Next Article in Journal
Influence of Reinforcing Agents on the Mechanical Properties of Denture Base Resin: A Systematic Review
Next Article in Special Issue
Enhanced Efficiency of Dye-Sensitized Solar Cells Based on Polymer-Assisted Dispersion of Platinum Nanoparticles/Carbon Nanotubes Nanohybrid Films as FTO-Free Counter Electrodes
Previous Article in Journal
Jawbones Scaffold Constructed by TGF-β1 and BMP-2 Loaded Chitosan Microsphere Combining with Alg/HA/ICol for Osteogenic-Induced Differentiation
Previous Article in Special Issue
Enhancing Thermo-Mechanical Properties of Epoxy Composites Using Fumed Silica with Different Surface Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

LDPE/Bismuth Oxide Nanocomposite: Preparation, Characterization and Application in X-ray Shielding

Nuclear Science Research Institute, King Abdulaziz City for Science and Technology, Riyadh 11442, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Polymers 2021, 13(18), 3081; https://doi.org/10.3390/polym13183081
Submission received: 24 August 2021 / Revised: 9 September 2021 / Accepted: 10 September 2021 / Published: 13 September 2021
(This article belongs to the Special Issue Advances in Multifunctional Polymer-Based Nanocomposites)

Abstract

:
Recently developed polymer-based composites could prove useful in many applications such as in radiation shielding. In this work, the potential of a bismuth oxide (Bi2O3) nanofiller based on an LDPE polymer was developed as lead-free X-ray radiation shielding offering the benefits of lightness, low-cost and non-toxic compared to pure lead. Three different LDPE-based composites were prepared with varying weight percentages of Bi2O3: 5%, 10% and 15%. The characterizations were extended to include structural properties, physical features, mechanical and thermal properties, and radiation shielding efficiency for the prepared nanocomposites. The results revealed that the incorporation of the Bi2O3 nanofiller into an LDPE improved the density of the composites. There was also a slight increase in the tensile strength and tensile modulus. In addition, there was a clear improvement in the efficiency of the shield when fillers were added to the LDPE polymer. The LDPE + Bi2O3 (15%) composite needed the lowest thickness to attenuate 50% of the incident X-rays. The LDPE + Bi2O3 (15%) polymer can also block around 80% of X-rays at 47.9 keV. In real practice, a thicker shield of the proposed composite materials, or a higher percentage of the filler could be employed to safely ensure the radiation is blocked.

Graphical Abstract

1. Introduction

The overall unique properties of various composite materials have recently attracted a wide range of scientific research regarding radiation protection due to the materials’ low weight, low manufacturing cost, mechanical strength, flexibility, and chemical stability [1]. Lead (Pb) has traditionally been employed as a shield since it has a high density and high efficiency in form of sheets, plates, foils, laminates, bricks, and blocks [2,3]. Nevertheless, the use of Pb is limited and it is unsuitable for some specific applications that require low cost, flexibility, chemical stability, mechanical strength, and lightness [4]. Polymer composites are an attractive option for radiation shielding because of their many advantages including being environmentally friendly, light, non-toxic, and flexible [5]. Consequently, diverse researchers have used various types of polymers as a matrix and have included fillers that provide reinforcement, depending on their overall application. The most used composites in X- and γ-rays shields are polymeric materials reinforced by metal and metal oxide [6,7,8,9]. As noted in the literature, metal–polymer composites generally combine two dissimilar components forming a light lattice together with Z particles [10]. The increased shielding properties of the metal–polymer are because of the uniform distribution of metal and metal oxide particles within the defined matrix [11].
Investigators have focused and reported several polymer matrices that can be used as X- and γ-ray shields. These polymer matrices include bismuth oxide (Bi2O3) filled poly (methyl methacrylate) composites, a high-density polyethylene (HDPE) composite that is loaded with tungsten (W), molybdenum sulfide (MoS2), boron carbide (B4C), micro-and nanosized tungsten oxide (WO3) dispersed emulsion polyvinyl chloride (EPVC) polymer composites, lead oxide filled isophthalic resin polymer composites, silicone rubber composites that contain bismuth content, polymer bricks, polyester composites that are reinforced with zinc, composites of HDPE with zinc oxide, lead oxide, and cadmium oxide [12,13,14,15,16,17,18,19,20,21]. The research attention has increasingly turned toward the overall effects of the presence of nanoparticles within the shielding materials, because of the novel uses of the materials [22]. The quantum effects and increase in the surface-to-volume ratio factor of nanoparticles were reported as the main factors that bring about dissimilar behaviors of the nano and microparticles. The outlined parameters affect the mechanical, thermal and shielding properties of the materials in some cases [23]. Because of induced agglomeration, the dispersion of nanoparticles into the overall polymer matrix is known to be more challenging than the dispersion of micro-sized particles [24]. Accordingly, the prominent features known to nano-scale particles than microparticles are known to cause an impressive increase in the resultant attenuation coefficient. As asserted in various studies, combining two phases of matrix and reinforcement is more complex and diverse than each individual phase [25].
Many researchers have performed studies on Pb-free composite shields. For example, one study designed and fabricated composites that consisted of HDPE mixed with micro-sized and nano-sized cadmium oxide particles for attenuation of photon beams with energy that ranged from 59.53 keV up to 1408.01 keV [19]. It was established that the nanoscale reinforced HDPE enhanced the overall shielding properties, particularly at lower photon energies. Another study was conducted on a new polyvinyl alcohol (PVA)/WO3 composite based on the Monte Carlo N-Particle (MCNP) simulation code [26]. It was noted that the PVA/WO3 composite could be considered as a shield for photon energy at the levels of 662, 778, 964, 1112, 1170, 1130, and 1407 keV. In addition, Atashi et al. conducted some fabrication of flexible silicone rubber/W/Bi2O3 using the technique of open mold cast [5]. The final composites possessed higher attenuation coefficients. Furthermore, it was established that increasing Bi2O3 in composites decreases the agglomeration of fillers. However, the composite of PVA containing 0–40 wt% filler loading (Bi2O3 and WO3) used in X-ray shielding is reported to play an important role in determining the density and the thickness of the composite sample [9].
Varying the quantity of tungsten nanopowder yielded improvement in thermophysical, radiation shielding and mechanical properties by [27]. In addition, concentration ranges of 30–70 wt% Bi2O3 were dispersed in carboxylated nitrile butadiene rubber latex films [28]. The films effectively attenuated low-energy photon beams. The overall effect of density on the composites’ suitability as radiation shield was studied using X-rays [29]. The attenuation performance against radiation was studied by varying the amounts of powdered fillers lead oxide and WO3 added to low-density polyethylene (LDPE). The results of the study indicated that samples with higher filler loads showed good attenuation performance against radiation [30].
In the current study, LDPE resin was chosen as a matrix duet to its superior mechanical and chemical properties. Bi2O3 was chosen for an alternative X-ray protective filler since it has key potential properties such as high density, high melting point, low conductivity, and is available in fine powder form. Therefore, the primary experiments aimed at furthering the understanding of LDPE nanocomposites. LDPE + Bi2O3 composites with different loadings of nano Bi2O3 were prepared via melt compounding. The study then conducted a study of X-ray attenuation, structure, mechanical, and thermal characteristics of Bi2O3 filled polymer composites. The results from this work would not only offer exciting possibilities for new, effective and safe X-ray protective LDPE nanocomposites but also highlight the potential of these new composite materials to be further developed for radiation shielding applications.

2. Materials and Methods

2.1. Materials

A commercial sample of LDPE (density: 0.93 g cm−3) was obtained from the Saudi Arabian Basic Industries Corporation (SABIC),(Riyadh, Saudi Arabia), under the trade name LDPE HP 0722N. Bismuth oxide nano powder was obtained from (Alfa Aesar, Kandel, Germany) with particle size smaller than 100 nm.

2.2. Nanocomposite Preparation

The Xplore conical twin-screw extruder was used to prepare the following samples: Pure LDPE, 5 wt%, 10 wt% and 15 wt% of Bi2O3 for LDPE + Bi2O3 nanocomposites. Firstly, an electrical balance (Sartorius Analytical, Karlsruhe, Germany) weighted LDPE and Bi2O3 with an accuracy ± 0.0001 g. Subsequently, 20 g of LDPE with a specified weight percentage of Bi2O3 was fed into a mini twin-screw extruder at 170 °C, for 10 min. The rotator speed was set to 100 rounds per minute (rpm) to ensure a uniformly mixed composite.
The fully mixed sample was then put into a stainless-steel frame (100 × 100 × 1 mm3) to be hot presses between two layers. The pressing was carried out by using a hydraulic press preheated at 170 °C for 3 min. The pressure was then gradually raised to 100 kN for another 10 min. The sample was left in the press for 20 min to cool down to room temperature. Finally, the resulting sheet was taken out of the mold and cut into circular samples of 25 mm in diameter to perform radiation-shielding tests and into dumbbell shape for tensile testing according to ASTM D638.

3. Characterization

3.1. X-ray Differentiation

X-ray diffraction (XRD) analysis was carried out using JOEL instruments (Tokyo, Japan) using Cu Kα radiation (λ = 0.1540 nm, a tube operated at 40 kV and a filament current of 40 mA, the Bragg’s angle (2θ) was in the range of 5°–80° using 0.01 step with and 1 s time count). All data were recorded and analyzed using the machine software. The crystal size was calculated by the Scherer equation:
D = K   λ β   cos   θ
where D is the crystallite size (nm), K is the crystallite shape factor (K = 0.9), λ is the X-ray wavelength of Cu (equal =0.154 nm) and β is the full width half maximum (FWHM) of XRD diffraction peak in radians.

3.2. Scanning Electron Microscope (SEM)

The surface morphology and dispersion of the nanocomposite were studied by SEM using JEOL (Tokyo, Japan). The samples were prepared as follows: the specimen was stuck on a 10 mm diameter carbon tab attached to the top of the aluminum pin tube. Then, the sample was coated with gold using a sputter coater. The coating exposure was 2 min.

3.3. Density Measurements

The experimental and calculated densities were evaluated using the Archimedes method (water as an immersing medium) for the mixture. The relative density of pure Bi2O3 and LDPE were calculated to 8.9 and 0.93 g.cm−3 respectively using the Mettler Toledo XS204 instrument (Greifensee, Switzerland).

3.4. Tensile Testing

Tensile testing was conducted by a tensile machine (Instron 5982, Grove city, PA, USA) according to ASTM D638, (dumbbell samples cut from pressed sheets 1 mm thickness) at crosshead speed 50 mm min−1. The results are the average of at least five measurements. The tensile strength, Young’s modulus and elongation at the break of the composite were calculated.

3.5. Thermal Analysis by TGA

The thermal behavior of the sample was evaluated using a thermogravimetric analyzer (TGA 1, Perkin Elmer, Shelton, CT, USA). Each sample was heated from room temperature to 700 °C at a rate of 10 °C min−1.

4. Results and Discussion

4.1. Density Measurements

In the present study, density was used to represent the physical property of LDPE composites. A high-density metal oxide nanofiller was used to improve some properties of the respective thermoplastic composition and density. Table 1 illustrates the test results via the resultant density of LDPE composites taken as a function of Bi2O3 content. Considering the rule of mixture, the overall density of any given particulate-filled composite is known to be related to the density of its constituent particles. As noted in Table 1, the overall density of the composite increased as the amount loading of Bi2O3 increased. The increase can be linked to the density range of Bi2O3 which is 8.9 g cm−3: much higher than LDPE’s density of 0.93 g cm−3. The density range of LDPE composites was between 0.961–1.060 g cm−3. Ambika et al. in which the density of unsaturated polyester (UP) was found to increase the overall increase in the Bi2O3 filler content reported a similar phenomenon [31]. This was related to the fact that Bi2O3 has a higher density compared to UP. As explained by Pavlenko et al., increasing the Bi2O3 nanoparticle filler content within the polyimide (PI) composites results in an increase in the density of PI, since Bi2O3 is known to have a higher density compared to other formulation contents [32]. In the experiment, the highest reported density was 1.07 g cm−3 for an LDPE with 15 wt% of Bi2O3, yielding an overall increment of approximately 13% in density in comparison to pure LDPE. On the other hand, composites of an LDPE with 5 wt% Bi2O3content showed the lowest density, at about 0.961 g cm−3.

4.2. Thermal Stability

In this study, an overall determination of thermal characteristics was carried out between 25–700 °C. Figure 1 illustrates the TGA curves of pure LPDE and those of the composites containing up to 15 wt% of Bi2O3. Figure 1 affirms that pure LDPE is stable up to 459.1 °C without the general loss of mass on the respective TGA curve. Complete thermal decomposition of the LDPE occurs at 500 °C. The introduction of inorganic filler results in a significant increase in the overall thermal stability of the respective polymers [33,34,35]. This is exemplified clearly by LDPE + Bi2O3 (15%). In the end, an increase in the Bi2O3 content results in a decrease in the rate of mass loss for the composites. Nonetheless, a slight mass loss in the composites starts at 80–120 °C. This is related to the loss of sorbed-water and hydroxyl water that is contained in the Bi2O3 [36]. The data on the thermal stability of various composites are outlined in Table 2. It can be concluded from the data in Table 2 and Figure 1 that the overall thermal stability of the LDPE is generally less than that of LDPE + Bi2O3 nanocomposites. This is a clear confirmation of the fact that the thermal stability of various inorganic compounds is greater than those of polymers. Furthermore, the overall increase in thermal stability of such composites could be linked to an increase in the density of Bi2O3. In the experiment, the char yield of LDPE + Bi2O3 nanocomposite was found to be 4.6%, 9.6%, and 13.2% (respectively) higher than that of respective pure LDPE that was 0.3% at 600 °C. The reported values affirm that the dispersion of Bi2O3 into the polymer was indeed great thus closer to the initial percentage of weight added. Table 3 outlines the burning characteristic of LDPE that contained various loadings of Bi2O3. The present study assessed whether the overall presence of metal oxide in relation to the composition was consistent with the outlined theoretical quantity of Bi2O3. Considering the range of experimental error, the estimated amount of metal oxide is in agreement with the loading weight percentage of Bi2O3 to LDPE.

4.3. Mechanical Properties

Table 4 illustrates mechanical properties such as tensile strength, elongation at break and Young’s modulus of LDPE composites taken as a function of Bi2O3 content, as shown in Figure 2. From Table 4, it can be inferred that all mechanical properties of various composites decreased with the overall increase in the Bi2O3 filler content excluding LDPE + Bi2O3 (10%). This could be attributed to the fact that the low tensile strength of the LPDE compound being used in the experiment. The highest tensile strength of 15.51 MPa was achieved with the LDPE composite that had 10 wt% Bi2O3 nanofiller. However, an increase in Bi2O3 filler loading over 10 wt% reduced the tensile strength. This could be due to the poor adhesion that exists between filler particles and the requisite LDPE matrix which weakened the resultant interfacial zone between the polymer and the particles of Bi2O3. The outlined weak zone increased with an increase in filler content, thus decreasing the tensile strength and elongation at the break of the composite [37]. Young’s modulus was computed from the slope at zero percent within the tensile curve. The outlined trend was similarly observed within the LDPE + Bi2O3 (10%): there was a slight increase in Young’s modulus with the adding 10 wt% filler content. In essence, the incorporation of Bi2O3 nanoparticles can enhance the stiffness of LDPE but an additional increase in particle loadings did not lead to a substantial enhancement to the resultant Young’s modulus. Ideally, agglomerations take place at higher particle loading thus reducing the total surface area of the respective nanoparticles [38].

4.4. X-ray Diffraction

The XRD patterns for pure Bi2O3, pure LDPE, and several quantities of Bi2O3 with LDPE are displayed in Figure 3. The appearance of diffraction peaks/shoulders at 2θ = 27.80°, 31.73°, 32.70°, 47.20◦, 55.22°, 56.48°, 57.74°, and 76.47° is consistent with the dominant diffraction 2θ angles of the β-phase of tetragonal Bi2O3 crystal structure [4]. Figure 3 outlines numerous peaks that correspond to the orthorhombic unit cell of polyethylene. For these peaks, the lattice parameters include a=7.39 Å, b=4.93 Å, and c=2.52 Å. The unit cell structures of 110, and the reflection of 200 planes, have two main peaks at 21° and 24° respectively. The results that were obtained in this experiment confirm the results of the previous study [4]. On the other hand, two other peaks are known to be located at 30° and 36°, which correspond to reflection planes of 210 and 020, respectively [39,40]. What is more, other weak peaks are located ranging from 40° to 60° [40]. As such, the role of Bi2O3 with LDPE can be assumed affirmatively. In essence, the addition of Bi2O3seems under slight variation in its position relative to the main peaks at 21° and 24°. Even so, the accumulation of Bi2O3 affected the overall original LDPE structure significantly. There is a new peak at approximately 28° that is characteristic of a high quantity of bismuth concentration and is linked to the reflection plane 120. Table 5 outlines the values of crystallite size computed using Scherrer’s equation. From the calculations, the crystalize size of Bi2O3 within the LDPE matrix as computed by Scherrer’s equation ranges from 18.39 nm to 18.49 nm. A closer look at the results reveals no considerable difference in the respective crystallite size of the filler particles. The noted slight difference in the overall crystallite size can be attributed to the crystallite size effect.

4.5. Morphology

The surface morphology was determined by SEM. The morphology of various LDPE composites that have 0 wt%, 5 wt%, 10 wt% and 15 wt% Bi2O3 addition is shown in Figure 4 (Details regarding the density of Bi atoms in the nanocomposites according to the analysis of SEM-EDS images are shown in the Supplementary Materials, Figures S1–S4). There are no observable particles on the surface of the pure LDPE composite; nonetheless, in the experiment, there was an observed even distribution of particles that dispersed more densely with increased loading, especially in the case of LDPE composites that had 5 wt%, 10 wt% and 15 wt% added Bi2O3. The dispersion quality of the nanoparticles into the LDPE was observed to differ based on the overall concentration of the nanoparticle. It is possible that some particles could have settled down as the mixture settled because of the difference in chemical structure and physical properties of both the polymer and nanoparticles despite the mixing process being uniform. This could further be attributed to higher density and/or lack of interaction or bonding with various polymer pellets during the heating process [41]. Nevertheless, the composite that had a 15 wt% filler load showed the agglomeration of filler particles thus forming larger particles. This must be minimized for the composites to perform better.

5. Shielding Properties and Application

5.1. Shielding Efficiency of the Polymer Composites in Low Energies Applications

In Equation (2), the Lambert–Beer law is used to determine the attenuation characteristics of a shielding material by calculating the linear attenuation coefficient (µ) (cm−1) or mass attenuation coefficient (µm or μ/ρ) (cm2 g−1)
I I o = e ( μ ρ   ) ρ x
where I0 is the initial intensity and I is the remaining radiation intensity (I) after traversing a layer (x) of the composite material considering the density of the absorber ( ρ ) . The experiment evaluated µ and µm of the Bi2O3 composite samples based on the LDPE polymer and the impact of the concentration of the Bi2O3 materials on the polymers. We used an X-ray source and an ionization chamber detector configuration (Figure 5) in a sequence of irradiation. The discs sample holder was placed in between to estimate the intensity of the attenuated beams. Three filler concentrations of LDPE (5%, 10%, and 15%) were prepared in discs with a diameter of 2 cm each. The irradiation procedures were conducted using 1.80 cm2 collimator size and 30 s acquisition time. The µ values for the composite samples were determined using the intensity values with no disks and with discs for Io and I determination, respectively.
Table 6 shows the X-ray irradiation qualities used using a narrow beam spectrum condition allowing only primary photons to pass through the attenuating material to contribute to the detected signal [42]. Using SpekCalc software and for illustration, Figure 6 confirms the effective energy stated in ISO-4037 for the applied voltage of 120 kVp with filters provided in Table 6.
To validate our outcomes, the experimental values of the µm were obtained for the studied samples against energy (E) and were then compared to the calculated values (XCOM database) [43,44].
The attention efficiency of studied samples could be further examined by the half-value layer (HVL) and mean free path (MPF). The definition of these radiation protection quantities could be the thickness of the examined material needed to reduce the primary radiation to 50% and 36.8% respectively. They depend highly on energy and it can be said that efficient shielding has low HVL and MFP values.
H V L = 0.693 μ
M F P = 1 μ
The radiation protection efficiency (RPE) could be also used as an indication of the shielding ability of the composite samples by knowing the intensity values measured with and without samples [14]:
RPE = ( 1 I I o ) × 100       ( % )

5.2. Shielding Ability Investigation of the Polymer Composites

The shielding ability of composites based on the LDPE polymer was examined in attenuating 40.9–248 keV X-ray beams. The composite materials were developed, mixed with different percentages of Bi2O3 as filler, and eventually shaped into discs. The setup in Figure 5 used experimentally using narrow-spectrum X-ray beams to determine Io and I when the samples were placed, allowing to calculate the corresponding µ at different incident beam energies using Equation (2). Figure 7 shows the behavior of the measured µ against energy for the studied composites based on LDPE polymer prepared in different filler percentages of 5%, 10% and 15%. The µ values were observed to decrease with energy and the composite coded LB-15 containing 15% of the Bi2O3 possesses the highest attenuation across all energies of the incident beams compared to other composites. It is clear that the fillers have enhanced the efficiency compared to pure LDPE especially in low energy below 200 keV region where the predominant mechanism is photoelectric absorption which has a strong dependency on the incoming photon energy (1/E3) and the atomic number of the material (Z4.5) [45].
For verification purposes of the experimentally measured results, Table 7 illustrates the measured values of µm and the calculated results of µm extracted from the XCOM calculator based on the NIST database for the present composite samples. It was revealed that the measured results of µm and the calculated µm were close with an average error of 7.4%. The disagreements could result from experimental errors such as disc placement, the composites densities and the dispersion quality of the fillers in the matrix affecting the evaluation of µ values and eventually µm.
The attenuation ability of a shield could be illustrated using the HVL and MFP values in Figure 8a and Figure 8b. In this examination, the outcomes show that a thinner thickness of fillers is needed to attenuate the X-ray beams compared to the pure LDPE polymer. Particularly, the thickness of pure LDPE required to absorb 50% of the X-rays is almost four times the thickness of the LDPE + Bi2O3 (15%) composite.
The results of RPE in Table 8 show that around a 1.60 cm thick shield of the composite with 15% filler could reduce approximately 40% of the incident beam at 100 keV. Increasing the thickness of the samples would allow them to be a potential choice for X-ray shielding in diagnostic radiology departments where the energy is used commonly below 100 keV at hospitals using kilovoltage X-rays.
The material will have better attenuation ability if it has lower HVL and MFP values. Figure 8 clearly shows that the HVL and MFP are energy-dependent. Therefore, the HVL and MFP values of the composite sample with 15 wt% of Bi2O3 with the best performance against radiation were chosen for comparison purposes with other materials or composites in the literature at 100 keV. The comparison results in Table 9 showed the Bi2O3-LDPE composites possess low values of HVL and MFP meaning better attenuation ability among other materials.

6. Conclusions

This study aimed to characterize a potential Pb-free and light-weighted shielding material based on LDPE polymer. The different weight percentages of the components of Bi2O3 within the final products could be linked to the variation in densities of the composites. The LDPE + Bi2O3 (15%) that had 15 wt% had higher bismuth composition compared to its LDPE counterparts. This can be affirmed by the SEM images which showed that the interaction between the polymer and the respective bismuth oxide can result in homogenous distribution and dispersion. Analysis using XRD revealed that the crystallite size ranged from 18.39 nm to 18.49 nm as determined based on Scherrer’s equation. The resultant minor difference could be attributed to the crystallite size effect. In the TG curves of the entire composites, one-stage degradation is present. The first degradation temperature is 459 °C that shows thermal stability until 600 °C.
The attenuation efficiency of the proposed materials was tested in terms of µ, µm, HVL, and MFP by measuring the transmitted radiation through the proposed shields using a photon energy range up to 248 keV. Clear enhancement of the attenuation ability was observed when the percentage of filler to the polymer was increased specifically below 150 keV. A future direction of this study could be utilizing the proposed materials to determine their performance in higher energy ranges and neutrons. In addition, increasing the thickness of the proposed composite samples or increasing the percentage of the filler in the polymer could offer a potential shield for low-energy radiation application.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/polym13183081/s1, Figure S1: SEM-EDS images of pure LDPE, Figure S2: SEM-EDS images of LDPE composite containing a 5 wt% loading of Bi2O3, Figure S3: SEM-EDS images of LDPE composite containing a 10 wt% loading of Bi2O3, Figure S4: SEM-EDS images of LDPE composite containing a 10 wt% loading of Bi2O3.

Author Contributions

Conceptualization, M.A.(Mohammed Alsuhybani) and S.A.(Saad Alshahri); methodology, M.A.(Mohammed Alsuhybani), A.A.( Alhanouf Alrwais), S.A.(Salha Alotaibi) and M.A.(Mohammed Alsuhybani); Data analysis; M.A.(Mohammed Alsuhybani), E.A.(Eid Alosime), A.A. and S.A.(Salha Alotaibi); characterization; M.A.(Mohammed Alsuhybani), A.A. and S.A.(Salha Alotaibi); writing—original draft preparation, all authors; writing—review and editing, S.A.(Saad Alshahri), M.A.(Mohammed Alsuhybani), E.A.(Eid Alosime) and M.A.(Mansour Almurayshid); visualization, E.A (Eid Alosime), A.A. and S.A.(Salha Alotaibi); supervision, M.A.(Mohammed Alsuhybani) and S.A.(Saad Alshahri). All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The authors would like to thank King Abdulaziz City for Science and Technology for providing all the necessary lab facilities and equipment for this work. We are very much grateful to Alshammari for providing us access to the facilities in the polymer-processing lab. The authors are also grateful to the radiation calibration lab team for their help in measuring the shield performance, Fadhl Alfadhl for SEM imaging, Raed BinKhnin for TGA analysis and Abdulrahman Alshahrani <[email protected]> Alshahrani for mechanical testing.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Alavian, H.; Tavakoli-Anbaran, H. Comparative study of mass attenuation coefficients for LDPE/metal oxide composites by Monte Carlo simulations. Eur. Phys. J. Plus 2020, 135, 82. [Google Scholar] [CrossRef]
  2. Harish, V.; Nagaiah, N.; Kumar, H. Lead oxides filled isophthalic resin polymer composites for gamma radiation shielding applications. Indian J. Pure Appl. Phys. 2012, 50, 847–850. [Google Scholar]
  3. Martin, J.E. Physics for radiation protection: A handbook. John Wiley & Sons, Inc.: Ann Arbor, MI, USA, 2006; pp. 367–421. [Google Scholar]
  4. Ambika, M.; Nagaiah, N.; Harish, V.; Lokanath, N.; Sridhar, M.; Renukappa, N.; Suman, S. Preparation and characterisation of Isophthalic-Bi2O3 polymer composite gamma radiation shields. Radiat. Phys. Chem. 2017, 130, 351–358. [Google Scholar] [CrossRef]
  5. Atashi, P.; Rahmani, S.; Ahadi, B.; Rahmati, A. Efficient, flexible and lead-free composite based on room temperature vulcanizing silicone rubber/W/Bi2O3 for gamma ray shielding application. J. Mater. Sci.: Mater. Electron. 2018, 29, 12306–12322. [Google Scholar] [CrossRef]
  6. Nambiar, S.; Yeow, J.T. Polymer-composite materials for radiation protection. ACS Appl. Mater. Interfaces 2012, 4, 5717–5726. [Google Scholar] [CrossRef] [PubMed]
  7. Shik, N.A.; Gholamzadeh, L. X-ray shielding performance of the EPVC composites with micro-or nanoparticles of WO3, PbO or Bi2O3. Appl. Radiat. Isot. 2018, 139, 61–65. [Google Scholar] [CrossRef]
  8. Fontainha, C.; Baptista-Neto, A.; Faria, L. Polymer-based Nanocomposites of P(VDF-TrFE)/Bi2O3 Applied to X-ray Shielding. J. Mater. Sci. 2016, 4, 16–24. [Google Scholar]
  9. Jamil, M.; Hazlan, M.H.; Ramli, R.M.; Azman, N.Z.N. Study of electrospun PVA-based concentrations nanofibre filled with Bi2O3 or WO3as potential x-ray shielding material. Radiat. Phys. Chem. 2019, 156, 272–282. [Google Scholar] [CrossRef]
  10. Khan, W.S.; Hamadneh, N.N.; Khan, W.A. Polymer nanocomposites–synthesis techniques, classification and properties. In Science and Applications of Tailored Nanostructures; One Central Press (OCP): Cheshire, UK, 2016; pp. 50–67. [Google Scholar]
  11. Higgins, M.C.M.; Radcliffe, N.A.; Toro-González, M.; Rojas, J.V. Gamma ray attenuation of hafnium dioxide-and tungsten trioxide-epoxy resin composites. J. Radioanal. Nucl. Chem. 2019, 322, 707–716. [Google Scholar] [CrossRef]
  12. Plionis, A.; Garcia, S.; Gonzales, E.; Porterfield, D.; Peterson, D. Replacement of lead bricks with non-hazardous polymer-bismuth for low-energy gamma shielding. J. Radioanal. Nucl. Chem. 2009, 282, 239–242. [Google Scholar] [CrossRef]
  13. El-Fiki, S.; El Kameesy, S.; Nashar, D.; Abou-Leila, M.; El-Mansy, M.; Ahmed, M. Influence of bismuth contents on mechanical and gamma ray attenuation properties of silicone rubber composite. Int. J. Adv. Res. 2015, 3, 1035–1041. [Google Scholar]
  14. Aghaz, A.; Faghihi, R.; Mortazavi, S.; Haghparast, A.; Mehdizadeh, S.; Sina, S. Radiation attenuation properties of shields containing micro and Nano WO3 in diagnostic X-ray energy range. Int. J. Radiat. Res. 2016, 14, 127–131. [Google Scholar] [CrossRef]
  15. Mahmoud, M.E.; El-Khatib, A.M.; Badawi, M.S.; Rashad, A.R.; El-Sharkawy, R.M.; Thabet, A.A. Fabrication, characterization and gamma rays shielding properties of nano and micro lead oxide-dispersed-high density polyethylene composites. Radiat. Phys. Chem. 2018, 145, 160–173. [Google Scholar] [CrossRef]
  16. Mahmoud, M.E.; El-Khatib, A.M.; Badawi, M.S.; Rashad, A.R.; El-Sharkawy, R.M.; Thabet, A.A. Recycled high-density polyethylene plastics added with lead oxide nanoparticles as sustainable radiation shielding materials. J. Clean. Prod. 2018, 176, 276–287. [Google Scholar] [CrossRef]
  17. Afshar, M.; Morshedian, J.; Ahmadi, S. Radiation attenuation capability and flow characteristics of HDPE composite loaded with W, MoS2, and B4C. Polym. Compos. 2019, 40, 149–158. [Google Scholar] [CrossRef] [Green Version]
  18. Lsayed, Z.; Badawi, M.S.; Awad, R.; Thabet, A.A.; El-Khatib, A.M. Study of some γ-ray attenuation parameters for new shielding materials composed of nano ZnO blended with high density polyethylene. Nucl. Technol. Radiat. Prot. 2019, 34, 342–352. [Google Scholar] [CrossRef] [Green Version]
  19. El-Khatib, A.M.; Abbas, M.I.; Abd Elzaher, M.; Badawi, M.S.; Alabsy, M.T.; Alharshan, G.A.; Aloraini, D.A. Gamma attenuation coefficients of nano cadmium oxide/high density polyethylene composites. Sci. Rep. 2019, 9, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Cao, D.; Yang, G.; Bourham, M.; Moneghan, D. Gamma radiation shielding properties of poly (methyl methacrylate)/Bi2O3 composites. Nucl. Eng. Technol. 2020, 52, 2613–2619. [Google Scholar] [CrossRef]
  21. Kaçal, M.; Polat, H.; Oltulu, M.; Akman, F.; Agar, O.; Tekin, H. Gamma shielding and compressive strength analyses of polyester composites reinforced with zinc: An experiment, theoretical, and simulation based study. Appl.Phys. A. 2020, 126, 1–15. [Google Scholar] [CrossRef]
  22. Saboori, A.; Dadkhah, M.; Fino, P.; Pavese, M. An overview of metal matrix nanocomposites reinforced with graphene nanoplatelets; mechanical, electrical and thermophysical properties. Metals 2018, 8, 423. [Google Scholar] [CrossRef] [Green Version]
  23. Eid, G.A.; Kany, A.; El-Toony, M.; Bashter, I.; Gaber, F. Application of epoxy/Pb3O4 composite for gamma ray shielding. Arab. J. Nucl. Sci. Appl. 2013, 46, 226–233. [Google Scholar]
  24. Kim, J.; Seo, D.; Lee, B.C.; Seo, Y.S.; Miller, W.H. Nano-W Dispersed Gamma Radiation Shielding Materials. Adv. Eng. Mater. 2014, 16, 1083–1089. [Google Scholar] [CrossRef]
  25. Buzea, C.; Pacheco, I.I.; Robbie, K. Nanomaterials and nanoparticles: Sources and toxicity. Biointerphases 2007, 2, MR17–MR71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Kazemi, F.; Malekie, S. A Monte Carlo study on the shielding properties of a novel polyvinyl alcohol (PVA)/WO3 composite, against gamma rays, using the MCNPX code. J. Biomed. Phys. Eng. 2019, 9, 465. [Google Scholar] [CrossRef] [PubMed]
  27. Gavrish, V.; Baranov, G.; Chayka, T.; Derbasova, N.; Lvov, A.; Matsuk, Y. Tungsten nanoparticles influence on radiation protection properties of polymers. In IOP Conference Series: Materials Science and Engineering; IOP publishing: Tomsk, Russia, 2015; Volume 110, pp. 1–6. [Google Scholar]
  28. Liao, Y.-C.; Xu, D.-G.; Zhang, P.-C. Preparation and characterization of Bi2O3/XNBR flexible films for attenuating gamma rays. Nucl. Sci. Tech. 2018, 29, 1–12. [Google Scholar] [CrossRef]
  29. Azman, N.N.; Siddiqui, S.; Hart, R.; Low, I.-M. Effect of particle size, filler loadings and x-ray tube voltage on the transmitted x-ray transmission in tungsten oxide—epoxy composites. Appl. Radiat. Isot. 2013, 71, 62–67. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Belgin, E.E.; Aycik, G. Preparation and radiation attenuation performances of metal oxide filled polyethylene based composites for ionizing electromagnetic radiation shielding applications. J. Radiational Nucl. Chem. 2015, 306, 107–117. [Google Scholar] [CrossRef]
  31. Ambika, M.; Nagaiah, N.; Suman, S. Role of bismuth oxide as a reinforcer on gamma shielding ability of unsaturated polyester based polymer composites. J. Appl. Polym. Sci. 2017, 134, 1–7. [Google Scholar] [CrossRef]
  32. Pavlenko, V.; Cherkashina, N.; Yastrebinsky, R. Synthesis and radiation shielding properties of polyimide/Bi2O3 composites. Heliyon 2019, 5, e01703. [Google Scholar] [CrossRef] [Green Version]
  33. Cherkashina, N.; Pavlenko, A. Synthesis of polymer composite based on polyimide and Bi12SiO20 sillenite. Polym. Plast. Technol. Eng. 2018, 57, 1923–1931. [Google Scholar] [CrossRef]
  34. Roy, P.; Surekha, P.; Raman, R.; Rajagopal, C. Investigating the role of metal oxidation state on the degradation behaviour of LDPE. Polym. Degrad. Stab. 2009, 94, 1033–1039. [Google Scholar] [CrossRef]
  35. Gupta, R.; Badel, B.; Gupta, P.; Bucknall, D.G.; Flynn, D.; Pancholi, K. Flexible Low-Density Polyethylene–BaTiO3 Nanoparticle Composites for Monitoring Leakage Current in High-Tension Equipment. ACS Appl. Nano Mater. 2021, 4, 2413–2422. [Google Scholar] [CrossRef]
  36. Klinkova, L.; Nikolaichik, V.; Barkovskii, N.; Fedotov, V. Thermal stability of Bi2O3. Russ. J. Inorg. Chem. 2007, 52, 1822–1829. [Google Scholar] [CrossRef]
  37. Sheela, M.; Kamat, V.A.; Kiran, K.; Eshwarappa, K. Preparation and characterization of bismuth-filled high-density polyethylene composites for gamma-ray shielding. Radiat. Prot. Environ. 2019, 42, 180–186. [Google Scholar]
  38. Chee, C.Y.; Song, N.; Abdullah, L.C.; Choong, T.S.; Ibrahim, A.; Chantara, T. Characterization of mechanical properties: Low-density polyethylene nanocomposite using nanoalumina particle as filler. J. Nanomater. 2012, 2012, 215978. [Google Scholar] [CrossRef]
  39. Inci, B.; Wagener, K.B. Decreasing the alkyl branch frequency in precision polyethylene: Pushing the limits toward longer run lengths. J. Am. Chem. Soc. 2011, 133, 11872–11875. [Google Scholar] [CrossRef]
  40. Trujillo, M.; Arnal, M.; Müller, A.; Laredo, E.; Bredeau, S.; Bonduel, D.; Dubois, P. Thermal and morphological characterization of nanocomposites prepared by in-situ polymerization of high-density polyethylene on carbon nanotubes. Macromolecules 2007, 40, 6268–6276. [Google Scholar] [CrossRef]
  41. Sabri, J.H.; Mahdi, K.H. A Comparative Study for Micro and Nano shield of (PbO) composite for gamma Radiation. Energy Procedia 2019, 157, 802–814. [Google Scholar] [CrossRef]
  42. ISO 4037-1:2019. Radiological Protection—X and Gamma Reference Radiation for Calibrating Dosemeters and Doserate Meters and for Determining Their Response as a Function of Photon Energy—Part 1: Radiation Characteristics and Production Methods; ISO: Geneva, Switzerland, 2019; ISC, 17, 17.240. [Google Scholar]
  43. Berger, M.J.; Hubbell, J.H.; Seltzer, S.M.; Chang, J.; Coursey, J.S.; Sukumar, R.D.; Zucker, S.; Olsen, K. XCOM: Photon Cross Sections on a Personal Computer; NIST, PML, Radiation Physics Division: Washington, DC, USA, 1987; pp. 87–3597. [Google Scholar] [CrossRef]
  44. El-Sayed, A.; Ali, M.A.M.; Ismail, M.R. Natural fibre high-density polyethylene and lead oxide composites for radiation shielding. Radiat. Phys. Chem. 2003, 66, 185–195. [Google Scholar] [CrossRef]
  45. Almurayshid, M.; Alsagabi, S.; Alssalim, Y.; Alotaibi, Z.; Almsalam, R. Feasibility of polymer-based composite materials as radiation shield. Radiat. Phys. Chem. 2021, 183, 109425. [Google Scholar] [CrossRef]
  46. Alavian, H.; Tavakoli-Anbaran, H. Study on gamma shielding polymer composites reinforced with different sizes and proportions of tungsten particles using MCNP code. Prog. Nucl. Energy 2019, 115, 91–98. [Google Scholar] [CrossRef]
  47. Gurler, O.; Akar-Tarim, U. Determination of radiation shielding properties of some polymer and plastic materials against gamma-rays. Acta Phys. Pol. A 2016, 130, 236–238. [Google Scholar] [CrossRef]
Figure 1. TGA thermograms for pure LDPE and LDPE/ Bi2O3 composites.
Figure 1. TGA thermograms for pure LDPE and LDPE/ Bi2O3 composites.
Polymers 13 03081 g001
Figure 2. Effect of wt% of Bi2O3 on the mechanical properties of LDPE.
Figure 2. Effect of wt% of Bi2O3 on the mechanical properties of LDPE.
Polymers 13 03081 g002
Figure 3. XRD patterns of pure LDPE, Bi2O3 and its composites.
Figure 3. XRD patterns of pure LDPE, Bi2O3 and its composites.
Polymers 13 03081 g003
Figure 4. The surface morphology of LDPE nanocomposites (a) pure LDPE (b) 5 wt% Bi2O3 (c) 10 wt% Bi2O3 and (d) 15 wt% Bi2O3.
Figure 4. The surface morphology of LDPE nanocomposites (a) pure LDPE (b) 5 wt% Bi2O3 (c) 10 wt% Bi2O3 and (d) 15 wt% Bi2O3.
Polymers 13 03081 g004
Figure 5. The experimental setup used for shielding efficiency proposed composite materials.
Figure 5. The experimental setup used for shielding efficiency proposed composite materials.
Polymers 13 03081 g005
Figure 6. The distribution of the number of X-rays for 120 kVp extracted from SpekCalc with no filters and with ISO filter.
Figure 6. The distribution of the number of X-rays for 120 kVp extracted from SpekCalc with no filters and with ISO filter.
Polymers 13 03081 g006
Figure 7. The measured linear attenuation coefficient against incident beam energy.
Figure 7. The measured linear attenuation coefficient against incident beam energy.
Polymers 13 03081 g007
Figure 8. (a) HVL and (b) MFP values against energy for the composite samples prepared in this study.
Figure 8. (a) HVL and (b) MFP values against energy for the composite samples prepared in this study.
Polymers 13 03081 g008
Table 1. The density results of LDPE and its composites.
Table 1. The density results of LDPE and its composites.
Sample CodeComposition
(wt%)
Density (g cm−3)
(Experimental)
Density (g cm−3)
(Calculated)
Error (%)
Pure LDPEPure LDPE0.9260.9300.430
LB-5LDPE (95%) + Bi2O3 (5%)0.9610.9731.293
LB-10LDPE (90%) + Bi2O3 (10%)1.0101.0213.270
LB-15LDPE (85%) + Bi2O3 (15%)1.0601.0748.762
Table 2. Summary of the TGA thermograms for pure LDPE and LDPE + Bi2O3 composites.
Table 2. Summary of the TGA thermograms for pure LDPE and LDPE + Bi2O3 composites.
SampleBi2O3 Loading (wt%)Onset Temp (°C)High Peak Temp (°C)Weight Loss at 600 °C (%)
Pure LDPE04595000.3
LB-55459500.24.6
LB-1010461501.29.6
LB-1515454500.613.2
Table 3. Burn test of pure LDPE and LDPE + Bi2O3 composites.
Table 3. Burn test of pure LDPE and LDPE + Bi2O3 composites.
SampleBi2O3 Loading (wt%)Error (%)
The Theoretical ValuesThe Experimental Values
Pure LDPE000
LB-55.004.4910.2
LB-1010.09.168.37
LB-1515.014.53.39
Table 4. Results of mechanical properties of LDPE and its composites.
Table 4. Results of mechanical properties of LDPE and its composites.
SampleBi2O3 Loading
(wt%)
Tensile Strength
(MPa)
Elongation @ Break
(%)
Young’s Modulus
(MPa)
Pure LDPE014.77 ± 0.60378 ± 19355 ± 36
LB-5514.19 ± 1.29356 ± 18355 ± 13
LB-101015.51 ± 0.43374 ± 22378 ± 16
LB-151513.09 ± 0.98335 ± 48350 ± 14
Table 5. Crystallite size of the LDPE composites.
Table 5. Crystallite size of the LDPE composites.
SampleScherrer Crystallite
Size (nm)
FWHM
Pure LDPE219.520.8400
Pure Bi2O327.818.410.4400
LB-527.418.390.4400
LB-1027.418.390.4400
LB-1527.818.410.4400
Table 6. Narrow X-ray spectrum qualities used (Adopted from ISO-4037 [42]).
Table 6. Narrow X-ray spectrum qualities used (Adopted from ISO-4037 [42]).
Shortened NameTube Potential (kV)Effective Energy
(keV)
Additional Filtration Thickness
mm Pbmm Snmm Cumm Al
N-606047.9--0.6313.912
N-10010083.3--5.0273.920
N-120120100-1.0135.0273.950
N-150150118-2.605-3.903
N-2002001651.0283.0042.0323.901
N-2502502073.0992.062-3.925
N-3003002485.1523.016-3.929
Table 7. The experimental and calculated values of the mass attenuation coefficient.
Table 7. The experimental and calculated values of the mass attenuation coefficient.
Energy (keV)Pure LDPELB-5LB-10LB-15
Exp.Cal.Exp.Cal.Exp.Cal.Exp.Cal.
47.900.2094 ± 0.00350.19870.5589 ± 0.00410.57750.9164 ± 0.00620.95621.2369 ± 0.00771.3350
100.000.1671 ± 0.00240.16870.3560 ± 0.00300.40860.5061 ± 0.00300.64850.6696 ± 0.00320.8883
118.000.1535 ± 0.00220.16200.2619 ± 0.00270.31710.4067 ± 0.00280.47210.5520 ± 0.00290.6271
165.000.1520 ± 0.00170.14780.2010 ± 0.00190.21090.2745 ± 0.00250.27390.3287 ± 0.00240.3369
207.000.1432 ± 0.00110.13780.1768 ± 0.00170.17160.1824 ± 0.00250.20530.2515 ± 0.00230.2391
248.000.1382 ± 0.00200.12980.1403 ± 0.00140.15010.1579 ± 0.00220.17040.1849 ± 0.00200.1908
Table 8. The RPE values in % for the samples prepared in this study.
Table 8. The RPE values in % for the samples prepared in this study.
Energy (keV)Pure LDPELB-5LB-10LB-15
47.9017.1144.1776.5279.16
100.0014.0223.1746.9540.24
118.0013.9031.0255.0857.22
165.0012.0623.9047.4350.34
207.0012.7319.6635.2134.08
248.0014.3716.8425.0527.31
Table 9. Comparison of the measured HVL and MFP outcomes with other materials in the literature at 100 keV.
Table 9. Comparison of the measured HVL and MFP outcomes with other materials in the literature at 100 keV.
StudyCompositesHVL (cm)MFP (cm)
This studyLDPE 85% + Bi2O3 15%1.031.50
Almurayshid, et al. 2021 [45]HDPE 85% + W 15%1.181.63
Almurayshid, et al. 2021 [45]HDPE 85% + Mo 15%2.293.30
(Alavian and Tavakoli-Anbaran, 2019) [46]LDPE 99% + W 1%-4.27
(Gurler and Akar-Tarim, 2016) [47]Concrete (NBS)1.812.62
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Alshahri, S.; Alsuhybani, M.; Alosime, E.; Almurayshid, M.; Alrwais, A.; Alotaibi, S. LDPE/Bismuth Oxide Nanocomposite: Preparation, Characterization and Application in X-ray Shielding. Polymers 2021, 13, 3081. https://doi.org/10.3390/polym13183081

AMA Style

Alshahri S, Alsuhybani M, Alosime E, Almurayshid M, Alrwais A, Alotaibi S. LDPE/Bismuth Oxide Nanocomposite: Preparation, Characterization and Application in X-ray Shielding. Polymers. 2021; 13(18):3081. https://doi.org/10.3390/polym13183081

Chicago/Turabian Style

Alshahri, Saad, Mohammed Alsuhybani, Eid Alosime, Mansour Almurayshid, Alhanouf Alrwais, and Salha Alotaibi. 2021. "LDPE/Bismuth Oxide Nanocomposite: Preparation, Characterization and Application in X-ray Shielding" Polymers 13, no. 18: 3081. https://doi.org/10.3390/polym13183081

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