5.2
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
5.3
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
View/Download PDF

Translate this page into:

Original Article
2025
:18;
4392025
doi:
10.25259/AJC_439_2025

Structural and radiation attenuation properties of modified barium lithium bismovanadium borate glasses

Department of Architecture, Faculty of Engineering and Architecture, Umm Al-Qura University, Makkah 24381, Saudi Arabia.
Department of Physics, College of Science, University of Tabuk, 71474 Tabuk, Saudi Arabia
Department of Physical Sciences, physics Division, College of Science, Jazan University, P.O. Box. 114, Jazan 45142, Kingdom of Saudi Arabia
Department of Chemistry, Al-Qunfudah University College, Umm Al-Qura University, Al-Qunfudah 1109, Saudi Arabia.
College of Science, Chemistry Department, Imam Mohammad Ibn Saud Islamic University (IMSIU), Riyadh, 11623, Saudi Arabia
Center for Innovation and Entrepreneurship, Imam Mohammad Ibn Saud Islamic University (IMSIU), Riyadh 11623, Saudi Arabia
Department of Chemistry, Faculty of Science, Umm Al-Qura University, Makkah 24230, Saudi Arabia
Department of Chemistry, Faculty of Science, Mansoura University, El-Gomhoria Street 35516, Egypt

* Corresponding author: E-mail address: n_elmetwaly00@yahoo.com (N. El-Metwaly)

Licence
This is an open-access article distributed under the terms of the Creative Commons Attribution-Non Commercial-Share Alike 4.0 License, which allows others to remix, transform, and build upon the work non-commercially, as long as the author is credited and the new creations are licensed under the identical terms.

Abstract

The presented work investigates the structural and radiation attenuation properties of modified barium lithium bismuth vanadium borate glasses with varying compositions of B₂O₃, Li₂O, BaO, Bi₂O₃, and V₂O₅. The glasses were fabricated using the melt quenching method, and their physical, structural, and shielding characteristics were systematically analyzed. Fourier transform infrared spectroscopy (FTIR) revealed distinct vibrational modes associated with BO₃, BO₄, Bi-O, and V-O structural units, indicating significant network modifications with increasing Bi₂O₃ content. Density and molar volume measurements demonstrated a direct correlation with compositional changes, while the packing density and free volume provided insights into the structural compactness and ion mobility. Radiation shielding performance was evaluated across a broad photon energy range (0.015–15 MeV), with calculation of key parameters such as the mass attenuation coefficient (MAC), linear attenuation coefficient (LAC), half-value layer (HVL), and effective atomic number (Zₑff). The results showed enhanced shielding efficiency at higher Bi₂O₃ concentrations, particularly near the Bi K-edge (0.1 MeV), where the photoelectric effect (PE) dominated. The sample with 45 mol% Bi₂O₃ exhibited the highest shielding effectiveness, attributed to the high atomic number and density of bismuth. Conversely, the inclusion of V₂O₅ reduced shielding performance, highlighting the critical role of composition optimization. These findings underscore the potential of bismuth-rich borate glasses as effective, non-toxic alternatives for radiation shielding applications in medical, nuclear, and industrial settings. The study provides a foundation for tailoring glass properties to meet specific shielding requirements while maintaining structural integrity.

Keywords

Bismuth oxide
Borate glass
Density and molar volume
Radiation shielding
Vanadium oxide

1. Introduction

Humans are constantly exposed to radiation, both from natural sources and from human activities. Radiotherapy applications often involve the use of γ-rays and X-rays, which are radioactive. Nuclear power plants utilize fissionable radionuclides to generate energy, while research reactors employ alpha sources to produce ionizing radiation and neutrons for various purposes such as research, teaching, and training [1]. However, if ionizing radiation is not handled carefully, it can pose significant risks to both human health and the environment. Consequently, several materials have been suggested as potential radiation protection agents to mitigate these hazards [2]. Glasses are extensively utilized in diverse fields, ranging from the building industry to medical facilities, military applications, and research laboratories, due to their remarkable attributes such as transparency, high water resistance, homogeneity, and exceptional durability in chemically corrosive environments [3].

Radiation-shielding glasses have become indispensable in architectural design, protecting against ionizing radiation while preserving natural light transmission. In healthcare settings such as hospitals, diagnostic imaging facilities, and laboratories, specialized lead-based or heavy metal oxide glasses are increasingly used in doors, walls, and observation windows to ensure safety for patients and personnel [4,5]. Beyond healthcare, these materials are being adopted in airports, nuclear research facilities, and select residential and commercial buildings, balancing safety with aesthetic and functional transparency. The construction industry has responded to growing radiation exposure concerns by developing innovative radiation-resistant materials, including advanced glass panels, bricks, and composites enriched with heavy elements. These materials not only offer superior shielding against natural and artificial radiation sources but also enhance structural integrity. By incorporating such cutting-edge solutions into public and industrial infrastructure, engineers and architects are addressing modern radiation risks while promoting safer, more resilient built environments [6,7].

The principle of “shielding” revolves around the capability of a specific medium to mitigate the impact of radiation through processes such as scattering and absorption. The effectiveness of radiation shielding provided by the medium is contingent upon factors like the medium’s thickness, density, and the energy level of the radiation [8,9]. Lead is a widely used material for shielding purposes to provide protection against radiation and reduce the effective dose [10]. However, lead is associated with various disadvantages, with its toxicity being a major concern. As a result, the use of lead in radiation shielding applications presents a critical challenge due to its hazardous nature. Therefore, there is a crucial need to explore inexpensive, safe, and non-toxic alternative materials for radiation shielding objectives [11,12]. The most important glass forms typically include silicon oxide (SiO2), borate oxide (B2O3), and phosphorus oxide (P2O5). B2O3-based glasses are particularly notable for their remarkable characteristics, such as high transmittance, low melting point, thermal stability, and exceptional glass-forming ability. These glass structures are widely utilized in various applications, ranging from piezoelectric actuators and microelectronics to opto-acoustical electronics, solid-state laser materials, and luminescent materials and microelectronics [13,14].

Extensive research has been conducted on bariumborate-based glass systems containing transition metal oxides due to their immense potential in radiation shielding and optical applications. The addition of barium oxide (BaO) to borate oxide (B2O3) leads to the formation of barium borate (BaO-B2O3) glass, which possesses the desired characteristics for effective radiation shielding [8,9]. These glasses can be further enhanced by incorporating dopants and adjusting their concentrations within the glass matrix. The unique properties of pure borate glasses, such as their sensitivity to dopants, wide transparency window, and low phonon energies, contribute to the overall performance of these glass systems [15]. Barium Vanadium Lithium Borate (BVLB) glasses play a significant role in radiation shielding due to their unique properties. These glasses are a type of optical material with high refractive indices and low dispersion, rendering them appropriate for a wide range of uses, such as radiation protection. BVLB glasses are composed of barium oxide (BaO), vanadium oxide (V2O5), lithium oxide (Li2O), and boron oxide (B2O3). The combination of these elements results in a material with excellent radiation shielding capabilities. The primary reason for this is the presence of barium, which has a high atomic number (Z = 56) and is highly effective in absorbing γ- and X-ray radiation. When exposed to radiation, BVLB glasses efficiently absorb and scatter the radiation, reducing the intensity of the radiation that reaches undesirable areas. This property makes them ideal for use in applications such as radiation protection windows, X-ray tubes, and other radiation shielding components. In summary, BVLB glasses contribute significantly to radiation shielding due to their composition and the ability to effectively absorb and scatter radiation, thus protecting individuals and equipment from potentially hazardous exposure.

The physical and chemical characteristics of bismuth barium borate glasses are intricately linked to their structural properties. The manner in which atoms and bonds are arranged within the glass network directly affects properties such as transparency, thermal stability, and mechanical strength. To cater to specific applications, it is crucial to have a comprehensive understanding of the structural features exhibited by these glasses [16]. When analyzing the radiation shielding properties of composite materials, such as glass, it is essential to consider various metrics. The parameters include the mass attenuation coefficient (MAC), half value layer (HVL), effective atomic number, and mean free path (MFP). Among these metrics, the MAC holds particular significance as it provides a comprehensive measure for understanding the interactions between γphotons and the shielding material.

The main purpose of this study is to assess the fundamental properties of the vanadium barium lithium borate glass, as well as the variations observed in glass compositions with varying concentrations of Bi2O3. This investigation will be conducted using optical and FTIR analysis, which will provide insights into the influence of bismuth ions on the optical and vibrational characteristics of these glasses. Additionally, the study will involve density measurements and compositional analysis to determine the shielding attenuation parameters. This study also introduces a novel barium borate glass system co-doped with Bi₂O₃ and V₂O₅, engineered to synergistically enhance optical and structural properties. The tailored reduction in optical band gap (2.45–3.25 eV) via vanadium incorporation, coupled with FTIR-validated borate network modifications, demonstrates unprecedented tunability. Induced disorder provides new insights into defect engineering. The work pioneers these glasses as low-cost, solution-processable candidates for near-infrared optical devices.

2. Materials and Methods

Modified barium borate glass samples were synthesized by incorporating varying concentrations of bismuth oxide (Bi₂O₃) and vanadium oxide (V₂O₅) via the conventional melt-quenching technique. The glass matrix was composed of boron oxide (B₂O₃), lithium oxide (Li₂O), barium oxide (BaO), bismuth oxide (Bi₂O₃), and vanadium oxide (V₂O₅) in molar percentages, as detailed in Table 1. High-purity raw materials, including boric acid (H₃BO₃) as the precursor for B₂O₃, lithium carbonate (Li₂CO₃) for Li₂O, and barium carbonate (BaCO₃) for BaO, were employed. Bismuth oxide (Bi₂O₃) and vanadium oxide (V₂O₅) were used in their pure forms. Stoichiometric quantities of the starting materials were precisely weighed and homogenized via mechanical mixing to ensure compositional uniformity. The batch mixtures were melted in 50 mL porcelain crucibles within an electric muffle furnace. An initial heat treatment at 400°C for 2 h was applied to facilitate the decomposition of carbonates and the release of volatile species (H₂O and CO₂). Subsequently, the temperature was gradually increased to 1100°C and maintained for 2 h to ensure complete melting and homogenization. To enhance homogeneity and minimize bubble formation, the melt was periodically stirred by rotating the crucible during the melting process. Upon achieving a homogeneous melt, the glass was rapidly cast into a preheated stainless-steel mold (preheated to 450°C to mitigate thermal stress) and immediately transferred to an annealing furnace maintained at 450°C. The furnace was then switched off and allowed to cool slowly to room temperature to prevent thermal shock and cracking. The annealed glass samples were stored in a desiccator until further characterization.

Table 1. Chemical compositions of the studied glasses (mol. %).
Samples B2O3 Li2O BaO Bi2O3 V2O5
Bi40V5 30 10 15 40 5
Bi41V4 30 10 15 41 4
Bi42V3 30 10 15 42 3
Bi43V2 30 10 15 43 2
Bi44V1 30 10 15 44 1
Bi45V0 30 10 15 45 0

2.1. Physical measurements and parameters

2.2.1. FTIR spectroscopy

The prepared glass samples would undergo FTIR spectral analysis, which is a standard procedure for studying the infrared absorption characteristics of materials. Typically, pellets or thin films are prepared from the glass samples to facilitate the measurement of infrared absorption in the spectral analysis process. The FTIR spectra of the glass samples were deconvoluted using the PeakFit software (v4.12) to identify overlapping vibrational bands. Gaussian-Lorentzian (Voigt) functions were employed for peak fitting, with the baseline corrected iteratively. The deconvolution process involved optimizing the number of peaks via second-derivative analysis and residual minimization, ensuring accurate assignment of functional groups. The full width at half maximum (FWHM), peak position, and area were quantified for each resolved band.

2.2.2. Density and molar volume

The precise determination of a substance’s density hinges on possessing the mass and volume measurements of the material in question. Through the provision of these fundamental values, one can accurately and dependably calculate the density of the substance, also used to calculate molar volume using the Eqs. (1,2)

(1)
ρ t h e o r = i n i ρ i

(2)
V m = i n i M i ρ S

where ni is the molar fraction, Mi is the molecular weight of the sample component, and pi is the density for each component.

The calculation of free volume, denoted as Vf, is a fundamental aspect of molecular studies. It allows scientists to quantify the space available for molecular movement and helps in predicting the behavior of molecules in different environments. By utilizing Eqs. (3,4), researchers can accurately determine the free volume and further explore the implications of molecular mobility within a network.

(3)
V f = V m i X i V i

(4)
V i = 4 π N A 3   ( b   r A 3 + c     r B 3 )

Where Xi is the oxide molar ratio, Vi is the corresponding volume, rA is the radius of the element ion, rB is a radius of the oxygen ion and b, c are the equivalence.

The concept of packing density (Pd) refers to the relationship between the effective volume of a glass material and the minimum volume required for the ions within the glass to occupy. This proportion is calculated using Eq. (5) [17].

(5)
P d = i X i V i V m

2.2.3. Radiation shielding characterizations

The Phys-x program was employed to analyze the radiation shielding properties of the produced glasses over a broad range of photon energies, spanning from 15 keV to 15 MeV. Key parameters, including MAC, linear attenuation coefficient (MAC), HVL, MFP, atomic (ACSs) and electronic cross sections (ECSs), and effective atomic number, were computed. These parameters are essential in evaluating the efficiency of radiation shielding [18].

The Beer-Lambert law is applicable in elucidating the penetration of γ-rays through a glass, as demonstrated by the utilization of Eq. (6) [19]:

(6)
I = I 0 e μ x

In the context of photon interactions with glass, (I0) signifies the maximum number of photons, while (I) represents the photon count that has passed through glass of thickness (x). The linear attenuation coefficient (LAC), denoted as (μ) and expressed in (cm-1), describes how photons interact with the glass material. The MACs quantify the interactions of incident photons per unit mass of absorber glasses, obtained by dividing the LAC values by the density of the glasses. Through the application of the mixture rule, Eq. (7) provides a means to easily determine the attenuation coefficients [20].

(7)
μ ρ = i w i ( μ ρ ) i

The MAC (μ⁄ρ) represents the relationship between the mass attenuation of the ith constituent element and its weight percentage (wi) in the glass sample.

The effectiveness of glass as a shield is determined by the HVL. According to Eq. (8) [21,22], the HVL represents the thickness at which 50% of the incident gamma rays are attenuated.

(8)
H V L = ln 2 μ

The MFP serves as a crucial metric in understanding the behavior of photons when they encounter glass, enabling the assessment of the glass’s ability to attenuate light. It quantifies the average distance that photons traverse before experiencing scattering or absorption phenomena. To ascertain this parameter, one can employ Eq. (9) [23].

(9)
M F P = 1 L A C

The parameter known as the effective atomic number plays a crucial role in characterizing the response of multi-element structures to ionizing radiation. In the current study, the direct method was employed to ascertain the effective atomic number through the assessment of both atomic and ECSs [24].

The calculation of the total ECS ( σ e ) (Eq. 10) for each element is determined by employing the subsequent equation:

(10)
  σ e = 1 N A f i A i Z i   ( μ ρ ) e n

Where f i = n / i n i   is used to indicate the proportionate abundance of a specific element i in relation to the total number of atoms, while i n f i = 1   denotes the atomic number assigned to the ith element.

The calculation of the effective ACS ( σ a ) (Eq. 11) is determined by employing a specific equation.

(11)
           σ a = 1 N A f i A i ( μ ρ ) en

Where ( μ ρ ) e n is the mass energy absorption coefficient of the soft tissue, fi is the fraction by weight of element i, and Ai is used to signify the atomic weight of element i

The relationship between the effective atomic number, Zeff, (Eq. 12) of a composite material and the σ e and σ a can be expressed as follows:

(12)
Z eff =  σ a  σ e

3. Results and Discussion

3.1. Fourier transform infrared spectroscopy (FTIR)

The examination of the FTIR spectrum offers valuable insights into the bonding characteristics and local structural environment of the glass system. Figure 1 demonstrates that bands may be observed within the 1400-1200 cm⁻1 range, which are a result of the asymmetric stretching vibrations of B-O bonds in the trigonal BO3 units that form the borate network [25]. The frequency range between 800-600 cm⁻1 is typically associated with the bending vibrations of B-O-B linkages or the stretching vibrations of Bi-O bonds. Conversely, the spectral range of 600-400 cm⁻1 often exhibits bands that correspond to the stretching vibrations of various metal-oxygen (M-O) bonds, such as V-O, Ba-O, and Bi-O [26].

FTIR of the studied glasses with varying vanadium content.
Figure 1.
FTIR of the studied glasses with varying vanadium content.

The bands’ presence and density offer valuable information regarding the local coordination environments of metal ions and the strength of metal-oxygen bonds. Moreover, the position and intensity of bands related to B-O, V-O, and other metal oxygen vibrations can exhibit variations based on the composition and structural alterations induced by different concentrations of V2O5 and Bi2O3 [27].

The observed modifications exhibit a correlation with the trends observed in the physical properties, such as density, molar volume, and polaron radius. The wavenumber of 957 cm-1 falls within the typical range associated with the stretching vibrations of B-O bonds in borate glasses. This particular wavenumber can be attributed to the stretching vibrations of tetrahedral BO4 units or the asymmetric stretching vibrations of trigonal BO3 units present in the borate network [28]. The wavenumber of 853 cm-1 could potentially be attributed to Bi−O bonds in [BiO3] or the stretching vibrations of specific metal-oxygen (M-O) bonds found in the glass system. In borate glasses that contain vanadium, this particular band might be associated with the stretching vibrations of V-O bonds or a combination of V-O-B vibrations [29].

The FTIR spectra of lithium barium borate glasses containing bismuth and vanadium oxides present significant analytical challenges due to extensive peak overlapping from multiple structural units. This spectral complexity arises from several sources: borate network units (both BO3 trigonal units with vibrations in 1200-1500 cm⁻1 and 650-700 cm⁻1 ranges, and BO4 tetrahedral units with bands around 850-1100 cm⁻1), bismuth-oxygen vibrations (primarily in the 400-600 cm⁻1 region), vanadium-oxygen vibrations (V=O stretching around 900-1000 cm⁻1 and V-O-V linkages in the 500-800 cm⁻1 region), and modifier effects from Li⁺ and Ba1⁺ cations. To address these challenges, deconvolution analysis must be performed through a systematic approach including proper baseline correction (especially important due to bismuth’s background absorption effects), initial band identification based on literature for individual components [30,31], selection of appropriate peak fitting functions (typically Gaussian functions for these glasses due to structural disorder), and an iterative fitting process [31]. With varying Bi2O3 and V2O5 content, deconvolution would reveal network transformation (conversion of BO3 to BO4 units), non-bridging oxygen formation (evident through new bands around 900-950 cm⁻1), vanadium coordination changes (affecting the ratio of VO4 to VO5 units), and alterations in network connectivity. Deconvoluted spectral data of the studied samples have been presented in Figure 2 and Table 2. The calculation of N4 is determined by analyzing the shift in population proportions between BO4 and BO3 triangular units. Notably, the levels of N4 are significantly impacted by the quantity of BO4 units that are present in the system Figure 3. Moreover, the number of bridging oxygens per network structure of the studied borate glass systems containing different concentrations of Bi2O3 and V2O5 shown in Figure 4. It is evident from the figure that as the concentration of Bi2O3 increases (from 40 mol% to 45 mol%) and the concentration of V2O5 decreases (from 5 mol% to 0 mol%), the number of bridging oxygens per networker decreases monotonically. This trend suggests that the addition of Bi2O3 and the removal of V2O5 lead to the depolymerization of the glass network structure.

Deconvolution of studied glasses.
Figure 2.
Deconvolution of studied glasses.
Table 2. Band position and area obtained from such deconvolution studied glasses.
Peak No. Bi40V5
Bi41V4
Bi42V3
Bi43V2
Bi44V1
Bi45V0
Center R.A Center R.A Center R.A Center R.A Center R.A Center R.A
1 470 10.78 457 14.96 455 10.89 458 3.85 467 12.81 470 11.27
2 571 10.69 573 9.31 573 9.84 528 4.75 569 16.40 568 18.61
3 723 30.30 678 41.39 682 45.53 578 7.12 721 37.31 712 50.44
4 843 19.82 910 78.76 902 71.57 674 42.54 860 58.33 858 65.95
5 921 70.94 994 76.61 1000 75.92 882 14.46 935 66.69 937 79.41
6 1032 150.10 1067 28.97 1070 22.77 946 15.65 1027 115.42 1030 100.44
7 1197 20.58 1196 3.02 1204 4.05 1031 13.98 1199 8.35 1217 8.86
8 1267 16.66 1327 9.56 1327 13.36 1201 2.76 1330 7.88 1324 11.39
9 1322 10.52 1391 2.52 1399 3.08 1331 4.53 1391 2.97 1401 1.05
N4 of the studied glasses.
Figure 3.
N4 of the studied glasses.
The number of bridging oxygens per networker of the studied glasses.
Figure 4.
The number of bridging oxygens per networker of the studied glasses.

Bismuth oxide (Bi2O3) is known to act as a network modifier in borate glasses, breaking the bridging oxygens and creating non-bridging oxygens (NBOs). On the other hand, vanadium oxide (V2O5) is a conditional glass former that can participate in the glass network structure, forming bridging oxygens. The observed decrease in the number of bridging oxygens per networker with increasing Bi2O3 content can be attributed to the disruption of the borate network by the introduction of NBOs, which weakens the overall connectivity of the glass structure. Conversely, the decrease in V2O5 content reduces the number of bridging oxygens contributed by the vanadium ions, further contributing to the depolymerization of the glass network.

The Weak absorption bands observed in the 2000–4000 cm⁻1 region were attributed to overtone/combination modes of borate networks (e.g., 2B–O stretching near 2850 cm⁻1) and residual hydroxyl (O–H) stretching vibrations (∼3400 cm⁻1) from adsorbed moisture. The deconvolution process confirmed these low-intensity features through second-derivative analysis and residual minimization, with FWHM, peak positions, and relative areas quantified for each resolved band [32].

3.2. Physical parameters calculation

Calculating physical parameters for lithium barium borate glasses containing bismuth and vanadium oxides provides crucial insights into their structure-property relationships and potential applications. Such physical properties have been shown in Figure 5 and summarized in Table 3. Density measurements directly reflect compositional changes and serve as a foundation for deriving other parameters, while molar volume indicates how tightly the glass components are packed together, revealing network expansion or contraction. Packing density quantifies the efficiency of space filling within the glass structure, helping identify open or compact arrangements created by different modifier oxides. Free volume calculations reveal the unoccupied spaces available for ion migration, directly correlating with ionic conductivity properties. Polaron radius and inter-ionic distances (such as boron-boron separation) characterize charge carrier mobility and network connectivity, respectively [33,34], which influence electrical and thermal transport. The number of ions per unit volume affects various properties including refractive index and electrical conductivity. Together, these parameters establish quantitative relationships between glass composition and technological properties, enabling prediction of radiation shielding effectiveness based on electron density and atomic packing. For radiation shielding applications specifically, these physical parameters help explain why bismuth-rich compositions typically outperform vanadium-rich counterparts, as the higher electron density, atomic number, and packing efficiency of bismuth-containing glasses enhance their photoelectric absorption capabilities, particularly near absorption edges. This comprehensive physical characterization guides the rational design of optimized glass compositions for specific applications in radiation protection, optoelectronics, and other advanced technological fields.

(a-d) Representation of physical parameters including density, molar volume, packing density B-B distance of the studied glasses.
Figure 5.
(a-d) Representation of physical parameters including density, molar volume, packing density B-B distance of the studied glasses.
Table 3. Physical properties of the studied glasses
Parameters Glass Code
Bi40V5 Bi41V4 Bi42V3 Bi43V2 Bi44V1 Bi45V0
Density (ds) gcm-3±0.0002 5.53 5.58 5.64 5.69 5.75 5.80
Molar volume (Vm) cm3/mol ±0.0001 43.86 43.94 44.01 44.08 44.15 44.22
Packing density (Pd) 0.50 0.49 0.49 0.49 0.49 0.48
Free volume (Vf) 22.06 22.23 22.39 22.56 22.72 22.88
Average mol.wt. (MAv) (g) 242.35 245.19 248.03 250.87 253.71 256.56
Ion concentration (N) (10+21 ions) 5.49 5.62 5.75 5.87 5.99 6.13
Polaron radius (rp) (A˚) 2.28 2.27 2.25 2.23 2.22 2.20
Inter-nuclear distance (ri) (A˚) 5.67 5.63 5.58 5.54 5.50 5.47
Field strength (F) 1017(g mol-1cm-2) 89.00 90.00 92.00 93.00 94.00 96.00
Molar volume of the boron atoms (Vb) 31.33 31.38 31.44 31.49 31.54 31.58
Average boron–boron distance10-8 (dB–B) (nm) 3.73 3.73 3.74 3.74 3.74 3.74

The seemingly paradoxical simultaneous increase in both density and molar volume in the studied as x decreases from 5 to 0 can be explained by examining the atomic and molecular characteristics of bismuth and vanadium oxides. This trend can be explained as bismuth oxide (Bi2O3) has significantly higher molecular weight (465.96 g/mol) compared to vanadium oxide (V2O5, 181.88 g/mol), while their molar volumes differ less dramatically. As V2O5 is progressively replaced by Bi2O3, the much heavier bismuth atoms substantially increase the overall density, despite not occupying proportionally smaller spaces in the glass network. The increase in molar volume alongside density indicates that bismuth ions create a slightly more expanded network structure than vanadium ions, potentially due to the larger ionic radius of Bi3⁺ (103 pm) compared to V⁵⁺ (54 pm) and bismuth’s stereochemically active lone pair of electrons that requires additional space.

The near-constant packing density suggests that these compositional changes maintain similar efficiency in space filling within the glass structure, with the slight increase in free volume (22.06 to 22.8) indicating that bismuth incorporation creates marginally more unoccupied space. This could be attributed to bismuth’s tendency to form asymmetric coordination environments due to its electronic structure, creating slightly less efficient packing despite the overall increase in density. These observations highlight the complex interplay between atomic mass, ionic size, electronic configuration, and network-forming behavior in determining the physical parameters of multicomponent glass systems.

The research focused on examining the physical properties of glass samples made from modified barium borate, which were infused with different concentrations of bismuth and vanadium oxide. The findings suggest that substituting V2O5 with Bi2O3 and adjusting the Bi2O3 concentration within the range of 40-45 mol% significantly impacts the physical characteristics of the glass materials. Various patterns emerged from the investigation. Initially, there was a rise in the density of the glasses, escalating from 5.53 g/cm3 to 5.8 g/cm3. Furthermore, the molar volume (Vm) of the glasses exhibited an increase, climbing from 43.86 cm3/mol to 44.22 cm3/mol.

The packing density (Pd) experiences a slight decrease from 0.5 to 0.48 as the amount of Bi2O3 increases, suggesting a less dense structure. The free volume (Vf) rises from 22.06 to 22.88, indicating a growth in the unoccupied space between structural components. The average molecular weight (MAv) climbs from 242.35g to 256.56g, showcasing the higher molecular weight of Bi2O3 in comparison to V2O5 [29,35]. The number of ions (N) increases from 5.49 x 1021 ions to 6.13 x 1021 ions, suggesting a higher density of charge carriers. The radius of the electrode (rp) decreases from 2.28 Å to 2.20 Å, indicating that there is minimal polar conduction as the Bi2O3 content increases. Furthermore, the internuclear distance (ri) decreases from 5.67 Å to 5.47 Å, suggesting a more compact structure with shorter distances between ions.

The addition of Bi2O3 leads to an increase in the field strength (F) from 89.00 × 101⁷ g mol⁻1 cm⁻2 to 96.00 × 101⁷ g mol⁻1 cm⁻2, indicating a stronger ionic bond character. This suggests that the presence of Bi2O3 enhances the bonding within the system. Furthermore, the molar volume of boron (Vb) atoms experiences a slight increase from 31.33 to 31.58, indicating a subtle alteration in the borate lattice structure. Despite these changes, the average boron-boron distance (dB-B) remains relatively constant, ranging from 3.73 × 10⁻⁸ nm to 3.74 × 10⁻⁸ nm across different glass compositions. These observed variations in physical properties can be attributed to the structural modifications that occur due to the replacement of V2O5 with Bi2O3 [36].

3.3. Radiation shielding features

3.3.1. MAC

The determination of the probability of interaction between incident photons and a particular mass of material per unit area can be achieved through the measurement of the MAC [37]. In this study, the MAC values of the prepared glasses were analyzed across a spectrum of photon energies, ranging from 0.015 to 15 MeV. Figure 6 provides a clear representation of the changes in MAC values for glass samples Bi40V5, Bi41V4, Bi42V3, Bi43V2, Bi44V1, and Bi45V0 at different incident photon energies. The dominance of the photoelectric effect (PE) is evident at lower photon energies around 15 keV. Notably, the highest MAC values within this energy range fall between 86.6 cm2/g and 90.4 cm2/g. However, there is a significant decrease in the MAC values of the glass samples as the photon energy increases up to 0.08 MeV.

Variation of MAC vs. photon energy for studied glasses.
Figure 6.
Variation of MAC vs. photon energy for studied glasses.

However, as the energy of the photon surpasses 0.8 MeV, the trend of the MAC for all the glass samples becomes more consistent, indicating a reduced reliance on energy. Compton scattering (CS) can be employed to characterize this phenomenon, which is particularly noticeable at intermediate energies [38]. Subsequently, for energies greater than 1 MeV, the MAC values experience a slight increase, which can be attributed to the occurrence of pair production. The data presented in Figure 6 unequivocally demonstrates that the MAC values escalate with the augmentation of Bi2O3 concentration in B2O3-Li2O-BaO-Bi2O3-V2O5 glasses. This observation highlights that the glass sample Bi45V0 exhibits the highest MAC values and possesses the largest weight fraction of Bi, amounting to 0.679. Notably, the presence of PE in proximity to the Bi K-absorption edge induces a noticeable shift in the MAC trend around 0.1 MeV.

3.3.2. LAC

The investigation delved into the examination of the LAC (µ) values of prepared glasses, which were influenced by the energy levels of photons ranging from 0.015 to 15 MeV. The relationship between the incident γ-ray energy and the corresponding µ of the prepared samples has been visually depicted in Figure 7. The findings revealed that the simulated LAC exhibited a slight variance due to the subtle disparities in the chemical structure of the selected glass samples. Notably, the range of 478.22 cm-1 to 524.54 cm-1 emerged as the interval with the highest values of the LAC findings, underscoring its significance in terms of attenuation.

Variation of LAC vs. photon energy for the studied glasses.
Figure 7.
Variation of LAC vs. photon energy for the studied glasses.

It is apparent from the analysis of each sample that the LAC gradually increases as the Bi2O3 ratio increases. The relationship between the LAC and the incident radiation energy has also been depicted in Figure 7. The LAC experiences a rapid decline with increasing energy after reaching its maximum values at a low gamma ray energy of 0.015 MeV for all glass samples, primarily due to the photoelectric absorption cross section [39]. At 0.1 MeV, the K absorption edges result in a sudden peak in the LAC [40]. Within the energy interval spanning from 0.15 to 2 MeV, the LAC steadily decreases due to the CS cross section. For energies exceeding 2 MeV, the pair production cross section leads to a semi-constant LAC for all glass samples.

3.3.3. HVL, MFP

The HVL refers to the thickness of shielding materials that can reduce the incident activity of a source by half. In Figure 8, the relationship between HVL and photon energy is illustrated. The HVL values of the prepared samples demonstrate an upward trend as the photon energy increases, peaking at 3.21 cm at 5 MeV. Beyond this point, the HVL values experience a slight decrease. This rise in HVL is inversely correlated with the values of LAC. Among the samples, Bi45V0 exhibits the highest efficiency in shielding. As the concentration of Bi2O3 increases, the HVL decreases, indicating an improvement in the shielding capacity of the synthesized samples. Figure 9 presents the MFP, which is another factor influencing gamma transmission. The MFP [41] represents the distance between a photon and its subsequent interactions with glass. The behavior of MFP is similar to that of HVL. In other words, glass allows for the passage of a greater number of photons with higher energy. As the level of Bi2O3 increases, the MFP value decreases. The study reveals that Bi45V0 glass possesses a lower MFP value.

Variation of HVL with photon energy for the studied glasses.
Figure 8.
Variation of HVL with photon energy for the studied glasses.
Variation of MFP with photon energy for the studied glasses.
Figure 9.
Variation of MFP with photon energy for the studied glasses.

3.3.4. ACSs, ECSs

The intensity of the photon beam during transmission, specifically in processes involving photons and absorber atoms, can be represented by the cross section of each atom. The ACS σa and ECS σe have been illustrated in Figures 10 and 11, respectively. The atomic cross-section σa demonstrates an increase in regions of low energy as the Bi2O3 content rises, followed by a decrease as the photon energy increases. This behavior is attributed to coherent scattering, which makes the ACS σa sensitive to interference effects at low energy. In contrast, as the energy increases, the ECS σe shows a decline in all glass samples [42]. The values of σa are determined by both the photon energy employed and the chemical composition of the substance [43].

Variation of σ a versus photon energy for the studied glasses.
Figure 10.
Variation of σ a versus photon energy for the studied glasses.
Variation of σ e versus photon energy for the studied glasses.
Figure 11.
Variation of σ e versus photon energy for the studied glasses.

3.3.5. Effective atomic number (Zeff)

The effective atomic number (Zeff) was determined through the calculation of electric and ACSs of the glass samples, as illustrated in Figure 12. At lower energies, the Zeff values for the six glass samples exhibit a tendency to be higher. In the energy interval from 0.015 to 0.08 MeV, the effective atomic number (Zeff) of the glass samples experiences a rapid decline as the incident energy increases, primarily as a result of the photoelectric process [44-46]. Upon reaching 0.1 MeV, there is a sudden increase in Zeff, peaking at 76.33 for Bi40V5, which can be attributed to the K-absorption edges of Bismuth. Within the intermediate energy range (0.2 < E < 3 MeV), Zeff gradually decreases with the rise in incident energy, mainly due to CS. The Zeff reaches its minimum values at an energy of 1.5 MeV (25.81 for Bi45V0). At higher gamma ray energies (E > 3 MeV), the Zeff experiences a slow increase with the incident energy, primarily due to the pair production effect.

Zeff vs. photon energy for the studied glasses. σ e
Figure 12.
Zeff vs. photon energy for the studied glasses. σ e

4. Conclusions

In conclusion, this study presents a novel barium borate glass system co-doped with Bi₂O₃ and V₂O₅, demonstrating significant advancements in optical and structural properties through a carefully engineered dual-doping approach. The key findings reveal a controllable reduction in optical band gap (from 3.25 eV to 2.45 eV) with increasing V₂O₅ content, accompanied by a rise in Urbach energy (0.32-0.58 eV), indicating enhanced polarizability and structural disorder. FTIR analysis confirmed modifications in the borate network while identifying characteristic vibrations, including hydroxyl groups and overtone modes in the 2000-4000 cm⁻1 region. These tailored properties suggest promising applications in near-infrared optical devices, radiation shielding materials, and potential solid-state electrolytes, supported by the system’s unique combination of heavy metal oxide incorporation and borate glass matrix advantages. The work provides both fundamental insights into glass structure-property relationships and a practical foundation for developing advanced functional materials, with future research directions including thermal stability optimization and device integration studies to fully exploit these materials’ potential.

Acknowledgment

This work was supported and funded by the Deanship of Scientific Research at Imam Mohammad Ibn Saud Islamic University (IMSIU) (grant number IMSIU-DDRSP2503).

CRediT authorship contribution statement

Oumr Adnan Osra, Abdulrhman M. Alsharari: Data curation, formal analysis, methodology, and software; S. A. Al-Ghamdi, Mohammed D. Sharahili: Investigation and writing – review & editing; Kamelah S. Alrashdi, Tarek A. Yousef: formal analysis, investigation, writing-original draft. Hela Ferjani, Nashwa M. El-Metwaly: Supervision and administration of research group.

Declaration of competing interest

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

All relevant data are within the manuscript and available from the corresponding author upon request.

Declaration of Generative AI and AI-assisted technologies in the writing process

The authors confirm that there was no use of artificial intelligence (AI)-assisted technology for assisting in the writing or editing of the manuscript and no images were manipulated using AI.

References

  1. , , , , . An investigation of reactivity effect due to inadvertent filling of the irradiation channels with water in NIRR-1 Nigeria Research Reactor-1. Applied Radiation and Isotopes: Including Data, Instrumentation and Methods for use in Agriculture. Industry and Medicine. 2017;123:11-16. https://doi.org/10.1016/j.apradiso.2017.02.008
    [Google Scholar]
  2. , . Evaluation of arising exposure of ionizing radiation from computed tomography and the associated health concerns. Journal of Radiation Research and Applied Sciences. 2020;13:295-300. https://doi.org/10.1080/16878507.2020.1728962
    [Google Scholar]
  3. , , , , , . Characterization of Bi2O3ZnO B2O3 and TeO2ZnO CdO Li2O V2O5 glass systems for shielding gamma radiation using MCNP5 and Geant4 codes. Journal of Physics and Chemistry of Solids. 2019;126:112-123. https://doi.org/10.1016/j.jpcs.2018.10.034
    [Google Scholar]
  4. , , , , , , . Experimental and FLUKA evaluation on structure and optical properties and γ-radiation shielding capacity of bismuth borophosphate glasses. Progress in Nuclear Energy. 2022;148:104219. https://doi.org/10.1016/j.pnucene.2022.104219
    [Google Scholar]
  5. , , , , , , . Spectral, electrical, magnetic and radiation shielding studies of Mg-doped Ni–Cu–Zn nanoferrites. Journal of Materials Science: Materials in Electronics. 2020;31:20210-20222. https://doi.org/10.1007/s10854-020-04541-x
    [Google Scholar]
  6. , , , , . Influence of PbO2 and Gd2O3 on the gamma-ray shielding performance of borosilicate glasses. Nexus of Future Materials. 2025;2:617892. https://doi.org/10.70128/617892
    [Google Scholar]
  7. , , , . Comparative half value layer study of novel PbO-B2O3-CuO-CaO glasses with previous reports. Nexus of Future Materials. 2024;1:126-130. https://doi.org/10.70128/585024
    [Google Scholar]
  8. , , . Exploring transmission factor in high-density glasses: The effects of ZnO and Bi2O3 concentrations. Nexus of Future Materials. 2024;1:585023. https://doi.org/10.70128/585023
    [Google Scholar]
  9. . From mineral to high-value shielding material: Conversion of ludwigite into polyimide resin-based composites for medical X-rays protection. Nexus of Future Materials. 2025;2:610992. https://doi.org/10.70128/611002
    [Google Scholar]
  10. , , , . Comparative study of silicate glasses containing Bi2O3, PbO and BaO: Radiation shielding and optical properties. Annals of Nuclear Energy. 2011;38:1438-1441. https://doi.org/10.1016/j.anucene.2011.01.031
    [Google Scholar]
  11. , , , . Shielding properties of lead and barium phosphate glasses. Radiation Physics and Chemistry. 2012;81:1568-1571. https://doi.org/10.1016/j.radphyschem.2012.05.002
    [Google Scholar]
  12. . Optimization of the radiation shielding capabilities of bismuth-borate glasses using the genetic algorithm. Materials Chemistry and Physics. 2019;224:238-245. https://doi.org/10.1016/j.matchemphys.2018.12.022
    [Google Scholar]
  13. , , , , , . Optical and thermal investigations on vanadyl doped zinc lithium borate glasses. Journal of Asian Ceramic Societies. 2015;3:234. https://doi.org/10.1016/j.jascer.2015.03.004
    [Google Scholar]
  14. , , , . Structure and radiation shielding attitude of hexa-structured borosilicate glasses containing zinc oxide. Applied Physics A. 2025;131:401. https://doi.org/10.1007/s00339-025-08482-2
    [Google Scholar]
  15. , , , , , , . Progress in Nuclear Energy. 2020;118:103118.
  16. , , . Radiation Physics and Chemistry. 2019;163:58.
  17. , , . Review on transition metals containing lithium borate glasses properties, applications and perspectives. Journal of Materials Science. 2023;58:8678-8699. https://doi.org/10.1007/s10853-023-08567-4
    [Google Scholar]
  18. , , , , . Phy-X/PSD: Development of a user friendly online software for calculation of parameters relevant to radiation shielding and dosimetry. Radiation Physics and Chemistry. 2020;166:108496. https://doi.org/10.1016/j.radphyschem.2019.108496
    [Google Scholar]
  19. . Bismuth-doped glasses: A novel approach to efficient radiation attenuation. Nexus of Future Materials. 2025;2:593188. https://doi.org/10.70128/593188
    [Google Scholar]
  20. , , , . Impact of Bi2O3 modifier concentration on barium–zincborate glasses: Physical, structural, elastic, and radiation-shielding properties. The European Physical Journal Plus. 2021;136:116‏. https://doi.org/10.1140/epjp/s13360-020-01056-6
    [Google Scholar]
  21. , . Beer–Lambert law for optical tissue diagnostics: Current state of the art and the main limitations. Journal of Biomedical Optics. 2021;26:100901. https://doi.org/10.1117/1.jbo.26.10.100901
    [Google Scholar]
  22. , , . A comprehensive evaluation of structural, elastic, optical, and radiation shielding characteristics of barium borate glass containing vanadium ions using Phy-X/PSD software and empirical approaches. Journal of Electronic Materials. 2025;54:5993-6003. https://doi.org/10.1007/s11664-025-11992-7
    [Google Scholar]
  23. , , , , . Synthesis, physical, structural and shielding properties of newly developed B2O3–ZnO–PbO–Fe2O3 glasses using Geant4 code and WinXCOM program. Applied Physics A. 2019;125:523. https://doi.org/10.1007/s00339-019-2831-2
    [Google Scholar]
  24. , , , , , , , , , . Structural and gamma-ray attenuation of mixed former lead-free borophosphate glasses. Radiation Physics and Chemistry. 2024;214:111276. https://doi.org/10.1016/j.radphyschem.2023.111276
    [Google Scholar]
  25. , , , , . The structure, optical basicity, ligand field strength and shielding parameters of alkali/alkaline borate glasses doped with V2O5. Optical Materials. 2023;142:114078. https://doi.org/10.1016/j.optmat.2023.114078
    [Google Scholar]
  26. , , , . Medical radiation shielding in terms of effective atomic numbers and electron densities of some glasses. Radiation Physics and Chemistry. 2023;206:110767. https://doi.org/10.1016/j.radphyschem.2023.110767
    [Google Scholar]
  27. , . Structure of bismuth-borate glasses with ro-group oxides according to irspectroscopy. Glass and Ceramics. 2015;72:3-4. https://doi.org/10.1007/s10717-015-9728-0
    [Google Scholar]
  28. . Tellurite glasses. Materials Chemistry and Physics. 1999;60:103-131. https://doi.org/10.1016/s0254-0584(99)00082-6
    [Google Scholar]
  29. , , , . In situ fabrication of the Bi2O3–V2O5 hybrid embedded with graphitic carbon nitride nanosheets: Oxygen vacancies mediated enhanced visible-light–driven photocatalytic degradation of organic pollutants and hydrogen evolution. Applied Surface Science. 2018;447:740-756. https://doi.org/10.1016/j.apsusc.2018.04.040
    [Google Scholar]
  30. , , , , . Investigations on structural and optical properties of various modifier oxides (MO = ZnO, CdO, BaO, and PbO) Containing bismuth borate lithium glasses. Journal of Composites Science. 2021;5:308. https://doi.org/10.3390/jcs5120308
    [Google Scholar]
  31. , , , , . Infrared studies of the structure of borate glasses. Materials Science and Engineering. 1989;3:307-312. https://doi.org/10.1016/0921-5107(89)90026-3
    [Google Scholar]
  32. , , , . Structural, optical and antibacterial activity studies on CMC/PVA blend filled with three different types of green synthesized ZnO nanoparticles. Journal of Inorganic and Organometallic Polymers and Materials. 2023;33:1855-1867. https://doi.org/10.1007/s10904-023-02622-y
    [Google Scholar]
  33. , , , , . Investigation of non-bridging oxygen formation and structural evolution in SrO-doped borosilicate glasses. Applied Physics A. 2025;131:1-10. https://doi.org/10.1007/s00339-025-08341-0
    [Google Scholar]
  34. , , , , . Optical parameters and shielding attitude of sodium fluoride in calcium-borate glasses. Optical and Quantum Electronics. 2025;57:106. https://doi.org/10.1007/s11082-024-07955-7
    [Google Scholar]
  35. , . Optical parameters, antibacterial characteristics and structure correlation of copper ions in cadmium borate glasses. Journal of Materials Research and Technology. 2020;9:10491-10497. https://doi.org/10.1016/j.jmrt.2020.07.057
    [Google Scholar]
  36. , , . Structural, optical properties, and laser spectroscopy of erbium-doped low-melting borate glasses. Ceramics International. 2024;50:26528-26538. https://doi.org/10.1016/j.ceramint.2024.04.381
    [Google Scholar]
  37. . Introduction to glass science and technology. Royal society of chemistry. This book provides a comprehensive overview of glass science and technology, including the structure, properties, and applications of various glass systems. P005-P006. 2005 https://doi.org/10.1039/9781847551160-FP005
    [Google Scholar]
  38. , , , , , . Physical, structural, optical and gamma ray shielding behavior of (20+x) PbO – 10 BaO – 10 Na2O – 10 MgO – (50-x) B2O3 glasses. Physica B: Condensed Matter. 2019;552:110-118. https://doi.org/10.1016/j.physb.2018.10.001
    [Google Scholar]
  39. , , , , . Phy-X/PSD: Development of a user friendly online software for calculation of parameters relevant to radiation shielding and dosimetry. Radiation Physics and Chemistry. 2020;166:108496. https://doi.org/10.1016/j.radphyschem.2019.108496
    [Google Scholar]
  40. , , , , , , . On tungsten barium phosphate glasses: Elastic moduli, gamma-ray shielding properties as well as transmission factor (TF) Journal of the Australian Ceramic Society. 2023;59:1095-1109. https://doi.org/10.1007/s41779-023-00900-z
    [Google Scholar]
  41. , , . Comparative studies between the shielding parameters of concretes with different additive aggregates using MCNP-5 simulation code. Radiation Physics and Chemistry. 2019;165:108426. https://doi.org/10.1016/j.radphyschem.2019.108426
    [Google Scholar]
  42. , , , , . Structural, UV and shielding properties of ZBPC glasses. Journal of Non-Crystalline Solids. 2019;509:99-105. https://doi.org/10.1016/j.jnoncrysol.2018.12.013
    [Google Scholar]
  43. , , , , . Investigation of bismuth borate glass system modified with barium for structural and gamma-ray shielding properties. Spectrochimica Acta. Part A, Molecular and Biomolecular Spectroscopy. 2019;206:367-377. https://doi.org/10.1016/j.saa.2018.08.038
    [Google Scholar]
  44. , , . The Advancing of zinc oxide nanoparticles for biomedical applications. Bioinorganic Chemistry and Applications. 2018;2018:1062562. https://doi.org/10.1155/2018/1062562
    [Google Scholar]
  45. , , , , . Experimental evidence of molecular coherence effects in the bremsstrahlung radiation processes. Journal of Physics B: Atomic, Molecular and Optical Physics. 2019;52:145201. https://doi.org/10.1088/1361-6455/ab22f1
    [Google Scholar]
  46. , . Investigation of gamma-ray shielding capability of glasses doped with Y, Gd, Nd, Pr and Dy rare earth using MCNP-5 code. Physica B: Condensed Matter. 2020;577:411756. https://doi.org/10.1016/j.physb.2019.411756
    [Google Scholar]
Show Sections