Next Article in Journal
Effect of Ultrafine Fly Ash and Water Glass Content on the Performance of Phosphorus Slag-Based Geopolymer
Previous Article in Journal
Recent Advances in Adsorptive Nanocomposite Membranes for Heavy Metals Ion Removal from Contaminated Water: A Comprehensive Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigation of Structural, Physical, and Attenuation Parameters of Glass: TeO2-Bi2O3-B2O3-TiO2-RE2O3 (RE: La, Ce, Sm, Er, and Yb), and Applications Thereof

by
Nehal Elkhoshkhany
1,2,
Samir Marzouk
3,
Mohammed El-Sherbiny
2,
Heba Ibrahim
2,
Bozena Burtan-Gwizdala
4,
Mohammed S. Alqahtani
5,6,
Khalid I. Hussien
5,7,*,
Manuela Reben
8 and
El Sayed Yousef
9,10,*
1
Physics Department, College of Arts and Sciences at Tabrjal, Jouf University, Sakaka 72388, Saudi Arabia
2
Department of Material Science, Institute of Graduate Studies and Researches, Alexandria University, 163 Horreya Avenue, Shatby, Alexandria 21526, Egypt
3
Department of Basic and Applied Science, Faculty of Engineering and Technology, Arab Academy of Science and Technology, Cairo 11511, Egypt
4
Institute of Physics, Cracow University of Technology, ul. Podchorazych 1, 30-084 Cracow, Poland
5
Department of Radiological Sciences, College of Applied Medical Sciences, King Khalid University, Abha 61421, Saudi Arabia
6
BioImaging Unit, Space Research Centre, Department of Physics and Astronomy, University of Leicester, Leicester LE1 7RH, UK
7
Department of Medical Physics and Instrumentation, National Cancer Institute, University of Gezira, Wad Medani 2667, Sudan
8
Faculty of Materials Science and Ceramics, AGH—University of Science and Technology, al. Mickiewicza 30, 30-059 Cracow, Poland
9
Physics Department, Faculty of Science, King Khalid University, Abha 61413, Saudi Arabia
10
Research Center for Advanced Materials Science (RCAMS), King Khalid University, Abha 61413, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Materials 2022, 15(15), 5393; https://doi.org/10.3390/ma15155393
Submission received: 20 July 2022 / Revised: 28 July 2022 / Accepted: 3 August 2022 / Published: 5 August 2022

Abstract

:
A novel series of glass, consisting of B2O3, Bi2O3, TeO2, and TiO2 (BBTT) containing rare earth oxide RE2O3, where RE is La, Ce, Sm, Er, and Yb, was prepared. We investigated the structural, optical, and gamma attenuation properties of the resultant glass. The optical energy bands, the linear refractive indices, the molar refractions, the metallization criteria, and the optical basicity were all determined for the prepared glass. Furthermore, physical parameters such as the density, the molar volume, the oxygen molar volume, and the oxygen packing density of the prepared glass, were computed. Both the values of density and optical energy of the prepared glass increased in the order of La2O3, Ce2O3, Sm2O3, Er2O3, and then Yb2O3. In addition, the glass doped with Yb2O3 had the lowest refractive index, electronic polarizability, and optical basicity values compared with the other prepared glass. The structures of the prepared glass were investigated by the deconvolution of infrared spectroscopy, which determined that TeO4, TeO3, BO4, BO3, BiO6, and TiO4 units had formed. Furthermore, the structural changes in glass are related to the ratio of the intensity of TeO4/TeO3, depending on the type of rare earth. It is also clarified that the resultant glass samples are good attenuators against low-energy radiation, especially those that modified by Yb2O3, which exhibited superior shielding efficiency at energies of 622, 1170, and 1330 keV. The optical and gamma ray spectroscopy results of the prepared glass show that it is a good candidate for nonlinear optical fibers, laser solid material, and optical shielding protection.

1. Introduction

Rare earth (RE) elements, such as La, Ce, Pr, Nd, Sm, Eu, Gd, Tb, Dy, Ho, Er, Tm, and Yb have many multilateral applications in advanced technology [1,2]. Recently, glass doped with rare earth (RE) ions has attracted interest due to its utility in various applications in the development of opto-electronic devices such as planar waveguides, fluorescent display devices, optical fibers, visible lasers, optical detectors, and optical amplifiers [3]. Tellurite glass, especially, has good optical characteristics such as a high dielectric constant, high refractive index, wide optical transmission windows, good semiconducting properties [4], low melting point, and low phonon energy ( ~ 700   cm 1 ) , as compared to borate glass which has high phonon energy and exists at a range of 1300–1500 cm 1 [4,5,6]. Glass based on TeO2 has a weak Te–O bond as compared to other composite glass; the former can easily be broken, which is useful for accommodating metal oxides and rare earth ions [7]. Combined B2O3 and TeO2 (borotellurite) glass exhibits high thermal stability, low phonon energy, easy fabrication, and chemical durability [8]. Borotellurite glass is widely used in several applications, particularly in opto-acoustics, radiation shielding, and micro-electronics [8]. Glass that contains heavy metal oxides (HMO), such as Bi2O3, has high nonlinear second/third-harmonic generation, exhibits high thermal expansion, high density, and IR transmission [8,9,10]. TiO2 incorporated into the glass matrix improves covalent bond formation and reinforces a continuous network consisting of TiO6 with an increase in the amount of bridging oxygen (BO) [10,11,12,13]. Moreover, TiO2 improves the properties of chemical resistance and thermal stability when added to tellurite glass. Nupur Gupta et al. [10] reported that the addition of both TiO2 and Bi2O3 to borotellurite glass leads to a decrease in the optical energy gap due to the formation of nonbridging oxygen (NBO), and the glass transition temperature, Tg, was increased. The addition of La2O3 to tellurite glass enhances the stability of the glass against crystallization. Cerium-ion-doped glass holds several applications in biosensors, solid oxide fuel cells, dielectric materials, and blue luminescent optical systems [14]. Depending on the excitation wavelength, Sm2O3-doped glass displays and emits a strong orange-red luminescence in the visible range [15], associated with 4G5/2, 6H9/2 transition. Er2O3-doped glass is considered to be a candidate for use in the manufacturing of optical amplifiers (EDFAS) and in the field of optical communications [16]. Furthermore, several researchers [17,18,19,20,21] have investigated tellurium-based glass for possible uses in nuclear-radiation-shielding applications. The shielding effectiveness for tellurium-based glass has been studied, with the results showing that glass containing 60 Mol % of lead oxide and heavy oxide, and recording the highest density resulted in higher linear attenuation values and superior shielding material protection with perfect shielding efficiency. In the same trend, our research group [22] developed a computation tool (MIKE) for estimating and analyzing the shielding and optical parameters for different types of shielding materials. Therefore, the purpose of the present research was to study the physical parameters as well as the optical and attenuation properties of novel tellurite glass structures modified with various rare earths ions: La3+, Ce3+, Sm3+, Er3+, and Yb3+.

2. Experimental Section

The prepared glass (BBTTER) with a composition of 25B2O3–20Bi2O3–45TeO2–7TiO2 (BBTT), modified by 3RE2O3 in mol % , where RE is La; Ce; Sm; Er; or Yb, was manufactured using the melt-quenching method. Chemical powders of TeO2, B2O3, Bi2O3, TiO2, and RE2O3 (La2O3, Ce2O3, Sm2O3, Er2O3, and Yb2O3) were homogenized and melted at 930 ° C for 30 min in an electric furnace. We used a platinum crucible while melting to obtain a homogeneous mixture, stirring the mixture several times. Then, each melt was put into a polished stainless-steel container and annealed at 300 ° C . The prepared glass samples are coded as: BBTTLa, BBTTCe, BBTTSm, BBTTEr, and BBTTYb, as shown in Table 1. Al2O3 powder of 600 grade was used to polish the glass samples. The value density of the samples was determined according to Archimedes’ method. The powder X-ray diffraction pattern (XRD) was determined using a Philips PW (1140) diffractometer, and a copper target (Kα = 1.54 Å) was used to study the amorphous properties of prepared glass. A double-beam UV–Visible spectrophotometer (JASCO Corp, v-570, Rel-00, Tokyo, Japan) was used to determine the optical absorption spectrum of the glass. The refractive index was determined with a prism coupler (Metricon Model 2010, Pennington, NJ, USA). Structural characterization of the glass was carried out using FTIR absorbance spectra (Perkin-Elmer spectrometer). For the prepared glass, the half-value layer, both the linear and mass attenuation coefficients, as well as the mean free pass were measured using a NaI detector system (SPECTECH-NaI 1.5 PX 1.5/2.0 IV, S/N 010723-6), with various gamma sources (Am241-5µCI-59.5 keV, Cs137-5µCI- 662 keV, Co60-5µCI-1170, and 1330 keV), which was connected to a computer and based on the multichannel analyzer. Figure 1 shows a collimated beam which was produced at the detector level using a variety of gamma sources (Am241-5µCI-59.5 keV, Cs137-5µCI-662 keV, Co60-5µCI-1170, and 1330 keV), according to the technique described in [4].

3. Results and Discussion

3.1. XRD, Physical Parameters, and UV–VIS–NIR Spectra

The X-ray diffraction (XRD) patterns of the glass samples were measured to investigate the nature of the glass samples, as shown in Figure 2. The absence of any discrete or sharp diffraction peaks in these profiles and the existence of broad bands proves that all the prepared glass samples had an amorphous nature. The density value ( ρ ) was calculated using Equation (1) [23].
ρ = ( W a W a W t ) · ρ t   ( gm / cm 3 )
The weight of the glass sample in air is “Wa”, whereas the weight of the glass immersed in reference liquid toluene is “Wt”, where ρ t = 0.864   g · cm 3 . The calculated density value of the prepared glass increased from 5.67 g·cm−3 to 6.32 g·cm−3, which corresponded to BBTTLa and BBTTYb glass, respectively. The results are shown in Table 1.
The molecular weights, Mwt, of rare earth compounds are ordered as follows: La2O3 < Ce2O3 < Sm2O3 < Er2O < Yb2O3, corresponding to 325.5, 328.24, 348.72, 382.52, and 394.08 g·mol−1, respectively. Thus, the density increased in the same trend, which means that the highest value of density occurred with Yb2O3 incorporated into the glass matrix; here, the network of the glass is more compact. It is possible to determine the glass sample’s molecular volume (Vm) and its oxygen molar volume (Vo) using Equations (2) and (3), respectively. These strongly depend on the value of the densities of the glass samples, tabulated in Table 2.
V m = x i   M wti ρ
where “Mwti” is the molecular weight and “xi” is the fraction ratio of each oxide. The oxygen molar volume, VO, can be estimated by the relationship as follow:
  V O = ( x i   M wti / ρ ) ( 1 / ( x i   n i )
where “ni” is the number of oxygen atoms in each oxide [24]. The number of bonds per unit volume, nb, of the prepared glass and the average force constant ( F ¯ ), calculated from Equations (4) and (5), respectively [25,26].
  n b = N a V m   x i   n f
where nf is the cation coordination number and Na is the Avogadro’s number.
F ¯ = x i   n f   f i   x i   n f
The stretching force, fi, of the oxide, i, can be measured using the following formula:
  f i = 1.7 r 3
The molar volume, Vm, decreased from 34.9 cm3/mol to 31.6 cm3/mol, which corresponded to BBTTLa and BBTTYb, respectively, and the glass structures became more compact due to a reduction in interatomic space or bond length between atoms (ri). The value of V0 decreased from 14.1 to 12.7 cm3·mol−1 by increasing both values of nb from 6.5 to 7.14 × 1022 m−3 and F ¯ from 301.21 to 303.04 Nm−1, when replacing the modified rare earth oxide La2O3 by Yb2O3. Furthermore, the decrease in Vo may indicate a decrement in the formation of NBO atoms reported here. The values of Vm, V0, nb, and F ¯ of the prepared glass are shown in Table 2. Equation (7) was used to compute the oxygen packing density, OPD, [23]:
OPD = 1000 · C · ( ρ M )
where “C” is the number of oxygen atoms per formula unit. The increase in the value of OPD from 71.06258 mol/liter to 78.40199 mol/liter of BBTTLa and BBTTYb, respectively, was associated with increases in the nb,   F ¯ , and Mwt values of rare earth oxides.
The optical absorption spectra of BBTTRE glass are shown in Figure 3. The results on the absorption edge provide important information on the transitions of the band structure of amorphous materials [27]. The absorption bands of BBTTEr glass were detected at 1512, 973, 800, 650, 550, 520, and 490 nm, which corresponded to the transitions from 4I15/2 to 4I13/2, 4I11/2, 4I9/2, 4F9/2, 4S3/2, 2H11/2, and 4F7/2, respectively.
The absorbance spectra of BBTTSm glass exhibited bands at 1620, 1551, 1485, 1379, 1230, 1085, 950, 476, 404, and 368 nm, attributed to the absorption ground state 6H5/2 to the excited states, 6H15/2, 6F1/2, 6F3/2, 6F5/2, 6F7/2, 6F9/2, 6F11/2, 4I13/2+ 4M15/2, 4P5/2, and 4P7/2, respectively. In addition, there was a strong absorption band transition of the level 2F7/2 to 2F5/2 of the Yb3+-ion-modified BBTTYb glass. The absorption coefficient, α(ν), of the fabricated glass was calculated using the absorbance spectra and the relationship shown below [28]:
α ( ν ) = 2.303 · A ·   d 1
where A represents the absorbance and d is the glass sample’s thickness in cm.
In amorphous material, optical transitions that occur at the absorption edge can be divided into two mechanisms: firstly, direct transitions are where the momentum of the electron from the valance to conduction band is preserved; secondly, there are indirect transitions where it is necessary to cooperate with the absorb/release phonon [29]. Mott and Davis [29] suggested a relationship between photon energy (hν) and absorption coefficient ( α ) to determine the indirect optical band gap, Eopt, as shown in Equation (9).
( α · h ν ) = B   ( h ν Eopt ) r
where B is a constant known as the band tailing parameter, r, which depends on the type of mechanism transition (r = 2) associated with the allowed indirect transitions [30] . A graph was plotted for ( α h ν ) 1 / 2 versus ( h ν ) to determine the indirect optical band gap, Eopt, as shown in Figure 4. The calculated values of the indirect optical band gap, E opt , can be obtained by the extrapolation of the linear range of the curve with a linear axis at (Y-axis = 0), which represents the photon energy ( h ν ) [31] and the value of Eopt for the prepared glass, as evaluated in Table 3. In the BBTTRE glass system, the formation of TeO4 caused oxygen anions to be tightly bound to the host materials; thus, the Eopt increased with a decrease in the number of NBO [31]. The value of Eopt depends on the structure of the prepared glass TBBT modified by 3RE2O3 in mol % , where RE is La, Ce, Sm, Er, or Yb. From the results presented in Table 3, the values of Eopt increase from 2.1 to 2.81 eV. The Eopt value increased as a result of more bridging oxygen being present (BO) and the decrease in the number of NBO, as confirmed in the FTIR results of the prepared glass discussed here. The value of the refractive index, n, molar polarizability, α m , molar refraction, R m , oxide ion polarizability, ( α o 2 ), and the value of optical basicity, ( Λ ), are important parameters for the fabrication of optical devices, especially fiber optic and laser material. Therefore, we determine these parameters of the studied glass by using the subsequent equations [32]:
  α m = ( 3 4 π N A ) R m
  R m = ( n 2 1 n 2 + 2 ) V m
α o 2 = [ V m 2.52 ( 1 Eopt 20 ) p α i ] q 1
Λ = 1.67 [ 1 1 α o 2 ]
where N A is Avogadro’s number, p is the cation number, and q denotes the number of ions of oxygen. The values of n, Rm,   α m , and Λ depend on the polarizability of ions; the type of RE was La, Ce, Sm, Er, or Yb, and the prepared BBTT glass was modified with 3RE2O3 in mol % . The values of n, Rm,   α m , and Λ decreased when the host glass network BBTT was modified with regard to the free ion polarizability due to internal contact and the polarizability of oxide rare earth ions decreased. High oxide ion polarizability and optical basicity are also closely related to the superior optical characteristics of tellurite glass. Herein, rare earth oxides had an order of polarizability of cation, αi, of (La2O3 = 1.32 Ă3), (Ce2O3 = 1.28 Ă3), (Sm2O3 = 1.16 Ă3), (Er2O3 = 0.89 Ă3), and (Yb2O3 = 0.86 Ă3) [33]. Furthermore, the values of optical basicity, Λ , in TeO2-based glass, i.e., Λ TeO40 = 0.99, Λ TeO4 = 1.23, and Λ TeO3 = 0.82, were estimated by Dimitrov and Komatsu [33]. This demonstrated the polarizability of the TeO3 unit to be substantially lower that of the TeO4 unit, supporting the order of Λ TeO4 > Λ TeO40 > Λ TeO3. Hence, highly distorted TeO4 tbp units with NBOs and different Te–O bond lengths should produce significant electrical polarizabilities. It is therefore of interest to clarify the fraction of TeO4 units created in the order La2O3 > Ce2O3 > Sm2O3 > Er2O3 > Yb2O3 for modifying the BBTT glass matrix.
The metallization criterion, M, for BBTTRE glass was estimated as follow:
M = 1 R m V m
The change in rare earth La2O3 → Ce2O3 → Sm2O3 → Er2O3 → Yb2O3 → La2O3 modified host matrix TBBT, causing a decrease in the width of the valance band and an increase in M, and consequently, an increase in the optical energy band gap. As shown in Table 3, the BBTTLa glass had the largest value of M and the smallest value of Eopt. In contrast, the BBTTYb glass had the smallest value of M and the highest value of Eopt.

3.2. Structural Categorization of Glass Using FTIR Spectra

The FTIR spectra of the investigated glass were measured; consequently, these were deconvoluted using Gaussian fitting into several Gaussian peaks marked as (a − x) bands, as shown Figure 5. Bands of 370–400 cm−1 have been linked to the stretching vibration mode of Bi–O–Bi linkages [34]. The bands observed at around 430, 440, 458, 425, and 430 cm−1 are attributable to La–O, Ce–O, Sm–O, Er–O, and Yb–O stretching vibrations, respectively [35,36]. Clearly visible peaks in the range of 463–480 cm−1 can be attributed to the bridging anion modes of Bi–O–Bi vibrations in distorted BiO6 octahedral units [37,38]. As a result of the combination of the corners of the TeO4, TeO3+1, and TeO3 units, the bands 494–512 cm−1 were aligned and matched with vibration connections Te–O–Te or O–Te–O [39,40,41]. The different IR peaks observed in the range of 500–800 cm−1 in the investigated glass may be related to the anti-symmetrical and symmetrical vibrations of TeO2 [41]. The peaks located in the range of 552–563 cm−1 were attributable to Bi–O bending vibration in BiO6 octahedral units [42]. The vibration of a continuous TeO4 trigonal bipyramid network was associated with bands detected in the 576–600 cm−1 range (tbp). These indicated the glass network’s more compacted connections [43]. The bands between 616 and 623 cm−1 were related to Ti–O bending vibrations [44]. The occurrence of Te–Oax vibrations in the TeO4 tetrahedral units was linked to the development of the very strong band which occurred in the region of 648 to 654 cm−1. The vibrations of BO in TeO3/TeO3+1 units were responsible for the peaks between 671 and 674 cm−1 that were seen in all glass samples. The B–O–B connections in the borate network’s mode of vibration were responsible for the bands positioned in the range of 692–695 cm−1 [42]. The bands between 712 and 721 cm−1 were attributable to the NBO stretching modes present in TeO3 units [43]. The TeO3 trigonal pyramid (tp) and (TeO3+1) polyhedral units (Teeq–O)as and (Teeq–O)S vibrational modes were responsible for the bands in the 759–772 cm−1 range [42,43]. The bending vibrations of BO4 at 600 to 800 cm−1 and the B–O bond stretching vibrations of BO4 tetrahedral units are responsible for the bands that developed between 800 and 1200 cm−1.
The IR in the regions between 1200 and 1600 cm−1 were attributable to vibrations of B–O bonds from BO3 trigonal units [39]. Absorption from 912 to 925 cm−1 could be related to the stretching vibrations of B–O bond in BO4 units from diborate groups [44,45]. The peaks observed in the region of 990 to 1001 cm−1 may be attributable to the stretching vibrations of B–O–Bi linkages [37,38]. The IR peaks observed in the ranges of 1062–1067 cm−1 may be due to the stretching vibrations of B–O bonds in BO4 units from tri-, tetra-, and penta-borate groups [46]. The two IR peaks in the region of 1118 to 1120 cm−1 and 1162 to 1168 cm−1 in our investigated samples were attributable to TiO4 tetrahedral units [47]. Other IR peaks in the spectral ranges of 1241–1248 cm−1 were associated with the presence of asymmetrical stretching vibrations of B–O bonds in BO3 triangular units from pyro-borate groups [47]. The peaks which appeared in the range of 1280–1285 cm−1 were related to the B–O asymmetrical stretching vibration of (BO3)−3 units in meta- and ortho-borate groups [45]. The next absorption bands in the spectral ranges of 1348–1349 cm−1 were attributable to symmetrical stretching vibrations of B–O bonds in triangular BO3 units from meta-, pyro-, and ortho-borate groups [47,48]. The IR peak at 1375 cm−1 in all glass samples may have been due to the asymmetrical stretching vibrations of B–O bonds in triangular BO3 units [40]. The IR peaks in the ranges of 1401–1404 cm−1 can be attributed to the asymmetrical stretching vibrations of B–O triangles with BO3, B2O, and stretching vibrations of borate triangles with NBO in various borate groups. The stretching vibrations of the B–O bonds in BO3 units obtained from different forms of borate groups were responsible for the bands seen between 1428 and 1430 cm−1 [40]. The peak at 1461 cm−1 may have been the result of three NBO oxygens in B–O–B links stretching in an anti-symmetrical manner [46]. All the IR bands in Table 4 were attributed to deconvolution FTIR spectra, as shown in Figure 5. Additionally, the ratio of TeO4 (tbp) to TeO3(tp) conversion was determined using the FTIR spectra. The ratio values of TeO4/TeO3 were 0.473, 0.504, 0.51, 0.52, and 0.53, corresponding to BBTTLa, BBTTCe, BBTTSm, BBTTEr, and BBTTYb, respectively. The ratio of transferring BO4 into BO3 was also determined from the deconvoluted FTIR spectra. Values of the ratio BO4/BO3 were 0.415, 0.5, 0.51, 0.52, and 0.53, which corresponded to BBTTLa, BBTTCe, BBTTSm, BBTTEr, and BBTTYb, respectively. The increases in the ratios of TeO4/TeO3 and BO4/BO3 show that the glass became more resistant as a result of the formation of additional bridging oxygens (BOs). Thus, the formation of BO sites with increased atomic numbers of rare earth elements resulted in a strictly dense glass structure that confirmed the increment in the ρ and decrement of Vm values for BBTTRE in the order similar to that La2O3 → Ce2O3 → Sm2O3 → Er2O3 → Yb2O3 → La2O3 glass samples, which was consistent with the parameter changes.

3.3. Attenuation Parameters

The total mass attenuation coefficient, μ m = ln I o I ρ d , and the linear attenuation coefficient, μ = ln I o I d , are calculated using the ratio between the intensities of the measured incident, I0, and the transmission radiation, I; d is the thickness of the shielding material. HVL = 0.693 μ , and the MFP parameter was calculated as MFP = 1 μ . Figure 6 shows the calculated linear attenuation coefficient (LAC) of the prepared glass for different energies (59.5, 622, 1170, and 1330 keV) compared with commercially available glass shielding materials, namely, RS360 and RS 520. For instance, the LAC value for BBTTYb glass exhibited the best shielding properties at the energies of 622, 1170, and 1330 keV at 59.5 keV, as compared with RS 360 and RS520 glass. Furthermore, the prepared glass (BBTTER) with a composition of 25B2O3 –20Bi2O3–45TeO2–7TiO2 (BBTT) modified by 3RE2O3 in mol % , where RE was La, Ce, Sm, Er, or Yb, was better than that reported in other glass systems modified with rare earth elements, such as 39B2O3–30PbO–20MO–10Bi2O3–1Eu2O3 (where M is K, Na, Ca, Sr, or Ba) [49], B2O3–CaO–TeO2–ZnO–ZnF2–Sm2O3 [50], and B2O3–SrCO3–Nb2O3–BaCO3–Dy2O3 [51]. Table 5 and Table 6 present the measured mass attenuation coefficients (MACs) of the prepared samples in comparison with the calculated theoretical values using the MIKE and WinXcom software. The HVL parameter signifies the material thickness that reduces the intensity of radiation by half. Herein, the values for HVL and MFP of the prepared glass were lower than that reported for commercial materials, such as window glass, serpentine, concrete, SCHOOT glass RS253, hematite serpentine, Ilmenite, and SCHOOT glass RS323 [52,53,54]. Figure 7 shows the measured values of MAC, LAC, HVL, and MFP for the BBTTEr glass at 59.5, 622, 1170, and 1330 keV, compared with the theoretical values calculated using MIKE software. The results showed good agreement between the measured mass attenuation coefficients and calculated values using MIKE software. Hence, the experimental attenuation results for the investigated prepared glass showed superior radiation shielding performance. Finally, we can estimate that the shielding parameters increased with the increasing ratios of TeO4/TeO3 and BO4/BO3 with bridging oxygens (BOs) of oxide glass, representing a candidate for the fabrication of superior shielding material.

4. Conclusions

Incorporating the rare earth ions La+3, Ce+3, Sm+3, Er+3, and Yb+3 as glass matrix modifiers, resulting in 25B2O3–20Bi2O3–45TeO2–7TiO2, it was found that the density of the studied glass increased from 5.67 to 6.31 gm.cm−3, nb increased from 6.5 to 7.14 × 1022 m−3, and the OPD increased from 71.1 to 78.4 mol/L−1 with an increased atomic number of incorporated rare earth ions. This is due to the increased amount of bridging oxygen (BO) and decreased number of NBOs in the prepared glass, which also led to the increased Eopt, from 1.71 to 2.8 eV, when La2O3 was replaced by Yb2O3. In addition, the molar polarizability, αm, decreased from 9.35 to 7.84 (Ă3), Λ decreased from 1.15 to 1.03, and the refractive index decreased from 2.69 to 2.45: this was due to the good agreement with the replacement of modifiers of La2O3 = 1.32 Ă3, Ce2O3 = 1.28 Ă3, Sm2O3 = 1.16 Ă3, Er2O3 = 0.89 Ă3, Yb2O3 = 0.86 Ă3, Λ TeO40 = 0.99, Λ TeO4 = 1.23, and Λ TeO3 = 0.82, in the order of La+3 → Ce+3 → Sm+3 → Er+3 → Yb+3. The high refractive index, electronic polarizability, and optical basicity of the prepared glass containing La2O3 led to the achievement of significant third-order optical susceptibility. This glass may be used to produce high-quality optical nonlinear devices. The FTIR spectra confirmed the existence of TeO4, TeO3, BO4, BO3, BiO6, and TiO4 in the glass matrix. The glass containing Yb3+ ions had a high value of the TeO4 phase with BO. When compared with prepared glass, Yb3+-ion-containing glass exhibited higher MAC values and lower HVL values, which was directly related to its high shielding characteristics. This glass is an excellent choice for use in low-energy diagnostic applications as a transparent shielding material.

Author Contributions

N.E.: conceptualization, methodology, investigation, writing—original draft, and writing—review and editing; S.M.: conceptualization, methodology, formal analysis, investigation, and writing—original draft; M.E.-S.: methodology, writing—review and editing, and visualization; H.I.: formal analysis, review and editing, and visualization; B.B.-G.: formal analysis, visualization, funding acquisition, and writing—review and editing; M.S.A.: formal analysis, investigation, writing—original draft, writing—review and editing, and visualization; K.I.H.: methodology, formal analysis, writing—review and editing, and visualization; M.R.: methodology, formal analysis, writing—review and editing, and visualization; E.S.Y.: conceptualization, methodology, investigation, funding acquisition, writing—review and editing, and visualization. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Deputyship for Research & Innovation, Ministry of Education, Saudi Arabia, grant number IFP-KKU-2020/7.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Data Availability Statement

Not applicable.

Acknowledgments

The authors extend their appreciation to the Deputyship for Research & Innovation, Ministry of Education, Saudi Arabia, for funding this research through project number IFP-KKU-2020/7.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xinyu, Z.; Xiaoli, W.; Hai, L.; Zhiqiang, W. Electronic polarizability and optical basicity of lanthanide oxides. Phys. B Condens. Matter 2017, 392, 132–136. [Google Scholar]
  2. Fudzi, F.M.; Kamari, H.M.; Latif, A.A.; Noorazlan, A.M. Linear Optical Properties of Zinc Borotellurite Glass Doped with Lanthanum Oxide Nanoparticles for Optoelectronic and Photonic Application. J. Nanomater. 2017, 2017, 4150802. [Google Scholar] [CrossRef]
  3. Selvi, S.; Marimuthu, K.; Muralidharan, G. Structural and luminescence behavior of Sm3+ ions doped lead boro-telluro-phosphate glasses. J. Lumin. 2015, 159, 207–218. [Google Scholar] [CrossRef]
  4. Hussein, K.I.; Alqahtani, M.S.; Alzahrani, K.J.; Alqahtani, F.F.; Zahran, H.Y.; Alshehri, A.M.; Yahia, I.S.; Reben, M.; Yousef, E.S. The Effect of ZnO, MgO, TiO2, and Na2O Modifiers on the Physical, Optical, and Radiation Shielding Properties of a TeTaNb Glass System. Materials 2022, 15, 1844. [Google Scholar] [CrossRef]
  5. Selvaraju, K.; Marimuthu, K. Structural and spectroscopic studies on concentration dependent Sm3+ doped boro-tellurite glasses. J. Alloys Compd. 2013, 553, 273–281. [Google Scholar] [CrossRef]
  6. Golis, E.; Yousef, E.S.; Reben, M.; Kotynia, K.; Filipecki, J. Measurements of defect structures by positron annihilation lifetime spectroscopy of the tellurite glass TeO2–P2O5–ZnO–LiNbO3 doped with ions of rare earth elements: Er3+, Nd3+ and Gd3+. Solid State Sci. 2015, 50, 81–84. [Google Scholar] [CrossRef]
  7. Hussain, N.S.; Hungerford, G.; El-Mallawany, R.; Gomes, M.J.M.; Lopes, M.A.; Ali, N.; Santos, J.D.; Buddhudu, S. Absorption and Emission Analysis of RE3+(Sm3+ and Dy3+): Lithium Boro Tellurite Glasses. J. Nanosci. Nanotechnol. 2009, 9, 3672–3677. [Google Scholar] [CrossRef]
  8. Elkhoshkhany, N.; Abbas, R.; El-Mallawany, R.; Hathot, S.F. Optical properties and crystallization of bismuth boro-tellurite glasses. J. Non-Cryst. Solids 2017, 476, 15–24. [Google Scholar] [CrossRef]
  9. Boda, R.; Shareefuddin, M.D.; Chary, M.N.; Sayanna, R. FTIR and optical properties of europium doped lithium zinc bismuth borate glasses. Mater. Today 2016, 3, 1914–1922. [Google Scholar] [CrossRef]
  10. Gupta, N.; Kaur, A.; Khanna, A.; Gonzàlez, F.; Pesquera, C.; Iordanova, R.; Chen, B. Structure-property correlations in TiO2-Bi2O3-B2O3-TeO2 glasses. J. Non-Cryst. Solids 2017, 470, 168–177. [Google Scholar] [CrossRef]
  11. Sapian, I.N.; Yusof, M.I.M.; Yahya, A.K. Elastic and Structural Properties of (95-x) TeO2-5La2O3-xTiO2 Lanthanum Tellurite Glass System. Chalcogenide Lett. 2014, 11, 471–484. [Google Scholar]
  12. Maheshvaran, K.; Linganna, K.; Marimuthu, K. Composition dependent structural and optical properties of Sm3+ doped boro-tellurite glasses. J. Lumin. 2011, 131, 2746–2753. [Google Scholar] [CrossRef]
  13. Jamalaiah, B.; Kumar, M.V.; Gopal, K.R. Fluorescence properties and energy transfer mechanism of Sm3+ ion in lead telluroborate glasses. Opt. Mater. 2011, 33, 1643–1647. [Google Scholar] [CrossRef]
  14. Singh, G.P.; Parvinder, K.; Simranpreet, K.; Singh, D.P. Investigation of structural, physical and optical properties of CeO2–Bi2O3–B2O3 glasses. Phys. B Condens. Matter 2012, 407, 4168–4172. [Google Scholar] [CrossRef]
  15. Pisarska, J.; Agnieszka, K.; Marta, S.; Agata, G.; Ewa, P.; Pisarski, W.A. Spectroscopy and energy transfer in Tb3+/Sm3+ co-doped lead borate glasses. J. Lumin. 2018, 195, 87–95. [Google Scholar] [CrossRef]
  16. Hegazy, H.H.; Almojadah, S.; Reben, M.; Yousef, E.S. Absorption spectra and Raman gain coefficient in near-IR region of Er3+ ions doped TeO2–Nb2O5–Bi2O3–ZnO glasses. Opt. Laser Technol. 2015, 74, 138–144. [Google Scholar]
  17. Sayyed, M.I.; Laariedh, F.; Kumr, A.; Al-Buriahi, M.S. Experimental studies on the gamma photons-shielding competence of TeO2–PbO–BaO–Na2O–B2O3 glasses. Appl. Phys. A 2020, 126, 4. [Google Scholar] [CrossRef]
  18. Kamislioglu, M.; Guclu, E.A.; Tekin, H.O. Comparative evaluation of nuclear radiation shielding properties of x TeO2+ (100 − x) Li2O glass system. Appl. Phys. A 2020, 126, 1–16. [Google Scholar] [CrossRef]
  19. Rammah, Y.S.; Kavaz, E.; Perişanoğlu, U.; Kilic, G.; El-Agawany, F.I.; Tekin, H.O. Charged particles and gamma-ray shielding features of oxyfluoride semiconducting glasses: TeO2-Ta2O5-ZnO/ZnF2. Ceram. Int. 2020, 46, 25035–25042. [Google Scholar] [CrossRef]
  20. Lakshminarayana, G.; Kaky, K.M.; Baki, S.O.; Ye, S.; Lira, A.; Kityk, I.V.; Mahdi, M.A. Concentration dependent structural, thermal, and optical features of Pr3+-doped multicomponent tellurite glasses. J. Alloys Compd. 2016, 686, 769–784. [Google Scholar] [CrossRef]
  21. Gerward, L.; Guilbert, N.; Jensen, K.; Levring, H. WinXCom—A program for calculating X-ray attenuation coefficients. Radiat. Phys. Chem. 2004, 71, 653–654. [Google Scholar] [CrossRef]
  22. Hussein, K.I.; Alqahtani, M.S.; Algarni, H.; Zahran, H.; Yaha, I.S.; Grelowska, I.; Reben, M.; Yousef, E.S. MIKE: A new computational tool for investigating radiation, optical and physical properties of prototyped shielding materials. J. Instrum. 2021, 16, T07004. [Google Scholar] [CrossRef]
  23. Swapna; Upender, G.; Prasad, M. Vibrational, Optical and EPR studies of TeO2-Nb2O5-Al2O3-V2O5 glass system doped with vanadium. Opt. Int. J. Light Electron. Opt. 2016, 127, 10716–10726. [Google Scholar] [CrossRef]
  24. Yousef, E.S.; Elokr, M.; AbouDeif, Y. Optical, elastic properties and DTA of TNZP host tellurite glasses doped with Er3+ ions. J. Mol. Struct. 2016, 1108, 257–262. [Google Scholar] [CrossRef]
  25. El-Moneim, A.A.; El-Latif, L.A. Theoretically calculated elastic moduli of tellurite, phosphate and silicate glasses. Phys. Chem. Glasses 2003, 44, 446–453. [Google Scholar]
  26. Sidkey, M.A.; El-Moneim, A.A. Ultrasonic studies on ternary TeO2–V2O5–Sm2O3 glasses. Mater. Chem. Phys. 1999, 61, 103–109. [Google Scholar] [CrossRef]
  27. Laila, S.; Suraya, A.K.; Yahya, A.K. Effect of glass network modification on elastic and structural properties of mixed electronicionic 35V2O5-(65-x) TeO2-(x) Li2O glass system. Chalcogenide Lett. 2014, 11, 91–104. [Google Scholar]
  28. Luo, M.; Sha, X.; Chen, B.; Zhang, X.; Yu, H.; Li, X.; Zhang, J.; Xu, S.; Cao, Y.; Wang, Y.; et al. Optical transition properties, internal quantum efficiencies, and temperature sensing of Er3+ doped BaGd2O4 phosphor with low maximum phonon energy. J. Am. Ceram. Soc. 2022, 105, 3353–3363. [Google Scholar] [CrossRef]
  29. Davis, E.A.; Mott, N.F. Conduction in non-crystalline systems V. Conductivity, optical absorption and photoconductivity in amorphous semiconductors. Philos. Mag. 1970, 22, 179. [Google Scholar] [CrossRef]
  30. Elkhoshkhany, N.; Marzouk, S.Y.; Shahin, S. Synthesis and optical properties of new fluoro-tellurite glass within (TeO2-ZnO-LiF-Nb2O5-NaF) system. J. Non-Cryst. Solids 2017, 472, 39–45. [Google Scholar] [CrossRef]
  31. Charfi, B.; Damak, K.; Alqahtani, M.S.; Hussein, K.I.; Alshehri, A.M.; Elkhoshkhany, N.; Assiri, A.L.; Alshehri, K.F.; Reben, M.; Yousef, E.S. Luminescence and Gamma Spectroscopy of Phosphate Glass Doped with Nd3+/Yb3+ and Their Multifunctional Applications. Photonics 2022, 9, 406. [Google Scholar] [CrossRef]
  32. Ahmadi, F.; Hussin, R.; Ghoshal, S. Spectroscopic attributes of Sm3+ doped magnesium zinc sulfophosphate glass: Effects of silver nanoparticles inclusion. Opt. Mater. 2017, 73, 268–276. [Google Scholar] [CrossRef]
  33. Komatsu, T.; Dimitrov, V. Features of electronic polarizability and approach to unique properties in tellurite glasses. Int. J. Appl. Glas. Sci. 2019, 11, 253–271. [Google Scholar] [CrossRef]
  34. Arunkumar, S.; Marimuthu, K. Concentration effect of Sm3+ ions in B2O3–PbO–PbF2–Bi2O3–ZnO glasses–structural and luminescence investigations. J. Alloy Compd. 2013, 565, 104–114. [Google Scholar] [CrossRef]
  35. Hager, I.Z.; El-Mallawany, R. Preparation and structural studies in the (70 − x) TeO2–20WO3–10Li2O–xLn2O3 glasses. J. Mater. Sci. 2010, 45, 897. [Google Scholar] [CrossRef]
  36. Berwal, N.; Kundu, R.S.; Nanda, K.; Punia, R.; Kishore, N. Physical, structural and optical characterizations of borate modified bismuth–silicate–tellurite glasses. J. Mol. Struct. 2015, 1097, 37–44. [Google Scholar] [CrossRef]
  37. Berwal, N.; Dhankhar, S.; Sharma, P.; Kundu, R.; Punia, R.; Kishore, N. Physical, structural and optical characterization of silicate modified bismuth-borate-tellurite glasses. J. Mol. Struct. 2017, 1127, 636–644. [Google Scholar] [CrossRef]
  38. Rada, S.; Culea, M.; Culea, E. Structure of TeO2·B2O3 glasses inferred from infrared spectroscopy and DFT calculations. J. Non-Cryst. Solids 2008, 354, 5491–5495. [Google Scholar] [CrossRef]
  39. Kundu, R.; Dhankhar, S.; Punia, R.; Nanda, K.; Kishore, N. Bismuth modified physical, structural and optical properties of mid-IR transparent zinc boro-tellurite glasses. J. Alloys Compd. 2014, 587, 66–73. [Google Scholar] [CrossRef]
  40. Kaur, A.; Khanna, A.; Aleksandrov, L.I. Structural, thermal, optical and photo-luminescent properties of barium tellurite glasses doped with rare-earth ions. J. Non-Cryst. Solids 2017, 476, 67–74. [Google Scholar] [CrossRef]
  41. Pascuta, P.; Pop, L.; Rada, S.; Bosca, M.; Culea, E. The local structure of bismuth borate glasses doped with europium ions evidenced by FT-IR spectroscopy. J. Mater. Sci. Mater. Electron. 2007, 19, 424–428. [Google Scholar] [CrossRef]
  42. Hung, W.-C.; Fu, S.-H.; Tseng, J.-J.; Chu, H.; Ko, T.-H. Study on photocatalytic degradation of gaseous dichloromethane using pure and iron ion-doped TiO2 prepared by the sol–gel method. Chemosphere 2007, 66, 2142–2151. [Google Scholar] [CrossRef]
  43. Abdelghany, A.M. Combined DFT, deconvolution analysis for structural investigation of copper–doped lead borate glasses. Open Spectrosc. J. 2012, 6, 9–14. [Google Scholar] [CrossRef]
  44. Elkhoshkhany, N.; Khatab, M.; Kabary, M.A. Thermal, FTIR and UV spectral studies on tellurite glasses doped with cerium oxide. Ceram. Int. 2018, 44, 2789–2796. [Google Scholar] [CrossRef]
  45. Elkhoshkhany, N.; El-Mallawany, R.; Syala, E. Mechanical and thermal properties of TeO2–Bi2O3–V2O5–Na2O–TiO2 glass system. Ceram. Int. 2016, 42, 19218–19224. [Google Scholar] [CrossRef]
  46. Kaur, A.; Khanna, A.; González, F.; Pesquera, C.; Chen, B. Structural, optical, dielectric and thermal properties of molybdenum tellurite and borotellurite glasses. J. Non-Cryst. Solids 2016, 444, 1–10. [Google Scholar] [CrossRef]
  47. Eevon, C.; Halimah, M.; Zakaria, A.; Azurahanim, C.; Azlan, M.; Faznny, M. Linear and nonlinear optical properties of Gd3+ doped zinc borotellurite glasses for all-optical switching applications. Results Phys. 2016, 6, 761–766. [Google Scholar] [CrossRef] [Green Version]
  48. Wang, M.; Cheng, J.; Li, M. Effect of rare earths on viscosity and thermal expansion of soda-lime-silicate glass. J. Rare Earths 2010, 28, 308–311. [Google Scholar] [CrossRef]
  49. Divina, R.; Naseer, K.A.; Marimuthu, K.; Alajerami, Y.S.M.; Al-Buriahi, M.S. Effect of different modifier oxides on the synthesis, structural, optical, and gamma/beta shielding properties of bismuth lead borate glasses doped with europium. J. Mater. Sci. Mater. Electron. 2020, 31, 21486–21501. [Google Scholar] [CrossRef]
  50. Chen, Q.; Naseer, K.; Marimuthu, K.; Kumar, P.S.; Miao, B.; Mahmoud, K.; Sayyed, M. Influence of modifier oxide on the structural and radiation shielding features of Sm3+-doped calcium telluro-fluoroborate glass systems. J. Aust. Ceram. Soc. 2020, 57, 275–286. [Google Scholar] [CrossRef]
  51. Sathiyapriya, G.; Naseer, K.A.; Marimuthu, K.; Kavaz, E.; Alalawi, A.; Al-Buriahi, M.S. Structural, optical and nuclear radiation shielding properties of strontium barium borate glasses doped with dysprosium and niobium. J. Mater. Sci. Mater. Electron. 2021, 32, 8570–8592. [Google Scholar] [CrossRef]
  52. Chanthima, N.; Kaewkhao, J. Investigation on radiation shielding parameters of bismuth borosilicate glass from 1 keV to 100 GeV. Ann. Nucl. Energy 2013, 55, 23–28. [Google Scholar] [CrossRef]
  53. Chanthima, N.; Kaewkhao, J.; Limkitjaroenporn, P.; Tuscharoen, S.; Kothan, S.; Tungjai, M.; Kaewjaeng, S.; Sarachai, S.; Limsuwan, P. Development of BaO–ZnO–B2O3 glasses as a radiation shielding material. Radiat. Phys. Chem. 2017, 137, 72–77. [Google Scholar] [CrossRef]
  54. Cheewasukhanont, W.; Limkitjaroenporn, P.; Sayyed, M.I.; Kothan, S.; Kim, H.J.; Kaewkhao, J. High density of tungsten gadolinium borate glasses for radiation shielding material: Effect of WO3 concentration. Radiat. Phys. Chem. 2022, 192, 109926. [Google Scholar] [CrossRef]
Figure 1. Experimental setup used for measuring the shielding parameters of the prepared samples.
Figure 1. Experimental setup used for measuring the shielding parameters of the prepared samples.
Materials 15 05393 g001
Figure 2. XRD profile of prepared glass.
Figure 2. XRD profile of prepared glass.
Materials 15 05393 g002
Figure 3. Absorbance spectra of prepared glass.
Figure 3. Absorbance spectra of prepared glass.
Materials 15 05393 g003
Figure 4. Relationship between (αhν)1/2 and hν of the prepared glass.
Figure 4. Relationship between (αhν)1/2 and hν of the prepared glass.
Materials 15 05393 g004
Figure 5. Deconvoluted FTIR spectra of the prepared glass.
Figure 5. Deconvoluted FTIR spectra of the prepared glass.
Materials 15 05393 g005aMaterials 15 05393 g005b
Figure 6. The calculated LAC for the glass samples compared with standard glass materials at energies: (A) 59.5; (B) 622; (C) 1170; and (D) 1330 keV.
Figure 6. The calculated LAC for the glass samples compared with standard glass materials at energies: (A) 59.5; (B) 622; (C) 1170; and (D) 1330 keV.
Materials 15 05393 g006
Figure 7. The measured and theoretical shielding parameters for the BBTTEr glass system at different photon energies (59.5, 622, 1170, and 1330 keV): (A) MAC; (B) LAC; (C) HVL; and (D) MFP.
Figure 7. The measured and theoretical shielding parameters for the BBTTEr glass system at different photon energies (59.5, 622, 1170, and 1330 keV): (A) MAC; (B) LAC; (C) HVL; and (D) MFP.
Materials 15 05393 g007
Table 1. Compositions and codes of glass systems (45TeO2–25B2O3–20Bi2O3–7TiO2–3RE2O3) in mol%.
Table 1. Compositions and codes of glass systems (45TeO2–25B2O3–20Bi2O3–7TiO2–3RE2O3) in mol%.
Sample NameGlass Composition (mol%)Sample Color
TeO2B2O3Bi2O3TiO2La2O3Ce2O3Sm2O3Er2O3Yb2O3
BBTTLa45252073 Materials 15 05393 i001
BBTTCe45252073 Materials 15 05393 i002
BBTTSm45252073 Materials 15 05393 i003
BBTTEr45252073 Materials 15 05393 i004
BBTTYb45252073 Materials 15 05393 i005
Table 2. Density ( ρ ), molar volume (Vm), oxygen molar volume (VO), number of bonds (nb), average stretching force constant ( F ¯ ), and oxygen packing density (OPD) of prepared glass samples.
Table 2. Density ( ρ ), molar volume (Vm), oxygen molar volume (VO), number of bonds (nb), average stretching force constant ( F ¯ ), and oxygen packing density (OPD) of prepared glass samples.
Sample Name ρ
(g/cm3) ± 0.001
Vm
(cm3/mol) ± 0.0056
Oxygen Molar
Volume, Vo (cm3/mol) ± 0.2980
nb × 1022 (m−3) F ¯ (Nm−1) OPD
(mol/L) ± 0.078
BBTTLa5.673514.16.5301.2171.1
BBTTCe6.08332.513.17.05300.276.2
BBTTSm6.1732.1312.957.08301.7977.1
BBTTEr6.2132.1112.947.09301.4877.2
BBTTYb6.3131.6312.77.14303.0478.4
Table 3. Optical energy gap, Eopt; refractive index, n; molar polarizability, αm; molar refraction, Rm; oxide ion polarizability, α o 2 ; metallization, M; and optical basicity, Λ, of the prepared glass.
Table 3. Optical energy gap, Eopt; refractive index, n; molar polarizability, αm; molar refraction, Rm; oxide ion polarizability, α o 2 ; metallization, M; and optical basicity, Λ, of the prepared glass.
Sample NameEopt (eV)
±0.01
n
±0.0001
Rm (cm3)
±0.0466
αm,3)
±0.0182
α o 2 3)
± 0.0065
M
± 0.0008
Λ
± 0.0025
BBTTLa1.712.6923.69.353.220.3241.15
BBTTCe2.472.6121.58.522.880.3391.09
BBTTSm2.572.4920.48.12.710.3651.05
BBTTEr2.682.4620.28.0082.680.3711.04
BBTTYb2.82.4519.777.842.610.3751.03
Table 4. The location of FTIR absorption bands corresponding to the structural bonds of the prepared glass samples.
Table 4. The location of FTIR absorption bands corresponding to the structural bonds of the prepared glass samples.
SymbolIR Bands Wavenumber (cm−1)Assignments
a370–400Stretching mode of vibration of Bi–O–Bi linkages
b430
440
458
425
430
Stretching vibration of La–O
Stretching vibration of Ce–O
Stretching vibration of Sm–O
Stretching vibration of Er–O
Stretching vibration of Yb–O
c463–480Bi–O–Bi vibration in distorted BiO6 octahedral units
b494–512Symmetrical stretching or bending vibrations of Te–O–Te or O–Te–O linkages
e552–563Bending vibration of Bi–O in BiO6 units
f576–600Vibration of the continuous network consisting of TeO4 tbp
g616–623Ti–O bending vibration
h648–654Symmetrical stretching vibration of Te–Oax in TeO4 tetrahedral units
i671–674Stretching vibrations tellurium with BO of TeO3/TeO3+1 units
j692–695Bending vibrations of B–O–B linkages in the borate network
k712–721Stretching modes of NBO found on TeO3 and TeO3+1 units
l759–772Symmetrical and asymmetrical vibration of (Teeq–O) in TeO3+1 polyhedra or trigonal pyramid TeO3 (tp) units
m912–925Stretching vibrations of B–O bond in BO4 units from diborate groups
n990–1001Stretching vibrations of B–O–Bi linkages
o
p
(1023–1028),
(1062–1067)
Stretching vibrations of B–O bond in BO4 units from tri-, tetra- and penta-borate groups
q
r
(1118–1120),
(1162–1168)
TiO4
s
t
(1247–1248),
(1280–1285)
Asymmetric stretching vibrations of B–O bond in BO3 triangular units from meta-, pyro-, and ortho-borate groups
u
v
(1317–1321),
(1348–1349)
Symmetrical stretching vibrations of B–O bond in BO3 triangular units from meta-, pyro-, and ortho-borate groups
w1375Asymmetrical stretching vibrations of B–O bond in BO3 triangular units
x1401–1404Asymmetrical stretching vibrations of B–O triangle with BO3, B2O and stretching vibration of borate triangle with (NBO) in various borate groups
y1428–1430Stretching vibration of B–O bond in BO3 units from varied types of borate groups
z1461Anti-symmetric stretching vibrations with 3 NBO of B–O–B linkages
Table 5. The measured mass attenuation coefficients of BBTTLa, BBTTCe, and BBTTSm samples in comparison with the values calculated using MIKE software and theoretical estimated values (WinXcom).
Table 5. The measured mass attenuation coefficients of BBTTLa, BBTTCe, and BBTTSm samples in comparison with the values calculated using MIKE software and theoretical estimated values (WinXcom).
Energy (keV)Mass Attenuation Coefficient
BBTTLa BBTTCe BBTTSm
ExpWinXComMIKEExpWinXComMIKEExpWinXComMIKE
59.54.5284.7174.79244.60704.7414.81514.6544.8314.9056
6620.0840.0900.09120.08390.0910.09130.08410.09070.0914
11700.0520.0570.05800.05210.0580.05800.05220.05790.0580
13300.0460.0530.05340.04630.0530.05340.04640.05340.0535
Table 6. The measured mass attenuation coefficients of BBTTEr and BBTTYb samples in comparison with the values calculated using MIKE software and theoretical estimated values (WinXcom).
Table 6. The measured mass attenuation coefficients of BBTTEr and BBTTYb samples in comparison with the values calculated using MIKE software and theoretical estimated values (WinXcom).
Energy (keV)Mass Attenuation Coefficient
BBTTEr BBTTYb
ExpWinXComMIKEExpWinXComMIKE
59.54.82454.9885.06824.30904.47104.5282
6620.08440.0910.09170.08450.09120.0918
11700.05240.0580.05810.05260.05800.0581
13300.04670.0530.05350.04680.05350.0535
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Elkhoshkhany, N.; Marzouk, S.; El-Sherbiny, M.; Ibrahim, H.; Burtan-Gwizdala, B.; Alqahtani, M.S.; Hussien, K.I.; Reben, M.; Yousef, E.S. Investigation of Structural, Physical, and Attenuation Parameters of Glass: TeO2-Bi2O3-B2O3-TiO2-RE2O3 (RE: La, Ce, Sm, Er, and Yb), and Applications Thereof. Materials 2022, 15, 5393. https://doi.org/10.3390/ma15155393

AMA Style

Elkhoshkhany N, Marzouk S, El-Sherbiny M, Ibrahim H, Burtan-Gwizdala B, Alqahtani MS, Hussien KI, Reben M, Yousef ES. Investigation of Structural, Physical, and Attenuation Parameters of Glass: TeO2-Bi2O3-B2O3-TiO2-RE2O3 (RE: La, Ce, Sm, Er, and Yb), and Applications Thereof. Materials. 2022; 15(15):5393. https://doi.org/10.3390/ma15155393

Chicago/Turabian Style

Elkhoshkhany, Nehal, Samir Marzouk, Mohammed El-Sherbiny, Heba Ibrahim, Bozena Burtan-Gwizdala, Mohammed S. Alqahtani, Khalid I. Hussien, Manuela Reben, and El Sayed Yousef. 2022. "Investigation of Structural, Physical, and Attenuation Parameters of Glass: TeO2-Bi2O3-B2O3-TiO2-RE2O3 (RE: La, Ce, Sm, Er, and Yb), and Applications Thereof" Materials 15, no. 15: 5393. https://doi.org/10.3390/ma15155393

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