Next Article in Journal
Bridging the Gap: Assessing Material Performance of Laboratory Specimens and Concrete Structures
Next Article in Special Issue
Nanocluster Evolution in D9 Austenitic Steel under Neutron and Proton Irradiation
Previous Article in Journal
Physical and Thermal Characterizations of Newly Synthesized Liquid Crystals Based on Benzotrifluoride Moiety
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress in Gd-Containing Materials for Neutron Shielding Applications: A Review

1
School of Materials Science and Engineering, Shanghai Jiao Tong University, Shanghai 200240, China
2
State Key Laboratory of Metal Matrix Composites, Shanghai Jiao Tong University, Shanghai 200240, China
3
Innovation Academy for Microsatellites of Chinese Academy of Sciences, Shanghai 201203, China
4
School of Physics and Lectronic Information College, Huaibei Normal University, Huaibei 235000, China
5
Institute of Alumics Materials, Shanghai Jiao Tong University (Anhui), Huaibei 235000, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(12), 4305; https://doi.org/10.3390/ma16124305
Submission received: 10 May 2023 / Revised: 5 June 2023 / Accepted: 8 June 2023 / Published: 10 June 2023
(This article belongs to the Special Issue Advanced Characterization Techniques on Nuclear Fuels and Materials)

Abstract

:
With the rising demand for nuclear energy, the storage/transportation of radioactive nuclear by-products are critical safety issues for humans and the environment. These by-products are closely related to various nuclear radiations. In particular, neutron radiation requires specific protection by neutron shielding materials due to its high penetrating ability to cause irradiation damage. Herein, a basic overview of neutron shielding is presented. Since gadolinium (Gd) has the largest thermal neutron capture cross-section among various neutron absorbing elements, it is an ideal neutron absorber for shielding applications. In the last two decades, there have been many newly developed Gd-containing (i.e., inorganic nonmetallic-based, polymer-based, and metallic-based) shielding materials developed to attenuate and absorb the incident neutrons. On this basis, we present a comprehensive review of the design, processing methods, microstructure characteristics, mechanical properties, and neutron shielding performance of these materials in each category. Furthermore, current challenges for the development and application of shielding materials are discussed. Finally, the potential research directions are highlighted in this rapidly developing field.

1. Introduction

Over the past few decades, the increasing application of nuclear energy has brought a great deal of convenience to the world. Nuclear energy, as a sustainable and environmentally friendly energy source, plays a critical role in worldwide power generation. It provides around 10% of total global energy output while releasing less greenhouse gas emissions than many other energy sources [1]. Nuclear technology has effectively solved the contradiction between energy demand and environmental pollution. Despite these benefits, the application of nuclear energy is still restricted by a variety of significant challenges. For instance, nuclear power plants simultaneously generate a large amount of highly radioactive by-products (i.e., spent nuclear fuels (SNFs)) in addition to electricity [2]. Herein, SNFs refer to the irradiated nuclear fuel in a nuclear reactor, and they contain various types of radiation (i.e., alpha (α)-rays, beta (β)-rays, gamma (γ)-rays, and neutrons) [3,4,5]. Among these nuclear radiations, neutron radiation should receive priority attention due to its extremely high penetration and irradiation damage ability [4]. Generally, the relative penetrative ability of different radiations is given in Figure 1.
Exposure to neutron radiation can pose serious risks to human health, potentially causing cancer, leukemia, and even genetic mutations [6,7]. Furthermore, the impact of radiation on the natural environment is far-reaching, affecting not only wildlife but also soil and water quality [8]. Given the potential risks of neutron radiation exposure, promoting research and development of neuron shielding materials becomes necessary for protecting humans from the hazards of radiation damage.
Unlike shielding α-, β-, and γ-ray radiation, materials with high atomic numbers and high density are ineffective in neutron shielding. Since neutrons are subatomic particles with no charge and can interact only with the nucleus of target atoms, the likelihood of a specific type of interaction depends on two key factors: the neutron energy and the intrinsic properties of the target nuclide [9]. Neutrons are normally confined within the atomic nucleus. The liberation process necessitates surpassing the binding energy inherent to neutrons, which typically ranges between 7 and 9 MeV for most isotopes. Thus, these free neutrons are initially born with high velocities. According to neutron energy levels, there are typically slow neutrons (energies < 1 MeV), fast neutrons (energies > 1 MeV), and high-energy neutrons (energies > 20 MeV). In particular, thermal neutrons have an energy of 0.0253 eV, equivalent to a velocity of 2200 m/s, and are in thermalized equilibrium with the surrounding atoms at room temperature [10]. Moderation, a vital process, involves the reduction of the initial high speed and kinetic energy of free neutrons through elastic or inelastic scattering. These thermal neutrons are then absorbed by a target nucleus. Neutron absorption is a nuclear process in which an atomic nucleus collides and mixes with one or more neutrons to generate a new nucleus. In brief, the interaction between neutrons and matter is divided into two main parts, namely, moderation (elastic or inelastic scattering) and absorption [11]. It is worth noting that the absorption of fast/high-energy neutrons is nearly impossible due to their high energy and velocity. Comparably, slow neutrons (particularly thermal neutrons) can be easily captured by most neutron absorbers. A schematic diagram of the neutron shielding process is given in Figure 2. In the neutron shielding process, fast neutrons must be moderated/slowed down by elastic or inelastic scattering to become thermal neutrons, which is the process of thermalization or moderation. Subsequently, the thermalized neutrons can be captured by neutron absorbers.
Here, neutron absorption plays an important role in the whole shielding process. The probability that a nucleus will absorb a neutron is defined as the neutron capture cross-section (σ), which has the unit 10−24 cm2 or “barns” [12]. Neutron shielding materials are often composed of low-atomic-number elements (e.g., H), which have a strong neutron scattering cross-section and may effectively moderate or thermalize the incoming fast neutrons. In addition to this, neutron absorbing elements are generally integrated into neutron shielding materials to absorb the thermalized neutrons. The effectiveness of shielding materials in protecting against neutron radiation is primarily determined by neutron absorbing elements and their concentrations. Normally, a larger neutron capture cross-section corresponds to a higher neutron capture ability. Table 1 summarizes several typical neutron absorbing elements (boron (B), cadmium (Cd), samarium (Sm), and gadolinium (Gd)) with a large neutron capture cross-section.
Among these neutron absorbing elements, Cd is a good thermal neutron absorber, comprising various stable isotopes. 113Cd, one of the isotopes of Cd, has a large thermal neutron capture cross-section (σ = 20,600 barns). Previously, Cd was widely utilized in nuclear power reactors due to its favorable processability and low cost. Currently, its application has been constrained due to its high biological toxicity [14]. Additionally, 149Sm possesses a higher thermal neutron capture cross-section of 40,000 barns. Florez et al. [15] observed that adding Sm2O3 to Portland cement substantially boosted thermal neutron attenuation. However, its inherent radioactivity and rapid burn-up render it ineffective in thermal neutron absorption, thus limiting its applications as neutron shielding materials [16].
In contrast, 10B is a chemically stable and cost-effective element with outstanding shielding performance as a neutron absorber. It can effectively absorb thermal neutrons through the 10B(n, α)7Li transmutation reaction [17]. Without the secondary radioactivity caused by neutron absorption, B has gained popularity and is commonly employed in SNF storage [18]. Boron alloy steel [19], hexagonal boron nitride (BN) [20], and boron carbide (B4C) [21] are well-known examples of B-containing shielding materials. The addition of B-containing phases can greatly enhance the neutron absorbing performance of the shielding materials, but it comes at the cost of deteriorated plasticity [22]. Moreover, the neutron absorption processes that occur in B-containing phases have the potential to generate helium (He), which may lead to material swelling and cracking [23]. Considering the limitations of conventional boron materials, researchers are looking into potential alternatives to improve the mechanical properties and neutron absorbing ability of shielding materials. Currently, it is still an open question whether to improve the neutron shielding performance of materials with the limited addition of neutron absorbing elements.
One promising strategy is to introduce elements with a larger neutron capture cross-section in the shielding materials, which can effectively reduce the addition of neutron absorbers while simultaneously improving the mechanical properties of the materials. As early as 1935, Dunning et al. [24] discovered that the rare earth element Gd possessed a very high neutron capture cross-section, which was confirmed in subsequent studies [25,26]. For instance, 157Gd has the highest neutron capture cross-section of 254,000 barns, which is ~67 times higher than the 10B isotope. Another stable isotope, 155Gd, also exhibits an extremely high neutron capture cross-section of 62,540 barns. Notably, the total abundance of both isotopes is up to 30%. As a result, Gd has received increased attention in the development of neutron shielding materials. For instance, Burke et al. [27] manufactured gadolinium oxide paint to serve as the neutron attenuation/absorption medium in order to enhance radiation shielding efficiency. The neutron absorption processes of 155Gd and 157Gd are realized through 155Gd(n, γ) and 157Gd(n, γ) reactions [28].
With the ever-present safety concerns surrounding the advancement of nuclear energy, the field of neutron shielding research has gained increasing interest (Figure 3a). Among a vast number of neutron absorbers, gadolinium (Gd) has an exceptional ability to shield thermal neutrons. As a result, in recent years, there has been a notable upsurge in the publication of novel shielding materials containing Gd, specifically tailored for neutron shielding applications (Figure 3b).
Despite the growing interest in neutron shielding research, systematic reviews focusing specifically on Gd-containing shielding materials remain scarce. Considering the broad prospects of Gd in the field of neutron shielding, we believe that a systematic summary of Gd-containing shielding materials is necessary. Therefore, we present an in-depth review of recent advancements in the design, synthesis, and characterization of Gd-containing shielding materials. This in-depth examination covers three main material categories, including inorganic nonmetallic-based materials (i.e., concrete, glass, and ceramics), polymer-based materials, and metal-based materials (i.e., Fe-based, Al-based, and other metal-based materials). Through this review, we plan to offer a comprehensive diagram of the mechanical and functional properties of the novel Gd-containing neutron shielding materials, their practical challenges, and future prospects in neutron shielding.

2. Gd-Containing Inorganic Nonmetallic Neutron Shielding Materials

Inorganic nonmetallic materials (i.e., concrete, glass, and ceramics) are a collective term for materials excluding organic polymers and metals. Comparably, inorganic nonmetallic materials provide considerable economic, chemical stability, and durability benefits, making them highly promising candidates in the field of building and construction (e.g., nuclear reactor buildings). For their applications in nuclear facilities, inorganic nonmetallic materials must possess a certain level of strength and radiation shielding ability. This section focuses on the recent scientific advances in the use of Gd as a neutron absorber in the form of concrete-, glass-, and ceramic-based shielding materials.

2.1. Concrete-Based Neutron Shielding Materials

Concrete is a commonly used radiation shielding material mainly against high-energy γ-rays due to its high density, but its shielding ability for thermal neutrons is relatively low [29,30]. Typically, the thickness and weight of the concrete must be quite considerable to enhance radiation shielding performance, which constrains the development of shielding concretes at the same time. Unfortunately, the modification of the water-to-cement ratio of concretes proved unfeasible to improve the shielding performance for incident neutrons [31].
Alternatively, several reports indicated that neutron shielding performance can be improved by incorporating shielding components (e.g., Gd) into the concrete [32]. For instance, using a melt-spinning process, Park et al. [33] prepared the amorphous Fe-B-Mo-Gd ribbons with ~10 wt%Gd, and used them as reinforcements to the concrete matrix. Their findings revealed that the neutron shielding efficiency of the reinforced concrete increased by 86%. Such superior performance can be attributed to the uniform distribution of Gd-containing ribbons within the concrete matrix, effectively absorbing incident neutrons. Furthermore, the reinforced concrete exhibited higher ductility than the original matrix, which was a desirable property for practical applications. In addition, Seshadri et al. [34] performed the transmission measurements of neutrons through Gd-containing concrete, indicating that the dose transmission factor was reduced by ~30% in the 1 wt%Gd/concrete. Nevertheless, with a total shield thickness of 50 cm, the dose rates of Gd-containing concrete were 2.5 times greater than those of ordinary concrete, owing to the excessive generation of γ-rays during the neutron absorption process of Gd. Given the complex nature of harmful radiation, protection against mixed types of radiation, particularly γ-rays and neutrons with strong penetrating properties, has become a key research direction.
Gd2O3 is a typical oxidized form of the Gd element. As a ceramic reinforced particle, Gd2O3 is found to be an effective additive as a thermal neutron absorber. For instance, Piotrowski et al. [35] evaluated the neutron shielding properties of ordinary and heavy-weight magnetite concrete modified by epoxy resin and Gd2O3. Unfortunately, with Gd2O3 alone, the shielding effectiveness of concrete against fast neutrons was limited. In comparison, a combination of Gd2O3 and epoxy resin can effectively improve shielding abilities against both fast and thermal neutrons [36]. In their subsequent research, the comprehensive effect of Gd2O3 addition on Portland cement was discussed in detail [37]. Although Gd2O3 generated a slight retardation of cement hydration, there was an increased heat intensity released by the acceleration of aluminate activity. Furthermore, Gd-containing concretes not only exhibited enhanced neutron shielding ability, but also showed improved long-term (56 days) compressive and flexural strength, despite a minor drop in strength at the early stage (3 days). Long-term (56 days) compressive and flexural strengths were 4% and 15% higher than the reference specimens, respectively.
To summarize, the integration of Gd into concrete has shown significant potential in the area of neutron shielding. Considering the limited shielding ability of Gd against fast neutrons, polymer materials with a high H content are often used together with Gd-containing shielding concretes to fill this gap. This approach provides a viable future option for improving the neutron shielding performance of Gd-containing concretes.

2.2. Glass-Based Neutron Shielding Materials

Glass is a non-crystalline and amorphous solid that has many advantages, such as high optical transparency, high chemical stability, and low cost. Glasses with high transparency outperform other materials for doors, windows, and screens in scintillators, nuclear power plants, and medical diagnostic or treatment rooms [38,39]. Gd is a paramagnetic element with a high atomic number (64) and density (7.895 g/cm3). From an overall standpoint, Gd is a good choice for the application of radiation shielding glasses. In recent studies [40,41,42,43,44,45,46,47,48,49,50,51], Gd-containing shielding glasses have mainly focused on γ-ray radiation shielding ability, which should be discussed in the following section. Considering that Gd-containing shielding materials tend to release highly hazardous secondary γ-rays while absorbing neutrons, this type of work should be helpful to improve the comprehensive shielding performance of Gd-containing neutron shielding glasses.
Kaewjang et al. [40] reported Gd2O3/B2O3-SiO2-CaO glasses for γ-ray radiation shielding. All Gd2O3/glass samples showed a lower half-value layer (HVL, the thickness of a material that can attenuate 50% of the initial intensity of radiation) than ordinary concrete, X-ray shielding windows, and commercial windows. Here, low HVL represents good γ-ray shielding performance. Furthermore, the increased Gd2O3 content led to a linear increase in the glass density due to its high molecular weight [41,42]. Glasses with a higher density are useful in radiation shielding against γ-rays [43]. When it comes to neutron shielding performance, xGd2O3/B2O3-10SiO2-10CaO (x = 15, 20, 25, 30, and 35 mol%) fabricated by melt quenching exhibited different neutron shielding performances depending on the neutron energy [42]. The 15 mol%Gd2O3/glass is a superior shielding material for neutron energy of 1 eV, and the 35 mol%Gd2O3/glass becomes the optimum choice at >2 MeV neutron energy.
In addition, Al-Buriahi et al. [44] surveyed the impact of Gd2O3 addition on γ-rays and neutron shielding capacities of TeO2-ZnO-Nb2O5 glasses and found that the fast neutron removal cross sections (∑R) of glasses were improved with the Gd2O3 addition, indicating their good fast neutron attenuation ability. They attributed this improvement to the increased density due to the replacement of Nb2O5 by Gd2O3. Nevertheless, the thermal neutron shielding properties of Gd2O3/TeO2-ZnO-Nb2O5 glasses were not studied here. Similarly, Al-Hadeethi et al. [45] observed that ∑R value was raised with the Gd2O3 addition to TeO2-ZnO-Nb2O5 glasses, but this improvement was rather limited. They held the view that the improvement of fast neutron attenuation was derived from the density change of shielding glasses.
Moreover, Ibrahim et al. [46] fabricated ZnO-Bi2O3-B2O3 glasses doped with rare earth oxides (e.g., La2O3, CeO2, Nd2O3, and Gd2O3) by a melt quenching process. Among these glasses, Gd2O3/glass had the highest oxygen packing density, revealing that the structure of the Gd2O3/glass composite was tightly packed. Besides, Gd2O3/glass has the characteristics of higher density, lower molar volume, and higher γ-ray radiation shielding efficiency. The fast neutron removal cross-section (∑R) followed the order ∑R (Gd2O3/glass) > ∑R (Nd2O3/glass) > ∑R (Ce2O3/glass) > ∑R (La2O3/glass) > ∑R (B2O3/glass). The best fast neutron attenuation ability was achieved for Gd2O3/glass with ∑R = 0.1 cm−1, which was higher than that of ordinary concretes, hematite serpentine concretes, and ileminite-limonite concretes. Their results revealed that Gd2O3 had distinct advantages in the field of radiation shielding over other rare earth oxides (La2O3/CeO2/Nd2O3). According to Lakshminarayana’s report [47], the 37.5 mol%Gd2O3/SiO2-B2O3 glass showed a large total neutron cross-section (σT) of 646.171 cm−1, indicating that Gd2O3 plays an important role as a thermal neutron shielding component. Overall, the Gd2O3 addition dramatically improved the thermal neutron shielding performance of glasses, but its influence on attenuating fast neutrons required further investigations [44,45,48]. Critically, the addition of Gd2O3 has a minimal impact on the light transparency of radiation-shielding glasses, as evidenced by previous studies [40,49,50]. For instance, gadolinium sodium borosilicate glasses demonstrated a negligible reduction in light transmission with increasing Gd2O3 content [50], as visually observed in Figure 4. These findings highlight the significant potential of Gd2O3-containing glasses in the field of radiation shielding, as they retain excellent light transparency properties.
Apart from Gd2O3-doped shielding glasses, gadolinium oxyfluoride (Gd4O3F6)-doped glass is also an applicable radiation shielding glass system. However, both radiation attenuation parameters and HVL values of Gd4O3F6/glass are inferior to those of Gd2O3/glass [51], leading to its limited use in radiation shielding. Nevertheless, both shielding glass systems exhibited good γ-ray shielding properties and high light transmittance. However, there is relatively little research focusing on the neutron shielding properties of Gd-containing glasses. For applications, Gd can effectively absorb thermal neutrons, but it has a limited contribution to fast neutron attenuation. In addition, it is known that the γ-ray shielding performance of Gd-containing glasses improves with increased Gd content. Nevertheless, little research has been undertaken to comprehensively analyze the influence of excessive Gd on glass systems.

2.3. Ceramic-Based Neutron Shielding Materials

Ceramics are a group of hard, heat- and corrosion-resistant materials created under heat and/or pressure conditions, consisting of at least two elements (one of which must be non-metallic) [52]. Considering their applications as structural components for fusion reactors, cladding for gas-cooled fission reactors, and immobilized forms for radioactive waste [53], it is imperative to design ceramics with good neutron and radiation shielding performance.
Currently, there are several related studies on Gd-containing radiation shielding ceramics. For example, when Ge et al. [54] prepared Gd2O3/Al2O3 ceramics by pressure-less sintering, they found that the γ-ray radiation shielding property of materials improved with the increased Gd2O3 content. However, both mechanical properties and thermal conductivities decreased significantly when the Gd2O3 content exceeded 15 wt%. Besides, excessive new phases (i.e., Gd2Si2O7, Al2Gd, and Al5Gd3O12) formed under this condition, hindering the densification of ceramics. The conflicts between shielding performance and mechanical properties should require more attention in further research. Hernandez et al. [55] fabricated Gd2O3/bauxite ceramics by sintering procedures and found that their macroscopic neutron capture sections (∑) were linearly correlated with the Gd2O3 incorporation. The 10 wt%Gd2O3/bauxite ceramics exhibited the highest neutron capture section of 1.5 × 105 barns, and their corresponding mechanical properties were also excellent compared with other ceramics.
Perovskite metal oxides (e.g., ABO3 structures) have attracted significant attention recently due to their potential use in various applications (e.g., catalytic responses, piezoelectricity, and photoelectricity sensors) [56]. Gadolinium borate (GdBO3) is an important binary compound in the Gd2O3-B2O3 phase diagram, and also a proper candidate for neutron shielding. Dong et al. [57] synthesized the GdBO3 ceramics by pressure-less sintering from 950 to 1100 °C using Gd2O3 and boric acid. The results indicated that GdBO3 ceramics possessed high thermal stability, low thermal conductivity, and good neutron shielding performance. Additionally, their neutron shielding ability was highly correlated with Gd2O3 content, sintering temperature, and sample thickness. Moreover, the B content can also contribute to the neutron shielding performance of GdBO3 due to its neutron absorbing ability. Li et al. [58] evaluated the structural stability and neutron-shielding capacity of GdBO3/Al18B4O33 ceramics fabricated by a pressureless sintering process. In comparison to Gd-free ceramics, the 6 wt%Gd2O3 addition brought a striking decrease in neutron transmission rate (i.e., from 65% to 26%) to the shielding composite at a fixed sample thickness of 0.5 cm. This dramatic downward trend can be attributed to the Gd2O3 incorporation with an optimal addition (6 wt%) to improve the structural stability and mechanical properties of ceramics. Nevertheless, further Gd2O3 addition should lead to excessive GdBO3 formation, resulting in intergranular micro-cracks. This was detrimental to the comprehensive properties of neutron shielding ceramics.
High-entropy ceramics are solid solutions of five or more cation or anion sublattices, with a high configuration entropy. Recently, these systems have become increasingly popular due to their diversity in compositional design, which can be tailored to meet specific needs in different application scenarios [59]. For instance, Zhang et al. [60] designed and synthesized a novel monoclinic-type high-entropy ceramic (La0.2Ce0.2Gd0.2Er0.2Tm0.2)2 (WO4)3 powder by high-temperature sintering and ball milling processes. Due to the improved phase stability offered by the high entropy effect, the powder was less likely to fail when exposed to radiation. Gd acted as an effective neutron absorber in this high-entropy system. Then, they fabricated the shielding ceramic by dispersing (La0.2Ce0.2Gd0.2Er0.2Tm0.2)2(WO4)3 powder in epoxy resin, and the ceramic composite showed an acceptable tensile strength (=19.25 MPa), relatively lower thermal conductivity (<0.3 W·m−1·K−1), and density (<1.5 g·cm−3). Importantly, such a composite system with the highest content of high-entropy ceramic powders can absorb most thermal neutrons.
For fabricating ceramic shielding composites with unique designs and customized requirements, the additive manufacturing technique offers more freedom compared with the traditional milling, curving, and injection molding processes. Based on these considerations, Wang et al. [61] manufactured the Gd2O3 ceramics by vat photopolymerization 3D printing (Figure 5a). They found that both linear shrinkage (Figure 5b) and mechanical properties (Figure 5c) of as-printed ceramics increased with higher sintering temperatures. With a sintering temperature of 1600 °C, the as-printed ceramics achieved a density of 58%, a bending stress of 40 MPa, and a flexural elastic modulus of 20.2 GPa. Although there are still some problems with the formability of as-printed ceramics, this study provides a good strategy for the design and preparation of ceramic-based shielding materials.

3. Gd-Containing Polymeric Neutron Shielding Materials

A polymer is a substance or material consisting of macromolecules, which are made up of many repeating subunits. The H element is an important component for neutron shielding since it can easily thermalize the incident neutrons with high energy and velocity through the elastic scattering process [62]. Polymer matrix composites (PMCs) are high-H-content materials. They can effectively attenuate fast neutrons, which makes them an attractive candidate for neutron shielding [63]. Generally, PMCs have remarkable advantages, such as being flexible, cost-effective, lightweight, and corrosion-resistant [64]. Most importantly, the neutron shielding performance of PMCs can be improved by introducing neutron absorbers (e.g., B or Gd) as shielding fillers in the polymer matrix [65]. Therefore, they have great potential for several radiation shielding applications (e.g., wearable protective devices for the nuclear or aerospace industries) (Figure 6) [66]. For instance, Yastrebinskii et al. [67] successfully designed a neutron-protective PMC cover with a high concentration of Gd filler for transporting packing containers of SNFs. This section focuses on Gd-containing PMCs and their potential in neutron shielding applications, as well as the effect of Gd particle size on the neutron shielding performance of these PMCs. In addition, some notable research advances will be mentioned in this area.

3.1. Traditional Polymer-Based Shielding Neutron Materials

In recent years, various polymers have been developed as neutron shielding materials, among which epoxy resin (ER), polyvinyl alcohol (PVA), and polyethylene (PE) are notable examples [64]. Normally, polymer materials can attenuate fast neutrons through the scattering process, achieving the desired neutron shielding efficiency at the required thickness of shielding materials. However, the thermal neutron shielding properties of polymer materials are relatively limited. Therefore, the incorporation of Gd, which has the highest thermal neutron capture cross-section, is a promising approach to improving the shielding performance of polymer materials.
PVA, ([CH2CH(OH)]n), which is a H- and O-rich polymer macromolecule with self-healing ability, can effectively attenuate fast neutrons [68]. Based on this point, several PVA-based shielding composites were developed. To improve their thermal neutron shielding performance, Tiamduangtawan et al. [69] introduced modifications to PVA hydrogels with either Sm2O3 or Gd2O3 fillers. Their findings revealed that both Gd2O3/PVA and Sm2O3/PVA hydrogels exhibited substantially enhanced thermal neutron capture abilities. Furthermore, both hydrogels displayed autonomous self-healing capacity on damaged/fractured surfaces. Additionally, these hydrogels showed increased tensile modulus and strength. Nevertheless, a small decrease in elongation occurred with the addition of Gd2O3 and Sm2O3 fillers. In terms of thermal neutron shielding effectiveness and tensile properties, Gd2O3/PVA hydrogels outperformed Sm2O3/PVA hydrogels under the same filler concentration and thickness of shielding composites.
Additionally, Wang et al. [70] synthesized poly(MMA-co-Gd(MAA)3) through the self-polymerization process of gadolinium methacrylate (Gd(MAA)3). The results indicated that poly(MMA-co-Gd(MAA)3) outperformed poly(methyl methacrylate) (PMMA) in thermal neutron shielding. To achieve a thermal neutron shielding efficiency of 90%, ~70 mm thick PMMA was required. Comparably, to achieve such a similar effect, the poly(MMA-co-Gd(MAA)3 only needed the thickness at ~14 mm (0.38 wt%Gd), ~5 mm (1.91 wt%Gd), and ~2.4 mm (3.81 wt%Gd). Moreover, Gd(MAA)3 can also be applied to polyacrylonitrile (PAN) as a thermal neutron absorber [71].
In a practical nuclear radiation environment, high-energy neutron and γ-ray radiation normally occur simultaneously. Besides, Gd also releases secondary γ-ray radiation when it absorbs neutrons. Thus, the design of Gd-containing neutron shielding must consider γ-ray radiation protection as well. Some investigations of Gd-containing neutron shielding materials often neglect this aspect, which needs more attention in further studies. Recently, Wang et al. [72] synthesized Gd2O3/polystyrene-block-poly(ethylene-ran-butylene)-block-polystyrene (SEBS) composites for thermal neutron and γ-ray radiation shielding applications. In their research, the Gd2O3 particles were homogeneously and tightly connected to the matrix via a melting and blending process. Additionally, the Gd2O3/SEBS composites exhibited high flexibility. By comparing the neutron transmittance rates of Gd2O3/SEBS and B4C/SEBS composites, they found that Gd2O3 was a superior neutron absorber in the 0.025–0.15 eV range over B4C. Experimental results revealed a positive correlation between neutron shielding performance and Gd2O3 addition, as evidenced by a decrease in neutron transmittance from 59.1% to 35.3% as Gd2O3 content increased from 3.9 wt% to 54 wt%. In their study, they found that the Gd2O3/SEBS composite had the best shielding efficiency for γ-rays at 59 keV. Specifically, γ-ray transmittance reduced with increasing Gd2O3 concentration, and the transmittance may be lowered from 90.08% to 28.42% with the introduction of Gd2O3 for 59 keV γ-rays.
The idea of self-healing is a very promising strategy for increasing the lifespan of some polymers by enabling them to repair physical damage on their own [73]. For instance, both Gd2O3/PVA and Sm2O3/PVA hydrogels designed by Tiamduangtawan et al. [69] exhibited autonomously self-healing properties at fractured surfaces. These hydrogels were able to recover a portion of their mechanical strengths over time, but the recovery efficiency was reduced with increased concentrations of Sm2O3 or Gd2O3. For self-healing shielding materials, Poltabtim et al. [74] proposed a Gd2O3-reinforced natural rubber (NR) composite. Such a Gd2O3/NR composite exhibited remarkable self-healing capability at fractured surfaces due to the incorporation of a reversible ionic supramolecular network. The addition of Gd2O3 to NR led to a substantial enhancement in neutron shielding performance. Even a minor addition of 25 parts per hundred rubber (phr) of Gd2O3 to pristine NR resulted in a drastic reduction in both thermal neutron attenuation rate (I/I0) and half-value layer values (HVL) over the NR matrix. Specifically, I/I0 and HVL were reduced from 97% and 84 mm in pure NR to a mere 55% and 2.3 mm in the 25 phr Gd2O3/NR composite (both I/I0 and HVL values were compared using 2 mm thick samples). However, the Gd2O3/NR composite showed decreased mean tensile strength and elongation with increasing Gd2O3 addition, but increased hardness. In summary, the available results demonstrate that the self-healing shielding materials that have been investigated thus far must have a better recovery efficiency. Since shielding materials are subjected to a variety of damages during service, the development of neutron shielding PMCs with good self-healing capability is of great importance.

3.2. Novel Polymer-Based Shielding Neutron Materials

MXenes are a new family of 2D transition metal carbides, carbonitrides, and nitrides. Due to their favorable 2D layered structure, highly tunable composition, and surface functional groups, MXenes have drawn significant interest in a variety of applications, including sensors, energy storage, and shielding from electromagnetic interference [75]. Inspired by fish scales, Zhu et al. [76] utilized the spin coating technique to produce Gd@MXene/PVA films, as formed by PVA and Gd@MXene nanoflakes (Figure 7). The Gd@MXene nanoflakes were synthesized through a hydrothermal reaction and consisted of 2D Ti3C2Tx-MXene nanoflakes and 0D Gd nanoparticles. Their design aligned the random distribution of Gd nanoparticles into an oriented 2D pattern, creating fish scale-like barrier walls. These walls effectively scattered neutrons multiple times between the nanoflakes, enhancing the neutron shielding efficiency of Gd nanoparticles. In Figure 7, the Gd@MXene/PVA composite showed synergistic effects with improved thermalization and absorption efficiency towards incident neutrons over both Gd/PVA and MXene/PVA composites. Since MXene has limited neutron shielding ability, neutrons can directly permeate the MXene/PVA layer. In comparison, the incorporation of Gd@MXene with the oriented 2D fish scale-like barrier walls facilitated multiple scatterings of slow neutrons. Moreover, the Gd@MXene/PVA composites showed improved neutron shielding properties with increased Gd@MXene contents. The thermal neutron attenuation rate of 20 wt%Gd@MXene/PVA was ~10.44%, which is 3.36 times higher than pure PVA.
Metal–organic frameworks (MOFs) are made by linking inorganic and organic units with strong bonds to create open crystalline frameworks with high porosity [77]. Hu et al. [78] prepared the Gd-MOF-reinforced epoxy resin (EP) composite using a solution casting approach. The hardness, stiffness, and thermal neutron capture ability of the composite increased remarkably with the addition of Gd-MOF fillers. However, the composite showed a slight reduction in tensile strength, potentially attributed to the weak interfacial adhesion between the Gd-MOF fillers and the matrix. Moreover, the incorporation of Gd-MOF fillers improved the thermal stability of the Gd-MOF/EP composite. In particular, the 10 wt%Gd-MOF/epoxy composite showed an acceptable tensile strength of 66.6 MPa and good thermal neutron shielding performance. Similarly, a recent study revealed that Gd-MOF fillers have a high potential for use as polyimide (PI) matrix reinforcements in improving neutron shielding performance [79]. The fast neutron penetration rate of the 3 wt% Gd-MOF/PI composite was 9.6% when the specimen thickness was 0.2 cm, while the thermal neutron penetration rate was only 0.9%. Compared with the original PI material, the introduction of Gd-MOF greatly improved the neutron shielding performance of the polymer matrix. In contrast to pure PI, the Gd-MOF/PI composite had a higher glass transition temperature (Tg), making it a suitable choice for engineering applications. Meanwhile, the Gd-MOF/PI composite exhibited improved thermal stability, which in turn increased the service temperature of the polymer shielding material. Generally, these examples provide valuable insights for designing high-performance neutron shielding materials based on polymers (which exhibit better fast neutron attenuation capability) and Gd fillers (which have effective thermal neutron shielding properties).

3.3. Improving Neutron Shielding and Mechanical Properties of PMCs with Gd Nanofillers

The advent of nanotechnology has greatly facilitated the application of nanomaterials as highly effective fillers/reinforcements, resulting in noteworthy enhancements in the mechanical properties and radiation shielding efficiency of composites [67,80]. There are considerable research efforts devoted to the applications of Gd-based nanofillers in PMCs for radiation shielding purposes [63,76]. In this section, the influence of Gd particle size on both radiation shielding and the mechanical properties of PMCs is discussed.
For instance, Li et al. [81] investigated the impact of particle size on the radiation shielding performance of epoxy resin (ER) matrix composites infused with dispersed micro- and nano-Gd2O3 fillers. The results indicated that nano-Gd2O3/ER composites had a superior ability to shield X-/γ-rays than their micro-Gd2O3/ER counterparts. This can be attributed to the fact that the nanoparticles have a smaller size and uniform distribution, increasing the likelihood of interaction between X-/γ-rays and particles. Interestingly, the size effect is strongest at low particle concentrations and declines as particle concentration increases. Prabhu et al. [67] also claimed that the enhancement of radiation shielding efficiency of composites was predominant with Gd nanofillers with decreased size at very low loadings. Although the shielding composites were not tested for neutron absorption in the above two studies [67,81], their findings have offered a new perspective on the role of particle size in material design.
In İrim’s work [82], the high-density polyethylene (HDPE) was reinforced by hexagonal boron nitride (h-BN) and Gd2O3 nanoparticles using the melt-compounding technique. The as-fabricated h-BN@Gd2O3/HDPE nanocomposites demonstrated impressive advancements in both neutron and γ-ray radiation shielding performance, leading to enhancements of 200–280% and 14–52%, respectively. Notably, Gd2O3 served the dual purpose of shielding against both neutron and γ-ray radiation. Simultaneously, the mechanical properties (i.e., tensile strength and modulus) were enhanced by the addition of nanoparticles. Likewise, Baykara et al. [83] incorporated Gd components into PMCs by creating polyimide nanocomposites doped with h-BN and Gd2O3 nanofillers. The shielding composites exhibited improved mechanical properties and radiation shielding performance, highlighting the potential of Gd nanofillers as versatile reinforcements for the design and development of advanced shielding materials.
The integration of Gd nanofillers in PMCs has brought remarkable improvements in both radiation shielding performance and mechanical properties. Nevertheless, recent studies revealed that excessive Gd content may have unfavorable impacts, i.e., a linear decrease in thermal stabilities of GdxFe2O3/PVA composites [84]. Likewise, He et al. [85] reported that the addition of nano-Gd2O3 is unfavorable to the thermal stability of the salicylic acid polyphenylene sulfide (SAPPS) composites. Additionally, the mechanical properties of nano-Gd2O3/SAPPS composites were adversely affected by the excessive nanofillers. Given the high temperatures encountered in operation, thermal stability is a key consideration when selecting radiation shielding materials. To improve thermal stability, Huo et al. [86] introduced surface-modified micro- and nano-Gd2O3 fillers to reinforce HDPE composites. They discovered that the modified fillers enhanced the thermal stability of composites by improving interfacial compatibility, which restricted the movement of HDPE chains. Furthermore, nano-Gd2O3 had a larger interaction probability with radiative particles (neutrons and γ-rays) due to its high dispersibility in the matrix (HDPE) than micro-Gd2O3. In addition, the uniform dispersion of nano-Gd2O3 filler in the matrix reduced the stress concentration effect and increased the tensile strength. Thus, the modified nano-Gd2O3/HDPE composites showed superior neutron and γ-ray radiation shielding performance and mechanical properties over the modified micro-Gd2O3/HDPE composites. Both particle size and surface modification improved the radiation shielding performance of composites. To summarize, the size of particles is a crucial factor in enhancing the neutron shielding properties of PMCs. The significant enhancements can be attributed to the highly synergistic interaction between the polymer matrix and nanoparticles. This interaction results in a reinforcing effect that surpasses that of most conventional microparticles. However, the improvement of nanoparticle homogeneity in PMCs remains a topic of ongoing research. Overall, the utilization of Gd nanoparticles as fillers in PMCs offers novel prospects for the design and processing of high-performance shielding materials with targeted properties.
In conclusion, the use of Gd offers significant potential for the advancement of polymer-based neutron shielding materials, and the reported findings are presented in Table 2. This should serve as a reference for the design and research of Gd-containing polymer neutron shielding materials.

4. Gd-Containing Metallic Neutron Shielding Materials

With the steady growth of nuclear energy, the development of neutron shielding structural materials for the storage and transportation of SNFs or radioactive wastes is a pressing issue in nuclear engineering and technology nowadays [87]. Practically, shielding materials have to operate under high-energy radiation exposure and multiple corrosive media in either dry or wet circumstances. Thus, ensuring the long-term stability of neutron shielding materials should necessitate not just adequate radiation shielding ability, but also good mechanical properties and corrosion resistance. In this regard, metallic materials offer an obvious advantage over inorganic nonmetallic materials and polymer materials. For example, polymer-based shielding materials cannot withstand higher temperatures (e.g., 200 °C) due to their high H content [88]. In contrast, metallic materials are widely used as structural materials in the storage and transportation of SNFs given their high thermal stability, mechanical strengths, and corrosion resistance [89].
Depending on their compositions, Gd-containing metallic materials can be classified into three categories: Fe-based alloys/composites, Al-based alloys/composites, and other metal-based alloys/composites. In metallic neutron shielding materials, Gd can also be introduced to the matrix in the metallic form as Gd-containing phases in addition to the common Gd2O3 ceramic reinforcements. Here, we will present the latest developments in metallic neutron shielding materials in each category.

4.1. Fe-Based Alloys/Composites

Fe-based alloys/composites, primarily utilizing stainless steel as the matrix, have been developed as neutron shielding materials. Stainless steel, known for its excellent corrosion resistance and mechanical properties, is a major structural material in nuclear reactor systems [90]. Incorporating neutron absorbers into stainless steel has been identified as a viable strategy for the development of Fe-based neutron shielding materials. Borated austenitic stainless steel, for example, has 0.2–2.25 wt%B and offers strong mechanical properties and neutron shielding performance for SNF storage applications [91]. However, due to the poor solubility of the B element in stainless steel, fabricating borated austenitic stainless steel with more than 2.25 wt%B is relatively difficult [92]. Furthermore, the B element would emit He bubbles after absorbing neutrons, which adversely affects its actual application [23].
The Gd-containing steel was first proposed by Kato [93], where Gd was utilized as a neutron shielding phase without the formation of He bubbles during the neutron absorption process. In addition, Gd has a far larger neutron capture cross-section than B. Therefore, the Gd-containing steel can effectively solve the trade-off between alloying contents and neutron shielding performance, allowing for higher neutron shielding performance with less alloying addition. Among available stainless steels, the duplex stainless steels (DSS), comprising both ferrite and austenite phases, exhibit a combination of superior strength and corrosion resistance. For this reason, Choi et al. [94] developed a 1 wt%Gd/DSS produced by melting, casting, hot-rolling, and solution treatment processes. The as-cast steel contained typical duplex phases (i.e., 31% ferrite and 69% austenite), and Gd-rich precipitates were found inside grains and at grain boundaries. After hot rolling to 6 mm thickness, the Gd/DSS exhibited good formability without cracks and showed improved strength and superior stress corrosion resistance. Later, Choi et al. [95] fabricated a 0.087 wt%Gd/DSS shielding material by casting, hot rolling, solution treatment, and cold rolling techniques. The ultimate tensile strength and elongation of the cold-rolled 0.087 wt%Gd/DSS were 694 MPa and 37.1%, which can be compared to the 1 wt%Gd/DSS with a higher content of Gd. A similar study by Baik et al. [96] indicated that the majority of Gd was detected as sub-micron precipitates of GdCrO3 in the 0.02 wt%Gd/DSS. When exposed to a neutron beam with a wavelength of 0.48 nm, the neutron absorption coefficient of a 2 mm thick 0.02% wt%Gd/DSS sheet was approximately three times greater than that of pure DSS.
The austenitic stainless steels (e.g., 304 and 316(L)), where austenite is the primary crystalline structure, are another prevalent type of stainless steel. Such stainless steels have higher Ni and Cr contents, offering exceptional formability, weldability, mechanical strength, and corrosion resistance [97]. Due to the low solubility of Gd in both ferrite and austenite phases, the Gd element is mostly formed in the intermetallic phase [98]. Robino et al. [99] designed a (0.1–10 wt%) Gd/316L steel for SNF storage, with the majority of Gd components forming an intermetallic phase of (Fe, Ni, Cr)3Gd. This intermetallic phase was identified to contain ~28 wt%Ni and ~3 wt%Cr, resulting in Ni depletion and Cr enrichment in the steel matrix, which in turn influenced the matrix stability. With increased Gd concentration, the formation of the (Fe, Ni, Cr)3Gd phase and the resulting shift in matrix structure from austenite to ferrite led to a rise in the hardness of the Gd/316L steel. However, DuPont et al. [100] discovered that the presence of this intermetallic phase in stainless steels posed challenges for both weldability and hot rolling process. Wang et al. [101] found that the Gd-rich phases with an average size of ≤3 μm were uniformly dispersed at ≤1.5 wt%Gd/316L steel after hot rolling. As Gd concentration increased to 2 wt%, some Gd-rich phases coarsened to ~6 μm and agglomerated, which had a detrimental effect on its corrosion resistance. However, there have been few investigations conducted to thoroughly investigate the structure and stability of Gd-rich phases to date.
The low solubility of Gd, similar to that of B, restricted its applications in Fe-based neutron shielding materials. However, non-equilibrium fabrication processes (e.g., mechanical alloying) can effectively deal with this issue [102]. Yang et al. [103] prepared the 7.87 wt%Gd/316L steel by mechanically alloying Gd2O3 and 316L powders, followed by hot isostatic pressing. Herein, the nanosized Gd-Si-O particles (~12.5 nm) were evenly distributed in the alloy through a dissolution and re-precipitation mechanism. The 7.87 wt%Gd/316L steel had much better thermal neutron shielding performance than 316L steel. Results showed that the 7.87 wt%Gd/316L alloy with a thickness of 0.2 mm exhibited the equivalent thermal neutron attenuation capability of a much thicker (33 mm) 316L alloy, achieving a 90% reduction in thermal neutron transmission. Their research provided a promising route for fabricating neutron shielding alloys enriched with a high Gd content through the process of mechanical alloying. Although the Gd addition can improve the neutron shielding performance of stainless steels, the relationships between the Gd concentrations, the thickness of shielding materials, and shielding efficiency remain unclear. Wan et al. [104] fabricated Gd2O3/316L composites through ball milling and spark plasma sintering processes. Using the Monte Carlo N-Particle Transport Code (MCNP) simulations, they established the constitutive equation of the 155/157Gd areal density and thermal neutron shielding rate and discovered that the thermal neutron absorption rate of Gd2O3/316L composite was ~99% when the 155/157Gd areal density surpassed 0.01 g/cm2 (Figure 8a,b). Meanwhile, Wang et al. [105] also confirmed that the thermal neutron capture rate of the Gd/316L composite was >99% when the 155/157Gd areal density exceeded 0.01 g/cm2 based on MCNP simulations. Qi et al. [106] recently undertook an analogous study to establish a mathematical link between the thermal neutron absorption rate and 155,157Gd areal density (A(155,157Gd)) (Figure 8c,d). The thermal neutron shielding performance can be directly estimated using 155,157Gd areal densities, which reduces the difficulty of taking numerous parameters into account. The simulation findings revealed that at a Gd concentration of 1.5 wt% and a thickness of 3 mm for such material, the thermal neutron shielding rate of Gd/316L steel was greater than 99.9%.
Theoretical modeling has greatly assisted in material design, but it has limitations. For example, recent studies by Wang et al. [105] and Qi et al. [106] revealed that the incorporation of Gd in 316L steel led to a reduction in mechanical properties and corrosion resistance with increased Gd content. This is not within the prediction range of theoretical models previously developed. Therefore, it is crucial to develop a model capable of identifying the optimal Gd concentration that can achieve a balance between neutron shielding performance, mechanical properties, and corrosion resistance.
Figure 9 presents the mechanical properties of the reported Fe-based Gd-containing neutron shielding materials [94,95,101,104,105,106]. Herein, the Fe-based alloys/composites have exhibited high tensile strength and/or elongation, which are promising for use as structural-functional integrated materials.

4.2. Al-Based Alloys/Composites

Al alloys (e.g., Al-Mg-Si alloy) and Al matrix composites (AMCs) are widely used in the automotive and aerospace industries due to their favorable mechanical properties and lightweight nature [107]. Furthermore, they can be strengthened or modified by various particles/phases for functional purposes, such as neutron shielding materials. Among these reinforcing particles/phases, B4C is the most commonly used material in the AMCs due to its high strength, low density, and excellent neutron shielding properties [108]. Normally, the neutron absorption ability of the B4C/Al composite is highly dependent on the B4C content and sample thickness [109]. Chen et al. [110] discovered that the B4C addition enhanced the neutron shielding performance of the B4C/6061Al composite. However, a high proportion of B4C adversely affected the plasticity of the shielding composite, bringing challenges during formation. Herein, the elongation of the 30 wt%B4C/6061Al composite was less than 4%, making it unsuitable for use as a structural material in neutron shielding applications. The composite plasticity can be improved by decreasing the B4C content [111], yet this leads to a reduction in neutron shielding effectiveness at the same time. As a result, thicker shielding plates are necessary to provide the required neutron shielding performance, which contradicts the principles of our design. Furthermore, it should be noted that simply optimizing the preparation process may not be sufficient to improve composite plasticity with a high B4C content [20].
As a strong neutron absorber, Gd was used as a substitute for B4C by Xu et al. [112]. They prepared (15 wt%B4C + 1 wt%Gd)/Al composite by the vacuum hot pressing method, achieving a uniform distribution of both B4C particles and Gd phases in the Al matrix. The addition of 1 wt%Gd to 15 wt%B4C/Al composite resulted in an almost equivalent thermal neutron shielding performance to the 30 wt%B4C/Al composite. Furthermore, the (15%B4C + 1%Gd)/Al composite showed improved elongation (~9%) by reducing B4C content. The obtained results indicated that the Gd addition can substantially improve the composite plasticity while maintaining its neutron shielding performance, opening up great potential for the development of highly effective Al-based neutron shielding materials with limited neutron absorber incorporation.
In recent years, there has been growing interest in hot-working techniques for processing AMCs (e.g., hot rolling, extrusion, and forging). Hot-working approaches can improve mechanical properties by modifying the microstructure of Al composites and regulating the size and distribution of reinforcements. To improve the interface bonding and achieve a more uniform distribution of Gd phases, Xu et al. [113] subjected the as-prepared (15 wt%B4C + 1 wt%Gd)/Al composite to a hot extrusion process. The as-extruded (15 wt%B4C + 1 wt%Gd)/Al composite demonstrated an impressive 15% increase in tensile strength over its 15 wt%B4C/Al counterpart and showed nearly comparable levels of plasticity. The reinforced phases in the (B4C + Gd)/Al composite were identified as Al2Gd3, Al5Gd3O12, and Al11GdO18. Besides, the interaction between Gd and the Al matrix in Gd-containing AMCs leads to the formation of Al-Gd phases. Furthermore, the occurrence of an oxide layer on the original Al and Gd powders during the preparation procedure contributed to the generation of Al-Gd-O phases. Both factors are responsible for the formation of three reinforcement phases in the composite. These in situ formed phases showed higher strength and modulus than Gd particles and Al matrix, making them suitable reinforcements along with B4C in the composite. In their subsequent research, Xu et al. [114] performed a hot rolling process on the as-prepared (B4C + Gd)/6061Al composite to evaluate the effect of hot rolling on the strengthening phase, microstructure, and mechanical properties of the composite. The results revealed that the dominant Gd-containing phase was Al5Gd3O12 and that it did not undergo any phase transition during the hot rolling process. The size of the Gd-containing phase decreased as the amount of hot rolling deformation increased. The hot extrusion process led to a more uniform distribution of Gd-containing phases, thus enhancing the mechanical properties of the (15 wt%B4C + 1 wt%Gd)/6061Al composite. Here, the Gd-containing phases functioned as strengthening components to improve the tensile properties of the (15 wt%B4C + 1 wt%Gd)/6061Al composite in comparison with the 15 wt%B4C/6061Al composite (Figure 10) [114]. Additionally, the presence of stacking faults, which were the primary defects of Gd-containing phases, allowed dislocations to slip, thereby improving the plasticity of the shielding composite.
On this basis, Jiang et al. [115] carried out a more comprehensive investigation to clarify the effect of hot rolling on the mechanical properties of the (1 wt%Gd + 15 wt%B4C)/6061Al composite. The study concluded that the hot rolling technique resulted in a significant improvement in the mechanical properties of composites. The mechanical properties of the (Gd + B4C)/6061Al composite can be enhanced by two factors: (1) The hot rolling process led to a more uniform distribution of the Gd-containing phases. (2) The hot rolling process enhanced interfacial bonding strength between Gd-containing phases and the Al matrix. Besides, the thermal neutron shielding coefficient of the (1%Gd + 15%B4C)/6061Al composite achieved 99.9% when the sample thickness reached 3 mm. The hot rolling technique substantially improved the thermal neutron shielding performance of the composite by achieving a more uniform distribution of Gd-containing phases. Their findings provided a promising approach for the development of structural-functional integrated shielding materials.
Casting is a common method for preparing AMCs. For B4C/Al composite, but the high B4C content could impair the formability during the casting process [116]. Incorporating a smaller amount of Gd instead of B4C has been shown to significantly enhance the plasticity and formability of AMCs while maintaining similar neutron shielding efficiency. Nevertheless, the large-sized Gd-containing phases are prone to aggregate along grain boundaries in the casting process due to the low solubility of Gd in the Al matrix, significantly worsening the mechanical properties of the corresponding composite [117]. To avoid the problems associated with conventional casting procedures, Chen et al. [118] utilized an ultrasound-assisted casting method to fabricate (10 wt%B4C + 3.6 wt%Gd)/6061Al composite, followed by hot extrusion and heat treatment. The use of ultrasonic stirring has been shown to improve the uniformity of Gd-containing precipitates during solidification in the molten Al matrix. The microstructure and mechanical properties of the as-cast composite were modified using both hot extrusion and heat treatment. During this process, larger Al3Gd particles were broken down into smaller ones. In addition, the hot-working technique induced the precipitation of a nano-sized β phase (Mg5Al2Si4), which acted as a strengthening phase in the Al matrix. These modifications led to a remarkable improvement in the mechanical properties of composites. In detail, the yield strength, ultimate tensile strength, and elongation of the modified composites increased to 301 MPa (improved by 219%), 342 MPa (improved by 191%), and 2.4% (improved by 189%), respectively. However, the plasticity of this shielding composite was rather limited. Nonetheless, this study laid a solid foundation for the subsequent large-scale production of (B4C + Gd)/Al composites via the casting process.
The incoherent interface between B4C and the Al matrix is prone to delamination during the deformation process, and the low fracture toughness of B4C may promote the initiation and propagation of cracks. These problems greatly hinder the development of B4C/Al composites with higher ductility [119]. In recent years, the in situ TiB2/Al composite has gained considerable attention due to its good mechanical properties and high interface coherence [120]. Besides, the presence of 10B has made TiB2/Al composites a potential candidate for neutron shielding applications. Yang et al. [121] fabricated a TiB2/Al-Mg-Gd composite with an integrated structural-functional design using powder metallurgy methods. This composite was designed with the goal of simultaneously improving both its mechanical properties and neutron shielding performance. Structurally, the TiB2/Al-Mg-Gd composite exhibited high yield strength (353 ± 5 MPa), tensile strength (464 ± 6 MPa), and elongation (15.6 ± 0.4%), surpassing the available Al-based counterparts. Here, both solution strengthening by Mg atoms and grain refinement strengthening contributed significantly to the overall mechanical performance. Functionally, the TiB2/Al-Mg-Gd composite exhibited a high thermal neutron shielding efficiency (99%) at a limited thickness.
Pure Gd is prone to being oxidized in the air, presenting itself in the form of Gd2O3 [37]. Generally, pure Gd is relatively expensive. Therefore, Gd2O3, with its low cost and high stability, is a commonly used reinforcement for neutron shielding materials. Thus, using Gd2O3 particles instead of B4C particles allows for a reduction in reinforcement addition without sacrificing the shielding effectiveness or mechanical strength of the corresponding neutron shielding composite. For instance, Zhang et al. [122] fabricated a Gd2O3/6061Al composite by spark plasma sintering and hot rolling. The Monte Carlo simulation and neutron shielding tests revealed that the 10 wt%Gd2O3/6061Al composite had comparable thermal neutron shielding performance with the 30 wt%B4C/6061Al composite. After the ball milling process, the composite microstructure exhibited a bimodal distribution of Gd2O3 particles, where some coarse particles were reduced to the nanoscale. The primary strengthening mechanisms of composites were synergistic effects from grain boundary enhancement of Gd2O3 microparticles and grain internal enhancement of Gd2O3 nanoparticles. The fracture mechanism observed in the 10 wt%Gd2O3/6061Al composite was a ductile fracture. Later, Li et al. [123] investigated the microstructure evolution and strengthening mechanisms of the hot-rolled Gd2O3/6061Al composite. As the deformation progressed, Gd2O3 particles were found dispersed evenly throughout the 6061Al matrix. It was confirmed that the Gd2O3 particles played a crucial role in promoting dynamic recrystallization nucleation while simultaneously inhibiting grain growth. The strengthening mechanism of the Gd2O3/6061Al composite was primarily attributed to Orowan strengthening, and the related fracture mechanisms shifted from brittle to ductile after hot rolling. Similarly, Cong et al. [124] designed and fabricated a Gd2O3/6063Al composite, which exhibited better thermal neutron shielding efficiency than the 30 wt%B4C/Al composite.
The neutron absorption process is obtained through the Gd(n, γ) reaction. The secondary γ-ray, a by-product of the Gd(n, γ) reaction, poses a great threat to the surroundings. Furthermore, the occurrence of high-energy neutron and γ-ray radiation is simultaneous in realistic nuclear radiation. However, the current Gd-containing neutron shielding materials are still inadequate for γ-ray radiation shielding. Thus, the design of Gd-containing neutron shielding materials may include effective γ-ray radiation protection as well.
Tungsten (W) is frequently employed as a reinforcing phase in radiation shielding composites considering its outstanding performance as γ-ray shielding components [125]. To overcome the limitations of current neutron shielding materials employed in the storage of γ-ray radiation protection, Cong et al. [126] designed a core–shell structure with a W outer layer and a Gd2O3 inner layer, and 6063Al was used as the matrix for structure support. They found that xW@25wt%Gd2O3/6063Al (x = 7, 15, 25, and 35 wt%) composite showed increased radiation shielding ability for both thermal neutrons (<0.15 eV) and γ-rays compared with their commercial 30 wt%B4C/Al counterparts. In-depth investigations have revealed that an extended sintering time and higher sintering temperature, coupled with an elevated W content, contributed to a substantial increase in the microhardness of the W@25wt%Gd2O3/6063Al composite. Similarly, Zhang et al. [127] fabricated the Al composite reinforced by particles consisting of micron-sized Gd2O3 cores and nano-sized W shells (Figure 11). In Figure 11, the outer W shell of Gd2O3@W composite can effectively protect against both secondary γ-rays in the shielding process and γ-rays in the radiation environment, filling the gap of low efficiency of Gd2O3/Al composite in γ-ray radiation shielding. Monte Carlo simulations demonstrated that the neutron shielding ability and the half-value layers for primary and secondary γ-rays met the necessary criteria for effective neutron shielding applications. The core–shell structure particles led to an improvement in both tensile strength and elongation of the shielding composite compared with the contrast sample. Both fine-grain strengthening and Orowan strengthening made significant contributions to the mechanical strength of the shielding composite. Notably, the tensile fracture morphology revealed that the trans-granular fracture of the composite was caused predominantly by the presence of grain-boundary Gd2O3@W particles, which hindered crack propagation.
It is widely acknowledged that Gd is an effective absorber of thermal neutrons, but its shielding effectiveness against fast neutron and γ-ray radiation is limited. Therefore, additional shielding components are necessary to design materials with multi-functional shielding applications. Recent studies by Cong et al. [126] and Zhang et al. [127] have demonstrated the potential of core–shell structures in shielding materials. Currently, the manufacturing of neutron and γ-ray co-shielding materials involves separate production and subsequent combination for practical use. Integrating different components into a functional unity and improving the connection stability between various components are crucial.
Pitting corrosion, which is the primary form of corrosion for Al-based composites used for the storage of SNFs, can be inhibited by Gd-containing particles (i.e., Al3Gd) in the W@Gd2O3/Al composite [128]. However, this study also revealed that Fe+ irradiation (used to simulate neutrons and γ-ray irradiation) led to the formation of numerous cavities on the surface layer of the composite, and these defects provided fast diffusion paths for corrosive ions, resulting in an accelerated corrosion rate. Therefore, resistance to corrosion and radiation damage should also be considered to ensure the reliability of shielding materials. Herein, Figure 12 summarizes the influence of particle size and distribution on the corrosion behavior of shielding composites [128]. The Gd-containing particles were coarsened significantly in the sample sintered at high temperatures, leaving large pure-Al regions exposed and leading to localized corrosion. During the corrosion process, the corrosive media diffused inward, and the concentration of corrosive ions decreased with depth. In addition, metallic Al was first oxidized to Al3+, and then hydrolyzed to form Al2O3 and AlO(OH) oxide layers.
Table 3 lists the reported Gd-containing AMCs designed for thermal neutron shielding. Herein, Gd was first used as a substitute for B4C in Al-based shielding materials [112,113,114,115]. Researchers found that adding extremely tiny quantities of Gd may effectively improve the neutron absorption ability of shielding composites, addressing the problem of poor plasticity in B4C/Al composites by replacing Gd for B while maintaining the same level of neutron shielding performance. Thereafter, the more cost-effective and stable Gd2O3 ceramic particle was proposed as an alternative neutron absorber in Al-based shielding materials [122,124]. Later, to fill the gap of Gd2O3 in both fast neutron attenuation and γ-ray shielding fields, a core–shell structure with the W shell and the Gd2O3 core was designed [126,127]. Gd2O3 particles can effectively shield the incident neutrons and γ-rays, improving the overall shielding performance of AMCs. However, there is a compromise between the mechanical properties and neutron shielding performance, and fabricating a structural material with high neutron shielding efficiency is an important area for future study.

4.3. Other Metal-Based Alloys

In the 2000s, many in-depth studies on Ni-Cr-Mo-Gd alloys for neutron shielding materials were conducted in the United States for the Yucca Mountain Project [129]. Initially, Dupont et al. [130] investigated the effects of Gd on the solidification behavior and weldability of a Ni-Cr-Mo alloy and discovered that the Ni-Cr-Mo-Gd alloy solidified in a way comparable to a binary eutectic system. The solidification of Ni-Cr-Mo-Gd alloy began with a primary L → γ reaction and ended with a eutectic type L → γ + Ni5Gd reaction at 1258 °C. In addition, the sensitivity to solidification cracking of Ni-Cr-Mo-Gd alloy was highest with 1 wt%Gd and declined with further Gd additions. Mizia et al. [131] found that the strength levels of Ni-Cr-Mo-Gd alloy were 10% higher in both longitudinal and transverse orientations than the base (Gd-free) Ni-Cr-Mo alloy. The corrosion resistance of Ni-Cr-Mo-Gd alloy was discovered to be related to the Gd amount, which determined the quantity of gadolinide (Ni5Gd) presented in the alloy. Likewise, Lister et al. [132] evaluated the corrosion resistance of Ni-Cr-Mo-Gd alloys. In electrochemical and longer-term immersion tests, the Ni5Gd phase was stable in both conventional and simulated test conditions. Nevertheless, further tests in more corrosive environments and at higher temperatures are required to fully ascertain its corrosion resistance and thermal stability.
However, the high Mo content in Ni-Cr-Mo-Gd alloys poses a limitation to their promotion and application considering their high manufacturing costs. Therefore, Zhang et al. [133] proposed a Ni-Cr-Fe-Gd alloy (i.e., Modified-690Gd), and this alloy was composed primarily of a γ-austenite matrix with a (Ni, Cr, Fe)5Gd phase featuring hexagonal crystal structure along grain boundaries. The (Ni, Cr, Fe)5Gd phase content increased with higher Gd concentration, resulting in reduced fracture elongation of the alloy. Despite this, the Modified-690Gd alloy showed good tensile strength exceeding 620 MPa and a minimum elongation of 50%, indicating excellent mechanical properties. Additionally, the Ni-Cr-Fe-Gd alloy displayed favorable corrosion resistance, and the (Ni, Cr, Fe)5Gd phase within the alloy was preferentially attacked and removed during the corrosion process. When the Gd content in the alloy is 2.0 wt% and the thickness exceeds 2 mm, the neutron transmittance can be significantly reduced to values below 0.01. It is worth noting that when subjected to identical neutron shielding test conditions, the Modified-690Gd alloy demonstrated a lower thickness in comparison with B-containing alloys, providing size advantages as neutron shielding components.
Based on the outstanding mechanical properties and corrosion resistance of Ti, Hur et al. [134] developed a novel Ti-Gd alloy with varying Gd contents. Herein, the Gd-rich phases were discovered randomly dispersed inside the Ti matrix, and the average chemical composition of the particles was 95.96 wt%Gd, 3.71 wt%Ti, and 0.32 wt%O. In the rolling direction, the elongated phases showed no signs of brittle fracture or gaps between the phases and the Ti matrix, suggesting excellent deformability of Gd-rich phases. In the simulated SNF pool water containing boric acid at 50 °C, the corrosion tests found that Ti-Gd alloy had lower corrosion potentials, higher corrosion current density values, and higher passive current density values than pure Ti. This study provided valuable information on the corrosion behavior of Ti-Gd alloy in wet storage. Unfortunately, the mechanical properties and neutron shielding performance of this alloy were not evaluated there.

5. Summary

Gd has outstanding neutron shielding ability, primarily against thermal neutrons, and it has been widely used as a neutron absorber in inorganic nonmetallic-based, polymer-based, and metallic-based shielding materials. In inorganic nonmetallic neutron shielding materials, the major Gd-containing form is Gd2O3. Among polymer neutron shielding materials, Gd can exist in the form of Gd-related structures (i.e., Gd-MOF and Gd@MXene). Furthermore, Gd is found in metal shielding materials as Gd-containing phases (e.g., (Fe, Ni, Cr)3Gd, Ni5Gd, and Al5Gd3O12).
Depending on the features of the matrix, different types of materials can be applied to various shielding scenarios. Inorganic nonmetallic materials are frequently employed in the building and construction industries. In contrast, given the lightweight and flexible nature of polymers, they are more suitable for use as wearable protective devices. Metallic materials are preferred for storing and transporting SNFs due to their high thermal stability, mechanical strength, and corrosion resistance. The incorporation of Gd substantially improves the neutron shielding performance of these materials, allowing for enhanced neutron shielding efficiency at reduced thickness and restricted geometrical size. After the review of the available reports on the novel Gd-containing shielding materials, there are still several challenges remaining in this area:
  • There is a mutually restrictive relationship between neutron shielding performance and the mechanical properties of Gd-containing shielding materials. Increasing the Gd content has the potential to improve the thermal neutron absorbing abilities of shielding materials, but excessive Gd content may lead to a reduction in mechanical properties (e.g., plasticity and strength). Furthermore, the high quantities of Gd may complicate the manufacture of shielding materials, e.g., the powder metallurgy method plus the deformation technique for metal-based materials. Therefore, the manipulations of the preparation procedures, neutron shielding performance, and mechanical properties are tough but critical. Possibly, a potential research direction is to establish theoretical models capable of determining the appropriate Gd addition for the controllable fabrication process, thus obtaining a trade-off between neutron shielding performance and mechanical properties.
  • Although Gd possesses a high thermal neutron absorbing ability, its shielding efficiency against fast neutrons is negligible. Moreover, the absorption of neutrons by Gd is accompanied by the release of secondary γ-rays, thereby posing a potential radiation hazard. In other words, both neutron and γ-ray radiation may occur simultaneously in practical applications. Therefore, shielding materials must have effective neutron shielding abilities for both fast and thermal neutrons, as well as adequate γ-ray radiation shielding performance. H-rich materials (e.g., polymers) and concretes are considered to be the optimal choice for attenuating fast neutrons. Thus, they can be used as the matrix to improve the overall neutron shielding performance of Gd-containing shielding material. For the deficiency of Gd-containing materials in shielding γ-ray radiation, incorporating components with high γ-ray radiation shielding properties (e.g., W, concrete, steel) is feasible. For instance, the Gd2O3@W core–shell structure with a W outer layer has proven to be promising for both neutron and γ-ray radiation shielding.
  • It is still challenging to improve the neutron shielding performance of Gd-containing materials. In this regard, regulating the form of Gd in the matrix and ensuring a homogenous distribution of Gd components is critical in the fabrication of Gd-containing shielding materials. Specifically, it is important to modify the preparation process of shielding materials to improve the homogeneity of the distribution of Gd-containing phases, promote the generation of reinforcing phases that are beneficial to the material properties, and avoid the appearance of harmful secondary phases. In addition, there are some innovative material preparation technologies, e.g., additive manufacturing, that allow for the fabrication of custom-designed samples with more freedom. This novel material processing method may give more options for meeting the demands of practical applications.
Currently, there are some problems in the development and application of Gd-containing shielding materials. Besides, it is noteworthy that the cost of metallic Gd remains relatively expensive in comparison to other neutron absorbers (e.g., B4C). Despite this, there is no doubt that the novel Gd-containing materials have broad application prospects as neutron shielding materials. We hope that our work will inspire and guide further research in this field.

Author Contributions

Conceptualization: K.W., C.Y., Z.B. and M.W.; Formal analysis, L.M., Z.B., D.Z. and S.C.; Investigation, K.W., L.M., C.Y., D.Z., S.C., Z.C. and X.L.; Writing—original draft preparation, K.W., L.M., D.Z., S.C., Z.C. and X.L.; Writing—review and editing, C.Y., Z.B. and M.W., Project administration, C.Y., Z.B. and M.W.; Funding acquisition, M.W. and Z.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research is financially supported by the National Natural Science Foundation of China (Grant Nos. 52071207, 51971137, 11875192, U1930101, and U22A20174), the Natural Science Foundation of Shanghai (China, Grant No. 22ZR1432800), and the Chinese Scholarship Council (Grant No. 202106235025). Bian would like to acknowledge the financial support from Special Research Assistant Program of CAS and the technical support from the 320 kV platform for multi-discipline research with highly charged ions at the Institute of Modern Physics, CAS.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. I.A.E. Agency. IAEA Annual Report 2021; I.A.E. Agency: Vienna, Austria, 2021. [Google Scholar]
  2. Ewing, R.C. Long-term storage of spent nuclear fuel. Nat. Mater. 2015, 14, 252–257. [Google Scholar] [CrossRef]
  3. Ewing, R.C.; Weber, W.J.; Clinard, F.W., Jr. Radiation effects in nuclear waste forms for high-level radioactive waste. Prog. Nucl. Energy 1995, 29, 63–127. [Google Scholar] [CrossRef]
  4. Amaldi, E.; Fano, U.; Spencer, L.V.; Berger, M.J. Neutrons and Related Gamma Ray Problems; Springer: Berlin/Heidelberg, Germany, 2012. [Google Scholar]
  5. Tyagi, G.; Singhal, A.; Routroy, S.; Bhunia, D.; Lahoti, M. Radiation Shielding Concrete with alternate constituents: An approach to address multiple hazards. J. Hazard. Mater. 2021, 404 Pt B, 124201. [Google Scholar] [CrossRef]
  6. Suzuki, K.; Yamazaki, S.; Iwata, K.-I.; Yamada, Y.; Morioka, T.; Daino, K.; Kaminishi, M.; Ogawa, M.; Shimada, Y.; Kakinuma, S. Lung-Cancer Risk in Mice after Exposure to Gamma Rays, Carbon Ions or Neutrons: Egfr Pathway Activation and Frequent Nuclear Abnormality. Radiat. Res. 2022, 198, 475–487. [Google Scholar] [CrossRef]
  7. Burgio, E.; Piscitelli, P.; Migliore, L. Ionizing Radiation and Human Health: Reviewing Models of Exposure and Mechanisms of Cellular Damage. An Epigenetic Perspective. Int. J. Environ. Res. Public Health 2018, 15, 1971. [Google Scholar] [CrossRef] [Green Version]
  8. Koo, Y.-H.; Yang, Y.-S.; Song, K.-W. Radioactivity release from the Fukushima accident and its consequences: A review. Prog. Nucl. Energy 2014, 74, 61–70. [Google Scholar] [CrossRef]
  9. Wood, J. Chapter 6—Neutron Attenuation. In Computational Methods in Reactor Shielding; Wood, J., Ed.; International Atomic Energy Agency: Pergamon, Turkey, 1982; pp. 192–269. [Google Scholar]
  10. Gibson, J.A.B.; Piesch, E.; Agency, I.A.E. Neutron Monitoring for Radiological Protection: A Manual; International Atomic Energy Agency: Pergamon, Turkey, 1985. [Google Scholar]
  11. Jing, H.; Geng, L.; Qiu, S.; Zou, H.; Liang, M.; Deng, D. Research progress of rare earth composite shielding materials. J. Rare Earth 2023, 41, 32–41. [Google Scholar] [CrossRef]
  12. Shusterman, J.A.; Scielzo, N.D.; Thomas, K.J.; Norman, E.B.; Lapi, S.E.; Loveless, C.S.; Peters, N.J.; Robertson, J.D.; Shaughnessy, D.A.; Tonchev, A.P. The surprisingly large neutron capture cross-section of Zr-88. Nature 2019, 565, 328. [Google Scholar] [CrossRef] [PubMed]
  13. Machiels, A.; Lambert, R. Handbook of Neutron Absorber Materials for Spent Nuclear Fuel Transportation and Storage Applications; Electric Power Research Institute, Inc.: Palo Alto, CA, USA, 2009; pp. 101–110. [Google Scholar]
  14. Robards, K.; Worsfold, P. Cadmium: Toxicology and analysis. A review. Analyst 1991, 116, 549–568. [Google Scholar] [CrossRef]
  15. Florez, R.; Colorado, H.A.; Giraldo, C.H.C.; Alajo, A. Preparation and characterization of Portland cement pastes with Sm2O3 microparticle additions for neutron shielding applications. Constr. Build. Mater. 2018, 191, 498–506. [Google Scholar] [CrossRef]
  16. Cuer, P.; Lattes, C.M.G. Radioactivity of Samarium. Nature 1946, 158, 197–198. [Google Scholar] [CrossRef] [PubMed]
  17. Barth, R.F.; Mi, P.; Yang, W. Boron delivery agents for neutron capture therapy of cancer. Cancer Commun. 2018, 38, 35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Jung, Y.-J.; Hwang, J.; Yeon, J.-W.; Boo, B.H.; Song, K. Dispersion properties of B4C microparticles as emergency neutron absorbers in Spent-Fuel pool water. Nucl. Sci. Eng. 2012, 172, 202–207. [Google Scholar] [CrossRef]
  19. Basturk, M.; Arztmann, J.; Jerlich, W.; Kardjilov, N.; Lehmann, E.; Zawisky, M. Analysis of neutron attenuation in boron-alloyed stainless steel with neutron radiography and JEN-3 gauge. J. Nucl. Mater. 2005, 341, 189–200. [Google Scholar] [CrossRef]
  20. Usta, M.; Tozar, A. The effect of the ceramic amount on the radiation shielding properties of metal-matrix composite coatings. Radiat. Phys. Chem. 2020, 177, 109086. [Google Scholar] [CrossRef]
  21. Li, Y.; Wang, W.; Zhou, J.; Chen, H.; Zhang, P. 10B areal density: A novel approach for design and fabrication of B4C/6061Al neutron absorbing materials. J. Nucl. Mater. 2017, 487, 238–246. [Google Scholar] [CrossRef]
  22. Abenojar, J.; Martinez, M.A.; Velasco, F. Effect of the boron content in the aluminium/boron composite. J. Alloys Compd. 2006, 422, 67–72. [Google Scholar] [CrossRef]
  23. Zhang, F.; Wang, X.; Wierschke, J.B.; Wang, L. Helium bubble evolution in ion irradiated Al/B4C metal metrix composite. Scr. Mater. 2015, 109, 28–33. [Google Scholar] [CrossRef] [Green Version]
  24. Dunning, J.R.; Pegram, G.B.; Fink, G.A.; Mitchell, D.P. Interaction of Neutrons with Matter. Phys. Rev. 1935, 48, 265–280. [Google Scholar] [CrossRef]
  25. Lapp, R.E.; VanHorn, J.R.; Dempster, A.J. The neutron absorbing isotopes in gadolinium and samarium. Phys. Rev. 1947, 71, 745–747. [Google Scholar] [CrossRef]
  26. Leinweber, G.; Barry, D.P.; Trbovich, M.J.; Burke, J.A.; Drindak, N.J.; Knox, H.D.; Ballad, R.V.; Block, R.C.; Danon, Y.; Severnyak, L.I. Neutron capture and total cross-section measurements and resonance parameters of gadolinium. Nucl. Sci. Eng. 2006, 154, 261–279. [Google Scholar] [CrossRef]
  27. Burke, D.S.; Byun, S.H. Feasibility of gadolinium oxide paint as neutron shielding. Nucl. Instrum. Methods Phys. Res. Sect. A Accel. Spectrometers Detect. Assoc. Equip. 2022, 1025, 166–175. [Google Scholar] [CrossRef]
  28. Kandlakunta, P.; Cao, L.R.; Mulligan, P. Measurement of internal conversion electrons from Gd neutron capture. Nucl. Instrum. Methods Phys. Res. Sect. A Accel. Spectrometers, Detect. Assoc. Equip. 2013, 705, 36–41. [Google Scholar] [CrossRef]
  29. Bashter, I.I. Calculation of radiation attenuation coefficients for shielding concretes. Ann. Nucl. Energy 1997, 24, 1389–1401. [Google Scholar] [CrossRef]
  30. Oto, B.; Gür, A.; Kavaz, E.; Çakır, T.; Yaltay, N. Determination of gamma and fast neutron shielding parameters of magnetite concretes. Prog. Nucl. Energy 2016, 92, 71–80. [Google Scholar] [CrossRef]
  31. Kharita, M.H.; Yousef, S.; AlNassar, M. The effect of the initial water to cement ratio on shielding properties of ordinary concrete. Prog. Nucl. Energy 2010, 52, 491–493. [Google Scholar] [CrossRef]
  32. Thomas, C.; Rico, J.; Tamayo, P.; Setién, J.; Ballester, F.; Polanco, J.A. Neutron shielding concrete incorporating B4C and PVA fibers exposed to high temperatures. J. Build. Eng. 2019, 26, 100859. [Google Scholar] [CrossRef]
  33. Park, J.S.; Kim, J.; Yi, S.H. Effects of Gadolinium in Fe based amorphous ribbons with high boron contents on the neutron shielding efficiency. Ann. Nucl. Energy 2017, 109, 365–369. [Google Scholar] [CrossRef]
  34. Seshadri, B.S. Transmission measurements of D-T neutrons through gadolinium loaded concrete and polypropylene. J. Nucl. Sci. Technol. 1989, 26, 881–886. [Google Scholar] [CrossRef]
  35. Piotrowski, T.; Tefelski, D.B.; Sokołowska, J.J.; Jaworska, B. NGS-concrete-new generation shielding concrete against ionizing radiation-the potential evaluation and preliminary investigation. Acta Phys. Pol. A 2015, 128, 9–13. [Google Scholar] [CrossRef]
  36. Zalegowski, K.; Piotrowski, T.; Garbacz, A.; Adamczewski, G. Relation between microstructure, technical properties and neutron radiation shielding efficiency of concrete. Constr. Build. Mater. 2020, 235, 117389. [Google Scholar] [CrossRef]
  37. Piotrowski, T.; Glinicka, J.; Glinicki, M.A.; Prochoń, P. Influence of gadolinium oxide and ulexite on cement hydration and technical properties of mortars for neutron radiation shielding purposes. Constr. Build. Mater. 2019, 195, 583–589. [Google Scholar] [CrossRef]
  38. Katagiri, M.; Sakasai, K.; Matsubayashi, M.; Kojima, T. Neutron/gamma-ray discrimination characteristics of novel neutron scintillators. Nucl. Instrum. Methods Phys. Res. Sect. A Accel. Spectrometers Detect. Assoc. Equip. 2004, 529, 317–320. [Google Scholar] [CrossRef]
  39. Tekin, H.O.; Altunsoy, E.E.; Kavaz, E.; Sayyed, M.I.; Agar, O.; Kamislioglu, M. Photon and neutron shielding performance of boron phosphate glasses for diagnostic radiology facilities. Results Phys. 2019, 12, 1457–1464. [Google Scholar] [CrossRef]
  40. Kaewjang, S.; Maghanemi, U.; Kothan, S.; Kim, H.J.; Limkitjaroenporn, P.; Kaewkhao, J. New gadolinium based glasses for gamma-rays shielding materials. Nucl. Eng. Des. 2014, 280, 21–26. [Google Scholar] [CrossRef]
  41. Ruengsri, S.; Kaewkhao, J.; Limkitjaroenporm, P.; Meejitpaisan, P.; Hongtong, W.; Cheewasukhanont, W. Development of gadolinium calcium phosphate oxyfluoride glass for radiation shielding materials. Integr. Ferroelectr. 2017, 177, 48–58. [Google Scholar] [CrossRef]
  42. Singh, V.P.; Badiger, N.M.; Kothan, S.; Kaewjaeng, S.; Korkut, T.; Kim, H.J.; Kaewkhao, J. Gamma-ray and neutron shielding efficiency of Pb-free gadolinium-based glasses. Nucl. Sci. Tech. 2016, 27, 103. [Google Scholar] [CrossRef]
  43. Mahmoud, K.A.; El-Agwany, F.I.; Rammah, Y.S.; Tashlykov, O.L. Gamma ray shielding capacity and build up factors of CdO doped lithium borate glasses: Theoretical and simulation study. J. Non. Cryst. Solids 2020, 541, 120110. [Google Scholar] [CrossRef]
  44. Al-Buriahi, M.S.; Tonguç, B.; Perişanoğlu, U.; Kavaz, E. The impact of Gd2O3 on nuclear safety proficiencies of TeO2–ZnO–Nb2O5 glasses: A GEANT4 Monte Carlo study. Ceram. Int. 2020, 46, 23347–23356. [Google Scholar] [CrossRef]
  45. Al-Hadeethi, Y.; Sayyed, M.I.; Agar, O. Ionizing photons attenuation characterization of quaternary tellurite–zinc–niobium–gadolinium glasses using Phy-X/PSD software. J. Non. Cryst. Solids 2020, 538, 120044. [Google Scholar] [CrossRef]
  46. Ibrahim, S.; El-Agawany, F.I.; Rammah, Y.S.; Ahmed, E.M.; Ali, A.A. ZnO-Bi2O3-B2O3 glasses doped with rare earth oxides: Synthesis, physical, structural characteristics, neutron and photon attenuation attitude. Optik 2021, 243, 167414. [Google Scholar] [CrossRef]
  47. Lakshminarayana, G.; Tekin, H.O.; Dong, M.G.; Al-Buriahi, M.S.; Lee, D.E.; Yoon, J.; Park, T. Comparative assessment of fast and thermal neutrons and gamma radiation protection qualities combined with mechanical factors of different borate-based glass systems. Results Phys. 2022, 37, 105527. [Google Scholar] [CrossRef]
  48. El-Khayatt, A.M.; Saudi, H.A. Recycling of waste porcelain into newly developed bismo-borate glass admixture with Gd3+ ions for nuclear radiation protection uses: An experimental and theoretical study. Radiat. Phys. Chem. 2023, 203, 110612. [Google Scholar] [CrossRef]
  49. Al-Hadeethi, Y.; Sayyed, M.I.; Kaewkhao, J.; Raffah, B.M.; Almalki, R.; Rajaramakrishna, R. An extensive investigation of physical, optical and radiation shielding properties for borate glasses modified with gadolinium oxide. Appl. Phys. A 2019, 125, 749. [Google Scholar] [CrossRef]
  50. Intachai, N.; Wantana, N.; Kaewjaeng, S.; Chaiphaksa, W.; Cheewasukhanont, W.; Htun, K.T.; Kothan, S.; Kim, H.J.; Kaewkhao, J. Effect of Gd2O3 on radiation shielding, physical and optical properties of sodium borosilicate glass system. Radiat. Phys. Chem. 2022, 199, 110361. [Google Scholar] [CrossRef]
  51. Shamshad, L.; Rooh, G.; Limkitjaroenporn, P.; Srisittipokakun, N.; Chaiphaksa, W.; Kim, H.J.; Kaewkhao, J. A comparative study of gadolinium based oxide and oxyfluoride glasses as low energy radiation shielding materials. Prog. Nucl. Energy 2017, 97, 53–59. [Google Scholar] [CrossRef]
  52. Barsoum, M. Fundamentals of Ceramics. In Fundamentals of Ceramics; CRC Press: Boca Raton, FL, USA, 2003; pp. 441–464. [Google Scholar]
  53. Akman, F.; Khattari, Z.Y.; Kaçal, M.R.; Sayyed, M.I.; Afaneh, F. The radiation shielding features for some silicide, boride and oxide types ceramics. Radiat. Phys. Chem. 2019, 160, 9–14. [Google Scholar] [CrossRef]
  54. Ge, R.; Zhang, Y.; Liu, Y.; Fang, J.; Luan, W.; Wu, G. Effect of Gd2O3 addition on mechanical, thermal and shielding properties of Al2O3 ceramics. J. Mater. Sci. Mater. Electron. 2016, 28, 5898–5905. [Google Scholar] [CrossRef]
  55. Hernández, M.F.; Herrera, M.S.; Anaya, R.; Martinez, J.M.; Cipollone, M.; Conconi, M.S.; Rendtorff, N.M. High macroscopic neutron capture cross section ceramics based on bauxite and Gd2O3. Sci. Sinter. 2020, 52, 387–403. [Google Scholar] [CrossRef]
  56. Mhareb, M.H.A.; Slimani, Y.; Alajerami, Y.S.; Sayyed, M.I.; Lacomme, E.; Almessiere, M.A. Structural and radiation shielding properties of BaTiO3 ceramic with different concentrations of Bismuth and Ytterbium. Ceram. Int. 2020, 46, 28877–28886. [Google Scholar] [CrossRef]
  57. Dong, C.; Li, Y.; Li, S.; Jia, W.; Chen, R.; Lao, D.; Xia, X. Thermally insulating GdBO3 ceramics with neutron shielding performance. Int. J. Appl. Ceram. Technol. 2022, 19, 1428–1438. [Google Scholar] [CrossRef]
  58. Li, Y.; Luo, H.; He, Z.; Xiang, R.; Jia, W.; Li, S.; Lao, D.; Zhou, Z.; Dong, C. Structural stability and neutron-shielding capacity of GdBO3-Al18B4O33 composite ceramics: Experimental investigation and numerical simulation. Ceram. Int. 2021, 47, 20935–20947. [Google Scholar] [CrossRef]
  59. Akrami, S.; Edalati, P.; Fuji, M.; Edalati, K. High-entropy ceramics: Review of principles, production and applications. Mater. Sci. Eng. R Rep. 2021, 146, 100644. [Google Scholar] [CrossRef]
  60. Zhang, X.; Li, Y.; Li, C.; Yang, F.; Jiang, Z.; Xue, L.; Shao, Z.; Zhao, Z.; Xie, M.; Yu, S. A novel (La0.2Ce0.2Gd0.2Er0.2Tm0.2)2(WO4)3 high-entropy ceramic material for thermal neutron and gamma-ray shielding. Mater. Des. 2021, 205, 109722. [Google Scholar] [CrossRef]
  61. Wang, G.; Zhang, J.; Shen, S.; Zhong, L.; Mei, L.; Tang, Z. 3D printing of gadolinium oxide structure neutron absorber. Ceram. Int. 2022, 48, 35198–35208. [Google Scholar] [CrossRef]
  62. Hayashi, T.; Tobita, K.; Nakamori, Y.; Orimo, S. Advanced neutron shielding material using zirconium borohydride and zirconium hydride. J. Nucl. Mater. 2009, 386–388, 119–121. [Google Scholar] [CrossRef]
  63. Abuali Galehdari, N.; Kelkar, A.D. Effect of neutron radiation on the mechanical and thermophysical properties of nanoengineered polymer composites. J. Mater. Res. 2017, 32, 426–434. [Google Scholar] [CrossRef]
  64. Prabhu, S.; Bubbly, S.G.; Gudennavar, S.B. X-Ray and γ-Ray Shielding Efficiency of Polymer Composites: Choice of Fillers, Effect of Loading and Filler Size, Photon Energy and Multifunctionality. Polym. Rev. 2022, 63, 246–288. [Google Scholar] [CrossRef]
  65. More, C.V.; Alsayed, Z.; Badawi, M.S.; Thabet, A.A.; Pawar, P.P. Polymeric composite materials for radiation shielding: A review. Environ. Chem. Lett. 2021, 19, 2057–2090. [Google Scholar] [CrossRef]
  66. Nambiar, S.; Yeow, J.T. Polymer-composite materials for radiation protection. ACS Appl. Mater. Interfaces 2012, 4, 5717–5726. [Google Scholar] [CrossRef]
  67. Yastrebinskii, R.N.; Bondarenko, G.G.; Pavlenko, V.I. Transport packing set for radioactive waste based on a radiation-protective polymeric matrix. Inorg. Mater. Appl. Res. 2015, 6, 473–478. [Google Scholar] [CrossRef]
  68. Gupta, S.C.; Baheti, G.L.; Gupta, B.P. Application of hydrogel system for neutron attenuation. Radiat. Phys. Chem. 2000, 59, 103–107. [Google Scholar] [CrossRef]
  69. Tiamduangtawan, P.; Wimolmala, E.; Meesat, R.; Saenboonruang, K. Effects of Sm2O3 and Gd2O3 in poly (vinyl alcohol) hydrogels for potential use as self-healing thermal neutron shielding materials. Radiat. Phys. Chem. 2020, 172, 108818. [Google Scholar] [CrossRef]
  70. Wang, C.; Wang, S.; Zhang, Y.; Wang, Z.; Liu, J.; Zhang, M. Self-polymerization and co-polymerization kinetics of gadolinium methacrylate. J. Rare Earth 2018, 36, 298–303. [Google Scholar] [CrossRef]
  71. Wang, C.H.; Hu, L.M.; Wang, Z.F.; Zhang, M. Electrospun and in situ self-polymerization of polyacrylonitrile containing gadolinium nanofibers for thermal neutron protection. Rare Met. 2019, 38, 252–258. [Google Scholar] [CrossRef]
  72. Wang, B.; Guo, X.; Yuan, L.; Fang, Q.; Wang, X.; Qiu, T.; Lai, C.; Wang, Q.; Liu, Y. Micro gadolinium oxide dispersed flexible composites developed for the shielding of thermal neutron/gamma rays. Nucl. Eng. Technol. 2023, 55, 1763–1774. [Google Scholar] [CrossRef]
  73. Wang, S.; Urban, M.W. Self-healing polymers. Nat. Rev. Mater. 2020, 5, 562–583. [Google Scholar] [CrossRef]
  74. Poltabtim, W.; Thumwong, A.; Wimolmala, E.; Rattanapongs, C.; Tokonami, S.; Ishikawa, T.; Saenboonruang, K. Dual X-ray- and Neutron-Shielding Properties of Gd2O3/NR Composites with Autonomous Self-Healing Capabilities. Polymers 2022, 14, 4481. [Google Scholar] [CrossRef]
  75. Bu, F.; Zagho, M.M.; Ibrahim, Y.; Ma, B.; Elzatahry, A.; Zhao, D. Porous MXenes: Synthesis, structures, and applications. Nano Today 2020, 30, 100803. [Google Scholar] [CrossRef]
  76. Zhu, X.; Zhang, X.; Guo, S. Constructing oriented two-dimensional fish scale-like Gd@MXene barrier walls in polyvinyl alcohol to achieve excellent neutron shielding properties. Nanoscale 2022, 14, 10581–10593. [Google Scholar] [CrossRef]
  77. Furukawa, H.; Cordova, K.E.; O’Keeffe, M.; Yaghi, O.M. The chemistry and applications of metal-organic frameworks. Science 2013, 341, 1230444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Hu, C.; Zhai, Y.T.; Song, L.L.; Mao, X.D. Structure-thermal activity relationship in a novel polymer/MOF-based neutron-shielding material. Polym. Compos. 2020, 41, 1418–1427. [Google Scholar] [CrossRef]
  79. Hu, C.; Huang, Q.; Zhai, Y. Thermal, mechanical investigation and neutron shielding analysis for Gd-MOF/polyimide materials. RSC Adv. 2021, 11, 40148–40158. [Google Scholar] [CrossRef] [PubMed]
  80. Soltani, Z.; Beigzadeh, A.; Ziaie, F.; Asadi, E. Effect of particle size and percentages of Boron carbide on the thermal neutron radiation shielding properties of HDPE/B4C composite: Experimental and simulation studies. Radiat. Phys. Chem. 2016, 127, 182–187. [Google Scholar] [CrossRef]
  81. Li, R.; Gu, Y.; Wang, Y.; Yang, Z.; Li, M.; Zhang, Z. Effect of particle size on gamma radiation shielding property of gadolinium oxide dispersed epoxy resin matrix composite. Mater. Res. Express 2017, 4, 035035. [Google Scholar] [CrossRef]
  82. İrim, Ş.G.; Wis, A.A.; Keskin, M.A.; Baykara, O.; Ozkoc, G.; Avcı, A.; Doğru, M.; Karakoç, M. Physical, mechanical and neutron shielding properties of h-BN/Gd2O3/HDPE ternary nanocomposites. Radiat. Phys. Chem. 2018, 144, 434–443. [Google Scholar] [CrossRef]
  83. Baykara, O.; İrim, Ş.G.; Wis, A.A.; Keskin, M.A.; Ozkoc, G.; Avcı, A.; Doğru, M. Polyimide nanocomposites in ternary structure: “A novel simultaneous neutron and gamma-ray shielding material”. Polym. Adv. Technol. 2020, 31, 2466–2479. [Google Scholar] [CrossRef]
  84. Kavun, Y.; Eskalen, H.; Kerlï, S.; Kavgaci, M. Fabrication and characterization of GdxFe2O3(100-x)/PVA (x = 0, 5, 10, 20) composite films for radiation shielding. Appl. Radiat. Isot. 2021, 177, 109918. [Google Scholar] [CrossRef]
  85. He, Y.; Li, P.; Ren, H.; Zhang, X.; Chen, X.; Yan, Y. The fabrication and neutron shielding property of polyphenylene sulfide containing salicylic acid/Gd2O3 composites. High Perform. Polym. 2022, 34, 105–114. [Google Scholar] [CrossRef]
  86. Huo, Z.; Zhao, S.; Zhong, G.; Zhang, H.; Hu, L. Surface modified-gadolinium/boron/polyethylene composite with high shielding performance for neutron and gamma-ray. Nucl. Mater. Energy 2021, 29, 101095. [Google Scholar] [CrossRef]
  87. Macfarlane, A. The problem of used nuclear fuel: Lessons for interim solutions from a comparative cost analysis. Energy Policy 2001, 29, 1379–1389. [Google Scholar] [CrossRef]
  88. Castley, D.; Goodwin, C.; Liu, J. Computational and experimental comparison of boron carbide, gadolinium oxide, samarium oxide, and graphene platelets as additives for a neutron shield. Radiat. Phys. Chem. 2019, 165, 108435. [Google Scholar] [CrossRef]
  89. Lee, S.-W.; Ahn, J.-H.; Moon, B.-M.; Kim, D.; Oh, S.; Kim, Y.-J.; Jung, H.-D. Preliminary study on Fe-Gd Alloy. as binary Alloy. and master Alloy. for potential spent nuclear fuel (SNF) application. Mater. Des. 2020, 194, 108906. [Google Scholar] [CrossRef]
  90. Sahin, S.; Uebeyli, M. A review on the potential use of austenitic stainless steels in nuclear fusion reactors. J. Fusion. Energy 2008, 27, 271–277. [Google Scholar] [CrossRef]
  91. Sun, W.-Q.; Hu, G.; Yu, X.-H.; Shi, J.; Xu, H.; Wu, R.-J.; He, C.; Yi, Q.; Hu, H.-S. Study on a High-Boron-Content Stainless Steel Composite for Nuclear Radiation. Materials 2021, 14, 7004. [Google Scholar] [CrossRef] [PubMed]
  92. Loria, E.A.; Isaacs, H.S. Type 304 Stainless Steel With 0.5% Boron for Storage of Spent Nuclear Fuel. JOM 1980, 32, 10–17. [Google Scholar] [CrossRef]
  93. Kato, H.; Armantrout, C.; Barstow, W.; Copeland, M.I. Stainless Steel-Gadolinium Alloy; U.S. Dept. of the Interior, Bureau of Mines: Washington, DC, USA, 1965; 29p.
  94. Choi, Y.; Moon, B.M.; Sohn, D.S. Fabrication of Gd containing duplex stainless steel sheet for neutron absorbing structural materials. Nucl. Eng. Technol. 2013, 45, 689–694. [Google Scholar] [CrossRef]
  95. Choi, Y.; Baik, Y.; Moon, B.M.; Sohn, D.S. Corrosion and Wear Properties of Cold Rolled 0.087% Gd Lean Duplex Stainless Steels for Neutron Absorbing Material. Nucl. Eng. Technol. 2016, 48, 164–168. [Google Scholar] [CrossRef]
  96. Baik, Y.; Choi, Y.; Moon, B.M.; Sohn, D.S.; Bogdanov, S.G.; Pirogov, A.N. Effect of gadolinium addition on the corrosion, wear, and neutron absorbing behaviors of duplex stainless steel sheet. Phys. Met. Metallogr. 2015, 116, 1135–1142. [Google Scholar] [CrossRef]
  97. Niu, G.; Zurob, H.S.S.; Misra, R.D.K.; Tang, Q.; Zhang, Z.; Nguyen, M.-T.; Wang, L.; Wu, H.; Zou, Y. Superior fracture toughness in a high-strength austenitic steel with heterogeneous lamellar microstructure. Acta Mater. 2022, 226, 117642. [Google Scholar] [CrossRef]
  98. Jang, J.H.; Kang, J.Y.; Kim, S.D. Development of GD-containing austenitic stainless steel. J. Nucl. Mater. 2023, 574, 154197. [Google Scholar] [CrossRef]
  99. Robino, C.V.; DuPont, J.N.; Mizia, R.E.; Michael, J.R.; Williams, D.B.; Shaber, E. Development of Gd-enriched Alloy. for spent nuclear fuel applications—Part 1: Preliminary characterization of small scale Gd-enriched stainless steels. J. Mater. Eng. Perform. 2003, 12, 206–214. [Google Scholar] [CrossRef]
  100. DuPont, J.N.; Robino, C.V.; Michael, J.; Mizia, R.E.; Williams, D. Physical and welding metallurgy of Gd-enriched austenitic Alloys for spent nuclear fuel applications—Part I: Stainless steel Alloys. Weld. J. 2004, 83, 289-S–300-S. [Google Scholar]
  101. Wang, H.; Wang, T.; Peng, J. Effect of Gadolinium Addition on Microstructure, Mechanical Properties and Corrosion Resistance of 316L Austenitic Stainless Steels. Phys. Met. Metallogr. 2022, 122, 1640–1647. [Google Scholar] [CrossRef]
  102. Suryanarayana, C. Mechanical alloying and milling. Prog. Mater. Sci. 2001, 46, 1–184. [Google Scholar] [CrossRef]
  103. Yang, X.; Song, L.; Chang, B.; Yang, Q.; Mao, X.; Huang, Q. Development of Gd-Si-O dispersed 316L stainless steel for improving neutron shielding performance. Nucl. Mater. Energy 2020, 23, 100739. [Google Scholar] [CrossRef]
  104. Wan, S.; Wang, W.; Chen, H.; Zhou, J.; Zhang, Y.; Liu, R.; Feng, R. 155/157Gd areal density: A model for design and fabrication of Gd2O3/316L novel neutron shielding composites. Vacuum 2020, 176, 109304. [Google Scholar] [CrossRef]
  105. Wang, W.; Zhang, J.; Wan, S.; Zhang, T. Design, fabrication and comprehensive properties of the novel thermal neutron shielding Gd/316L composites. Fusion Eng. Des. 2021, 171, 112566. [Google Scholar] [CrossRef]
  106. Qi, Z.D.; Yang, Z.; Yang, X.G.; Wang, L.Y.; Li, C.Y.; Dai, Y. Performance study and optimal design of Gd/316L neutron absorbing material for Spent Nuclear Fuel transportation and storage. Mater. Today Commun. 2023, 34, 105342. [Google Scholar] [CrossRef]
  107. Wang, Z.; Wang, M.; Li, Y.; Xiao, H.; Chen, H.; Geng, J.; Li, X.; Chen, D.; Wang, H. Effect of pretreatment on microstructural stability and mechanical property in a spray formed Al-Zn-Mg-Cu alloy. Mater. Des. 2021, 203, 109618. [Google Scholar] [CrossRef]
  108. Park, J.-J.; Hong, S.-M.; Lee, M.-K.; Rhee, C.-K.; Rhee, W.-H. Enhancement in the microstructure and neutron shielding efficiency of sandwich type of 6061Al–B4C composite material via hot isostatic pressing. Nucl. Eng. Des. 2015, 282, 1–7. [Google Scholar] [CrossRef]
  109. Zubcak, M.; Soltes, J.; Zimina, M.; Weinberger, T.; Enzinger, N. Investigation of Al-B4C Metal Matrix Composites Produced by Friction Stir Additive Processing. Metals 2021, 11, 2020. [Google Scholar] [CrossRef]
  110. Chen, H.S.; Wang, W.X.; Li, Y.L.; Zhang, P.; Nie, H.H.; Wu, Q.C. The design, microstructure and tensile properties of B4C particulate reinforced 6061Al neutron absorber composites. J. Alloys Compd. 2015, 632, 23–29. [Google Scholar] [CrossRef]
  111. Zhang, P.; Li, Y.; Wang, W.; Gao, Z.; Wang, B. The design, fabrication and properties of B4C/Al neutron absorbers. J. Nucl. Mater. 2013, 437, 350–358. [Google Scholar] [CrossRef]
  112. Xu, Z.G.; Jiang, L.T.; Zhang, Q.; Qiao, J.; Gong, D.; Wu, G.H. The design of a novel neutron shielding B4C/Al composite containing Gd. Mater. Des. 2016, 111, 375–381. [Google Scholar] [CrossRef]
  113. Xu, Z.G.; Jiang, L.T.; Zhang, Q.; Qiao, J.; Wu, G.H. The microstructure and influence of hot extrusion on tensile properties of (Gd+B4C)/Al composite. J. Alloys Compd. 2017, 729, 1234–1243. [Google Scholar] [CrossRef]
  114. Xu, Z.G.; Jiang, L.T.; Zhang, Q.; Qiao, J.; Zhang, L.C.; He, P.; Ma, J.; Wu, G.H. The formation, evolution and influence of Gd-Containing phases in the (Gd+B4C)/6061Al composites during hot rolling. J. Alloys Compd. 2019, 775, 714–725. [Google Scholar] [CrossRef]
  115. Jiang, L.T.; Xu, Z.G.; Fei, Y.K.; Zhang, Q.; Qiao, J.; Wu, G.H. The design of novel neutron shielding (Gd+B4C)/6061Al composites and its properties after hot rolling. Compos. Part B Eng. 2019, 168, 183–194. [Google Scholar] [CrossRef]
  116. Xue, W.; Jiang, L.; Kang, P.; Yang, W.; Zhao, Q.; Xu, Z.; Zhang, N.; Liu, H.; Wu, G. Design and Fabrication of a Nanoamorphous Interface Layer in B4C/Al Composites to Improve Hot Deformability and Corrosion Resistance. ACS Appl. Nano Mat. 2020, 3, 5752–5761. [Google Scholar] [CrossRef]
  117. Khorshidi, R.; Honarbakhsh-Raouf, A.; Mahmudi, R. Microstructural evolution and high temperature mechanical properties of cast Al–15Mg2Si–xGd in situ composites. J. Alloys Compd. 2017, 700, 18–28. [Google Scholar] [CrossRef]
  118. Chen, M.; Liu, Z.; Yang, C.; Xiao, P.; Yang, W.; Hu, Z.; Sun, L.; Jia, Y. A novel (B4Cp+Gd)/Al6061 neutron absorber material with desirable mechanical properties. Mater. Sci. Eng. A 2022, 861, 144376. [Google Scholar] [CrossRef]
  119. Topcu, I.; Gulsoy, H.O.; Kadioglu, N.; Gulluoglu, A.N. Processing and mechanical properties of B4C reinforced Al matrix composites. J. Alloys Compd. 2009, 482, 516–521. [Google Scholar] [CrossRef]
  120. Ma, Y.; Addad, A.; Ji, G.; Zhang, M.-X.; Lefebvre, W.; Chen, Z.; Ji, V. Atomic-scale investigation of the interface precipitation in a TiB2 nanoparticles reinforced Al–Zn–Mg–Cu matrix composite. Acta Mater. 2020, 185, 287–299. [Google Scholar] [CrossRef]
  121. Chen, Y.; Chen, Y.; Zhai, H.; Wang, F.; Wang, M.; Chen, Z.; Zhong, S.; Li, X.; Wang, H. Structural-functional integrated TiB2/Al–Mg-Gd composite with efficient neutron shielding phases and superior mechanical properties. Intermet 2022, 148, 107630. [Google Scholar] [CrossRef]
  122. Zhang, P.; Li, J.; Wang, W.X.; Tan, X.Y.; Xie, L.; Guo, F.Y. The design, microstructure and mechanical properties of a novel Gd2O3/6061Al neutron shielding composite. Vacuum 2019, 162, 92–100. [Google Scholar] [CrossRef]
  123. Li, J.; Zhang, P.; Wang, W.-X.; Tan, X.-Y.; Ma, Y.-F.; Jia, C.-P. Microstructure evolution and strengthening mechanism of Gd2O3/6061Al neutron shielding composite during rolling. Vacuum 2020, 172, 109098. [Google Scholar] [CrossRef]
  124. Cong, S.; Li, Y.; Ran, G.; Zhou, W.; Feng, Q. Microstructure and its effect on mechanical and thermal properties of Al-based Gd2O3 MMCs used as shielding materials in spent fuel storage. Ceram. Int. 2020, 46, 12986–12995. [Google Scholar] [CrossRef]
  125. Wu, Y.; Cao, Y.; Wu, Y.; Li, D. Mechanical Properties and Gamma-Ray Shielding Performance of 3D-Printed Poly-Ether-Ether-Ketone/Tungsten Composites. Materials 2020, 13, 4475. [Google Scholar] [CrossRef]
  126. Cong, S.; Ran, G.; Li, Y.; Chen, Y. Ball-milling properties and sintering behavior of Al-based Gd2O3–W shielding materials used in spent-fuel storage. Powder Technol. 2020, 369, 127–136. [Google Scholar] [CrossRef]
  127. Zhang, P.; Jia, C.; Li, J.; Wang, W. Shielding composites for neutron and gamma-radiation with Gd2O3@W core-shell structured particles. Mater. Lett. 2020, 276, 128082. [Google Scholar] [CrossRef]
  128. Cong, S.; Li, Y.; Ran, G.; Zhou, W.; Dong, S.; Feng, Q. High-temperature corrosion behavior of hot-pressed Al-based Gd2O3-W shielding materials used in spent fuel storage. Corros. Sci. 2021, 179, 109166. [Google Scholar] [CrossRef]
  129. Mizia, R.E.; Lister, T.E. Accelerated testing of neutron-absorbing Alloy. for nuclear criticality control. Nucl. Technol. 2011, 176, 9–21. [Google Scholar] [CrossRef]
  130. Dupont, J.N.; Robino, C.V.; Anderson, T.D. Influence of Gd and B on solidification behaviour and weldability of Ni-Cr-Mo alloy. Sci. Technol. Weld. Joining 2008, 13, 550–565. [Google Scholar] [CrossRef]
  131. Mizia, R.E.; Lister, T.E.; Pinhero, P.J.; Trowbridge, T.L.; Hurt, W.L.; Robino, C.V.; Stephens, J.J., Jr.; Dupont, J.N. Development and testing of an advanced neutron-absorbing gadolinium alloy for spent nuclear fuel storage. Nucl. Technol. 2006, 155, 133–148. [Google Scholar] [CrossRef]
  132. Lister, T.E.; Mizia, R.E.; Pinhero, P.J.; Trowbridge, T.L.; Delezene-Briggs, K. Studies of the corrosion properties in Ni-Cr-Mo-Gd neutron-absorbing Alloys. Corrosion 2005, 61, 706–717. [Google Scholar] [CrossRef]
  133. Zhang, C.; Pan, J.; Wang, Z.; Wu, Z.; Mei, Q.; Ding, Q.; Gao, J.; Xiao, X. Gd effect on microstructure and properties of the Modified-690 alloy for function structure integrated thermal neutron shielding. Nucl. Eng. Technol. 2023, 55, 1541–1558. [Google Scholar] [CrossRef]
  134. Hur, D.H.; Chun, Y.B.; Park, S.Y. Corrosion behavior of neutron absorbing Ti-Gd Alloy in simulated spent nuclear fuel pool water. Corros. Sci. 2022, 209, 110776. [Google Scholar] [CrossRef]
Figure 1. The relative penetrative ability diagram of different radiations [5]. Reproduced with permission from Tyagi et al., Journal of Hazardous Materials; published by Elsevier, 2021.
Figure 1. The relative penetrative ability diagram of different radiations [5]. Reproduced with permission from Tyagi et al., Journal of Hazardous Materials; published by Elsevier, 2021.
Materials 16 04305 g001
Figure 2. Schematic diagram of the neutron shielding process: scattering and absorption.
Figure 2. Schematic diagram of the neutron shielding process: scattering and absorption.
Materials 16 04305 g002
Figure 3. (a) The number of publications in neutron shielding and Gd-related neutron shielding for the period spanning 1935–2022, based on the Scopus database. (b) The recent trend of Gd-related neutron shielding research ranging from 2007 to 2022 (magnified view in the blue box from Figure 3a).
Figure 3. (a) The number of publications in neutron shielding and Gd-related neutron shielding for the period spanning 1935–2022, based on the Scopus database. (b) The recent trend of Gd-related neutron shielding research ranging from 2007 to 2022 (magnified view in the blue box from Figure 3a).
Materials 16 04305 g003
Figure 4. Photographs of several gadolinium sodium borosilicate glasses with Gd2O3 content varying from 0 mol% to 15 mol% [50]. Reproduced with permission from Intachai et al., Radiation Physics and Chemistry; published by Elsevier, 2022.
Figure 4. Photographs of several gadolinium sodium borosilicate glasses with Gd2O3 content varying from 0 mol% to 15 mol% [50]. Reproduced with permission from Intachai et al., Radiation Physics and Chemistry; published by Elsevier, 2022.
Materials 16 04305 g004
Figure 5. Preparation, characteristics, and mechanical properties of as-printed Gd2O3 ceramics [61]. (a) Schematic of the vat photopolymerization 3D printing process for Gd2O3 ceramics. (b) Diagrams illustrating relative densities and linear shrinkage to sinter temperature. (c) The bending strength curves of as-printed ceramics as a function of strain with the increasing sintering temperature. Reproduced with permission from Wang et al., Ceramics International; published by Elsevier, 2022.
Figure 5. Preparation, characteristics, and mechanical properties of as-printed Gd2O3 ceramics [61]. (a) Schematic of the vat photopolymerization 3D printing process for Gd2O3 ceramics. (b) Diagrams illustrating relative densities and linear shrinkage to sinter temperature. (c) The bending strength curves of as-printed ceramics as a function of strain with the increasing sintering temperature. Reproduced with permission from Wang et al., Ceramics International; published by Elsevier, 2022.
Materials 16 04305 g005
Figure 6. Filler-reinforced polymer composite materials for effective radiation protection [66]. Reproduced with permission from Nambiar et al., Applied Materials; Copyright (2019) American Chemical Society.
Figure 6. Filler-reinforced polymer composite materials for effective radiation protection [66]. Reproduced with permission from Nambiar et al., Applied Materials; Copyright (2019) American Chemical Society.
Materials 16 04305 g006
Figure 7. Diagram of the neutron shielding process of four modified PVA-based composites [76]: Gd/PVA, MXene/PVA, 5-Gd@MXene/PVA, and 20-Gd@MXene/PVA. Reproduced with permission from Zhu et al., Nanoscale; published by the Royal Society of Chemistry, 2022.
Figure 7. Diagram of the neutron shielding process of four modified PVA-based composites [76]: Gd/PVA, MXene/PVA, 5-Gd@MXene/PVA, and 20-Gd@MXene/PVA. Reproduced with permission from Zhu et al., Nanoscale; published by the Royal Society of Chemistry, 2022.
Materials 16 04305 g007
Figure 8. Neutron shielding performance prediction model: MCNP simulation of Gd2O3/316L composite) [104]. (a) 3D plots of MCNP results. (b) The 2D contour map of 155/157Gd areal density. Reproduced with permission from Wan et al., Vacuum; published by Elsevier, 2020.; MCNP simulation of Gd/316L steel [106]. (c) The 3D relationship between thermal neutron shielding rate, Gd content, and material thickness. (d) 2D contour maps of A(155,157Gd). Reproduced with permission from Qi et al., Materials Today Communications; published by Elsevier, 2023.
Figure 8. Neutron shielding performance prediction model: MCNP simulation of Gd2O3/316L composite) [104]. (a) 3D plots of MCNP results. (b) The 2D contour map of 155/157Gd areal density. Reproduced with permission from Wan et al., Vacuum; published by Elsevier, 2020.; MCNP simulation of Gd/316L steel [106]. (c) The 3D relationship between thermal neutron shielding rate, Gd content, and material thickness. (d) 2D contour maps of A(155,157Gd). Reproduced with permission from Qi et al., Materials Today Communications; published by Elsevier, 2023.
Materials 16 04305 g008
Figure 9. Mechanical properties of reported Gd-containing Fe-based neutron shielding materials, including: 1 wt%Gd/DSS (Hot rolled) [94], 0.087 wt%Gd/DSS (Cold rolled) [95], 0.5 wt%Gd/316L (Hot rolled) [101], 1.5 wt%Gd2O3/316L [104], 1 wt%Gd/316L [105], 1.5 wt%Gd/316L (Hot rolled) [106].
Figure 9. Mechanical properties of reported Gd-containing Fe-based neutron shielding materials, including: 1 wt%Gd/DSS (Hot rolled) [94], 0.087 wt%Gd/DSS (Cold rolled) [95], 0.5 wt%Gd/316L (Hot rolled) [101], 1.5 wt%Gd2O3/316L [104], 1 wt%Gd/316L [105], 1.5 wt%Gd/316L (Hot rolled) [106].
Materials 16 04305 g009
Figure 10. Tensile properties of the (15 wt%B4C + 1 wt%Gd)/Al composite and the 15 wt% B4C/6061Al composite prepared by hot rolling techniques [114]: (a) tensile strength; (b) yield strength; (c) elongation. Reproduced with permission from Xu et al., Journal of Alloys and Compounds; published by Elsevier, 2019.
Figure 10. Tensile properties of the (15 wt%B4C + 1 wt%Gd)/Al composite and the 15 wt% B4C/6061Al composite prepared by hot rolling techniques [114]: (a) tensile strength; (b) yield strength; (c) elongation. Reproduced with permission from Xu et al., Journal of Alloys and Compounds; published by Elsevier, 2019.
Materials 16 04305 g010
Figure 11. Gd2O3@W core–shell structure for double shielding against γ-ray and neutron radiation [127]: (a) schematic diagram of Gd2O3@W core–shell structure particles; (b) 6063 aluminum composite plate doped with Gd2O3@W core–shell structure particles. Reproduced with permission from Zhang et al., Materials Letters; published by Elsevier, 2020.
Figure 11. Gd2O3@W core–shell structure for double shielding against γ-ray and neutron radiation [127]: (a) schematic diagram of Gd2O3@W core–shell structure particles; (b) 6063 aluminum composite plate doped with Gd2O3@W core–shell structure particles. Reproduced with permission from Zhang et al., Materials Letters; published by Elsevier, 2020.
Materials 16 04305 g011
Figure 12. Schematic of the corrosion mechanism of Al-based Gd2O3-W shielding composites with different particle distributions [128]. Reproduced with permission from Cong et al., Corrosion Science; published by Elsevier, 2023.
Figure 12. Schematic of the corrosion mechanism of Al-based Gd2O3-W shielding composites with different particle distributions [128]. Reproduced with permission from Cong et al., Corrosion Science; published by Elsevier, 2023.
Materials 16 04305 g012
Table 1. Neutron capture cross-sections of several typical neutron absorbing elements [13].
Table 1. Neutron capture cross-sections of several typical neutron absorbing elements [13].
ElementCorresponding IsotopeAtomic NumberRelative Atomic MassNeutron Capture Cross-Sections/BarnsAbundance/wt%
Boron10B510.81384019.9
Cadmium113Cd48112.4120,60012.22
Samarium149Sm62150.3640,00013.8
Gadolinium155Gd64157.2562,54014.8
157Gd255,00015.65
Table 2. Neutron shielding performance of reported Gd-containing PMCs.
Table 2. Neutron shielding performance of reported Gd-containing PMCs.
YearMatrixChemical CompositionContent of Gd-Containing ComponentsNeutron Shielding Performance *Main FeaturesRef.
2018PMMApoly(MMA-coGd(MAA)3)1.91 wt%I/I0 = 0.1 (Sample thickness: 5 mm)Self-polymerization[70]
2018HDPEh-BN/Gd2O3/HDPE3 wt%
(with 7 wt%BN)
I/I0 ≈ 0.3 (Sample thickness: 5 cm)Hybridization of h-BN and Gd2O3[82]
2019PANGd(MAA)3/PAN12.11 wt%I/I0 = 0.01 (Sample thickness: 2 mm, by simulation)Self-polymerization[71]
2019EPGd-MOF/EP10 wt%I/I0 < 0.1 (Sample thickness: 5 cm, by simulation)MOF-based[78]
2020PVAGd2O3/PVA3.5 wt%I/I0 = 0.43 (Sample thickness: 5 mm, by simulation)Self-polymerization[69]
2021PIGd-MOF/PI3 wt%I/I0 = 0.096 (Sample thickness: 5 cm, for fast neutrons, by simulation)MOF-based[79]
I/I0 = 0.009 (Sample thickness: 0.2 cm, for thermal neutrons, by simulation)
2021HDPEGd2O3/B4C/HDPE10 wt%
(with 10 wt% B4C and 80 wt% HDPE)
I/I0 = 0.1 (Sample thickness: 9.1 cm)Surface modification; nanoparticles[86]
2022NRGd2O3/NR50 phrI/I0 = 0.55 (Sample thickness: 2 mm)Self-healing ability[74]
2022PVAGd@MXene/PVA15.1 wt%I/I0 = 0.169 (Sample thickness: 1 mm)MXene-based[76]
2022SAPPSGd2O3/SAPPS10 wt%I/I0 = 0.2 (Sample thickness: 2 cm)Nanoparticles[85]
* Here, the average counts of neutrons detected with and without specimens are noted as I and I0, respectively. I/I0 refers to the neutron transmission ratio, and the neutron shielding efficiency is higher in materials with a lower I/I0 ratio.
Table 3. Several Gd-containing AMCs developed for neutron shielding application.
Table 3. Several Gd-containing AMCs developed for neutron shielding application.
YearMatrixChemical Composition *,1Mechanical PropertiesThermal Neutron Shielding Performance *,2Ref.
2016Pure Al(1%Gd + 15%B4C)/AlUTS = 170 MPa, EL ≈ 9%I/I0 ≈ 0 (sample thickness: 3 mm)[112]
2017Pure Al(1%Gd + 15%B4C)/AlUTS = 260 MPa, EL ≈ 7%Not mentioned[113]
20196061Al(1%Gd + 15%B4C)/6061AlUTS = 421 MPa, EL ≈ 6.4%Not mentioned[114]
20196061Al(1%Gd + 15%B4C)/6061AlUTS = 380 MPa, EL = 5%I/I0 ≈ 0 (sample thickness: 3 mm)[115]
20196061Al10%Gd2O3/6061AlUTS = 240 MPa, EL = 16%I/I0 = 0.0036 (sample thickness: 10 mm)[122]
20206063Al25%Gd2O3/AlMicrohardness = 120 HV0.1I/I0 ≈ 0 (sample thickness: 0.75 mm)[124]
2020Pure Al5%Gd2O3@20%W/AlUTS = 310 MPa, EL ≈ 3%I/I0 = 0.01 (sample thickness: 3 mm, by simulation)[127]
20226061Al(10%B4C + 3.6%Gd)/6061AlUTS = 342 MPa, EL = 2.4%I/I0 = 0.002 (sample thickness: 3 mm, by simulation)[118]
2022AlMg1%Gd/6%TiB2/Al–6%MgUTS = 464 MPa, EL = 15.6%I/I0 ≈ 0 (sample thickness: 8 mm)[121]
*,1 All units in the chemical composition of shielding materials are provided in weight percent (wt%). *,2 Similarly, the average counts of neutrons detected with and without specimens are noted as I and I0, respectively. I/I0 refers to the neutron transmission ratio, and the neutron shielding efficiency is higher in materials with a lower I/I0 ratio.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, K.; Ma, L.; Yang, C.; Bian, Z.; Zhang, D.; Cui, S.; Wang, M.; Chen, Z.; Li, X. Recent Progress in Gd-Containing Materials for Neutron Shielding Applications: A Review. Materials 2023, 16, 4305. https://doi.org/10.3390/ma16124305

AMA Style

Wang K, Ma L, Yang C, Bian Z, Zhang D, Cui S, Wang M, Chen Z, Li X. Recent Progress in Gd-Containing Materials for Neutron Shielding Applications: A Review. Materials. 2023; 16(12):4305. https://doi.org/10.3390/ma16124305

Chicago/Turabian Style

Wang, Kangbao, Litao Ma, Chen Yang, Zeyu Bian, Dongdong Zhang, Shuai Cui, Mingliang Wang, Zhe Chen, and Xianfeng Li. 2023. "Recent Progress in Gd-Containing Materials for Neutron Shielding Applications: A Review" Materials 16, no. 12: 4305. https://doi.org/10.3390/ma16124305

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