic energy loss induced by Si3+ irradiation in Enucl-damaged LiTaO3 leads to complete disorder in the region at a depth of ~100 nm, demonstrating the additive effect of electronic and nuclear energy losses on damage production. Compared to pristine LiTaO3, the enhanced track damage response in the predamaged LiTaO3 upon electronic energy loss is also confirmed by TEM observations. As shown in Figure 4b,c, the lattice defects and defect clusters were first produced by nuclear energy loss via Au+ irradiation; then, upon subsequent 0.71 MeV/u Si3+ irradiation, the introduced Eele approximatively locates at the Bragg peak position, and remains constant (6.2 keV/nm) in the near surface region; finally, the induced track damage in the corresponding Enucl-damaged region is shown in Figure 4h,i. This damage is much more significant than the Eele-induced damage in pristine LiTaO3 (Figure 4e,f).
To understand the damage response of Au+-irradiated (Enucl-damaged) LiTaO3 upon subsequent Si3+ irradiation with intense electronic energy loss, the corresponding energy deposited to atoms and the consequent atomic temperature evolutions were further calculated using the iTS model [23,24,25]. In contrast to pristine LiTaO3, the lattice defects and defect clusters existing in the Au+-irradiated sample scatter electrons and phonons, decreasing thermal conductivities ( K e and K a ) of the electron and lattice subsystems and the electron-phonon mean free path λ , thus leading to an increasing electron–phonon coupling parameter g . Based on the previous research work [5,23,34,35,36], during the iTS calculation process for Au+-irradiated LiTaO3 crystals, K e and K a are assumed to reduce to 75% and g is assumed to be 40% larger compared to the original values of pristine LiTaO3. Comparing Figure 5a,c to Figure 5b,d, under 0.71 MeV/u Si3+ irradiation with an Eele of 6.2 keV/nm, the energy deposited to the lattice along the ion path increases from 0.68 to 1.07 eV/atom within ~10−13 s, and the corresponding atomic temperature increases from 2400 to 3860 K, demonstrating that the introduction of Enucl-induced lattice damage would effectively enhance the response intensity of the Eele-induced thermal spike. In addition, the melting temperature of the crystal decreases owing to the lattice defects and defect clusters embedded in the crystalline system, further promoting the formation of the melting phase and track damage via the subsequent rapid quenching process. Therefore, the experimentally observed additive impacts of the coupled effects of nuclear and electronic energy losses upon track damage production is theoretically demonstrated based on the evolution of energy deposition and atomic temperature in the lattice subsystem.

4. Conclusions

In the research field of ion–solid interactions and irradiation effects, a key challenge is to clarify the deposition of irradiation energy to electron and lattice subsystems and further understand the induced damage production of various materials in irradiation environments. In this work, under ion irradiation with 0.71–6.17 MeV/u with electronic energy loss ranging from 6.0 to 13.8 keV/nm, the experimentally observed latent tracks with different damage morphologies were theoretically discussed based on inelastic thermal spike analysis. The effect of electronic energy loss and absorption radius relative to ion velocity on irradiation energy deposition and exchange processes between electron and lattice subsystems and the further induced evolutions of electronic and atomic temperatures are determined, describing the formation of the lattice melting phase and latent track damage in LiTaO3 crystals. In addition, the lattice defects and defect clusters produced via nuclear energy loss (i) decrease the thermal conductivity and increase the electron-phonon coupling, leading to the enhanced thermal spike response under electronic energy loss and (ii) accordingly reduce the melting temperature of the crystalline system, similarly to [34,36,37,38], demonstrating the coupled mechanism by which electronic and nuclear energy losses promote track damage.

Author Contributions

P.L. and X.W. conceived and designed the experiments; X.H., Y.L., and Q.H. performed the irradiations, RBS/channeling, and TEM measurements; X.H. carried out the iTS-model calculations; M.L.C., E.Z., and P.L. performed the data analysis; P.L. and M.L.C. wrote the paper in consultation with all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (grant numbers 11875038 and 11775135), the Young Scholars Program of Shandong University and the Youth Innovation Promotion Association CAS (grant number 2019262).

Acknowledgments

P.L. acknowledges the support from the National Laboratory of Heavy Ion Accelerator in Lanzhou, and the State Key Laboratory of Nuclear Physics and Technology, Peking University; E.Z. was supported by the U.S. Department of Energy, Office of Science, Basic Energy Sciences, Materials Sciences and Engineering Division.

Conflicts of Interest

The authors declare no conflict interest.

References

  1. Luo, T.Y.; Oda, T.; Oya, Y.; Tanaka, S. IR observation on O-D vibration in LiNbO3 and LiTaO3 single crystal irradiated by 3 keV D2+. J. Nucl. Mater. 2008, 382, 46–50. [Google Scholar] [CrossRef]
  2. Tsuchiya, K.; Kikukawa, A.; Hoshino, T.; Nakamichi, M.; Yamada, H.; Yamaki, D.; Enoeda, M.; Ishitsuka, E.; Kawamura, H.; Ito, H.; et al. In situ tritium recovery behavior from Li2TiO3 pebble bed under neutron pulse operation. J. Nucl. Mater. 2004, 329–333, 1248–1251. [Google Scholar] [CrossRef]
  3. Levy, M.R.; Grimes, R.W.; Sickafus, K.E. Disorder processes in A3+B3+O3 compounds: Implications for radiation tolerance. Philos. Mag. 2004, 84, 533–545. [Google Scholar] [CrossRef]
  4. Glass, A.M.; Abrams, R.L. Study of piezoelectric oscillations in wideband pyroelectric LiTaO3 detectors. J. Appl. Phys. 1970, 41, 4455–4459. [Google Scholar] [CrossRef]
  5. Weber, W.J.; Zarkadoula, E.; Pakarinen, O.H.; Sachan, R.; Chisholm, M.F.; Liu, P.; Xue, H.; Jin, K.; Zhang, Y.W. Synergy of elastic and inelastic energy loss on ion track formation in SrTiO3. Sci. Rep. 2015, 5, 7726. [Google Scholar] [CrossRef] [PubMed]
  6. Krasheninnikov, A.V.; Nordlund, K. Ion and electron irradiation-induced effects in nanostructured materials. J. Appl. Phys. 2010, 107, 071301. [Google Scholar] [CrossRef]
  7. Chen, F.; Wang, X.L.; Wang, K.M. Development of ion-implanted optical waveguides in optical materials: A review. Opt. Mater. 2007, 29, 1523–1542. [Google Scholar] [CrossRef]
  8. Fassbender, J.; McCord, J. Magnetic patterning by means of ion irradiation and implantation. J. Magn. Magn. Mater. 2008, 320, 579–596. [Google Scholar] [CrossRef]
  9. Baglin, J.E.E. Ion beam nanoscale fabrication and lithography—A review. Appl. Surf. Sci. 2012, 258, 4103–4111. [Google Scholar] [CrossRef]
  10. Stepanov, A.L. Synthesis of silver nanoparticles in dialectric matrix by ion implantation: A review. Rev. Adv. Mater. Sci. 2010, 26, 1–29. [Google Scholar]
  11. Gonzalez-Martinez, I.G.; Bachmatiuk, A.; Bezugly, V.; Kunstmann, J.; Gemming, T.; Liu, Z.; Cuniberti, G.; Rümmeli, M.H. Electron-beam induced synthesis of nanostructures: A review. Nanoscale 2016, 8, 11340–11362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Chen, Y.F. Nanofabrication by electron beam lithography and its applications: A review. Microelectron. Eng. 2015, 135, 57–72. [Google Scholar] [CrossRef]
  13. Liu, P.; Zhang, Y.; Xue, H.; Jin, K.; Crespillo, M.L.; Wang, X.; Weber, W.J. A coupled effect of nuclear and electronic energy loss on ion irradiation damage in lithium niobate. Acta Mater. 2016, 105, 429–437. [Google Scholar] [CrossRef] [Green Version]
  14. Crespillo, M.L.; Agulló-López, F.; Zucchiatti, A. Cumulative approaches to track formation under swift heavy ion (SHI) irradiation: Phenomenological correlation with formation energies of Frenkel pairs. Nucl. Instrum. Methods Phys. Res. B 2017, 394, 20–27. [Google Scholar] [CrossRef]
  15. Sellami, N.; Crespillo, M.L.; Zhang, Y.; Weber, W.J. Two-stage synergy of electronic energy loss with defects in LiTaO3 under ion irradiation. Mater. Res. Lett. 2018, 6, 339–344. [Google Scholar] [CrossRef] [Green Version]
  16. Liu, Y.; Huang, Q.; Xue, H.; Crespillo, M.L.; Liu, P.; Wang, X.L. Thermal spike response and irradiation-damage evolution of a defective YAlO3 crystal to electronic excitation. J. Nucl. Mater. 2018, 499, 312–316. [Google Scholar] [CrossRef]
  17. Zhang, Y.; Gao, F.; Jiang, W.; McCready, D.E.; Weber, W.J. Damage accumulation and defect relaxation in 4H-SiC. Phys. Rev. B 2004, 70, 125203. [Google Scholar] [CrossRef]
  18. Zhang, Y.; Weber, W.J.; Jiang, W.; Wang, C.M.; Shutthanandan, V.; Hallén, A. Effects of implantation temperature on damage accumulation in Al-implanted 4H-SiC. J. Appl. Phys. 2004, 95, 4012–4018. [Google Scholar] [CrossRef]
  19. Ziegler, J.F.; Ziegler, M.D.; Biersack, J.P. SRIM-The stopping and range of ions in matter (2010) Methods. Phys. Res. B 2010, 268, 1818–1823. [Google Scholar] [CrossRef] [Green Version]
  20. Ziegler, J.F.; Biersack, J.P.; Ziegler, M.D. SRIM-The Stopping and Range of Ions in Matter. 2013. Available online: www.srim.Org (accessed on 1 April 2020).
  21. Velis, G.; Wendler, E.; Wang, L.L.; Zhang, Y.W.; Weber, W.J. Ion mass dependence of irradiation-induced damage accumulation in KTaO3. J. Mater. Sci. 2019, 54, 149–158. [Google Scholar] [CrossRef]
  22. Meldrum, A.; Boatner, L.A.; Weber, W.J.; Ewing, R.C. Amorphization and recrystallization of the ABO3 oxides. J. Nucl. Mater. 2002, 300, 242–254. [Google Scholar] [CrossRef]
  23. Xue, H.; Zarkadoula, E.; Liu, P.; Jin, K.; Zhang, Y.; Weber, W.J. Amorphization due to electronic energy deposition in defective strontium titanate. Acta Mater. 2017, 127, 400–406. [Google Scholar] [CrossRef] [Green Version]
  24. Toulemonde, M.; Assmann, W.; Dufour, C.; Meftah, A.; Studer, F.; Trautmann, C. Experimental phenomena and thermal spike model description of ion tracks in amorphisable inorganic insulators. Mat. Fys. Medd. 2006, 52, 263–292. [Google Scholar]
  25. Meftah, A.; Costantini, J.M.; Khalfaoui, N.; Boudjadar, S.; Stoquert, J.P.; Studer, F.; Toulemonde, M. Experimental determination of track cross-section in Gd3Ga5O12 and comparison to the inelastic thermal spike model applied to several materials. Nucl. Instrum. Methods Phys. Res. B 2005, 237, 563–574. [Google Scholar] [CrossRef]
  26. Han, X.Q.; Liu, Y.; Huang, Q.; Crespillo, M.L.; Liu, P.; Wang, X.L. Swift heavy ion tracks in alkali tantalate crystals: A combined experimental and computational study. J. Phys. D Appl. Phys. 2020, 53, 105304. [Google Scholar] [CrossRef]
  27. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef]
  28. Toulemonde, M.; Costantini, J.M.; Dufour, C.; Meftah, A.; Paumier, E.; Studer, F. Track creation in SiO2, and BaFe12O19 by swift heavy ions: A thermal spike description. Nucl. Instrum. Methods Phys. Res. B 1996, 116, 37–42. [Google Scholar] [CrossRef]
  29. Waligorski, M.P.R.; Hamm, R.N.; Katz, R. The radial distribution of dose around the path of a heavy ion in liquid water. Nucl. Tracks Radiat. Meas. 1986, 11, 309–319. [Google Scholar] [CrossRef] [Green Version]
  30. Toulemondea, M.; Assmannb, W.; Dufoura, C.; Meftahc, A.; Trautmann, C. Nanometric transformation of the matter by short and intense electronic excitation: Experimental data versus inelastic thermal spike model. Nucl. Instrum. Methods Phys. Res. B 2012, 277, 28–39. [Google Scholar] [CrossRef]
  31. Dufour, C.; Khomrenkov, V.; Wang, Y.Y.; Wang, Z.G.; Aumayr, F.; Toulemonde, M. An attempt to apply the inelastic thermal spike model to surface modifications of CaF2 induced by highly charged ions: Comparison to swift heavy ions effects and extension to some others material. J. Phys. Condens. Matter 2017, 29, 095001. [Google Scholar] [CrossRef]
  32. Liu, Y.; Crespillo, M.L.; Huang, Q.; Han, X.Q.; Wang, X.L.; Liu, P. Latent tracks and novel infrared waveguide formation in lithium tantalate irradiated with swift heavy ions. J. Phys. D Appl. Phys. 2019, 52, 175303. [Google Scholar] [CrossRef]
  33. Zhang, Y.; Lian, J.; Zhu, Z.; Bennett, W.D.; Saraf, L.V.; Rausch, J.L.; Hendricks, C.A.; Ewing, R.C.; Weber, W.J. Response of strontium titanate to ion and electron irradiation. J. Nucl. Mater. 2009, 389, 303–310. [Google Scholar] [CrossRef]
  34. Zarkadoula, E.; Pakarinen, O.H.; Xue, H.; Zhang, Y.W.; Weber, W.J. Predictive modeling of synergistic effects in nanoscale ion track formation. Phys. Chem. Chem. Phys. 2015, 17, 22538–22542. [Google Scholar] [CrossRef] [PubMed]
  35. Zarkadoula, E.; Xue, H.; Zhang, Y.W.; Weber, W.J. Synergy of inelastic and elastic energy loss: Temperature effects and electronic stopping power dependence. Scr. Mater. 2016, 110, 2–5. [Google Scholar] [CrossRef] [Green Version]
  36. Zarkadoula, E.; Jin, K.; Zhang, Y.W.; Weber, W.J. Synergistic effects of nuclear and electronic energy loss in KTaO3 under ion irradiation. AIP Adv. 2017, 7, 015016. [Google Scholar] [CrossRef] [Green Version]
  37. Lam, N.Q.; Okamoto, P.R. Generalized melting criterion for beam-induced amorphization. Surf. Coat. Technol. 1994, 65, 7–14. [Google Scholar] [CrossRef] [Green Version]
  38. Zarkadoula, E.; Zhang, Y.W.; Weber, W.J. Molecular dynamics simulations of the response of pre-damaged SrTiO3 and KTaO3 to fast heavy ions. AIP Adv. 2020, 10, 015019. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) Absorption radius αe of LiTaO3 crystal for energy deposition via electronic energy loss as a function of ion velocity. (b) Normalized distribution of irradiation energy deposited to electron subsystem as a function of radial distance from ion trajectory. The fraction of energy deposited to the electron subsystem versus radius is also shown in the inset. (c) Density distribution of energy deposited to the electron subsystem via ion irradiation with different ion velocities and electronic energy losses. (d) Irradiation energy deposition and subsequent energy exchange processes between electron and lattice subsystems calculated by inelastic thermal spike model. The radial temperature distributions of the electron and lattice subsystems at 2 × 10−15 s are also shown in the inset.
Figure 1. (a) Absorption radius αe of LiTaO3 crystal for energy deposition via electronic energy loss as a function of ion velocity. (b) Normalized distribution of irradiation energy deposited to electron subsystem as a function of radial distance from ion trajectory. The fraction of energy deposited to the electron subsystem versus radius is also shown in the inset. (c) Density distribution of energy deposited to the electron subsystem via ion irradiation with different ion velocities and electronic energy losses. (d) Irradiation energy deposition and subsequent energy exchange processes between electron and lattice subsystems calculated by inelastic thermal spike model. The radial temperature distributions of the electron and lattice subsystems at 2 × 10−15 s are also shown in the inset.
Crystals 10 00877 g001
Figure 2. (a,d,g,j,m) Energy deposited to atoms and (b,e,h,k,n) corresponding evolution of atomic temperature at different radii versus time under the action of ion irradiation with different ion velocities and electronic energy losses. (c,f,i,l,o) TEM-observed latent tracks with different morphologies in LiTaO3 under different irradiation conditions.
Figure 2. (a,d,g,j,m) Energy deposited to atoms and (b,e,h,k,n) corresponding evolution of atomic temperature at different radii versus time under the action of ion irradiation with different ion velocities and electronic energy losses. (c,f,i,l,o) TEM-observed latent tracks with different morphologies in LiTaO3 under different irradiation conditions.
Crystals 10 00877 g002
Figure 3. (a) SRIM-simulated nuclear energy loss and dpa profiles in Au+-irradiated LiTaO3 corresponding to a fluence of 1 × 1014 cm−2. (b) RBS/channeling spectra of LiTaO3 samples irradiated with Au+ at different fluences. (c) Disorder profiles of the Ta sublattice in Au+-irradiated LiTaO3 samples. (d) Relative disorder of the Ta sublattice at the damage peak versus local dose (dpa) and data fitting through the disorder accumulation model.
Figure 3. (a) SRIM-simulated nuclear energy loss and dpa profiles in Au+-irradiated LiTaO3 corresponding to a fluence of 1 × 1014 cm−2. (b) RBS/channeling spectra of LiTaO3 samples irradiated with Au+ at different fluences. (c) Disorder profiles of the Ta sublattice in Au+-irradiated LiTaO3 samples. (d) Relative disorder of the Ta sublattice at the damage peak versus local dose (dpa) and data fitting through the disorder accumulation model.
Crystals 10 00877 g003
Figure 4. (a) RBS/channeling spectra of pristine and predamaged LiTaO3 samples under subsequent 0.71 MeV/u Si3+ irradiation with an Eele of 6.2 keV/nm. Cross-sectional TEM images and corresponding FFT patterns of LiTaO3 under (bd) Au+ irradiation, (eg) Si3+ irradiation, and (hj) sequential Au+ and Si3+ irradiation, respectively.
Figure 4. (a) RBS/channeling spectra of pristine and predamaged LiTaO3 samples under subsequent 0.71 MeV/u Si3+ irradiation with an Eele of 6.2 keV/nm. Cross-sectional TEM images and corresponding FFT patterns of LiTaO3 under (bd) Au+ irradiation, (eg) Si3+ irradiation, and (hj) sequential Au+ and Si3+ irradiation, respectively.
Crystals 10 00877 g004
Figure 5. (a,b) Energy deposited to atoms and (c,d) corresponding evolution of atomic temperature at different radii versus time in damage-free and predamaged LiTaO3 under subsequent electronic energy loss.
Figure 5. (a,b) Energy deposited to atoms and (c,d) corresponding evolution of atomic temperature at different radii versus time in damage-free and predamaged LiTaO3 under subsequent electronic energy loss.
Crystals 10 00877 g005
Table 1. SRIM simulation results corresponding to different ion-irradiation conditions.
Table 1. SRIM simulation results corresponding to different ion-irradiation conditions.
Ion-Irradiation ConditionsSurface RegionEele-Peak Region
speciesfluxfluenceenergyvelocityEeledpadepthenergyvelocityEeledpa
(cm−2·s−1)(cm−2)(MeV)(MeV/u)(keV/nm) (μm)(MeV)(MeV/u)(keV/nm)
28Si3+1.9 × 10101 × 1013200.716.21.6 × 10−40200.716.21.6 × 10−4
40Ar12+6.5 × 10103 × 10122476.176.02.1 × 10−530.0461.158.46.4 × 10−5
58Ni19+8.1 × 10103 × 10123586.1712.02.9 × 10−518.01252.1513.88.0 × 10−5
Ion-Irradiation ConditionsSurface Regiondpa-Peak Region
speciesfluxfluenceenergyvelocityEnucldpadepthenergyvelocityEnucldpa
(cm−2·s−1)(cm−2)(MeV)(MeV/u)(keV/nm) (μm)(MeV)(MeV/u)(keV/nm)
197Au+8.3 × 10101 × 101410.0055.30.240.10.350.00185.40.71

Share and Cite

MDPI and ACS Style

Han, X.; Liu, Y.; Crespillo, M.L.; Zarkadoula, E.; Huang, Q.; Wang, X.; Liu, P. Latent Tracks in Ion-Irradiated LiTaO3 Crystals: Damage Morphology Characterization and Thermal Spike Analysis. Crystals 2020, 10, 877. https://doi.org/10.3390/cryst10100877

AMA Style

Han X, Liu Y, Crespillo ML, Zarkadoula E, Huang Q, Wang X, Liu P. Latent Tracks in Ion-Irradiated LiTaO3 Crystals: Damage Morphology Characterization and Thermal Spike Analysis. Crystals. 2020; 10(10):877. https://doi.org/10.3390/cryst10100877

Chicago/Turabian Style

Han, Xinqing, Yong Liu, Miguel L. Crespillo, Eva Zarkadoula, Qing Huang, Xuelin Wang, and Peng Liu. 2020. "Latent Tracks in Ion-Irradiated LiTaO3 Crystals: Damage Morphology Characterization and Thermal Spike Analysis" Crystals 10, no. 10: 877. https://doi.org/10.3390/cryst10100877

APA Style

Han, X., Liu, Y., Crespillo, M. L., Zarkadoula, E., Huang, Q., Wang, X., & Liu, P. (2020). Latent Tracks in Ion-Irradiated LiTaO3 Crystals: Damage Morphology Characterization and Thermal Spike Analysis. Crystals, 10(10), 877. https://doi.org/10.3390/cryst10100877

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