Next Article in Journal
Anisotropy of the Electric Field Gradient in Two-Dimensional α-MoO3 Investigated by 57Mn(57Fe) Emission Mössbauer Spectroscopy
Previous Article in Journal
The Impact of Full-Scale Substitution of Ca2+ with Ni2+ Ions on Brushite’s Crystal Structure and Phase Composition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Investigation of New Phosphate–Titanite Glasses According to Optical, Physical, and Shielding Properties

by
Khalid I. Hussein
1,2,*,
Mohammed S. Alqahtani
1,3,
Khloud J. Alzahrani
1,
Heba Y. Zahran
4,5,
Ali M. Alshehri
4,
Ibrahim S. Yahia
4,5,6,
Manuela Reben
7 and
El Sayed Yousef
4,5
1
Department of Radiological Sciences, College of Applied Medical Sciences, King Khalid University, Abha 61421, Saudi Arabia
2
Department of Medical Physics and Instrumentation, National Cancer Institute, University of Gezira, Wad Medani 2667, Sudan
3
BioImaging Unit, Space Research Centre, Department of Physics and Astronomy, University of Leicester, Leicester LE1 7RH, UK
4
Physics Department, Faculty of Science, King Khalid University, Abha 61413, Saudi Arabia
5
Research Center for Advanced Materials Science (RCAMS), King Khalid University, Abha 61413, Saudi Arabia
6
Nanoscience Laboratory for Environmental and Biomedical Applications (NLEBA), Semiconductor Laboratory, Department of Physics, Faculty of Education, Ain Shams University, Roxy, Cairo 11757, Egypt
7
Faculty of Materials Science and Ceramics, AGH–University of Science and Technology, Al. Mickiewicza 30, 30-059 Cracow, Poland
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(7), 941; https://doi.org/10.3390/cryst12070941
Submission received: 10 June 2022 / Revised: 1 July 2022 / Accepted: 1 July 2022 / Published: 4 July 2022

Abstract

:
The melt-quenching approach was used to prepare phosphate–titanite glasses with the composition P2O5-Na2O-CaO-8KF-CaCl2-xTiO2 (where x = 2, 4, and 6) in a mol %. The optical, physical, and shielding properties, such as the mass attenuation coefficient (MAC), half-value layer (HVL), effective electron density (Neff), and effective atomic number (Zeff), of the glasses were investigated at energies ranging between 15 and 200 keV. The shielding parameters were investigated using recently developed software (MIKE). The optical properties were examined using devices such as UV-Vis-NIR spectroscopy over wavelengths ranging between 190 and 2500 nm. The reported results showed that increasing the concentration of TiO2 led to an increase in the density from 2.657 to 2.682 g/cm3 and an increase in the OPD from 66.055 to 67.262 mol/L, while the molar volume (VM) and oxygen molar volume (VO) decreased from 39.21 to 39.101 cm3/mol and from 15.139 to 14.867 cm3/mol, respectively. The energy gap was found to decrease from 3.403 to 3.279 eV when the TiO2 concentration increased. Furthermore, as the surface plasmon resonance of TiO2 increases, so does its third-order susceptibility, non-linear refractive indices, linear attenuation, and mass attenuation. The shielding performance evaluation indicates that the most suitable energy range is between 15 and 50 keV. Based on the results, the PCKNT3 glass sample exhibits the highest attenuation performance of all of the samples tested.

1. Introduction

Metal oxides such as sodium oxide (Na2O), calcium oxide (CaO), magnesium oxide (MgO), phosphorus pentoxide (P2O5), thallium oxide (TiO2), and silicon dioxide (SiO2) have been discovered to be necessary components for glass-ceramics and glasses, as well as bioactive glasses [1,2]. Their physical and optical performance factors, such as their refractive-index- or energy-gap-based oxide ion polarizability, cation polarizability, and optical basicity, make them suitable candidates for different optical applications [2]. A considerable contribution to the existing literature on advanced medical applications is made by theoretical investigations into bioactive glasses and their radiation attenuation capacities, which are difficult or impossible to accomplish experimentally or clinically. By delivering biomaterial-driven regenerative medicine, bioactive glasses enable the discovery of new techniques in medicine. This has cleared the path for bioactive glasses to be used as implant materials in medical repair and replacement procedures. Bioactive materials have been employed in a variety of applications, from dental to soft tissue healing. The first bioactive glass-ceramic created with the chemical form of 45 wt% SiO2, 24.5 wt% CaO, 24.5 wt% Na2O, and 6.0 wt% P2O5 [3] showed good biocompatibility and bone-bonding ability. This sample is known as 45S5 bio-glass. During therapeutic and diagnostic operations, bioactive materials can interact with ionizing radiation. Because bioactive glasses are used inside the human body for the aforementioned purposes, they may be subjected to harmful radiation from X-ray and gamma-ray equipment, which are commonly employed in hospitals to detect and cure disorders in the human body. Titanium dioxide has been widely investigated as an implant material in dental and orthopedic implants. Results show that it has good mechanical and biocompatibility compared to other existing materials such as stainless steel and cobalt–chrome alloys [4,5,6,7]. A complete understanding of photon interactions with various phosphate–titanite glass compositions, including their X-ray and gamma photon interactions, is therefore critical and must be achieved. Several studies have been conducted to evaluate the effectiveness of phosphate–titanite materials in shielding applications as well as their mechanical and optical properties [8,9,10,11,12,13]. Al-Harbi et al. (2020) studied the shielding characteristics of a glass system comprising phosphate glasses with the composition SiO2–Na2O–P2O5–CaO–B2O3 using MCNP5 and Phys-X software. Good radiation protection was reported at the low energies of 15 to 40 keV. Alalawi et al. (2020) studied the shielding effectiveness of phosphate glasses as bioactive glass systems with the structureP2O5-Na2O-CaO-K2O-MgO at nuclear medicine energies. Their results indicate that the addition of K2O has a significant effect on shielding properties. Another factor that also should be considered is the radiation-induced point defects in oxides. Highly ionic MgO, partly covalent corundum (Al2O3), ferroelectric KNbO3, silica-based optical fibers, fiber-based devices, and optical fiber sensors were reported [14,15]. This material, if irradiated by energetic particles, leads to the displacement of an atom into an interstitial position, leaving a vacancy behind. Most radiation shielding materials are made of lead, which has a high atomic number and high shielding efficiency. On the other hand, the lead is considered to be a toxic, heavy, and electron-contaminated material that increases the dose that staff and patients are exposed to, especially when dealing when equipment for patient contact shielding, such as protective lead aprons and thyroid shields placed in contact with a patient’s skin to protect sensitive organs such as the chest, thyroid, and lenses of eyes. In this paper, we presented a new phosphate–titanite glass system with the composition 45P2O5-20CaO-15CaCl2-8KF-10Na2O-xTiO2 (where x = 2, 4, and 6), encoded by PCKNT1, PCKNT2, and PCKNT3 to replace the existing toxic shielding material used in dental and low-energy diagnostic applications. The shielding parameters include the linear attenuation coefficient (LAC), mass attenuation coefficient (MAC), half-value layer (HVL), mean free path (MFP), and effective atomic number (Zeff). In addition, the optical and physical properties of the prepared glasses were studied at low energies to replace the toxic shielding materials and to reduce the amount of exposure doses.

2. Materials and Methods

2.1. Sample Preparation

Phosphate–titanite glasses with the composition 45P2O5-20CaO-15CaCl2-8KF-10Na2O-xTiO2 (where x = 2, 4, and 6) in a mol percentage were prepared using the melt-quenching technique. The raw materials were put in a Pt crucible in the heating furnace at a temperature in the range of 1200 to 1250 °C for 30 min, depending on the composition. The melt was stirred, and when the viscosity was high, the melt was cast in the brass mold. The prepared sample was put in the annealing furnace for 2 h at 420 °C, and after that, it was switched off. A helium pycnometer (UltraPyc1200e) was used to measure the sample densities. The sample densities and the chemical compositions of the prepared samples together with the computed refractive index (n) are illustrated in Table 1.
The radiation parameters were calculated using a recently developed software package, MIKE [16]. The shielding performance of the glass samples was evaluated and compared to other phosphate–titanite glasses materials that are reported in the literature.

2.2. Optical Properties

The average molar weight of the mixtures, M ¯ , can be calculated from the mole fractions, xi, of the constituent elements and their molar masses, Mi [17]:
M ¯ = x i M i ,
where x i is the element molar fraction, and M i is the glassy composition molecular weight.
The change in the sample structure with respect to the molar composition can be better explained in terms of the molar volume rather than the density of the sample material, which expresses the oxygen distribution in the sample structure. The molar volume (VM) of glass materials can be calculated using the following equation [17]:
V M =   M ¯ ρ
where M is the average molar weight of the sample and ρ is the density of the sample. The parameter that measured the volume of glass in 1 mole of oxygen is known as the oxygen molar volume VO, which can be calculated using the following equation [17]:
V O =   V M ( 1 x i n i )
where VM is the molar volume of the glass material, xi is a molar fraction, and ni is the number of oxygen atoms in each oxide.
The oxygen packing density (OPD) of any glass material that characterizes the optical properties of the glass samples can be calculated using the following relationship [17]:
OPD = 1000 x i n i ( 1 V M )
where V M is the molar volume of the glass materials, x i is a molar fraction, and   n i is the number of oxygen atoms in each oxide.
The molar refraction (Rm) can be calculated using the following equation [17]:
R m = n 2 1 n 2 + 2 × V M
The reflection loss, RL in percentage, can be calculated using the following equation [17]:
R L = [ ( n 1 ) ( n + 1 ) ] 2
where n is the refractive index of the glass materials.
The molar electronic polarizability ( α m ) can be calculated using the following equation [17]:
α m = R m 2.52
The Miller coefficient estimated to determine the first-order nonlinear optical susceptibility of the isotropic medium, χ 1 , can be computed as follows: χ ( 1 ) = ( n 2 1 ) 4 π ; the third-order nonlinear optical susceptibility, χ ( 3 ) , can be determined by χ ( 3 ) = 1.7 × 10 10 [ χ ( 1 ) ] 4 esu; and the nonlinear refractive indices, n2, are calculated according to n 2 = 12 π χ ( 3 ) n .

2.3. Shielding Properties

When a mono-energetic photon beam with an initial intensity of I 0 travels through a   d cm thick barrier, its intensity is reduced according to the Beer–Lambert law [18]:
LAC = μ = ln I I 0 d
where I0, I, and μ represent the un-attenuated and attenuated photon intensity, and the linear attenuation coefficient, respectively.
On the other hand, the mass attenuation coefficient, μ m = ( μ / ρ ), is defined as the probability of photons interacting within the barrier and can be calculated using the following equation [19]:
MAC = μ ρ = wi ( μ ρ ) i
where ( wi ,   ( μ / ρ ) i   ( cm 2 / g ) ,   ρ )   represents the fractional weight and the mass attenuation coefficient of the individual components in each component, and ρ indicates the density of the material, respectively.
The total atom cross section ( σ a ) and total electronic cross section ( σ e ), which characterize the probability of photon interaction within the material, can be calculated using the following relationships [19,20]:
σ a = σ m 1 i   n i = ( μ / ρ ) target N A i w i A i
σ e = 1 N i (   μ ρ   ) i f i A i Z i
where NA, Ai, fi, and Zi represent the Avogadro constant, the atomic weight of the element, the fractional abundance, and the atomic number of the target element.
The effective atomic number (Zeff), which identifies photon interaction in terms of photoelectric absorption and Compton scatter, can be estimated from the total atom cross section and total electronic cross section using the following relation [21]:
Z eff = σ a σ e
The number of electrons per unit mass of the shielding material and the electron density can be calculated using the following relationship [21]:
N eff = N Z eff if i A i
The mean free path (MFP), which indicates the average distance that the photon is able to travel through the barrier, is inversely proportional to the linear attenuation coefficient, LAC. The MPF can be estimated using the following equation [22,23,24,25]:
MFP = 1 LAC ( μ )
The necessary thickness of the shielding material, which is characterized by the half-value layer (HVL) and the tenth value layer (TVL), is inversely proportional to the LAC. The following equations can be used to calculate the HVL and TVL [22,23,24,25]:
TVL = 2.302 LAC ( μ )
HVL = 0.693 LAC ( μ )

3. Results and Discussion

3.1. Physical and Optical Parameters

Table 1 shows the sample codes, compositions, measured densities, and refractive indices of the proposed phosphate–titanite glasses. Table 2 shows the molar volume, oxygen molar volume, oxygen packing density, energy gap, and Urbach energy of the proposed phosphate–titanite glasses. Table 3 shows the molar reflection (Rm), electronic polarizability (αm), and metallization (M) of the studied glasses. These are the parameters that are used to determine if the current network of glasses is dense or weak. Table 1 shows that increasing the TiO2 concentration from 2 to 6 mol percent raises the density from 2.657 to 2.682 g/cm3. As illustrated in Table 2, the VM and VO of the PCKNT glass sample decreased from 39.21 to 39.101 cm3/mol and from 15.139 to 14.867 cm3/mol, respectively. Each variable, VM and V0, is exactly proportional to the spatial distribution of the oxygen in the glass matrix. Whenever the ionic radius of the modifier ion is smaller than the size of the interstices of the glass network, an attraction to oxygen ions occurs, and the size of the interstices may decrease, as reported by Shebly [26]. This leads to a decrease in the values of VM and VO. Otherwise, by increasing the TiO2 content from 2 to 6 mol %, the OPD increased from 66.055 to 67.262 mol. Otherwise, upon increasing the TiO2 content from 2 to 6 mol %, the OPD increased from 66.055 to 67.262 mol·L−1 due to the increased oxygen atom number per unit volume [27].
The optical absorption spectra of crystal and non-crystalline materials are a useful tool for calculating optical band gap values in glass systems. As illustrated in Figure 1, appealing behavior can be observed over the wavelength range from 190 to 2500 nm. As shown in all spectra of the glasses, there is a maximal peak at the absorbance positions in the near ultraviolet range, in the wavelength range of 400–750 nm in the visible range.
As shown in Figure 1, the increase in observed absorbance peaks in this range (400–600 nm) can be attribed to the increase in the TiO2 content from 2 to 6 mol % in the prepared glass system. The observed absorbance peaks at around 497 nm indicate the presence of nanoparticle (NP) surface plasmon resonance (SPR). The SPR-related effects of the NPs of TiO2 in the prepared glass matrices are as reported by ref. [28], showing good agreement with our results. Monge et al. [29] estimated the band that appears in the optical spectroscopy of the MgO crystals with the photon excitation of the positively charged anion vacancies at 5.0 eV is equal to the wavelength = 1.24/5 = 248 nm. Moreover, PCKNT2 and PCKNT3 have their greatest absorbance values in the visible spectrum of light, implying that they are suitable for optical applications.
The transmission spectra of the investigated glasses are depicted in Figure 2. As shown in Figure 2, the transmission spectra were calculated in the ultraviolet (UV) range through the visible (Vis) to mid-infrared (MIR) ranges.
All of the glasses are characterized by similarly shaped transmission spectra in which the MIR absorption cut-off wavelength is equal to about 563 nm. The optical absorption coefficient α ( v ) was estimated from the Beer–Lambert law (Equation (1)) using the following equation:
α ( v ) = ln ( 10 ) · A / X
where A and X are the absorbance of the glass sample and the glass sample thickness, respectively. The optical band gap energy ( E opt ) for an indirect transition can be calculated using the equation derived by Mott and Davis as follows [30]:
α ( v ) = B   ( hv E opt ) n / ( hv )
where α , B, Hv, and E opt are the absorption coefficient, constant depending on the glass composition, photon energy, and optical band energy gap, respectively.
Figure 3 shows the plot of (αkv)1/2 versus the photon energy (hv), which was used to calculate the optical energy gap for indirect transitions; the optical energy gap (Eopt) for glass samples was estimated by extrapolating the linear region of (αkv)1/2 vs. (hv) (αkv)1/2 = 0, as shown in Figure 3. The energy gap was found to be decreased from 3.403 to 3.279 eV when the TiO2 concentration increased.
The refractive index values ( n ) were estimated from the energy gap values using the following equation [31]:
1 + E opt 20 = ( n + 1 ) ( n 1 ) n 2 + 2
The value of the refractive index (n) increases from 1.616 to 1.649 when the TiO2 concentration ranges from 2 to 6 mol %. The rise in the refractive index is due to the high polarity of the Ti + 3 ion, which has the ability to break the bridging oxygen (BO) with low polarity and to generate non-bridging oxygen (NBO) with high polarity. The increase in the concentration of non-bridging oxygen increases the value of the refractive index, which is consistent with the findings by others [32,33]. The values of indirect Eopt obtained for the prepared glasses were higher than those reported in other glass systems composed of 39B2O3-30PbO-20MO-10Bi2O3-1Eu2O3 (where M=K, Na, Ca, Sr, and Ba) [34], B2O3-CaO-TeO2-ZnO-ZnF2-Sm2O3 [35], and B2O3-SrCO3-Nb2O3-BaCO3-Dy2O3 [36]. Otherwise, the values of the density, refractive index, and optical packing density of the present glasses were lower compared to the glasses reported in Refs. [34,35,36,37].
The Urbach energy (ΔE), the width of the localized states, is utilized to measure the disorder degree of the atomic structure, which can be obtained by the following equation [33]:
α ( v ) = β   exp ( hv Δ E )
where ( β ) is constant.
Table 2 shows the Urbach energy values (ΔE). These values were estimated by taking the reciprocal of the slopes of the linear part from the plot of ln( α ) against ( hv ), as shown in Figure 4. It was found that the value of ( Δ E) increased from 0.2964 to 0.3031 eV by increasing the TiO2 concentration from 2 to 6 mol %.
Table 3 illustrates the molar refraction (Rm), molar polarizability ( α m ), metallization (M), third-order non-linear susceptibility χ(3), values and the nonlinear refractive indices, n2. The Rm and α m values were increased from 23.246 to 23.468 in cm3⋅mol−1 and from 9.254 to 9.313 in A03, respectively, while the value M decreased from 0.407 to 0.4 as the doped TiO2 increased from 2 to 6 mol %. The refractive index (n) value was found to be strongly dependent on the ratio α m / V m (i.e., the value of n increases as the ratio α m / V m increases). Both χ(3) and n2 increased from 4.61 to 5.97 × 10−14 (esu) and from 6.97 to 9.02 × 10−13 (esu), respectively, as the TiO2 increased in the prepared glass matrix, which was due to the hyperpolarizability of SPR of TiO2.

3.2. Radiation Shielding Properties

Figure 5 and Table 4 show the calculated mass and linear attenuation coefficients (MAC and LAC) for the prepared glasses at photon energies ranging between 15 keV and 200 keV using the MIKE program. As shown in Figure 5, the mass and linear attenuation coefficients show strong dependance on the photon energy. The value of the mass attenuation coefficient decreases rapidly up to a photon energy of 50 keV due to the influence of the photoelectric effect, which is the dominant interaction in low photon energies. For energies above 50 keV, the mass attenuation coefficient decreases slowly as the photon energy increases. It is clear that the reduction in the Na2O concentration and the increase in TiO2 result in an increase in the MAC values. The glass sample PCKNT3 has the highest MAC value among the glass samples under investigation.
The half-value layer (HVL) and mean free path (MPF) are considered important parameters used to evaluate the shielding effectiveness of the proposed material. The HVL is used to estimate the thicknesses that reduce the intensity of the radiation beam to half its initial values. They are often utilized in shielding calculations and are considered important since they quickly and directly indicate the ability of any barrier evaluated to reduce the ionizing radiation to a level that is fairly good for the environment. Table 5 and Figure 6a,b show the calculated HVL, TVL, and MFP of the prepared phosphate–titanite glasses at photon energies ranging between 15 and 200 keV. The values of HVL, TVL, and MFP decreased as the TiO2 concentration increased. These findings are consistent with the values reported for the MAC. Furthermore, the phosphate–titanite glasses under investigation were evaluated via comparison to commercially standard shielding materials such as those developed by Schott Co., Germany (RS-253 G18, RS-520, and RS-360) [38]. As shown in Figure 7a,b, the prepared glass samples show good performance at low energies. Although the standard materials showed better shielding performance compared to the proposed phosphate–titanite, the prepared glasses have advantages as light-weighted shielding materials, especially for digital dentistry applications. In such applications, a heavy shielding material such as a lead-based shielding material is not preferable due to its toxicity and heavy weight. As shown in Figure 7a, the prepared phosphate–titanite glasses showed an HVL of less than 0.5 cm at energies up to 50 keV, which makes them the material of choice in this energy range. For instance, the values recorded for PCKNT1, PCKNT2, and PCKNT3 at 45 keV were found to be 0.443 cm, 0.432 cm, and 0.424 cm, respectively. These findings are consistent with other findings in the literature [8,9].
Table 6 and Figure 8a,b depict the Zeff and Neff behaviors of prepared phosphate–titanite glass samples at energies ranging between 15 and 200 keV. As shown in Figure 8, the values of Zeff and Neff decrease as the photon energy increases up to 200 keV. The recorded values are found to be directly correlated with the TiO2 concentration in the glasses, following the order of PCKNT3 > PCKNT2 > PCKNT1. This was expected because the Zeff and Neff for all of the photons of any energy level are dependent on the mass attenuation coefficients of the constituent elements. The maximum values of Zeff and Neff were recorded at an energy of 15 keV. These findings are in agreement with other findings in the literature [39,40].

4. Conclusions

The physical, optical, and shielding properties of novel glass systems with the composition of P2O5-Na2O-CaO-8KF-CaCl2-xTiO2 (where x = 2, 4, and 6) in a mol % were developed. The reported results showed that increasing the concentration of TiO2 led to an increase in the density from 2.657 to 2.682 (g/cm3) and an increase in the OPD from 66.055 to 67.262 mol/L, while the molar volume (Vm) and oxygen molar volume (V0) decreased from 39.21 to 39.101cm3/mol and from 15.139 to 14.867 cm3/mol, respectively. The increase in both χ(3) and n2 from 4.61 to 5.97 × 10−14 (esu) and from 6.97 to 9.02 × 10−13 (esu), respectively, when the TiO2 in the prepared glass matrix increased, is due to the hyperpolarizability of SPR of TiO2. Furthermore, these glasses show good shielding properties, which makes them suitable for low energy applications at the energy range between 15 and 50 keV. Further work determining the biocompatibility of the proposed materials will be conducted to study the safety and compatibility of these material for low-energy diagnostic applications.

Author Contributions

K.I.H.: conceptualization, methodology, investigation, writing—original draft, and writing—review and editing; M.S.A.: methodology, writing—review and editing, and visualization; K.J.A.: conceptualization, methodology, formal analysis, investigation, and writing—original draft; H.Y.Z.: formal analysis, review and editing, and visualization; A.M.A.: methodology, formal analysis, investigation, writing—original draft, writing—review and editing, and visualization; I.S.Y.: 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

The Deputyship for Research and Innovation, Ministry of Education, Saudi Arabia, for funding this research work through project 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 and Innovation, Ministry of Education, Saudi Arabia, for funding this research work through project number IFP-KKU-2020/7.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ravindranadh, K. Bioactive glasses for technological and clinical applications. Int. J. Chem. Sci. 2016, 14, 1339–1348. [Google Scholar]
  2. Vesselin, D.; Takayuki, K. Classification of Simple Oxides: A Polarizability Approach. J. Solid-State Chem. 2002, 163, 100–112. [Google Scholar] [CrossRef]
  3. Larry, L.H. The story of Bioglass. J. Mater. Sci. Mater. Med. 2006, 17, 967–978. [Google Scholar]
  4. Kim, J.H.; Kim, D.K.; Lee, O.J.; Ju, H.W.; Lee, J.M.; Moon, B.M.; Park, H.J.; Kim, D.W.; Lee, J.H.; Park, C.H. Osteoinductive silk fibroin/titanium dioxide/hydroxyapatite hybrid scaffold for bone tissue engineering. Int. J. Biol. Macromol. 2016, 82, 160–167. [Google Scholar] [CrossRef] [PubMed]
  5. Itiravivong, P.; Promasa, A.; Laiprasert, T.; Techapongworachai, T.; Kuptniratsaikul, S.; Thanakit, V.; Heimann, R.B. Comparison of tissue reaction and osteointegration of metal implants between hydroxyapatite/Ti alloy coat: An animal experimental study. J. Med. Assoc. Thai. 2003, 86 (Suppl. S2), S422–S431. [Google Scholar]
  6. Awad, N.K.; Edwards, S.L.; Morsi, Y.S. A review of TiO2 NTs on Ti metal: Electrochemical synthesis, functionalization and potential use as bone implants. Mater. Sci. Eng. C Mater. Biol. Appl. 2017, 76, 1401–1412. [Google Scholar] [CrossRef]
  7. Ke, D.; Robertson, S.F.; Dernell, W.S.; Bandyopadhyay, A.; Bose, S. Effects of MgO and SiO2 on Plasma-Sprayed Hydroxyapatite Coating: An in Vivo Study in Rat Distal Femoral Defects. ACS Appl. Mater. Interfaces 2017, 9, 25731–25737. [Google Scholar] [CrossRef]
  8. Su, E.P.; Justin, D.F.; Pratt, C.R.; Sarin, V.K.; Nguyen, V.S.; Oh, S.; Jin, S. Effects of titanium nanotubes on the osseointegration, cell differentiation, mineralisation and antibacterial properties of orthopaedic implant surfaces. Bone Jt. J. 2018, 100-B (Suppl. SA1), 9–16. [Google Scholar] [CrossRef]
  9. Al-Harbi, N.; Al-Hadeethi, Y.; Bakry, A.B. Mechanical and radiation shielding features of bioactive glasses: SiO2-Na2O-CaO-P2O5-B2O3 for utilization in dental applications. J. Non-Cryst. Solids 2021, 552, 120489. [Google Scholar] [CrossRef]
  10. Alalawi, A.; Al-Buriahi, M.S.; Rammah, Y.S. Radiation shielding properties of PNCKM bioactive glasses at nuclear medicine energies. Ceram. Int. 2020, 46, 15027–15033. [Google Scholar] [CrossRef]
  11. Kilicoglu, O.; Tekin, H.O. Bioactive glasses with TiO2 additive: Behavior characterization against nuclear radiation and determination of buildup factors. Ceram. Int. 2020, 46, 10779–10787. [Google Scholar] [CrossRef]
  12. Kilicoglu, O.; Tekin, H.O. Bioactive glasses and direct effect of increased K2O additive for nuclear shielding performance: A comparative investigation. Ceram. Int. 2020, 46, 1323–1333. [Google Scholar] [CrossRef]
  13. Al-Hadeethi, Y.; Al-Buriahi, M.S.; Sayyed, M.I. Bioactive glasses and the impact of Si3N4 doping on the photon attenuation up to radiotherapy energies. Ceram. Int. 2020, 46, 5306–5314. [Google Scholar] [CrossRef]
  14. Tekin, H.O.; Kavaz, E.; Altunsoy, E.E.; Kilicoglu, O.; Agar, O.; Erguzel, T.T.; Sayyed, M.I. An extensive investigation on gamma-ray and neutron attenuation parameters of cobalt oxide and nickel oxide substituted bioactive glasses. Ceram. Int. 2019, 45, 9934–9949. [Google Scholar] [CrossRef]
  15. Kotomin, E.A.; Popov, A.I. Radiation-induced point defects in simple oxides. Nucl. Instrum. Methods Phys. Res. B 1998, 141, 1–15. [Google Scholar] [CrossRef]
  16. Girard, S.; Alessi, A.; Richard, N.; Martin-Samos, L.; de Michele, V.; Giacomazzi, L.; Agnello, S.; di Francesca, D.; Morana, A.; Winkler, B.; et al. Overview of radiation induced point defects in silica-based optical fibers. Rev. Phys. 2019, 4, 100032. [Google Scholar] [CrossRef]
  17. Hussein, K.I.; Alqahtani1, 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]
  18. Hubbell, J.H.; Seltzer, S.M. Tables of X-ray Mass Attenuation Coefficients and Mass Energy-Absorption Coefficients 1 keV to 20 MeV for Elements Z = 1 to 92 and 48 Additional Substances of Dosimetric Interest; 1995-PL; National Institute of Standards and Technology: Gaithersburg, MD, USA, 1995. [Google Scholar]
  19. Chilton, A.B.; Shultis, J.K.; Faw, R. Principle of Radiation Shielding, 1st ed.; Prentic-Halle: Englewood Cliffs, NJ, USA, 1984. [Google Scholar]
  20. Kıbrıslı, O.; Ersundu, A.E.; Ersundu, M.Ç. Dy3+ doped tellurite glasses for solid-state lighting: An investigation through physical, thermal, structural and optical spectroscopy studies. J. Non-Cryst. Solids 2019, 513, 125–136. [Google Scholar] [CrossRef]
  21. Agar, O. Study on gamma ray shielding performance of concretes doped with natural sepiolite mineral. Radiochim. Acta 2018, 106, 1009–1016. [Google Scholar] [CrossRef]
  22. Tekin, H.O.; Kavaz, E.; Papachristodoulou, A.; Kamislioglu, M.; Agar, O.; Altunsoy Guclu, E.E.; Kilicoglu, O.; Sayyed, M.I. Characterization of SiO2–PbO–CdO–Ga2O3 glasses for comprehensive nuclear shielding performance: Alpha, proton, gamma, neutron radiation. Ceram. Int. 2019, 45, 19206–19222. [Google Scholar] [CrossRef]
  23. Mostafa, A.M.A.; Issa Shams, A.M.; Sayyed, M.I. Gamma ray shielding properties of PbO–B2O3–P2O5 doped with WO3. J. Alloys Compd. 2017, 708, 294–300. [Google Scholar] [CrossRef]
  24. Rammah, Y.S.; Sayyed, M.I.; Ali, A.A.; Tekin, H.O.; El-Mallawany, R. Optical properties and gamma-shielding features of bismuth borate glasses. Appl. Phys. A 2018, 124, 83. [Google Scholar] [CrossRef]
  25. Sayyed, M.I.; Kawa, M.K.; Gaikwad, D.K.; Agar, O.; Gawai, U.P.; Baki, S.O. Physical, structural, optical and gamma radiation shielding properties of borate glasses containing heavy metals (Bi2O3/MoO3). J. Non-Cryst. Solids 2019, 507, 30–37. [Google Scholar] [CrossRef]
  26. Kyong-Soo, H.; Miae, K.; Myoung, G.H.; Jong, P.K.; Jang, -H.Y.; Jong, H.K.; Ho-Soon, Y.; Hyun, G.K. Red-emission properties and crystallization behavior in Eu2O3-TeO2 glasses. J. Non-Cryst. Solids 2019, 505, 400–405. [Google Scholar]
  27. Shelby, J.E. Introduction to Glass Science and Technology, 2nd ed.; The Royal Society of Chemistry: Cambridge, UK, 1997; pp. 138–161. ISBN 0-85-404639-9. [Google Scholar]
  28. Kumar, A.; Gaikwad, D.K.; Obaid, S.S.; Tekin, H.O.; Agar, O.; Sayyed, M.I. Experimental studies and Monte Carlo simulations on gamma ray shielding competence of (30+x)PbO–10WO3–10Na2O−10MgO–(40-x)B2O3-glasses. Prog. Nucl. Energy 2020, 119, 103047. [Google Scholar] [CrossRef]
  29. Ahmadi, F.; Ebrahimpour, Z.; Asgari, A.; Ghoshal, S.K. Insights into spectroscopic aspects of Er 3 + doped sulfophosphate glass embedded with titania nanoparticles. Opt. Mater. 2020, 111, 110650. [Google Scholar] [CrossRef]
  30. Monge, M.A.; Gonzalez, R.; Munoz Santiuste, J.E.; Pareja, R.; Chen, Y.; Kotomin, E.A. Photoconversion and dynamic hole recycling process in anion vacancies in neutron-irradiated MgO crystals. Phys. Rev. B 1999, 60, 3787–3791. [Google Scholar] [CrossRef]
  31. 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]
  32. Elkhoshkhany, N.; Reda, A.; Embaby, A.M. Preparation and study of optical, thermal, and antibacterial properties of vanadate–tellurite glass. Ceram. Int. 2017, 43, 15635–15644. [Google Scholar] [CrossRef]
  33. Upender, G.; Prasad, M. Vibrational, optical and EPR studies of TeO2-Nb2O5-Al2O3-V2O5glass system doped with vanadium. Optik 2016, 127, 10716–10726. [Google Scholar]
  34. 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]
  35. Chen, Q.; Naseer, K.A.; Marimuthu, K.; Kumar, P.S.; Miao, B.; Mahmoud, K.A.; Sayyed, M.I. Influence of modifier oxide on the structural and radiation shielding features of Sm3+-doped calcium telluro-fluoroborate glass systems. J. Aust. Ceram. Soc. 2021, 57, 275–286. [Google Scholar] [CrossRef]
  36. 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]
  37. Naseer, K.A.; Marimuthu, K.; Mahmoud, K.A.; Sayyed, M.I. Impact of Bi2O3 modifier concentration on barium–zincborate glasses: Physical, structural, elastic, and radiation-shielding properties. Eur. Phys. J. Plus 2021, 136, 116. [Google Scholar] [CrossRef] [PubMed]
  38. Hazlin, M.A.; Halimah, M.K.; Muhammad, F.D.; Faznny, M.F. Optical properties of zinc borotellurite glass doped with trivalent dysprosium ion. Physica B 2017, 510, 38–42. [Google Scholar] [CrossRef]
  39. Schott Co. Radiation Shielding Glasses. Available online: https://www.schott.com/advanced_optics/english/products/optical-materials/special-materials/radiation-shielding-glasses/index.html (accessed on 24 February 2022).
  40. Agar, O.; Khattari, Z.Y.; Sayyed, M.I.; Tekin, H.O.; Al-Omari, S.; Maghrabi, M.; Zaid, M.H.M.; Kityk, I.V. Evaluation of the shielding parameters of alkaline earth based phosphate glasses using MCNPX code. Results Phys. 2019, 12, 101–106. [Google Scholar] [CrossRef]
Figure 1. Absorbance spectroscopy for different compositions of PCKNT.
Figure 1. Absorbance spectroscopy for different compositions of PCKNT.
Crystals 12 00941 g001
Figure 2. UV-Vis-NIR transmission spectra of the TiO 2 -doped phosphate–titanite glasses.
Figure 2. UV-Vis-NIR transmission spectra of the TiO 2 -doped phosphate–titanite glasses.
Crystals 12 00941 g002
Figure 3. Plot of (αhv)1/2 as a function of the photon energy (hv) of prepared glasses.
Figure 3. Plot of (αhv)1/2 as a function of the photon energy (hv) of prepared glasses.
Crystals 12 00941 g003
Figure 4. Plot of ln(α) as a function of the photon energy (hv) of prepared glasses.
Figure 4. Plot of ln(α) as a function of the photon energy (hv) of prepared glasses.
Crystals 12 00941 g004
Figure 5. (a) The mass attenuation coefficient of prepared glasses; (b) the linear attenuation coefficient of prepared glasses.
Figure 5. (a) The mass attenuation coefficient of prepared glasses; (b) the linear attenuation coefficient of prepared glasses.
Crystals 12 00941 g005
Figure 6. The shielding parameters of TPNK systems: (a) HVL, (b) TVL, and (c) MFP.
Figure 6. The shielding parameters of TPNK systems: (a) HVL, (b) TVL, and (c) MFP.
Crystals 12 00941 g006aCrystals 12 00941 g006b
Figure 7. The shielding parameters for TPNK and standard materials: (a) HVL and (b) MFP.
Figure 7. The shielding parameters for TPNK and standard materials: (a) HVL and (b) MFP.
Crystals 12 00941 g007
Figure 8. The radiation shielding parameters: (a) effective atomic number (Zeff); (b) effective electron number (Neff).
Figure 8. The radiation shielding parameters: (a) effective atomic number (Zeff); (b) effective electron number (Neff).
Crystals 12 00941 g008
Table 1. The composition, density (ρ), and refractive index (n) of the PCKNT glass system.
Table 1. The composition, density (ρ), and refractive index (n) of the PCKNT glass system.
Sample CodeComposition
(mol%)
Density in
gcm−3 ± 0.037
Refractive
Index
PCKNT145P2O5-20CaO-15CaCl2-8KF-10Na2O-2TiO22.6571.616
PCKNT245P2O5-20CaO-15CaCl2-8KF-10Na2O-4TiO22.67921.637
PCKNT345P2O5-20CaO-15CaCl2-8KF-10Na2O-6TiO22.68271.649
Table 2. The molar volume (VM), oxygen molar volume (VO), optical packing density (OPD), energy gap (Eopt), and Urbach energy (ΔE) of the prepared glasses.
Table 2. The molar volume (VM), oxygen molar volume (VO), optical packing density (OPD), energy gap (Eopt), and Urbach energy (ΔE) of the prepared glasses.
Sample CodeVM (cm3/mol)VO (cm3/mol)OPD (mol−1)Energy Gap, Eopt (eV)Urbach Energy, ∆E (eV)
PCKNT137.46014.46469.1393.4030.2964
PCKNT237.28414.28570.0003.3240.2914
PCKNT337.36914.20870.3793.2790.3031
Table 3. The electronic polarizability (αm), molar reflection (Rm), and the metallization (M) of the studied glasses.
Table 3. The electronic polarizability (αm), molar reflection (Rm), and the metallization (M) of the studied glasses.
Sample Code Molar   Polarizability ,   α m ,   ( Å 3 ) Molar   Refraction ,   R m ,   ( c m 3   / mol ) Metallization (M) (±0.001)Third-Order Non-Linear Susceptibility χ(3) × 10−14 (esu)Nonlinear Refractive Indices, n2, × 10−13 (esu)
TPNK15.19513.0900.4074.616.97
TPNK25.31013.3830.4025.448.22
TPNK35.40213.6130.4005.979.02
Table 4. Mass and linear attenuation coefficient values for PCKNT glass systems.
Table 4. Mass and linear attenuation coefficient values for PCKNT glass systems.
Energy MAC   ( c m 2 / g )   LAC (cm−1)
(keV)PCKNT1PCKNT2PCKNT3PCKNT1PCKNT2PCKNT3
1510.9010211.1684911.4340428.9640129.9226130.67410
204.7748764.8935445.01136412.6868513.1107813.44399
301.5486141.5853031.6217294.1146684.2473424.350612
400.7512240.7669210.7825061.9960032.0547352.099228
450.5766250.5876830.5986621.5320931.5745211.606031
500.4640470.4721140.4801241.2329741.2648891.288029
600.3347670.3394340.3440670.8894770.9094110.923029
800.2275030.2294570.2313960.6044760.6147600.620767
1000.1848840.1858670.1868440.4912370.4979760.501247
1400.149130 0.1494700.1498080.3962380.4004610.401891
1500.1439280.1442000.1444700.3824170.3863410.387570
1600.1394970.1397160.1399340.3706430.3743270.375400
1700.1355840.1357610.1359380.3602470.3637320.364680
1800.1321660.1323120.1324560.3511650.3544890.355340
1900.1290590.1291790.1292990.3429100.3460980.346871
2000.1262310.1263310.1264300.3353970.3384660.339174
Table 5. Half-value layer, tenth value layer, and the mean free path of prepared PCKNT glasses.
Table 5. Half-value layer, tenth value layer, and the mean free path of prepared PCKNT glasses.
EnergyHVL (cm)TVL(cm)MFP(cm)
(keV)PCKNT1PCKNT2PCKNT3PCKNT1PCKNT2PCKNT3PCKNT1PCKNT2PCKNT3
150.0239260.0231600.0225920.0794090.0768650.0749820.0345260.033420.032601
200.0546240.0528570.0515470.1812900.1754280.1710800.0788220.0762730.074383
300.1684220.1631610.1592880.5589760.5415150.5286610.2430330.2354410.229853
400.3471940.3372700.3301211.1523031.1193661.0956410.5010010.4866810.476366
450.4523220.4401340.4314991.5012141.4607621.4321020.6527020.6351140.622653
500.5620560.5478740.5380311.8654091.8183411.7856740.8110470.7905830.776380
600.7791090.7620310.7507892.5857892.5291092.4917971.1242561.0996121.083390
801.1464481.1272681.1163613.8049493.7412953.7050951.6543261.6266501.610911
1001.4107261.3916331.3825534.6820624.6186954.5885612.0356792.0081281.995026
1401.7489481.7305061.7243505.8045885.7433835.7229512.5237342.4971232.488240
1501.8121571.7937521.7880636.0143745.9532905.9344102.6149452.5883872.580178
1601.8697221.8513211.8460306.2054286.1443556.1267952.6980122.6714592.663824
1701.9236831.9052491.9002986.3845176.3233396.3069042.7758772.7492782.742132
1801.9734291.9549251.9502456.5496216.4882076.4726752.8476612.8209602.814207
1902.0209402.0023251.9978626.7073066.6455226.6307102.9162202.8893582.882917
2002.0662092.0474702.0431986.8575466.7953566.7811752.9815422.9545032.948337
Table 6. Effective atomic number and effective electron density for PCKNT glasses.
Table 6. Effective atomic number and effective electron density for PCKNT glasses.
Energy Z e f f N e f f   ×   10 + 23
(keV)PCKNT1PCKNT2PCKNT3PCKNT1PCKNT2PCKNT3
1515.92916.08516.3474.24.224.24
2015.85016.00816.2734.184.214.22
3015.39415.55015.8224.064.084.11
4014.69614.84215.1173.883.903.92
4514.32114.46214.7343.783.803.82
5013.95814.09014.3573.693.713.73
6013.32013.43613.6853.523.543.55
8012.44712.53412.7473.293.303.31
10011.97512.04512.2303.163.173.17
14011.56611.61811.7753.053.053.06
15011.51511.56511.7193.043.043.04
16011.47411.52211.6733.033.033.03
17011.44211.48811.6373.023.023.02
18011.41511.46011.6063.013.013.01
19011.39211.43711.5812.972.993.01
20011.37511.41811.5612.952.972.99
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hussein, K.I.; Alqahtani, M.S.; Alzahrani, K.J.; Zahran, H.Y.; Alshehri, A.M.; Yahia, I.S.; Reben, M.; Yousef, E.S. The Investigation of New Phosphate–Titanite Glasses According to Optical, Physical, and Shielding Properties. Crystals 2022, 12, 941. https://doi.org/10.3390/cryst12070941

AMA Style

Hussein KI, Alqahtani MS, Alzahrani KJ, Zahran HY, Alshehri AM, Yahia IS, Reben M, Yousef ES. The Investigation of New Phosphate–Titanite Glasses According to Optical, Physical, and Shielding Properties. Crystals. 2022; 12(7):941. https://doi.org/10.3390/cryst12070941

Chicago/Turabian Style

Hussein, Khalid I., Mohammed S. Alqahtani, Khloud J. Alzahrani, Heba Y. Zahran, Ali M. Alshehri, Ibrahim S. Yahia, Manuela Reben, and El Sayed Yousef. 2022. "The Investigation of New Phosphate–Titanite Glasses According to Optical, Physical, and Shielding Properties" Crystals 12, no. 7: 941. https://doi.org/10.3390/cryst12070941

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