Skip to content
BY 4.0 license Open Access Published by De Gruyter Open Access September 29, 2023

Effect of gamma rays irradiation in the structure, optical, and electrical properties of samarium doped bismuth titanate ceramics

  • Mansour K. Gatasheh , Mahmoud S. Alkathy , Hamoud Abdoh Kassim EMAIL logo , Jakkapally Pundareekam Goud and Jose A. Eiras EMAIL logo
From the journal Open Chemistry

Abstract

Ceramics have enormous potential in several emerging technologies, including nuclear reactors. Materials with chemical inertness, high-temperature operation, and physical properties stability under applied radiation with high energy are all desired in this field of technology. Given these broad specifications, bismuth titanate ceramics may prove to be a valuable material. Regarding this task, the effect of gamma rays on the structural, optical, and ferroelectric properties of samarium-modified bismuth titanate ceramics was investigated. The Bi3.15Sm0.85Ti3O12 (BSmT) compound was irradiated for 0, 50, 100, and 200 kGy using a 60Co gamma source at a dose rate of 10 kGy h−1. The phase structure confirmed the orthorhombic, single-phase nature even after gamma irradiation. The results show that the unit cell volume decreases from 966.39 to 962.38 Å3 with an increase in gamma dose from 0 to 200 kGy. The X-ray photoelectron spectroscopy study shows an irradiation-induced defect in the host matrix. The results show that the bandgap energy, dielectric constant, Curie temperatures, and remnant polarization slightly decreased with an increase in gamma irradiation. According to the findings of this study, the BSmT exhibits adequate stability against gamma irradiation, which offers tremendous promise in their utilization in nuclear reactor technology.

Graphical abstract

To fully understand how irradiation affects the structural, optical, dielectric, and ferroelectric characteristics of BSmT ceramics, a systematic study is needed. The characteristics of BSmT ceramics were minimally affected by different gamma doses, and they demonstrated high stability against gamma irradiation, which offers considerable promise for their use in nuclear reactor technology.

1 Introduction

As the aerospace and nuclear sectors expand, there is a growing need for innovative memory devices and actuator materials that can withstand harsh conditions like high temperatures and radiation. The perovskite materials based on lead titanate (PZT) are utilized for these objectives due to their high remnant polarization, Curie temperature (T C), piezoelectric constant, and ability to tune their physical properties by doping. Unfortunately, some drawbacks can happen to these materials, such as their environmental toxicity, limiting their use in this objective [16]. Another piezoelectric can be used in place of PZT when operating at extremely high temperatures or after a large, accumulated radiation dose, despite being more expensive and generally less efficient (lower d33) than PZT [6]. Aurivillius family materials with bismuth layer-structured ferroelectric ceramics have been considered a promising candidate for the potential piezoelectric ceramic for high-temperature applications due to their remarkable characteristics, including a high Curie temperature (T C), high resistivity, low aging rate, good temperature stability, and the absence of the toxic element lead [714].

The BiT compound is one of the Aurivillius family materials that can be described by the general formula of (Bi2O2)2+ (A m−1B m O3m+1)2−, which consists of fluorite-like (Bi2O2)2+ layers and perovskite-like (A m−1B m O3m+1)2− blocks stacked alternately along the c-axis [1517]. Perovskite-like (A m−1B m O3m+1)2− block has excellent possibilities in component flexibility, which allows monovalent, divalent, trivalent ions or their mixtures at A-site with dodecahedral coordination, such as Na+, Ca2+, Sm3+, La3+, Nd3+, etc., and transition metal cations at B-site with octahedral coordination, such as Fe3+, Co3+, Ti4+, Nb5+, W6+, etc. The m varies from 1 to 6, representing the number of [BO6] octahedral layers in the perovskite-like blocks [1517]. Exposure of bismuth layer-structured ferroelectrics to a radiation environment is predicted to cause defects such as ionizing and displacement effects. Trapped charges alter the electrical characteristics of the bismuth layer-structured ferroelectric by producing a local electric field, and irradiation may create multiple electron–hole pairs on the host lattice due to ionizing effects.

When a lot of energy is used to expose something, the atoms in the lattice structure shift about from where they normally would be, displaced atoms may leave vacancies and contribute to interstitial defects by upsetting the positions of surrounding atoms, therefore facilitating the propagation of the impurity [1821], quick, heavy, and rapid ion irradiation has been established as a valuable strategy for changing the microstructures of ceramics [2224]. When a material is subjected to energetic ions, it produces a broad range of defect states. They affect the material’s structural, optical, and electrical characteristics [2527]. These variations are highly influenced by the incoming ion’s mass, the energy of the irradiation, and the radiation fluence [28]. The characteristics of structurally dependent materials may be changed in the presence of gamma radiation, affecting their performance. These modifications are strongly influenced by radiation factors such as linear energy transfer, dosage, and energy, as well as material characteristics and structure [29]. Since bismuth is one of the more volatile elements during the sintering process, rare-earth elements such as Sm3+ ions were chosen to combine with Bi3+ ions to reduce the metal’s volatility and increase sinterability [30]. There have been relatively few investigations on the impact of gamma radiation on modified bismuth titanate ceramics [31], and a comprehensive literature search revealed that no studies on gamma-ray irradiated Bi3.15Sm0.85Ti3O12 (BSmT) ceramics had been published. Systematic research is required to comprehend the influence of irradiation on the structural, optical, dielectric, and ferroelectric properties of BSmT ceramics. Various gamma dosages barely influenced the properties of BSmT ceramics and showed good stability against gamma irradiation, which holds great potential for their use in nuclear reactor technology.

2 Experimental details

The detailed preparation methods of Bi3.15Sm0.85Ti3O12, abbreviated as (BSmT) ceramics, are as follows. The polycrystalline material was synthesized using the traditional solid-state reaction technique, using Bi2O3 (Sigma Aldrich 99.9%), Sm2O3 (Sigma Aldrich 99.9%), and TiO2 (Sigma Aldrich 99.9%) powders as the starting ingredients. After determining the appropriate quantities of starting components based on their respective chemical formula, a wet milling process was conducted for 24 h using acetone and zirconia balls. The slurry that underwent milling was subjected to drying in an electric oven operating at 100°C. After drying, the slurry was further subjected to calcination at 850°C for 6 h. The powders that had undergone calcination were subjected to an additional 24 h re-ball milling process using high-energy ball milling. The milled powders were combined with 1 wt% polyvinyl alcohol aqueous solution, serving as a binder for the pressing process. Subsequently, the powders that had undergone the sieving process were subjected to uniaxial pressing to fabricate disc-shaped pellets measuring 10 mm in diameter. The pellets were subjected to a sintering process at a temperature of 1,150°C for 3 h. The sintered ceramic samples underwent irradiation with gamma rays emitted by a 60Co source. The irradiation was conducted at a dosage rate of 10 kGy/h, and the absorbed doses applied to the materials were 50, 100, and 200 kGy. The irradiation process was carried out using the Nordion GC-220 model. The sample’s phase structure was investigated using X-ray diffraction (XRD) analysis using an X-ray diffractometer (PAN analytical X’pert Pro) before and after irradiation. The surface images were performed using field emission scanning electron microscopy (JSM‐7600F). A UV-2600 (Shimadzu UV-spectroscopy) was used to test the optical properties in the 200–900 nm wavelength range. Defects were confirmed using X-ray photoelectron spectroscopy (XPS; JPS-9200). An Agilent E4980A precision LCR meter was used to evaluate the painted samples’ dielectric characteristics as a temperature function. Radiant Precision Premier II Analyzer was used for ferroelectric measurements at a 10 Hz frequency at room temperature.

3 Results and discussion

3.1 XRD analysis

The XRD patterns of BSmT ceramics recorded at room temperature before and after gamma irradiation are shown in Figure 1. The measured patterns demonstrated that all samples, before and after irradiation, exhibited a single orthorhombic phase that matched well with the standard crystal data corresponding to BiT Card no. 73-2181 in orthorhombic crystal symmetry and space group Fmmm [32,33]. There was no indication of secondary phases in examining the acquired data for irradiated ceramics, showing that neither gamma irradiation nor defects generated by significant structural disordering resulted in secondary phases. The XRD patterns indicated that the primary change was connected with peak broadening, lattice strain, and atomic positions due to an external effect such as lattice distortion induced by gamma irradiation. At the same time, the most diverse diffraction patterns were seen for materials irradiated with doses greater than 100 kGy. They displayed decreased intensity and moved to the upper 2θ sides, indicating a reduction in unit cell volume, as illustrated in Figure 1d. According to the literature, the interpretation of crystal lattice deformation may induce variations in structural characteristics and unit cell volume of the crystal lattice. Tensile stress causes deformation to increase or decrease during irradiation.

Figure 1 
                  XRD pattern of BSmT ceramic taken at room temperature before (represented by the black line) and after (shown by the colored lines) gamma irradiation.
Figure 1

XRD pattern of BSmT ceramic taken at room temperature before (represented by the black line) and after (shown by the colored lines) gamma irradiation.

Consequently, the crystal lattice’s properties and volume may shrink [34]. The shrinkage of the unit cell following irradiation may be the cause of this reduction in lattice parameter. The X-ray data were refined to Rietveld refinement analysis using the Fullprof program to confirm the integrity of the absence of impact of the crystal phase of the BSmT sample after radiation, as shown in Figure 2a. The correctness of the studied structure was determined by looking at the values of the R-factors R p%, R wp%, R exp%, and χ 2 . A plot discrepancy between the observed and computed patterns was also utilized to assess the Rietveld refinement’s quality. Table 1 summarizes the structural parameters estimated from this refinement. Peak widening is expected to rise with increasing gamma dosage, leading to a reduction in crystallite size. Figure 2b shows the obtained parameters. According to the acquired results, a change in crystallite size and lattice strain was seen in the case of the sample with radiation of more than 100 kGy, indicating a rise in crystal lattice distortion. These defects might cause the unit cell to shrink compared to the ideal BSmT crystal before irradiation.

Figure 2 
                  (a) XRD refinements performed on BSmT ceramic, with the observed data shown in black circles, the refinement profile in red, the discrepancy between the measured and estimated diffraction patterns in blue, and Bragg reflections indicated in pink. (b) Variation of lattice constant and unit cell volume with gamma dosage. (c) Variation of peak broadening and (117) peaks shifting with gamma dosage is represented. (d) iRepresents the variation of crystallite size (D
                     p) and lattice strain with gamma dosage performed on BSmT ceramic.
Figure 2

(a) XRD refinements performed on BSmT ceramic, with the observed data shown in black circles, the refinement profile in red, the discrepancy between the measured and estimated diffraction patterns in blue, and Bragg reflections indicated in pink. (b) Variation of lattice constant and unit cell volume with gamma dosage. (c) Variation of peak broadening and (117) peaks shifting with gamma dosage is represented. (d) iRepresents the variation of crystallite size (D p) and lattice strain with gamma dosage performed on BSmT ceramic.

Table 1

Rietveld refinement results of XRD data of BSmT ceramics irradiated with different gamma doses and measured at room temperature

Parameter @ 0 kGy @ 50 kGy @ 100 kGy @ 200 kGy
Crystal system Orthorhombic Orthorhombic Orthorhombic Orthorhombic
Space group Fmmm Fmmm Fmmm Fmmm
SG no. 69 69 69 69
a (Å) 5.44638 ± 0.01 5.44082 ± 0.01 5.44330 ± 0.01 5.40420 ± 0.01
b (Å) 5.40754 ± 0.01 5.40827 ± 0.01 5.40280 ± 0.01 5.404 ± 0.01
c (Å) 32.813 ± 0.01 32.813 ± 0.01 32.823 ± 0.01 32.772 ± 0.01
a/b 1.00718 1.00601 1.0075 1.00004
V (Å)3 966.39 965.52 965.29 962.38
Density (g/cm3) 7.771 7.79 7.864 8.005
Measured density (g/cm3) 7.541 7.543 7.524 7.336
Relative density 97.04028 96.82927 95.6765 91.64272
FWHM of (117) peak 0.069228 0.089187 0.094155 0.134634
2θ of (117) peak 30.071 30.079 30.080 30.093
The crystallite size (nm) 124.16 96.38 91.289 63.85
Lattice strain 0.0011 0.0014 0.0015 0.0022
R p (%) 10.8 10.5 10.4 11.2
R wp (%) 11.7 11.4 11.5 12.5
R ex (%) 8.47 8.70 8.60 8.41
χ 2 1.90 1.71 1.78 2.20

3.2 Morphological analysis

Figure 3 shows the typical pre- and post-gamma irradiation morphologies of polished and thermally etched BSmT samples. Grain size and density were shown to decrease with increasing gamma exposure. There was some grain size and orientation diversity among all samples, but overall, they had a plate-like shape. Grain length and thickness were both visibly reduced with increasing gamma dose. Gamma irradiation may significantly decrease the anisotropies of surface and grain boundary energies and mobilities. There were random-aligned grains visible in the samples. Grain size and grain orientation decreased and became more disordered in samples with a high gamma dose, such as 100 and 200 kGy. This finding agreed with the XRD result shown in Figure 2d, where the preferred orientation of (117)-type planes was found in BSmT ceramics. While looking at how gamma-ray irradiation affects grain size, the average length of grains developed along the c-axis was observed. These findings are consistent with previous research that has compared the morphology of ferroelectric thin films before and after being exposed to gamma rays [35,36].

Figure 3 
                  SEM images of surface morphology of BSmT ceramics before and after gamma-ray irradiation.
Figure 3

SEM images of surface morphology of BSmT ceramics before and after gamma-ray irradiation.

3.3 Optical properties analysis

Figure 4a displays the UV–visible spectra of successively irradiated base sodium borosilicate glasses. The optical spectra of unirradiated BSmT ceramics exhibit a central band at 400 nm, indicating a high charge transfer ultraviolet absorption. Visible-region absorption spectra show a more significant improvement with increasing gamma irradiation dosages. Many factors caused during the irradiation process [19,3739] may account for the observed increase in UV absorption intensity, as suggested by earlier publications. Ionizing radiation may cause many electron–hole pairs to form on BSmT, and the trapped charges influence the material’s electrical characteristics by creating a local electric field.

Figure 4 
                  (a) UV–vis absorption spectra, and (b) plots of (αE)2 versus E of the BSmT samples before and after gamma irradiation.
Figure 4

(a) UV–vis absorption spectra, and (b) plots of (αE)2 versus E of the BSmT samples before and after gamma irradiation.

Additionally, atoms in the lattice structure vary from their original location due to displacement effects when a sufficiently significant energy is applied. Atoms that move from their original sites because of the displacement effect may create vacancies there while also contributing to interstitial defects by disrupting the positions of neighboring atoms, causing the impurity to spread [19,3739]. In addition, gamma irradiation may prevent defects from spreading by trapping them on the sample’s exterior [40]. Changes in the induced band as a function of gamma-ray dosage [41] provide strong evidence for an increasing number of the aforementioned caused defects in the lattice. Hence, the improved absorption spectra in the irradiated samples may result from combining the abovementioned causes. The study of optical absorption is an intriguing approach to learn about the band structure and energy bandgap of the irradiated samples by examining optically induced transitions. The optical bandgap of the materials under investigation was calculated using Tauc’s formula [42]:

(1) ( α h ν ) = C ( h ν E g ) n ,

where C is a constant, hv is the energy of a photon, n is an exponent reflecting the band transition that relies on the type of optical transition (n = 1/2) yields the direct energy bandgap, and α is the absorption coefficient. Extrapolating the slope line of the (αhv)2 versus (hv) as shown in Figure 4b. The values obtained were 3.25 eV for samples exposed to no radiation, 3.15 eV for those exposed to 200 kGy, 3.04 eV for those exposed to 100 kGy, and 2.95 eV for those exposed to 600 kGy. The optical absorption edge thus shifts into the visible spectrum. Changes in the induced bandgap energy as a function of gamma-ray dosage [41] provide strong evidence for an increasing number of the aforementioned caused defects in the lattice.

To prove the presence of defects caused by gamma irradiation, post-irradiated and irradiated samples of 200 kGy by gamma rays were chosen. Figure 5(a)–(d), depicts high-resolution XPS spectra with associated fitting findings of Ti 2p and O 1s. The peak spectra of Ti4+ 2p3/2 and Ti4+ 2p1/2 show that the binding energy is indicated at 456.92, 457.62 eV and 462.42, 463.62 eV for the BSmT sample before and after gamma irradiation, respectively, and these banding energies are similar to those of TiO2. Additionally, the spin–orbital splitting energies between the Ti4+ 2p3/2 and Ti4+ 2p1/2 peaks are determined to be 5.5 and 6 eV for the sample before and after irradiation, confirming the oxidation state of Ti4+. Figure 5a shows no signal similar to Ti3+ found for the non-irradiated sample, which may be attributable to its low Ti3+ content. Further two signals are located in the irradiation sample with binding energies of 455.58 and 462.23 eV referred to as Ti3+ 2p3/2 and Ti3+ 2p1/2, respectively [43]. The concentrations of Ti3+ ions in Ti for the irradiation sample were evaluated using the equation described in ref. [44], which was revealed to be 22%.

Figure 5 
                  High-resolution XPS spectra of Ti 2p and O 1s in the BSmT sample: (a) and (b) before and (c) and (d) after gamma irradiation with 200 kGy dosage.
Figure 5

High-resolution XPS spectra of Ti 2p and O 1s in the BSmT sample: (a) and (b) before and (c) and (d) after gamma irradiation with 200 kGy dosage.

The primary peak at 529.54 and 529.49 eV in Figure 4(b) and (d) for the samples before and after gamma irradiation, which attributed to an oxygen lattice in the bulk Ti–O bond [45]. The oxygen vacancies are responsible for the other peak recorded at 530.53 and 530.93 eV. The oxygen vacancies content in the BSmT sample was determined to be 5.13 and 9.54% before and after 200 kGy gamma ray irradiation. As a result, it is inferred that the irradiated sample contains Ti3+ surface defects and oxygen vacancies, which may play an essential role in improving the absorption spectra and decreasing the bandgap energy, confirming the preceding discussion.

3.4 Dielectric permittivity analysis

Polarization and loss processes in materials are heavily influenced by the material’s composition, frequency, and temperature range. The three types are electronic or ionic conduction, dipolar relaxation, and atomic or electronic polarization. To investigate the effect of gamma rays on the dielectric properties of BSmT ceramics, post-irradiated and irradiated samples of 200 kGy by gamma rays were chosen, as depicted in Figure 6(a) and (b). The dielectric as a function of temperature was measured at 1 and 100 kHz. The dielectric constant values before and after irradiation rose with increasing temperature, with the highest value recorded on the temperature dependences for the samples investigated, corresponding to the ferroelectric to the paraelectric phase transition. Furthermore, a diffused phase transition is occurring, which may be due to the fact that the difference in ionic radii between Sm and Bi may cause a local structural distortion, preventing long-range dipole alignment and driving polar nanoregions. As a result, the observed diffused phase change is caused by the substitution of Bi3+. This inhomogeneous distribution causes compositional variability, resulting in microscopic heterogeneity with variable Curie points and observable diffuseness [46]. The Curie temperature value (T C) was shifted from 568 to 496oC after irradiation with a gamma dose of 200 kGy; also, from Figure 6, we can see that the dielectric constant decreased from 648 to 632 after gamma irradiation. The reason behind reducing the phase transition after gamma irradiation can be explained by resorting to structural features. Table 1 shows that the lattice deformation (a/b) decreased with increasing gamma dose, which may explain why the Curie temperature shifted to the lower temperature location. On the other hand, SEM images reveal a decrease in grain size with increasing gamma dose, which leads to a reduction in T C. The decreasing grain size may result in the formation of internal stress, which reduces the phase transition temperature [47]. Other researchers observed the same behavior in other ceramic materials [4850].

Figure 6 
                  Temperature dependence of (a) dielectric constant and (b) dielectric loss of BSmT samples measured at 1 kHz (black curve) and 100 kHz (red curve) before and after 200 kGy gamma irradiation.
Figure 6

Temperature dependence of (a) dielectric constant and (b) dielectric loss of BSmT samples measured at 1 kHz (black curve) and 100 kHz (red curve) before and after 200 kGy gamma irradiation.

3.5 Ferroelectric permittivity analysis

The ferroelectric properties at room temperature, both pre- and post-gamma irradiation, are shown in Figure 7(a). The gamma irradiation was conducted at a specific dosage, while the measurements were performed at a frequency of 10 Hz and a maximum applied field of 73 kV/cm. The exposure to gamma rays resulted in a slight decrease in saturated and remnant polarization levels. However, the ferroelectric properties remained intact after exposure to a total gamma-ray dose of 200 kGy. Figure 7(b) illustrates the changes in the P r and E c parameters due to augmented gamma-ray irradiation, reaching a maximum of 200 kGy. After irradiation, the rate of change of the 2P r value exhibited a slight reduction compared to the BSmT without irradiation. The decrease in ferroelectric polarization may be primarily due to mobile defects, such as oxygen vacancies and gamma-ray stored energy [51,52]. Mobile defects, such as oxygen vacancies, exhibit an interaction with the polarity of the domain, leading to their migration in a manner that minimizes the overall energy. Mobile defects have the potential to aggregate in the domain boundaries, creating a depolarization field inside the grain. Consequently, the interaction between these mobile defects and domain polarity might lead to degradation in the ferroelectric properties of the irradiated materials. Although both samples were fabricated using the same conventional solid-reaction technique, it is evident that they exhibit differences in microstructure, indicating a little but conceivable effect. The XPS investigation in Figure 5 confirms the specific mechanism by which irradiation generates defects in ferroelectrics. A comprehensive examination was conducted on many sources to ascertain potential defects from gamma irradiation. Based on the results of our research, it has been shown that gamma rays have the potential to induce atomic displacement inside the matrix when interacting with matter [53]. When the energy reaches a sufficient level, it may cause the removal of atoms from their original positions, forming point defects [54,55]. In Figure 4, it is shown that energy levels inside the bandgap may be introduced as recombination sites, which are characterized by a reduction in bandgap energy. Modifying material properties, such as conductivity, mobility, and electron density, may occur via interaction with radiation, with the extent and kind of radiation determining the specific effects [56].

Figure 7 
                  (a) Measurement results of P–E hysteresis loops before and after gamma-ray irradiation on BSmT ceramics. (b) Measurement results of the variation of the P
                     r and E
                     c values BSmT ceramics with gamma-ray irradiation.
Figure 7

(a) Measurement results of PE hysteresis loops before and after gamma-ray irradiation on BSmT ceramics. (b) Measurement results of the variation of the P r and E c values BSmT ceramics with gamma-ray irradiation.

4 Conclusions

In summary, this study demonstrates a novel approach for modifying the structural, optical, and ferroelectric characteristics of samarium-modified bismuth titanate ceramics. This study used the solid-state reaction approach to create BSmT ceramics for high-energy mechanical ball milling. A comprehensive investigation was carried out into the effects of gamma irradiation on structural, optical, dielectric, piezoelectric, and ferroelectric characteristics. A 60Co gamma source irradiated the BSmT compound for 0, 50, 100, and 200 kGy. Even after a high dosage of gamma irradiation, the phase structure revealed the orthorhombic single-phase nature. Increasing gamma irradiation dosage reduced the band gap from 3.25 to 2.95 eV, decreasing the dielectric constant. The rationale for this reduction in bandgap was explored using defect creation caused by irradiation, which was corroborated by the XPS analysis. The temperature-dependent dielectric studies demonstrate that BSmT ceramics have a low dielectric loss and temperature-dependent dielectric behavior, showing that they have outstanding dielectric property stabilities, which is crucial for high-temperature device applications. The ferroelectric investigation was conducted at room temperature, and the findings show that a higher gamma dosage causes a modest reduction in residual polarization. According to the research, lead-free compounds derived from irradiation BSmT are promising for photovoltaic applications.

  1. Funding information: The authors gratefully appreciate the funding from Researchers Supporting Project number (RSP2023R393), King Saud University, Riyadh, Saudi Arabia. The authors also thank Sao Paulo Research Foundation (FAPESP: Grant no# 2017/13769-1) and (FAPESP: Grant no# 2023/05716-6) for the financial support.

  2. Author contributions: M.K.G. – investigator and resources; M.A. – writing, correction, and Fullprof software analysis; H.K. – methodology, formal analysis, writing, and original draft preparation; J.P.G. – analysis and writing the original draft; J.A.E. – writing, correction, supervision, and approval of the final version.

  3. Conflict of interest: The authors declare that they have no conflicts of interest.

  4. Ethical approval: The conducted research is not related to either human or animal use.

  5. Data availability statement: All data that has been used is included in the manuscript.

References

[1] Augereau FP, Ferrandis JY, Villard JF, Fourmentel D, Dierckx M, Wagemans J. Effect of intense neutron dose radiation on piezoceramics. J Acoust Soc Am. 2008;123(5):3928.10.1121/1.2935976Search in Google Scholar

[2] Holbert KE, Sankaranarayanan S, McCready SS. Response of lead metaniobate acoustic emission sensors to gamma irradiation. IEEE Trans Nucl Sci. 2005;52(6):2583–90.10.1109/TNS.2005.860708Search in Google Scholar

[3] Holmes-Siedle A, Adams L. Handbook of radiation effects. New York: Oxford University Press; 2002.10.1093/oso/9780198507338.001.0001Search in Google Scholar

[4] Lee SC, Teowee G, Schrimpf RD, Birnie DP, Uhlmann DR, Galloway KF. Total-dose radiation effects on sol-gel derived PZT thin films. IEEE Trans Nucl Sci. 1992;39(6):2036–43.10.1109/23.211401Search in Google Scholar

[5] Benedetto JM, Moore RA, McLean FB, Brody PS, Dey SK. The effect of ionizing radiation on sol-gel ferroelectric PZT capacitors. IEEE Trans Nucl Sci. 1990;37(6):1713–7.10.1109/23.101181Search in Google Scholar

[6] Sinclair AN, Chertov AM. Radiation endurance of piezoelectric ultrasonic transducers – a review. Ultrasonics. 2015;57:1.10.1016/j.ultras.2014.10.024Search in Google Scholar PubMed

[7] Moure A, Castro A, Pardo L. Aurivillius-type ceramics, a class of high-temperature piezoelectric materials: drawbacks, advantages, and trends. Prog Solid State Ch. 2009;37(1):15–39.10.1016/j.progsolidstchem.2009.06.001Search in Google Scholar

[8] Li G, Tian G, Zhai Y, Du J, Zhao L, Kuai W, et al. Textured CaBi4Ti4O15 ceramics with large piezoelectricity, excellent thermal stability, and high resistivity. J Eur Ceram Soc. 2022;42(13):5584–90.10.1016/j.jeurceramsoc.2022.06.049Search in Google Scholar

[9] Kaymaz A. Ionizing radiation response of bismuth titanate-based metal-ferroelectric-semiconductor (MFS) type capacitor. Microelectron Reliab. 2022;133:114546.10.1016/j.microrel.2022.114546Search in Google Scholar

[10] Reinhardt BT, Tittmann BR. Use of the ferroelectric ceramic bismuth titanate as an ultrasonic transducer for high temperatures and nuclear radiation. Sensors. 2021;21(18):6094.10.3390/s21186094Search in Google Scholar PubMed PubMed Central

[11] Shameer Ahmed B, Namratha K, Nandaprakash MB, Somashekar R, Byrappa K. Effect of gamma irradiation on hydrothermally synthesized barium titanate nanoparticles. Radiat Eff Defects. 2017;172(3–4):257–70.10.1080/10420150.2017.1303835Search in Google Scholar

[12] Ferrandis JY, Gatsa O, Combette P, Fourmentel D, Destouches C, Radulovic V, et al. Acoustic instrumentation of the new generation of MTR: effect of nuclear radiation on modified bismuth titanate piezoelectric elements. In: EPJ Web of Conferences. EDP Sciences; 2020. Vol. 225, p. 04012.10.1051/epjconf/202022504012Search in Google Scholar

[13] Jardiel T, Caballero AC, Villegas M. Aurivillius cerâmicas: Bi4Ti3O12-based piezoelectrics. J Ceram Soc Jpn. 2008;116(1352):511–8.10.2109/jcersj2.116.511Search in Google Scholar

[14] Park BH, Kang BS, Bu SD, Noh TW, Lee J, Jo W. Lanthanum-substituted bismuth titanate for use in non-volatile memories. Nature. 1999;401(6754):682–4.10.1038/44352Search in Google Scholar

[15] Alkathy MS, Zabotto FL, Milton FP, Eiras JA. Bandgap tuning in samarium-modified bismuth titanate by site engineering using iron and cobalt co-doping for photovoltaic application. J Alloys Compd. 2022;908:164222.10.1016/j.jallcom.2022.164222Search in Google Scholar

[16] Alkathy MS, Zabotto FL, Lente MH, Eiras JA. Octahedral distortion and oxygen vacancies induced band-gap narrowing and enhanced visible light absorption of Co/Fe co-doped Bi3.25Nd0.75Ti3O12 ferroelectrics for photovoltaic applications. J Phys D: Appl Phys. 2020;53(46):465106.10.1088/1361-6463/aba930Search in Google Scholar

[17] Peláiz-Barranco A, González-Abreu Y. Ferroelectric ceramic materials of the Aurivillius family. J Adv Dielectr. 2013;3(04):1330003.10.1142/S2010135X1330003XSearch in Google Scholar

[18] Raneesh B, Saha A, Kalarikkal N. Effect of gamma radiation on the structural, dielectric and magnetoelectric properties of nanostructured hexagonal YMnO3. Radiat Phys Chem. 2013;89:28–32.10.1016/j.radphyschem.2013.03.040Search in Google Scholar

[19] Solanki A, Choudhary S, Satsangi VR, Shrivastav R, Dass S. Enhanced photoelectrochemical properties of 100 MeV Si8+ ion irradiated barium titanate thin films. J Alloys Compd. 2013;561:114–20.10.1016/j.jallcom.2013.01.154Search in Google Scholar

[20] Medhi N, Nath AK. Gamma-ray irradiation effects on the ferroelectric and piezoelectric properties of barium titanate ceramics. J Mater Eng Perform. 2013;22:2716–22.10.1007/s11665-013-0569-0Search in Google Scholar

[21] Kim BH, Yang SA, Kang SW, Choi GP, Cho SY, Han JK, et al. Change of electrical properties of (K0.5Na0.5)(Mn0.005Nb0.995)O3 thin films induced by gamma-ray irradiation. Curr Appl Phys. 2016;16(5):539–44.10.1016/j.cap.2016.01.009Search in Google Scholar

[22] GhaedRahmati H, Frounchi M, Dadbin S. Piezoelectric behavior of Gamma-radiated nanocomposite hydrogel based on PVP-PEG-BaTiO3. Mater Sci Eng B. 2022;276:115535.10.1016/j.mseb.2021.115535Search in Google Scholar

[23] Sharma S, Paliwal A, Tomar M, Singh F, Puri NK, Gupta V. Effect of ion beam irradiation on dielectric properties of BaTiO3 thin film using surface plasmon resonance. J Mater Sci. 2016;51:4055–60.10.1007/s10853-016-9725-xSearch in Google Scholar

[24] Zinkle SJ. 1.03-Radiation-induced effects on microstructure. Comprehensive Nucl Mater. 2012;1:65–98.10.1016/B978-0-08-056033-5.00003-3Search in Google Scholar

[25] Kumar P, Saxena N, Gupta V, Singh F, Agarwal A. Correlation between surface phonon mode and luminescence in nanocrystalline CdS thin films: an effect of ion beam irradiation. J Appl Phys. 2014;116(4):043517.10.1063/1.4891452Search in Google Scholar

[26] Gautam SK, Singh F, Sulania I, Singh RG, Kulriya PK, Pippel E. Micro-Raman study on the softening and stiffening of phonons in rutile titanium dioxide film: competing effects of structural defects, crystallite size, and lattice strain. J Appl Phys. 2014;115(14):143504.10.1063/1.4868079Search in Google Scholar

[27] Xu J, Lu Q, Lin J, Lin C, Zheng X, Lin T, et al. Enhanced ferro-/piezoelectric properties of tape-casting-derived Er3+-doped Ba0.85Ca0.15Ti0.9Zr0.1O3 optoelectronic thick films. J Adv Ceram. 2020;9:693–702.10.1007/s40145-020-0405-6Search in Google Scholar

[28] Singh F, Singh RG, Kumar V, Khan SA, Pivin JC. Softening of phonons by lattice defects and structural strain in heavy ion irradiated nanocrystalline zinc oxide films. J Appl Phys. 2011;110(8):043926.10.1063/1.3651638Search in Google Scholar

[29] Ashok K, Masin B, Shukla N, Gupta K, Gupta V, Ambati PK, et al. Synthesis and characterization of complex perovskite BMT ceramics and the effect of gamma radiation at microwave frequency. J Korean Ceram Soc. 2022;59:44–50.10.1007/s43207-021-00156-ySearch in Google Scholar

[30] Alkathy MS, Rahman A, Zabotto FL, Milton FP, Raju KJ, Eiras JA. Room-temperature multiferroic behavior in Co/Fe co-substituted layer-structured Aurivillius phase ceramics. Ceram Int. 2022;48(20):30041–51.10.1016/j.ceramint.2022.06.273Search in Google Scholar

[31] Cho SY, Choi GP, Kim EY, Lee GJ, Lee MK, Bu SD. Gamma-ray irradiation effects on electrical properties of bismuth layer-structured ferroelectric ceramics. Ceram Int. 2017 ;43(17):15694–8.10.1016/j.ceramint.2017.08.129Search in Google Scholar

[32] Warren WL, Dimos D, Tuttle BA, Nasby RD, Pike GE. Electronic domain pinning in Pb(Zr,Ti)O3 thin films and its role in fatigue. Appl Phys Lett. 1994;65(8):1018–20.10.1063/1.112211Search in Google Scholar

[33] Roy M, Bala I, Barbar SK, Jangid S, Dave P. Synthesis, structural and electrical properties of La and Nb modified Bi4Ti3O12 ferroelectric ceramics. J Phys Chem Solids. 2011;72(11):1347–53.10.1016/j.jpcs.2011.08.007Search in Google Scholar

[34] Al-Ghamdi H, Almuqrin AH, Kassim H. Effect of gamma irradiation on the structural, optical, electrical, and ferroelectric characterizations of bismuth-modified barium titanate ceramics. Materials. 2022;15(12):4337.10.3390/ma15124337Search in Google Scholar PubMed PubMed Central

[35] Yang SA, Kim BH, Lee MK, Lee GJ, Lee NH, Bu SD. Gamma-ray irradiation effects on electrical properties of ferroelectric PbTiO3 and Pb(Zr0.52Ti0.48)O3 thin films. Thin Solid Films. 2014;562:185–9.10.1016/j.tsf.2014.04.038Search in Google Scholar

[36] Glinšek S, Pečnik T, Cindro V, Kmet B, Rožič B, Malič B. Role of the microstructure in the neutron and gamma-ray irradiation stability of solution-derived Ba0.5Sr0.5TiO3 thin films. Acta Mater. 2015;88:34–40.10.1016/j.actamat.2015.01.028Search in Google Scholar

[37] Mhareb MH, Morsy MA, Almarri H, Sayyed MI, Alrammah I, Alonizan N, et al. Gamma-ray induced effect on the structural and optical properties and durability of neodymium-doped zinc–bismuth–borotellurite glasses and glass ceramics. Opt Mater. 2023;137:113572.10.1016/j.optmat.2023.113572Search in Google Scholar

[38] Saini L, Jani RK, Janu Y, Kumar M, Patra MK, Dixit A. Gamma radiation induced microwave absorption properties of ultra-thin barium titanate (BaTiO3) ceramic tiles over X-band (8.2–12.4 GHz). Ceram Int. 2021;47(16):22397–403.10.1016/j.ceramint.2021.04.249Search in Google Scholar

[39] Brewer SJ, Deng CZ, Callaway CP, Paul MK, Fisher KJ, Guerrier JE, et al. Effect of top electrode material on radiation-induced degradation of ferroelectric thin film structures. J Appl Phys. 2016;120(2):024101.10.1063/1.4955424Search in Google Scholar

[40] Kumar AR, Rao TG, Neeraja K, Reddy MR, Veeraiah N. Gamma ray induced changes on vibrational spectroscopic properties of strontium alumino-borosilicate glasses. Vib Spectrosc. 2013;69:49–56.10.1016/j.vibspec.2013.09.005Search in Google Scholar

[41] Fayad AM, Abd-Allah WM, Moustafa FA. Effect of gamma irradiation on structural and optical investigations of borosilicate glass doped yttrium oxide. Silicon 2018;10:799–809.10.1007/s12633-016-9533-6Search in Google Scholar

[42] Dang NV, Thanh TD, Hong LV, Lam VD. Structural, optical and magnetic properties of polycrystalline BaTi1−xFexO3 ceramics. J Appl Phys. 2011;110(4):043914.10.1063/1.3625235Search in Google Scholar

[43] Erdem B, Hunsicker RA, Simmons GW, Sudol ED, Dimonie VL, El-Aasser MS. XPS and FTIR surface characterization of TiO2 particles used in polymer encapsulation. Langmuir. 2001;17(9):2664–9.10.1021/la0015213Search in Google Scholar

[44] Liu J, Xu J, Cui B, Yu Q, Zhong S, Zhang L, et al. Colossal permittivity characteristics and mechanism of (Sr, Ta) co-doped TiO2 ceramics. J Mater Sci Mater Electron. 2020;31:5205–13.10.1007/s10854-020-03080-9Search in Google Scholar

[45] Li L, Lu T, Zhang N, Li J, Cai Z. The effect of segregation structure on the colossal permittivity properties of (La0.5Nb0.5)xTi1−xO2 ceramics. J Mater Chem C. 2018;6(9):2283–94.10.1039/C7TC05277DSearch in Google Scholar

[46] Filipič C, Kutnjak Z, Pirc R, Canu G, Petzelt J. BaZr0.5Ti0.5O3: lead-free relaxor ferroelectric or dipolar glass. Phys Rev B. 2016;93(22):224105.10.1103/PhysRevB.93.224105Search in Google Scholar

[47] Meng JF, Katiyar RS, Zou GT. Grain size effect on ferroelectric phase transition in Pb1−xBaxTiO3 ceramics. J Phys Chem Solids 1998;59(6–7):1161–7.10.1016/S0022-3697(97)00139-XSearch in Google Scholar

[48] Chattopadhyay S, Ayyub P, Palkar VR, Multani M. Size-induced diffuse phase transition in the nanocrystalline ferroelectric PbTiO3. Phys Rev B 1995;52(18):13177.10.1103/PhysRevB.52.13177Search in Google Scholar PubMed

[49] Lu R, Hu G, Zhao W, Liu T, Fan J, Ma C, et al. Effects of He-ion bombardment on the ferroelectric and dielectric properties of BaHf0.17Ti0.83O3 films. Appl Phys Lett. 2022;121(7).10.1063/5.0107438Search in Google Scholar

[50] Mudinepalli VR, Feng L, Lin WC, Murty BS. Effect of grain size on dielectric and ferroelectric properties of nanostructured Ba0.8Sr0.2TiO3 ceramics, J Adv Ceram. 2015;4:46–53.10.1007/s40145-015-0130-8Search in Google Scholar

[51] Lim J, Lee YS, Yang SA, Choi GP, Bu SD. Change of photoluminescence property of ferroelectric compounds with the gamma-ray irradiation. J Lumin. 2017;188:188–92.10.1016/j.jlumin.2017.04.030Search in Google Scholar

[52] Bu S, Shin E, Park G. Two distinct dielectric relaxation mechanisms in the low-frequency range in (K0.50Na0.50)2(Sr0.75Ba0.25)4Nb10O30 ceramics. Appl Phys Lett. 1998;73(10):1442–4.10.1063/1.122379Search in Google Scholar

[53] Claeys C, Simoen E. Radiation effects in advanced semiconductor materials and devices. Springer Science & Business Media; 2002. p. 21.10.1007/978-3-662-04974-7Search in Google Scholar

[54] Khamari SK, Dixit VK, Ganguli T, Porwal S, Singh SD, Kher S, et al. Effect of 60Co γ-ray irradiation on electrical properties of GaAs epilayer and GaAs p–i–n diode. Nucl Instrum Methods Phys Res. 2011;269(3):272–6.10.1016/j.nimb.2010.11.067Search in Google Scholar

[55] Atanassova E, Paskaleva A, Konakova R, Spassov D, Mitin VF. Influence of γ radiation on thin Ta2O5–Si structures. Microelectron J. 2001;32(7):553–62.10.1016/S0026-2692(01)00043-XSearch in Google Scholar

[56] Kumari M, Rana P, Chauhan RP. Modifications in structural and electrical properties of gamma irradiated CdSe nanowires. Nucl Instrum Methods Phys Res. 2014;753:116–20.10.1016/j.nima.2014.03.062Search in Google Scholar

Received: 2023-07-07
Accepted: 2023-08-11
Published Online: 2023-09-29

© 2023 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 23.5.2024 from https://www.degruyter.com/document/doi/10.1515/chem-2023-0117/html
Scroll to top button