Next Article in Journal
Organic Nonlinear Optical Crystals for Highly Efficient Terahertz-Wave Generation
Previous Article in Journal
Lyotropic Liquid Crystal System for Drug Delivery of Astaxanthin: Physical Characterization and Enhanced Antioxidant Potential
Previous Article in Special Issue
Impact of Natural Graphite Flakes in Mixed Fillers on the Irradiation Behavior of Fine-Grained Isotropic Graphite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Structure, Property, and Ion Irradiation Effects of Pyrochlores: A Comprehensive Review

1
School of Materials Science and Engineering, Harbin Institute of Technology, Harbin 150001, China
2
Nuclear Fuel Cycle Research Theme, Australian Nuclear Science and Technology Organisation, Locked Bag 2001, Sydney, NSW 2232, Australia
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(1), 143; https://doi.org/10.3390/cryst13010143
Submission received: 14 November 2022 / Revised: 27 December 2022 / Accepted: 9 January 2023 / Published: 13 January 2023

Abstract

:
Since the beginning of the use of nuclear energy, humans have been faced with the problem of radionuclide disposal. At present, a large amount of waste is stored in pools or dry tanks at reactor sites. With the development of the nuclear power generation industry worldwide, the high storage cost (including building, maintaining, and operating storage pools) is overwhelming and serious, and urgent radionuclide disposal problems have become increasingly difficult. Safe and economical strategies are urgently needed for long-term storage and disposal of nuclear waste, which has become among the core issues in the utilization of nuclear energy. Pyrochlore ceramics are able to immobilize a variety of radionuclides and have excellent irradiation stability, so they have received extensive attention as hosts of radionuclides waste. This review summarizes the structure, composition, synthesis process, properties, and irradiation stability of pyrochlore ceramics, focusing on the ion irradiation effect of pyrochlore. In general, the cation radii ratio rA/rB is a key parameter related to various properties of pyrochlores. Zirconate pyrochlore is more easily transformed from pyrochlore to defective fluorite, and leads to better irradiation resistance.

1. Overview

Throughout the world, there are currently 440 commercial nuclear power reactors in operation or suspension with ~390 GW(e) producing over 10% of the world’s electricity and 56 reactors under construction with ~58 GW(e) (data from https://pris.iaea.org/PRIS, accessed on 15 September 2022). Greenhouse gas, which is the main pollution from fossil fuel burning, has been reduced substantially as a result of nuclear utilization. However, the radiative waste produced by the nuclear power plants in the form of spent nuclear fuel (SNF) is still a big issue [1]. SNF is generated by nuclear fuel, mainly uranium, thorium, and small amounts of plutonium, through fission reactions in the reactor core. Approximately 30 t of SNF is generated by a typical 1 GW(e) light-water reactor. That means that more than 12,000 t of SNF per annum are produced and a cumulative repertory of 370,000 metric tons of heavy metal (MTHM) that have been stored either in pools or dry tanks at reactor sites [2,3]. The ultimate composition of SNF depends upon a lot of influencing factors, such as the reactor type, the neutron energy spectrum during the reaction and the burn-up (the degree of fission consumption of the fuel), as well as the origin fuel (type, chemical composition, and initial enrichment level 235U) [4]. Based on a general rule of thumb, in power generation, only approximately 5% of the uranium in the fuel is involved in fission, which also generates approximately 3–4 at.% fission products (137Cs, 131I, 90Sr, ……) and 1–2 at.% transmutation products such as 241Am, 239Pu and 237Np, with typical burn-ups in the range of 35–45 MWd/kgU [1]. Hence, the high-level wastes must be converted into more stable solid forms by indispensable treatments, which are suitable for sequential waste management, such as transportation, temporary storage, and geological disposal for long-term storage. This has become among the extremely crucial challenges for nuclear energy production.
Generally, the requirements for waste forms that are able to immobilize radionuclides are as follows [1,5,6,7]:
  • Stable waste forms must exhibit high radiation stability and excellent tolerance to radiation effects, such as transmutation to new elements, ballistic effects, ionic and electronic excitation, gas bubble and volume swelling effect, depending on the types of high-level radionuclides.
  • The immobilization system must have high waste loading for a sufficient amount of waste (typically 20–35%) to reduce volumes. Of course, higher waste loading is accompanied by higher irradiation damage, so a balance needs to be found between those.
  • The waste forms must be chemically flexible and be able to accommodate a mixture of radionuclides and other contaminant species. It is possible to separate and immobilize mixed radionuclides into different substrates, but this means more complex processes and higher costs in terms of energy, money and time.
  • The waste forms must be durable and resistant to aqueous dissolution to minimize the release of radiotoxic species, especially for those long half-life radionuclides such as 237Np (2.1 million years) and 239Pu (24,100 years).
  • The fabrication process of waste forms must be simple and feasible under reasonable conditions, such as low temperatures and in an air atmosphere, by using well-established methods and avoiding vacuums and special atmospheres.
  • The waste form are natural analogues. It is generally impossible to build confidence in predicted behavior from real and complete experiments of the waste forms in the laboratory directly (except some relatively short half-life nuclides, e.g., 244Cm 18 years).
  • The possibility of criticality for actinide waste forms must be carefully considered by reducing actinide concentrations or incorporating neutron absorbers, such as Gd and Hf, into the structure of the waste form.
Borosilicate glass has been comprehensively and thoroughly investigated as a waste form for the immobilization of high-level radionuclides [1,3,8,9,10,11,12,13]. However, borosilicate glass is not a universal waste form due to the restricted solubility of actinide oxides. Only approximately 3–10 wt.% actinide oxides can dissolve in the glass network [9,10,11]. Low solubility limits waste loadings to a relatively low level, which leads to large volumes of vitrified waste forms, and finally results in a significant increase in overall disposal costs. Glass-ceramics (GCs) with both crystalline and glass phases are a variation of a vitrified waste form, which is a chemically more flexible and less expensive compromise than the pure ceramic form. Generally, GCs can provide a higher chemical durability than glass, provided that the radionuclides are prevented from crystallizing into soluble species prior to incorporating into the matrix [14,15,16,17].
Based on a group of geochemically stable titanate minerals, synthetic rock (Synroc) was initially developed in the late 1970s in Australia for the immobilization of HLWs from Purex-type reprocessing of spent nuclear fuels [18]. Synroc is a polyphase ceramic waste form, which is composed of zirconolite (CaZrTi2O7), perovskite (CaTiO3), and hollandite (Ba(Al,Ti)2Ti6O16)), which demonstrates the ability to host thorium and uranium in the natural geological environment [18,19,20,21,22,23,24]. Apart from Synroc, many other ceramic waste forms were developed, such as pyrochlore [9,25,26,27]. For those ceramic waste forms, no matter single- or poly-phase ceramic, specific atomic sites in periodic structures of constituent crystalline phases may be occupied by radionuclides to form a dilute solid solution. Another form is to incorporate radionuclides through encapsulation, where the radionuclide-bearing phases, by simply drying the waste sludge, are encapsulated in another non-radionuclide-bearing material (e.g., ZrO2, TiO2 and Al2O3) to form a composite waste form [19,28,29].
In nuclear waste glasses, rare earth titanates were discovered to incorporate actinide nuclides and further identified as a pyrochlore structure, which is the earliest evidence of pyrochlore as a waste form [5,9,29]. Titanate wastes have been studied extensively since the 1980s, due to their high chemical durability [30,31]. Early research focused on monoclinic zirconolite, which is now regarded as a structural derivative of the isometric pyrochlore structure, and further studies reported a “cubic” zirconolite formed stably from higher actinide (or actinide analogue) concentrations, which is actually a disordered pyrochlore structure [32,33]. In addition, pyrochlore ceramics have excellent aqueous durability and irradiation resistance. Based on the available literature for various types of materials about radionuclide waste forms, we focus on an in-depth discussion of recent research on the pyrochlore waste forms. Although it is not possible for actual waste forms over the limited time scales by direct laboratory tests, ion irradiation experiments of the matrix with or without radionuclide analogues can provide important information for the long-term performance of waste forms. This comprehensive review will summarize the structure, the common synthesis processes and the main characteristics of pyrochlore in detail; introduce the main irradiation effect and damage model; and finally provide a summary of the irradiation resistance of pyrochlore ceramics.

2. Structure of Pyrochlore

Hitherto, pyrochlore phase materials encompass a wide range of chemical component that exists in over 500 naturally or artificially materials [34]. Pyrochlore structure is known as a derivative fluorite structure (CaF2, space No. 225, F m 3 ¯ m ), which forms an isometric crystal system with the ideal structural formula VIIIA2VIB2IVX6IVY’ (space No. 227, F d 3 ¯ m , Z = 8 and a = 9–12 Å), where A-sites and B-sites are occupied by metal cations, and X (= O2−) and Y (= O2−, OH, F) are anions [34,35,36]. The structure can be outlined in a simple and well-understood way by describing the shapes and topology of the coordination polyhedral (Figure 1a). All of the atoms in an ideal pyrochlore are on special positions, except the X-anion occupied 48f-site (X48f). The larger A-site cations (mainly trivalent rare earth elements, usually ~1 Å ionic radius) are eight coordinated (AX6Y2) and occupy16d (0.5, 0.5, 0.5) sites. The smaller B-site cations (~0.6 Å ionic radius, mainly tetravalent cations including Ti, Zr, Hf, Sn, and etc.) are six coordinated (BX6) which occupy 16c (0, 0, 0) sites. Herein, the X-anion occupies the 48f (x48f, 0.125, 0.125) and the Y-anion locates at 8b (0.375, 0.375, 0.375). Thus, for the ideal structure, the X-anion and Y-anion are situated the tetrahedron XA2B2 and YA4, respectively. Note that the center of the tetrahedron consisting of four B-sites (□B4) is blank with a marked 8a position (0.125, 0.125, 0.125). The BX6 octahedra are corner-linked with A-cations filling the interstices. Additionally, the structure can also be described as a structure in which the A-site and B-site coordinated polyhedral are connected with an edge sharing, as shown in Figure 1a. Moreover, the shapes of these polyhedrons are controlled by the position parameter x48f which suggests that X48f shifts towards the two of its B-cations. For x48f = 0.375, the A-site coordination polyhedron AX6Y2 is a regular cube and the B-site polyhedron BX6 is a trigonally-flattened octahedron (the topology of the fluorite structure). In this case, if A and B cations randomly occupy each other’s positions, materials have a defective fluorite structure (space No. 225, F m 3 ¯ m ), and the anions X and Y share the 7/8 of the 8c-site (0.125, 0.125, 0.125) in the new space group (Figure 1c). For x48f = 0.3125, the B-site polyhedron BX6 is a regular octahedron and the A-site coordination polyhedron AX6Y2 is a distorted trigonal scalenohedron, which is an ideal pyrochlore structure. Theory simulations of a wide range of pyrochlore compositions (A = Lu3+ to La3+; B = Ti4+ to Pb4+) using energy minimization techniques suggest that the x48f parameter also reflects some degree of disordering of the oxygen lattice for specific cation compositions [37,38,39]. Thus, the position parameter of X-anion, x48f, between 0.3125 and 0.375, indicating the degree of distortion of the polyhedron and the degree of order in the structure. Generally, x48f is nearly linearly associated with the cation radius ratios rA/rB [6,37,40]. So, the crystal structure of pyrochlore is mainly controlled by the cation radius ratios rA/rB, when rA/rB is between 1.46–1.78, it shows a pyrochlore structure, and when rA/rB is lower or higher than this range, showing defective fluorite structure and monoclinic structure, respectively [6,34,35].
In recent years, weberite-type structure, another derivative structure of fluorite and pyrochlore, has been paid more and more attention. It is considered to be an intermediate phase of irradiated-induced ordered-disordered transition from pyrochlore to fluorite, which is obtained by pyrochlore loss over part of the long-range order [36,41,42,43,44,45]. The chemical formula of the orthorhombic weberite-type structure (space No. 63, Ccmm) can be written as ABC2X7, where A and B are cations occupying 4a and 4b site, which are in octahedral (6-fold) and distorted cubic (8-fold) coordination, respectively. Whereas C is a random mixing of A and B atoms located at 8g site, which is distorted mono-capped octahedral (7-fold) coordination (Figure 1e). Anions X occupy 16h and three 4c sites, and there remains an ordered vacancy, as in the pyrochlore structure, but it is situated in another 4c site [46]. Figure 1d–f shows the structural relations between the pyrochlore, weberite-type, and disordered fluorite arrangement. Both pyrochlore and weberite-type structures are supercells of disordered fluorite with unit cell parameters of aP = aWB = 2aF and bWBcWB ≈ 21/2aF. During the transition from pyrochlore to weberite-type, in pyrochlore, half of A and B atoms randomly exchange position, occupying the 8g site in weberite-type; in addition, two O48f move to V8a vacant site along the direction of [1 0 0]. Further, when the cations and anions are completely disordered respectively, the structure transforms to disordered fluorite.
In natural pyrochlore minerals, A-site can also be occupied by smaller radius cations such as alkali metal cations or alkaline-earth metal cations and molecular water and hydroxyl are potentially part of the structure. Stanek et al. [47,48] point out that a stable pyrochlore structure can be formed through A3+ vacancies to compensate for additional tetravalent B-site cations and interstitial oxygen interstitial oxygen atoms are also conducive to structural stability, especially near the defective fluorite structure. Due to a series of defects, structures can be formed by loss of cations and anions or inequivalence substitution, the general formula of pyrochlore is A2−mB2X6−wY1−n·pH2O. The parameters m, w and n are used to balance the charge generated by incomplete substitution or unequivalence replacement of the corresponding site: A, X, and Y. Based on the structural formula, several different ions can therefore be substituted into the pyrochlore structural sites as presented in Table 1 [6,34,35,36]. Due to the wide variety of combinations of A-site and B-site cations, as well as tunable oxygen vacancies, there is a possibility of the pyrochlores that can incorporate trivalent lanthanides and tri- and tetra-valent actinides which have great potential for nuclear waste forms and attracted the greatest interest [3,5,6,29].

3. Synthesis of Pyrochlore

Solid-state reaction is one of the most common methods for preparing polycrystalline pyrochlore ceramics. First, the metal oxide powders are preheated to remove any absorbed impurities, such as H2O and CO2, and then anhydrous ethanol or acetone is used as a mixing medium to mix the raw powder evenly. Finally, the dried mixture is pre-compressed for sintering. The residues of initial oxides can be avoided by chosing the appropriate synthesis sintering temperature, which can deteriorate the isolation properties of waste forms. The synthesis temperature required depends on the composition of the system, such as replacing Zr4+ with Ti4+, which can reduce the synthesis temperature [49]. In addition to heating by air furnaces, the energy required for the reaction can be provided by different heat sources. Guo et al. [50] use the nozzle of an aerodynamic levitator with oxygen flow to levitate the green pellet and heat the sample (~3023 K) by controlling the laser power. Zhang et al. [51,52,53] use the heat generated during the redox reaction between Ti and Fe2O3/CuO to provide the energy required for the reaction. Among them, heating by furnace and laser heating can precisely control the cooling rate. When heating by furnace (except for spark plasma sintering), a two-step [42,44,54,55,56,57,58] or three-step [59,60,61,62,63,64] sintering method is more commonly used compared with one-step [65,66,67,68] sintering, which is conducive to the uniform distribution of metallic cations in pyrochlore ceramics by being reground, pelletized, and heated. Although it needs to be held at a higher temperature for a long time, it is also conducive to the formation of the pyrochlore structure [69]. Pyrochlore ceramics prepared by reactive heating takes no longer than 5 min for the combustion and densification process [52,53], but this method will introduce metallic simple substance, and the distribution of metallic cation in the ceramic is not completely uniform. Spark plasma sintering (SPS) can rapid synthesis pyrochlore ceramicswithout introducing any impurities. Chen et al. [70] prepare Gd2Zr2−xCexO7 (x = 0.0, 0.5, 1.0, 1.5 and 2.0) pyrochlore ceramics with only three minutes of holding time via SPS and the single phase maintains until the value of x reaches 1.0, also the bulk density of the ceramics are more than 92%. It is worth noting that the relative density of the pyrochlore ceramics prepared under atmospheric pressure is only about 90% [58,71,72] unless used in conjunction with laser heating. Laser heating can prepare pyrochlore ceramics with a relative density of 96.3% under atmospheric pressure by solidification method [50]. Yet the relative density of the ceramic can be increased when a certain pressure is applied in the sintering step, such as SPS [70] and hot pressing. Brykała et al. [73] prepared Gd2Zr2O7 pyrochlore ceramics with the relative density close to the theoretical value by HP under pressure of 30 MPa at 1723 K At the same time, the increase in the density of the ceramic is beneficial to improve its radiation resistance ability [53].
The operation of the solid-state reaction is relatively simple. Since most rare earth and transition metal elements exist in the form of oxides, the raw materials required for the preparation of various kinds of pyrochlore ceramics by the solid-state sinter are easy to obtain, which makes the solid-state sintering highly devisable. Due to the fact that of this, the high-entropy pyrochlore ceramics with complex compositions are mostly prepared by solid-state sintering [50,74,75,76,77]. Qi et al. [78,79,80,81,82] use oxides as raw materials to prepare single-phase ceramic powder by the above mentioned one-step method, the prepared powder is pressed into pellet to sinter at 2023 K for 8–10 h under vacuum and annealed at 1573 K in the air to prepare high entropy (Y0.2La0.2Gd0.2Yb0.2Dy0.2)2Zr2O7 and different kinds of A2B2O7 transparent ceramics, all the powders and ceramics have a defective fluorite structure. Additionally, Wang et al. [76] point that size disorder of about 5% is the boundary for the multi-component rare earth zirconate to form a single- or dual-phase region. The limitation of solid-state sintering is that it is difficult to achieve uniform distribution of cations and will introduce pollution during multiple steps of reground, pelletized, and heated.
The chemical-coprecipitation method is to react the mixed solution of the metal salt solution with a precipitant to obtain gel-like precipitates, and then the gel-like precipitates are washed, dried, calcined, ground, and the obtained powder is compacted and sintered to obtain the polycrystalline pyrochlore ceramics [83,84,85,86,87,88,89,90]. Compared with the solid-state sinter, the chemical-coprecipitation method has better dispersibility and uniformity, and the holding time required in the sintering process is also shortened.
The sol-gel method uses citric acid and ethylene glycol/anhydrous ethanol as chelator and solvent, respectively, and is mixed with the compound containing the desired cation (such as nitrate, etc.), and the mixture is evaporated to dryness to form the precursor. Pyrochlore ceramics powder can be prepared by calcining the precursor at a temperature of more than 1073 K, and the powder can be used to prepare polycrystalline pyrochlore ceramics through briquetting then sintering at about 1773 K [91,92,93,94,95]. Li et al. [96] modified sol-gel process by introducing NaCl and KCl, the salt addition is beneficial to control the abnormal particle growth, and the morphology of the Lu2Ti2O7 powder is affected greatly by the salt amounts and types. The sol-gel method enables uniform mixing of cations at the atomic level, which facilitates the diffusion of components, so the reaction can be carried out at lower temperatures. Compared with the solid-state sintering, the chemical-coprecipitation method and the sol-gel method are more complicated to operate and will generate more chemical waste.
Hydrothermal method can prepare polycrystalline pyrochlore ceramics at the low temperature of no more than 513 K under autogenous pressure [97,98,99,100,101]. The reacting temperature, holding time, and reaction environment (pH value) are the key factors affecting the final product of hydrothermal method. Increasing the reaction temperature and prolonging the holding time is conducive to promoting the formation of the pyrochlore structure, also improving the crystallinity of the reaction products. The pH value required for each reaction system is different, and the pyrochlore ceramics phase cannot be formed unless the reaction environment is at the proper pH value [97,98,100]. When Trujillano et al. [99] prepare Sm2Sn2O7 pyrochlore ceramics via hydrothermal method by microwave irradiation and compare this with the solid-state method, the results show that the reaction time (4h) can be greatly reduced by microwave irradiation compared with other heating methods, however, there are amorphous phase and hydrogen-carbonate impurities in the final product and the crystallinity is not as high as that of the solid-state method.
The single crystal titanate pyrochlore ceramics can be grown by flux technique [102,103]. A single crystal is grown at ~1508 K in a molten mixture consisting of oxides and PbF2 as raw material and flux, respectively. Although the PbF2 is an impurity and evaporates slowly during the synthesis period, but it will remain primarily superficially on the crystal and can be removed physically before analysis. The single crystal titanate pyrochlore ceramics can also be grown by floating zone technique [104,105,106,107].
In addition to the above bulk and powder pyrochlore ceramics, the molten salt and electrospinning method can be used to prepare nano pyrochlore ceramics. Li et al. have prepared nanowires [108], nanorods [109], and nanoplates [110] by molten salt method, and the nanowires, nanotubes, and wire-in-tube architectures are prepared by Yu et al. [111] through single-nozzle electrospinning technique.

4. Property of Pyrochlore

Due to the flexible composition and unique structure, pyrochlore has a variety of outstanding properties including thermophysical properties, enhanced electrical properties, irradiation resistance and optical properties, and have potential applications for thermal barrier coatings material, solid electrolytes, radionuclide waste forms, etc. [40,75,76,83,84]. The irradiation resistance of pyrochlore will be introduced in detail in Section 5, and thermophysical properties and ionic conductivity will be summarized preliminarily in this section.

4.1. Thermophysical Properties

Most studies about thermophysical properties of pyrochlore are based on the application of thermal barrier coatings [112,113]. In fact, the irradiation resistance and the radiation damage process are inseparable from thermophysical properties of pyrochlore, especially, the transient thermal process and recrystallization after ion irradiation (described in the thermal spike model in Section 5.2.1) [114].
Liu et al. [83,84,85,86,87,88] have studied various thermophysical properties of rare earth zirconate ceramics (A2Zr2O7), including heat capacity, thermal expansion and thermal conductivity, and reported that the thermal conductivities of rare earth zirconate ceramics first gradually decrease with increasing temperature, and then slightly increase above 1073 K due to the increased radiation contribution. The thermal conductivities of (YbxSm1−x)2Zr2O7, (YbxGd1−x)2Zr2O7, (SmxGd1−x)2Zr2O7, (SmxGd1−x)2Zr2O7, and SmxZr2−x/2O2 are located in the range of 1.40–1.99, 1.34–1.99, 1.20–1.99, 1.33–2.04, and 1.41–1.86 W·m−1·K−1, respectively [84,85,86,87,88]. The calculated minimum thermal conductivities of A2Zr2O7 and A-site doped zirconate pyrochlores by Density functional theory are ~1.2 W·m−1·K−1 [40,112,113,115,116], while that of B-site doped zirconate pyrochlores shows a decreasing trend, except Ti [113,115,116,117]. Recently, high entropy pyrochlore exhibits lower thermal conductivity attracted attention [50,118,119,120]. Guo et al. [50] report that thermal conduction 0.947 W·m−1·K−1 of (Ho1/5Y1/5Dy1/5Gd1/5Eu1/5)2Ti2O7 at room temperature, which is lower than their single rare earth composition pyrochlore. Similarly, Li et al. [118] report that six species high entropy pyrochlores with thermal conductivities are lower than 1 W·m−1·K−1.

4.2. Ionic Conductivity

The ionic conductivity of pyrochlore is linked to the migration ability of oxygen ions, which is mainly related to the concentration of oxygen vacancy and the activation energy of oxygen movement [75]. The ionic conductivity of pyrochlore shows a function of cation radius ratio rA/rB, and the total conductivity approaches its maximum value when the cation radius ratio is ~1.48 [75,89,121,122,123,124,125], as shown in Figure 2a. Due to the fact that as rA/rB decreases, the oxygen vacancy formation energy and migration energy barrier of oxygen decrease, that is, the migration ability of oxygen ions is enhanced, but at the same time, there is an order-disorder transition from pyrochlore to defective fluorite that causes the loss of the preferential migration path of oxygen ions [75,126]. Figure 2b shows the total conductivity of various pyrochlore including low/medium entropy and high entropy ceramics [75,122,123,124,125,127,128], suggesting that a high entropy strategy can effectively enhance the ionic conductivity of pyrochlores. Due to the fact that ionic conductivity is related to the disordered sensitivity of oxygen, to some extent, the high ionic conductivity suggests that the pyrochlore has a good anti-irradiation performance.

5. Radiation Stability

Radionuclides in HLW will undergo alpha (α) decay (e.g., Np, Am, Pu, and Cm) or beta (β) decay (e.g., 137Cs and 90Sr), or even both (e.g., U and Th), and release the energetic particles. β-decay produces high-energy β-particles (>0.5 MeV), very low energy β-recoil nuclei, and γ-rays, while, α-decay releases energetic α-particles (4.0–5.8 MeV), energetic α-recoil nuclei (70–100 keV), and some γ-rays. The energetic α-particle has a penetration depth of ~20 μm and loses most of its energy through inelastic collisions (ionization and electronic excitations) with surrounding atoms within this path, and produces 100–200 atomic displacements near the end of its track. While an α-recoil atom shows predominately nuclear stopping (elastic collisions) and produces a dense collision cascade with 1000–2000 atomic displacements within a short range of ~0.02 μm. However, β-decay causes on the average less than one atomic displacement per β-decay event hence only contributes a minor amount to radiation damage. There are also spontaneous fission and (α, n) reactions of some actinides but the rate of the reactions is very low, thus the contribution to the radiation field is small, and do not significantly contribute to the overall effect of radiation. In summary, >90 percent of the structural damage is caused by the alpha-recoil nucleus [7,26,129,130].
Generally, β-decay is the dominated radiation source in the first 500 years of storage, while α-decay is gradually dominant at longer time because of the long half-lives of the actinides and their daughter products. As shown in Figure 3a [8,29], for DOE (U.S. Department of Energy) high-level tank waste consisted of multiphase ceramics or glass-ceramics, the cumulative β-decay dose saturates at about 500 years, while the alpha decay dose is a near-linear function of storage time. For non-USA commercial HLW contained in ceramics endures doses of at least a time of 10 higher than those for DOE tank waste [8]. For several 239Pu concentrations in a ceramic waste form, the α-decay cumulative dose reaches a provisional plateau after 100,000 years, but increases over much longer time periods due to the 235U decay series, as shown in Figure 3b [7,8]. Additionally, the dose-time ranges for the contribution of self-irradiation damage in actinide-bearing glasses or ceramics are also shown in Figure 3b. The spontaneous fission half-life of 239Pu is 5.5 × 1015 years which is about 2 × 1011 times slower than the rate of α-decay (2.411 × 104 years). On the other hand, the spontaneous fission and α-decay half-lives of 244Cm are 1.35 × 107 and 18.10 years, respectively. This means that a rate of spontaneous fission for 244Cm is about 750 thousand times slower than that of α-decay [131]. Thus, spontaneous fission of 239Pu or 244Cm gives a negligible contribution to self-irradiation damage. Although in some materials, each spontaneous fission event can produce a significantly larger amorphous volume, and spontaneous fission can account for 2–3% of the amorphization rate [25]; however, the accumulation rate of radiation damage from spontaneous fission is a few orders of magnitude lower than that from α-decay, except 244Cm. Therefore, radiation effects of spontaneous fission are generally ignored.

5.1. Radiation Effects

The radiation effects of the interactions between ions and the structure, also called radiation damage, fall into two broad categories. The first one is the transfer of energy to electrons mainly by ionization and electronic excitations. The second one is the transfer of energy to atomic nuclei, primarily by a ballistic processes involving elastic (billiard-ball-like) collisions. Except for the transfer of energy, the particles from radioactive decay have a significant chemical effect on the host materials through deposition or incorporation into the host structure.
For ions, the mode of energy loss is controlled by the relative velocity of the incident ions and that of the orbital electrons of the target atoms. If the ion velocity is greater than that of the orbital electrons, ionization processes (also called electronic energy loss) will dominate. Otherwise, if the particle velocity is below, the proportion of ionization is small and most energy is lost by ballistic processes (also called nucleus energy loss). A widely accepted relationship between energy loss and incident particle energy is that if the energy of the incident particle is greater than 1 kV per nucleon (1 kV/u), ionization processes dominate; and if the energy of the ion falls below this limiting approximation elastic collisions with atomic nuclei are predominant; see Figure 4.

5.1.1. Ionization and Electronic Excitation

Energy loss by β-particles and α-particles is predominantly through Coulombic interactions. Interaction of γ-rays with matter is primarily through the photoelectric effect, Compton scattering, and electron-positron pair production. The high rate of energy absorption through ionization and electronic excitation in ceramics can result in self-heating. In the first few hundred years, the wasteforms were kept at temperatures of several hundred degrees Celsius due to significant radiogenic self-heating [132]. In addition, a large number of electron-hole pairs were produced by ionization and electronic excitations. This resulted in covalent and ionic bond rupture, charged defects, enhanced self-ion and defect diffusion, and localized electronic excitations. In some ceramics, such as zirconolite, pyrochlore, and monazite, as well as most glasses, permanent defects were formed in the structure generated by localized electronic excitations [8,17,133,134].
In recent years, swift heavy ions irradiation with ultra-high energy has been applied to study the effect of electron energy loss on the structure of materials. Lang et al. [135] performed ion irradiation experiments with a specific energy of 11.1MeV/u (atomic mass unit u) on the Gd2Zr2-xTixO7 binary polycrystalline sample at room temperature and under vertical incidence, using 58Ni, 101Ru, 129Xe, 181Ta, 197Au, 208Pb, and 238U ions. The electronic energy loss (dE/dx) of all ions was calculated by SRIM-2008 code [136] ranges from 10.9 keV/nm (58Ni) to 49.7 keV/nm (238U). Same as other Gd2Zr2-xTixO7 irradiation experiment results [60,137,138,139,140,141], as rA/rB increases (more Ti content), the irradiation damage track diameter is greater, which obeys a near-linear correlation. A concentric ring structure (also called core + shells) [141,142] was found as a representative morphology of damage of an individual ion tracks (cross-section); see Figure 5a. Generally, an ion track consists of an amorphous core surrounded by a disordered, defect-fluorite structured shell, and sometimes, a defect-rich pyrochlore halo would encircle the periphery as shown as Figure 5b. The total track size increases significantly over the energy-loss range studied, from ~5 nm for 101Ru ions (21.7 keV/nm) to ~11 nm for 197Au ions (40.1 keV/nm). Further, within the range of the ions studied, the intramural track structure changes noticeably: the small amorphous core gradually expands with the increase of dE/dx, and forming a predominantly amorphous tracks at high dE/dx [60,135,141]. The area ratio of amorphous core to the whole track r a 2 / r d f + a 2 is defined as the amorphous fraction f, and it has been plotted as a function of dE/dx in Figure 5b, which suggests an exponential growth. Ni tracks are intermittent, consisting of several relatively independent damage domains. That means that the critical energy loss for forming a track (dE/dx)c is very close to and only slightly higher than the energy loss of Ni ions of 10.9 keV/nm, which is in agreement with the result of Gd2Ti2O7 at depth of ~5 μm [143]. What is more, the diameter of Ru ion tracks shows drastic change along their trajectory, which suggests that energy loss of Ru approaches to the track formation threshold [135]. Therefore, for the specific energy of ion beams ~11 MeV/u, in Gd2Ti2O7, a consecutive tracks will arise when the energy loss surpasses (dE/dx)c ~11 keV/nm [135], which accords with experimental results (11–13 keV/nm) obtained by Moll et al. [60] via heavy ions with slightly smaller specific energy (4.5–9.3 MeV/u).
The structure of ion tracks in multiple concentric phases is generally considered as a result of rapid deposition of high-density energy, which leads to melting or even vaporization of the material along the incident direction, followed by quenching and recrystallization during the rapid cooling of the matrix [137,144,145]. Firstly on atto-/femto-second time scales, the incident ion inelastic interacts with the target and primary collisions electron, ionization and a Coulomb field are generated thereby; secondly, electron cascades to form a plasma state with thermalization of electrons and screening of Coulomb field on femtosecond time scales; then, the excited electrons transfer their energy to atoms through electron-phonon coupling. In this process, the atoms are rapidly heated instantaneously, resulting in local melting and vaporization. A pressure wave to the surrounding area is generated by the synchronous volume change on pico-/femto-second scales. Subsequently the melted or vaporized material self-quenching, recrystallization, and defect recovery occur over nanoscale structures or phases during rapid cooling on nano-/pico-second time scales [137,146,147,148]; see Figure 4. The thermal spike models [114,143,145,149,150,151,152,153] to describe the track formation processes will be discussed in Section 5.2.1.

5.1.2. Ballistic Processes

Ballistic processes cause direct atomic displacements through elastic scattering collisions and are responsible for the atomic-scale rearrangement of the structure. The kinetic energies of recoils produced in β-decay and γ-decay are not sufficiently to form permanently displaced atoms. Likewise, due to their low mass, β-particles only induce well separated single displacement events at high energies, in general, only 0.1 atomic displacements per β-decay event [25,29]. In the 16 to 22 μm, α-particles dissipate most of their energy through ionization processes, but as the energy dissipates, nucleus energy loss (elastic collisions, also called nuclear stopping) begins to dominate, undergoing a lot of elastic collisions along their path to produce several hundred isolated atomic displacements [7,29]. A similar process can be found in ion irradiation, where the largest number of displacement atoms occurs near the end of the incident ion range. Compared with the aforementioned particles, α-recoil particles are heavier but exhibit lower energy. They can produce almost all the total number of displacement atoms by ballistic processes. The α-recoil loses 80% or more of its energy in elastic collisions over a very short range (20–30 nm) that produce energetic recoils, which in turn lose 60% or more of their energy in elastic collisions, resulting in cascading collision. As a result, about 50% of the energy of the recoils is deposited as “damage energy” in a highly localized displacement cascade, where 500–2000 atoms are energetically displaced by elastic collisions [8]. The displacement of atoms within a cascade primarily produces Frenkel pairs, but the cascade may also result in the direct formation of topologically disordered or amorphous state in the core of the cascade. The energy density deposited over the initial portion (~5 nm) of an α-recoil cascade can be very high (up to 1 eV/atom) and occurs over a very short time (<10−12 s); this may lead to local melting in the cascade or additional atomic relaxations and rearrangements. Moll et al. [106] have compared the different damage processes between the electronic stopping at high energy and the nuclear stopping at low energy in Gd2Ti2O7 single crystal. At high energy (irradiated by 870 MeV Xe, Se = 29 keV/nm), amorphization occurs directly in individual ion tracks from the extreme electronic-energy deposition, and full amorphization results from the overlapping of these tracks as described by a direct impact model (cf. Section 5.2.2), while at low energy (irradiated by 4 MeV Au, Sn = 5.2 keV/nm), amorphization occurs via indirect processes and energy deposits by ballistic process. The amorphization accumulation process is described in the framework of both direct-impact/defect-stimulated model (cf. Section 5.2.3) and multi-step damage accumulation model (cf. Section 5.2.4) [106]. The above mentioned and other ion irradiation experience are summarized in Table A1 in Appendix A.
Both the partitioning of the energy loss by α-particles between ionization and elastic collisions and the distribution of displaced atoms can be determined by computational approaches based on the binary collision approximation (BCA), if the threshold displacement energies of atoms are known [154]. The scattering angle and energy transfer that results from each binary elastic collision computer can be determined by the SRIM codes [136] and a Monte Carlo approach, assuming a structureless medium (i.e., crystallinity is not considered). This method can also be used for charged-particle irradiation calculations. Molecular dynamics simulations have also been used to calculate radiation damage [137,155,156] and dislocation threshold energies [157,158,159]. Recently, ab-intio molecular dynamic (AIMD) has also been introduced into this field to obtain more accurate calculation results [160,161,162,163,164].

5.1.3. Transmutations and Gas Production

In addition to the energy transferred to the host, α-decay and β-decay also lead to the transmutation of radioactive parent nuclei into different chemical daughter elements. These elements must be accommodated into the structure and may tremendously impact the physicochemical properties of the host materials, especially ceramic waste forms. Being the mainly sources of transmutation in HLW, transmutation of 90Sr and 137Cs is accompanied by changes in both ionic radius and valence. 90Sr has a half-life of 29.1 years and it decays to 90Y by emission of a 546 keV β-particle without gamma photons. Additionally, 90Y, which has a 64 h half-life, emits another high-energy β-particle with a maximum energy of 2.28 MeV and transmutes to 90Zr which is stable (Figure 6a). Usually, the valence of Sr, Y, and Zr are +2, +3 and +4, respectively, and a final ionic radius decrease is 29% [165]. While the half-life of 137Cs is 30.17 years, it emits high energy beta particle transmutes into stable 137Ba in two ways, as shown in the Figure 6b and Cs1+ decays to Ba2+ with a decrease in ionic radius of 20%.
The α-particles capture of two electrons to form helium atoms. Additionally, He atoms must (i) be accommodated in the interstitial position of lattices, (ii) be trapped at internal defects, (iii) aggregate to form bubbles, or (iv) be released via the ceramic surface. For HLW, most of the He will be generated over long time periods near ambient temperature, and the concentration should not exceed 100 atomic parts per million (appm), which should be easily accommodated. In the case of ceramic waste forms for Pu immobilization and disposal, concentrations of He can become quite high and difficult to accommodate within the structure, i.e., exceeding 2 at.% after 100,000 years for ceramics containing 20 wt.% 239Pu [29]. For ion irradiation experiments using noble gas elements (He, Ne, Ar, Kr, Xe), bubbles formation is widely studied [58,166,167,168,169,170]. Taylor et al. [167] determined the critical He concentration (4.6 at.%) required to nucleate bubbles >1 nm in diameter in bulk Gd2Zr2O7 by combined irradiation of Au and He ions. Additionally, bubbles distributed in the matrix alone, or formed 10–30 nm “long chains” parallel to the surface of the sample. In addition, the individual bubbles slightly increase lattice swelling, but the formation of bubble chains can reduce lattice strain. Huang et al. [170] further found that as the average grain size ranges from 221 to 55 nm, the He concentration thresholds for helium bubble nucleation along grain boundaries, formation of helium bubble chains within the grain, and intragranular ribbon helium bubbles are 1.7% to 2.8%, 5.9% to 15.2%, and 19.2% to more than 25.8%, respectively. The lattice swelling effect of Gd2Zr2O7 is size-dependent, the smaller grain size has smaller lattice swelling. Holesinger et al. [58] concluded that the rapid amorphization of Gd2Ti2O7 can accommodate more Kr large bubbles in the matrix without generating cracks, while Gd2Zr2O7 maintains the crystalline structure under Kr irradiation but produces microcracks due to the accumulation of bubbles.

5.2. Damage Models

The irradiation-induced phase transformation in ceramics is of considerable interest, both in technological application and fundamental theory. Several phenomenological models were developed to decipher the damage accumulation induced by ion irradiation, as a function of temperature or flux. The amorphous fraction of Lu2Ti2O7 which irradiated by 400 keV Ne at 77 K, 2.7 MeV Ar at RT, or 6.5 MeV Xe at RT increase with dose can be ascribed to defect-accumulation-like amorphization model (not introduce in this paper), direct-impact/defect-stimulated model (D-I/D-S) model and direct-impact model, respectively [57,171,172]. This suggests that different ion beam irradiation may apply different model. Common models are direct-impact model, cascade-overlap model, direct-impact/defect-stimulated model (D-I/D-S), the multi-step damage accumulation model (MSDA) of Jagielski and Thome, the heterogeneous track overlap model (HTOM) and the thermal spike model (TS). The accumulated amorphous fraction derived from various models which widely used in the literature is compared in Figure 7.

5.2.1. Thermal Spike Model

The thermal spike model has been applied to explain irradiation phenomena for a long time [173]. Many scholars have been instrumental in the development of this model, and now, it can be described the track formation in semiconductors and insulators [114,145,149,173]. The basic assumption of the model is that a high-temperature region (the thermal spike) is formed around the trajectory of a high-energy incident ion [145]. It is assumed that a melt is created when the temperature exceeds the melting point of the crystal. In a cylindrical region with a characteristic time of the order of 10−12 s, the energy deposited on the electronic system is transferred to the phonon system. Due to its small diameter, when the melt solidifies, the quenching rate of the molten cylinder may reach 1013–1014 K/s, which results in an amorphous structure. Szenes et al. [145] assumes that the variation of lattice temperature ΔT(r, t) has a radial Gaussian distribution which evolves with time as following equation:
Δ T ( r , t ) = Q π a 2 ( t ) 1 exp r 2 / a 2 ( t )
where a2(t) depends on the thermal diffusivity. Additionally, the value of Q can be obtained from the balance of energy Q = (gSeLρπR2)(ρc)−1, where gSe is the fraction of energy deposited in the thermal spike, R = R(t) is the radius of the melted zone, and L, c and ρ are the latent heat of phase transition, the mean specific heat and the density, respectively. Another important assumption of the model is that the volume of the amorphous phase formed is proportional to the maximum volume of the melt region, and the proportionality factor taken is equal to 1 by the Szenes approach [145]. Yet as we all know now, the trajectory may be partially amorphous, or even with no amorphization, such as Gd2Zr2O7. Therefore, this assumption should be revised to that track radius is equal to the maximum melt region [60]. The following equations are derived from the model:
R 2 = a 0 2 ln ( S e S e t ) ,   for   1     S e / S et     2.7  
R 2 = a 0 2 S e 2.7 S e t ,   for   2.7     S e / S et  
S e t = ρ π c a 0 2 ( T m T i r r ) g
where a0 is a parameter which characterizes the initial width of the temperature distribution in the spike, Se and Set are the electronic stopping power and the threshold of electronic stopping power respectively, and Tm and Tirr are the melting and irradiation temperatures. Moll et al. [60] used TS model to fit experiment data and provide that a0 = 5.8 nm and Set = 13.2 keV/nm for Gd2Ti2O7; a0 = 4.9 nm and Set = 13.2 keV/nm for Gd2TiZrO7; and a0 = 4.1 nm and Set = 13.8 keV/nm for Gd2Zr2O7.
The thermal spike model developed by Szenes [145] only considers the process starting when the energy is already transferred to the atoms. If the process of thermalization of the deposited energy on the electron system through electron-electron interaction is also considered [150]. A set of coupled nonlinear differential equations in cylindrical geometry were governed to describe the time and space evolutions of electronic system and lattice temperatures [114,150]:
C e ( T e ) T e t = 1 r r [ r K e ( T e ) T e r ] g ( T e T a ) + A ( r )
C a ( T a ) T a t = 1 r r [ r K a ( T a ) T a r ] + g ( T e T a )
where Ce,a(r,t), Te,a(r,t), and Ke,a(r,t) are the heat capacity per unit volume (called specific heat), the temperature and the thermal conductivity of the electronic (e) and atomic (a) subsystem, respectively. A(r) is the energy deposited into the electronic system, and g is the electron-phonon coupling strength.
The thermal spike proposed by Szenes [145], also called an analytical thermal spike (a-TS) model, provides an analytical solution to the differential equation for the atomic system (4b) ignoring the way the energy is transferred to the atoms. A complete numerical solution of the differential Equation (4) is determined by the inelastic thermal spike (i-TS) model [114,151]. The electronic thermal conductivity is linked to electron diffusivity by De = Ke(Te)/Ce(Te), and the electron-phonon mean free time τ = Ce(Te)/g. So, the electron-phonon mean free path λ2 = De(Te) × τ = Ke(Te)/g and the λ will be the only free parameter in the solution of Equation (4) [151]. Yet a systematic study with several insulators has shown that λ decreases with the optical band gap, as shown in Figure 8a [174]. Based on such empirical correlation, λ has been determined for Gd2Ti2O7, Gd2TiZrO7, and Gd2Zr2O7 to be equal to 4.9, 4.3 and 3.8 nm, corresponding to a calculated band gap of 3.3, 7.4 and 11.5 eV, respectively [175]. For Gd2Ti2O7, using λ = 4.9 nm, the energy Et = 0.8 eV/at for creation of an ion track was determined [143]. Figure 8b presents i-TS calculations of Au ion (7 MeV/u), displaying the deposited energy density on the lattice as a function of (i) time after the ion impact (abscissa) and (ii) cylinder diameter (different lines) relative to the experiment data [135]. The horizontal lines correspond to the high-resolution TEM track diameters for Au (core 8.6 nm and core + shell 11.0 nm) and the energy densities to produce a molten track of given size is Et = 0.8 eV/at for Au ion, while it is yielding an average <Et> = 0.9 eV/at for Ru, Xe, Ta, and Au [135]. Additionally, this <Et> value agrees also with the results from other i-TS calculation for track formation in the same pyrochlore by U ions [143,176,177]. The variation of the calculated track diameters as functions of dE/dx based on <Et> = 0.9 eV/at is displayed in Figure 8c [135] (solid red curve) showing the great agreement with experimental results (round, red data points). The curve for Et = 0.56 eV/at, which corresponds to the sum of the energy required to reach the melting temperature of Gd2Ti2O7 (2123K), also plotted in Figure 8c via i-TS model (black dashed curve). The difference between 0.9 and 0.56 eV/at is about 360 kJ/mol supplying to melt latent heat, which is in good agreement with ~350 kJ/mol for the latent heat determined experimentally for La2Zr2O7 [178]. Additionally, deposited energy of formation amorphization Ea of Gd2Ti2O7 irradiated by Au is plotted by dashed line in Figure 8b, and that from other ion irradiations is between 1.5 and 1.0 eV/at [135].

5.2.2. Direct-Impact Model and Cascade-Overlap Model

Direct-impact model has been completely analyzed and published by Gibbons [147]. Under the conditions of dose and temperature that ensure the production of an amorphous layer, assume that heavy ions create a sufficiently large defect density around the track of the ion to ensure that individual damage clusters are amorphous. For the idealized case, every implanted ion produces an amorphous cluster with a projected area Ai (each cluster is a right circular cylinder), and the total amorphous area AA is the sum of the amorphous area of each implanted ion, which can be written as:
d A A d N = A i 1 A A A 0
where N is the total number of implanted ion, the A0 is the total area being implanted. The factor within parentheses is just the probability that a given ion will hit in the undamaged area. The solution of (5) is
A A = A 0 1 e A i D
where D is the dose or ion fluence. Usually, we do not care about the real area but relative rate or called amorphous fraction fa which represents the ratio of transformation to amorphous phase. Thus, the function (6) was rewritten as
f a = 1 exp ( σ a D )
where fa is the amorphous fraction and σa is the amophization cross-section.
For light ions, that cannot produce individual amorphous clusters without pre-damage, even though the approach to saturation is a function of dose. Thus another model cascade-overlap model was proposed. In this model, also assume that an implanted ion will create a cylindrical cluster with a project area Ai, but that only overlap of clusters can produce an amorphous zone. There are serial equations describing the process of the production of amorphous:
d A A d N = A i A D A 0
d A D d N = A i A u A 0 A D A 0
d A u d N = A i A u A 0
where AA, AD, and Au represent the amorphous area, the damage area but not amorphous and the undamaged area. The solution for AA is
A A = A 0 1 1 + A i D e A i D
An easy extension of this model can be developed for cases in which n-tuple cluster overlap is required for the production of amorphous zones and yields
A A = A 0 1 0 n ( A i D ) k k ! e A i D
Obviously, when n = 0, Equation (10) is the same as Equation (6), and when n = 1, it is the same as Equation (9), which is called single-cascade-overlap model (also called double-impact model). Usually, only two damage overlaps are considered (n = 2), so it is called double-cascade-overlap model, and rewritten in the form of the commonly used amorphous fraction
f a = 1 1 + σ t D + σ t 2 D 2 / 2 exp ( σ t D )
where fa is the amorphous fraction, σt is the total damage cross-section. The double-overlap model generally provides an excellent fit to data for a number of materials that exhibit a sigmoidal dependence of fa on dose.

5.2.3. Direct-Impact/Defect-Stimulated Model

In this model [179], amorphous fractions mainly come from two ways, (i) direct-impact production σaD, and (ii) stimulated amorphization, which takes into account the defect-stimulated, cascade-stimulated, or implanted-ion-stimulated growth of the amorphous production at crystalline/amorphous (c/a) interfaces. The mathematical representation of this model is generally expressed as
d f a d D = P a ( 1 f a ) + P s f a ( 1 f a )
where Pa and Ps are parameters for primary production of amorphous nuclei and stimulated amorphozation. This model has been numerically fitted to results for semiconductor materials, such as Si [179,180], Si1−xGex [181], and SiC [182,183]. Additionally, the solution of Equation (12) is given by the expression [144]
f a = 1 ( σ a + σ s ) / { σ s + σ a exp [ ( σ a + σ s ) D ] }
where σa and σs are cross-sections for primary amorphization and stimulated amorphization, respectively. Within this model, stimulated amorphization occurs at the c/a interfaces (the primary amorphization zone acts as the nucleus), so there must be some direct-impact amporphization (σa > 0) in order for continuous amorphization to “epitaxially” grow, since otherwise fa = 0 if σa = 0. Yang et al. [171] used this model to fit the experimental data for Lu2Ti2O7 irradiated by 2.7 MeV Ar11+ ions, with great fitted results for σs = 0.888 ± 0.102 nm2 and σa = 0.008 ± 0.003 nm2. Zhang et al. successfully used this model to fit the single-crystal Ho2Ti2O7 and Sm2Ti2O7 irradiated by 1MeV Au at room temperature (300 K), and the σa are both equal to 1.0 dpa−1 and σs are 56.8 and 66.0 dpa−1, respectively [104,184].
The above models mainly consider the displacement cascades at low energy where the nuclear energy loss by ballistic collision is dominant and does not take into account simultaneous recovery processes. Furthermore, these models describe the change in amorphous fraction, fa, with dose or time, and can be written in a unified form
d f a d D = i P a i g i ( f a )
where Pai is the amorphous production parameter for some i-th process and gi(fa) is the probability of the process to take place for a given value of fa. For example, for direct-impact model, parameter Pa is the represent the amorphization cross-section, σa, and the probability function gi(fa) is given by the term (1 − fa), and the solution for fa is given by Equation (7). If the recovery process is taken into account, the additional term should be added to the right of Equation (14) and the rate of amorphization is given by
d f a d D = i P a i g i ( f a ) j K j ( T ) h j ( f a )
where Kj(T) is the recovery rate at temperature T for some j-th process and hj(fa) is the probability for that recovery process to occur for a given value of fa. The different recovery mechanisms can have different forms. If the recovery process takes place in the bulk crystalline matrix, hj(fa) is equal to (1 − fa), while at c/a interfaces, hj(fa) is equal to fa(1 − fa), similar D-I/D-S model. If the case of recovery processes in the bulk of amorphous regions, hj(fa) is equal to fa [144].

5.2.4. Multi-Step Damage Accumulation Model

Based on the assumption that the damage accumulation is caused by multiple atomic recombination processes occurring in successive steps, a multi-step damage accumulation model (MSDA) is presented by Jagielski et al. [185]. In this model, each ion impact leads to the transformation of a given volume of the crystal into a new metastable atomic configuration which has the lowest free energy in the current conditions. Additionally, this metastable structure can be an intermediate structure or “nucleation site” in structural transformations during ion irradiation, so the direct amorphization process, the firstly disorder and then amorphization or other process containing intermediate phase transitions are all included. Although, the volume transformed by a single ion impact is comparable to that of a collision cascade for low-energy ion or to that of a damaged track for swift ion, short-range migration of defects and atoms (local atoms rearrangement) significantly affects the transition volume. That is, the previous ion impact will affect the next collision transition. This description is similar to the “overlap” and “stimulated effect” in the previous model, but the assumptions are totally different. The multi-step processes can be written in the following equations. Firstly, the perfectly crystal transforms to a partially damaged structure by a direct-impact mechanism, and therefore it can satisfy the following equation:
f d , 1 = f d , 1 sat 1 exp σ 1 Φ   for   0   <   Φ <   Φ 1
where f d , 1 sat is the saturated damage accumulation in the first step, σ1 is the cross-section of damage, and Φ is the irradiation fluence. Due to the destabilization of the crystal structure causing by implanted ions, a subsequent step of structural transformation occurs. The accumulated damage fd,2 increase gradually from the saturation at the first step f d , 1 sat to that at the end of the second step f d , 2 sat . Additionally, the corresponding effective influence should be replaced ΦΦ1 rather than Φ. So, the equation of stage 2 can be written as
f d , 2 = f d , 1 sat + f d , 2 sat f d , 1 sat 1 exp σ 2 ( Φ Φ 1 )   for   Φ 1   <   Φ <   Φ 2
According to the above considerations, the corresponding equation for stage i following
f d , i = f d , i 1 sat + f d , i sat f d , i 1 sat 1 exp σ i ( Φ Φ i 1 )   for   Φ i 1   <   Φ <   Φ i
The final equation which accounts for MSDA model is the sum of the individual steps m described by (16a–c):
f d = i = 1 m f d , i sat f d , i 1 sat G 1 exp σ i ( Φ Φ i 1 )
where G is a control function which keeps fd nonnegative (converts negative values to 0 while positive values remain unchanged).
Comparing MSDA and the earlier model supported by Gibbons [147] (model in Section 5.2.2), the m and n parameters do not have the same meaning in both descriptions. In MSDA, m is the number of steps required for the development of the entire sequence of damage accumulation, whereas Gibbons n is the number of ion impacts needed to create a permanent disorder in a given region of the crystal. What is more, the function for MSDA is a piecewise function that is a continuous function of the Gibbons model. Specifically, when m = 1, the MSDA model has the exact same form as the direct-impact model (n = 0). Figure 9 [185] shows a three step MSDA model fit well with the low-energy Cs ions irradiation experimental date for yttria-stabilized zirconia (YSZ) and magnesium-aluminate spinel (MAS) [186,187]

5.2.5. Heterogeneous Track Overlap Model

Due to the limitations of the direct-impact model in explaining the inhomogeneous ion tracks in Gd2(ZrxTi1−x)2O7, which coexist a pyrochlore → amorphous (P → A) transition, a pyrochlore → deficient fluorite (P → F) transition and even a deficient fluorite → amorphization (F → A) transition simultaneously. Sattonnay et al. [142] have developed a new phenomenological description, called the heterogeneous track overlap model (HTOM), to describe and explain the phase transformation accumulation (in fact, damage accumulation) when and where heterogeneous track formed.
In this model, an ion track is consisted of a mixture of an amorphous part and a defective-fluorite part, some of which may also have defective pyrochlore phase. The deficient fluorite part due to a P → F transition, but amorphization part comes from both P → A transition and F → A transition. The P → A transition occurs via directly in each individual ion impact, while a second impact into the previously transformed region (P → F region) will produce amorphization via F → A transition. In brief, the HTOM model can be regarded as the direct-impact mechanism coupled with a single-cascade-overlap process. Therefore, the variation in the transformed fraction with the ion fluence following the framework of the direct impact model, since one ion impact is most often sufficient to totally transform the matter inside a track,
f = f sat [ 1 exp ( σ D ) ]
where fsat is the saturated transformed fraction, σ is the section of the transformal cylinder.
Figure 10 provides a schematic for the processes of the HTOM. It assumed that an ion injecting into a pyrochlore region of a sample (marked P) creates a track consisted of both defective fluorite (marked F) and amorphous (marked A) regions. According to this assumption, the variation in the fraction of pyrochlore structure (fP) with the ion fluence follows the following equation:
d f P ( D ) d D = σ PF f P σ PA f P
where σPF and σPA are the cross sections of the cylinder directly transformed by each ion impinge into defective-fluorite or amorphous structures respectively.
Assumption two is that a second ion impact occurring in a disordered region will result in a transformation to an amorphous region. So the variation in the fraction of defective fluorite crystal (fF) is given by:
d f F ( D ) d D = σ PF f P σ FA f F
where the former term on the right-hand side of equation represents the transformation from pyrochlore to defective fluorite regions arising from a first ion impact, and the second term represents the transformation of the defective fluorite into amorphous regions which occurs in a second ion impact, and σFA is the cross sections of the F → A transition. Finally, the variation of amorphous fraction (fA) equal to the sum of P → A transition and F → A transition and can be written as:
d f A ( D ) d D = σ PA f P + σ FA f F
Let the total transformation cross section be σt, with σt = σPF + σPA. The resolution of the system of differential equations (Equation (19a–c)) is thus given by the following expressions:
f P ( D ) = exp ( σ t D )
f F ( D ) = σ PF exp ( σ t D ) [ 1 + exp ( σ t σ FA ) D ) ] σ t σ FA
f A ( D ) = 1 σ PA σ FA σ t σ FA exp ( σ t D ) σ PF σ t σ FA exp ( σ FA D )
It is worth noting that the direct-impact model can be regarded as the special case of HTOM model. When in Equation (20a–c), σPF = σFA = 0 and σT = σPA, that is, the tracks are totally amorphous, or σPA = σFA = 0 and σT = σPF, there is no amorphization occurs, the Equation (20a–c) are simplified to Gibbons equation form (Equation (7)).
Figure 11 [142] shows the experimental data and fitting line by Equation (20a–c) for phase transformation accumulation process for typical pyrochlore with irradiation under 120 MeV U ion, including Gd2Ti2O7, Gd2TiZrO7, Gd2Zr2O7, and Nd2Zr2O7. Among them, Gd2Ti2O7 and Gd2Zr2O7 are more special, the former only a pyochlore (P) → amorphous (A) transformation is observed, and the latter only a pyochlore (P) → defective fluorite (F) transformation occurs. The HTOM has also been successfully applied to Y2Ti2O7 irradiated by 93 MeV Xe, and σPA, σPF, and σFA are 46.2 nm2, 21.0 nm2 and 67.2 nm2, respectively [188].
The HTOM fitting results σi of pyrochlores with chemical composition which irradiated by 119 MeV U are plotted in Figure 12a, marked by solid symbols [61,62,142]. Both σPA, σT, and σFA clearly increase with increasing rA/rB, whereas σPF globally decrease as rA/rB increasing with a relatively stable value ~50 nm2 for rA/rB between 1.5 to 1.6. The consequences of Nd2Zr2O7 [62] and Y2Ti2O7 [188] which irradiated by 93 MeV Xe ions are also shown in Figure 12, using hollow marks. For Figure 12b, using the σi/σT instead of σi as the ordinate, that is, normalized treatment of σi, which excludes the effect of variation of σT on σPF. It is clearly demonstrated that the susceptibility to amorphization by electronic excitation of pyrochlore and fluorite phases increases with increasing rA/rB. For 93 MeV Xe ions irradiation, due to the low Se energy, the σi is smaller than that of corresponding pyrochlores irradiated by 119 MeV U ions, while the σPF/σT and σFA/σT are larger than that of ceramics irradiated by 119 MeV U ions. Due to the fact that the weaker electronic excitation (lower Se) lead to smaller track cross-sections and make more fluorite fraction for single ion damage process. This phenomenon conforms to the laws of the TS model, and many pyrochlores show similar rules after being ion-irradiated in a broad energy range, including Nd2Zr2O7 [62], Gd2(ZrxTi1−x)2O7 [135]. Further, the diameter of ion tracks (d) may be deduced from the value for σ according d = 2(σ/π)1/2. Thus, the total track diameters (dT) and amorphous core diameter (dA) are calculated and plotted vs. rA/rB in Figure 12c. Both dT and dA globally decrease as rA/rB decreases, but the variation rule of dT is nearly linear, while dA changes linearly with a greater slope during rA/rB from 1.75 to 1.5 and decreases rapidly below rA/rB = 1.5. The hollow marks in Figure 12c represent pyrochlores irradiated by 93 MeV Xe ion showing a similar rule. Moreover, these data also indicate that the radiation resistance of pyrochlores depends more on the cation radius ratio (rA/rB), both A and B cations play a significant role in the amorphization process, than on the B cation itself. Additionally, this rule is under a huge implication for the design of future matrices for waste forms. If the B cation is Zr and the A cation is Am, rA/rB = 1.51, so that the probability that Am2Zr2O7 amorphizes upon electronic excitation is high. A similar conclusion is obtained with A = Np (rNp/rZr = 1.60) or A = Cm (rCm/rZr = 1.54). In fact, 241Am2Zr2O7 shows excellent resistance to amorphization during self-irradiation (alpha decay) for 40 days (0.02 dpa) and 370 days (0.21 dpa), respectively [189].

5.3. Temperature Dependence

As previously stated, the processes of ion irradiation involve transient thermal process, quenching and self-annealing process. Each step involves heat diffusion, especially during self-annealing process which significantly affects the rearrangement of atoms and the restoration of structures. This also means that the ion irradiation damage process shows a strong temperature dependence.

5.3.1. The Critical Amorphous Temperature

Ordinarily, for ion-irradiation experiments of this special material, the higher the temperature is, a higher the ion dose required for the material to be completely amorphous. Even when the temperature arrives at a certain degree, the recrystallization rate is higher than the damage rate, thus, the complete amorphous cannot be approved even if the dose is higher. So, the critical amorhization fluence FC (or called critical amorphous dose DC) is defined as the ion fluence (dose) required for full amorphiozation at a special temperature [91,190,191]. Additionally, the critical amorphous temperature TC is defined as the lowest temperature above which no full amorphization occurs in the materials under ion-irradiation. In other words, at the critical amorphous temperature, the critical amorphous dose tends to infinity. Gd2Ti2O7 will occur fully amorphous transform by 1 MeV Kr at low dose, about 0.2 dpa during 320 to 500 K [139] which is similar to the dose for Cm-doped Gd2Ti2O7 [25], and this dose is equivalent to 4 × 1018 α decays per gram or 600 years of storage for a waste form containing 10 wt% 29Pu. However, the critical amorphous fluence FC increase with the increase of temperature, for Gd2Ti2O7, the FC increase sharply above 800 K, and the upper temperature limit for amorphization is about 1100 K. While for Gd2(Ti1−xZrx)2O7, with Ti substituted by Zr, the FC increase and the TC decrease. For Zr-rich composition (x = 0.5), there is no amorphization above 380 K, whereas highly enriched in Zr, for x > 0.75, amorphization is extremely difficult to achieve at ultra-low temperature (cryogenic temperatures), and only partial amorphization was observed (30% amorphous for x = 0.75) [139]. For end member, Gd2Zr2O7, irradiation to extremely high ion fluences, 1.5 MeV Xe at 25 K with equivalent dose 7 dpa or at room temperature with 15 dpa, showed no evidence of amorphous regions in the crystalline matrix [138,139,192]. For comparison, the dose of 15 dpa is equivalent to 3 × 1020 α decays per gram or the accumulated dose for a waste form containing 10 wt% 239Pu storing over 30 million years. It confirms that compositions in this regime will be extremely durable and radiation resistance to amorphization under repository conditions. In Zr-rich pyrochlores (e.g., Gd2Zr2O7, Gd2Zr1.8Mg0.2O6.8, Gd1.9Sr0.1Zr1.9Mg0.1O6.85, and Gd1.9Sr0.1Zr1.8Mg0.2O6.75, the pyrochlore-to-fluorite structure transformation at low temperatures is complete after a dose of about 0.4 dpa, whereas could not be amorphized even at temperature as low as 25 K [139]. Up to now, there are few reports of complete amorphization of rare earth zirconated pyrochlores, e.g., La2Zr2O7 with TC = 310 K under 1.5 MeV Xe irradiated and critical amorphous dose 1.19, 3.55, and 5.20 dpa for (La1−xCex)2Zr2O7 (x = 0, 0.1 and 0.2) with 1 MeV Kr at 25 K, respectively [65,91].
Wang et al. [193] developed an empirical model based on cascade “quenching” and epitaxial recrystallization to describe the temperature effect of the accumulation of the amorphous fraction during ion beam irradiation experiments. Weber also deduces the expression of the critical temperature, TC. The critical fluence (FC) versus temperature data are fitted to follow equation [193]:
F C = F C 0 1 exp [ E a k b × ( 1 T C 1 T ) ]
T C = E a k b ln ( ν F C 0 Φ )
where FC0 is the critical fluence at 0 K, Ea is the activation energy for recovery process, kb is the Boltzmann constant, T is the temperature of the irradiation experiment, Φ is the ion flux of ion irradiation and υ is the effective attempt frequency.
The critical temperature of a serial of pyrochlores was fitted by Equation (21) and plotted as a function of cation radii ratio in Figure 13 [55,66,68,91,93,133,139,190,191,192,194,195,196]. The triangle represents the data fitted from pyrochlores which are irradiated by 1 MeV Kr, while square, circle and rhombus represent that irradiated by 1.5 MeV Xe, 0.6 MeV Ar and 0.6 MeV Bi, respectively. A series of colors are used to distinguish the types of pyrochlores, including grey for actinide doped pyrochlores [66], red for rare earth titanate pyrochlores [68,93,133,190,191,192,195], green for A2(Ti1−xZrx)2O7 [55,139], purple for rare earth stannate pyrochlores [55,194], yellow for Y2(Ti1−xSnx)2O7 [55], orange for rare earth hafnate pyrochlores [55], and blue for La2Zr2O7 [91], respectively. The triangles with black edge are calculated by Lumpkin et al. [55] from the model (Equation (2) in Ref. [196]) which obtain by multiple linear regression method fitting data from ab-initio calculation. It is clear from Figure 13 that there is a certain correlation between the critical temperature (TC) and the ratio of cation radius (rA/rB), so the data obtained by 1 MeV Kr ion irradiation, except the three points with black arrow (Y2TiO5, Yb2TiO5 and YbYTiO5), are linearly fitted and drawn by a red solid line in the Figure 13. The fitting equation is following TC = 3335.9 rA/rB − 4792.7, and coefficient of determination R2 is equal to 0.85. The TC increases with the increase of rA/rB, obviously. However the rA/rB of data for fitting is located between 1.51 and 1.78, so the fitting formula is not suitable for rA/rB lower than 1.51. When the rA/rB of pyrochlores is lower than 1.52, the critical temperature decreases rapidly and approaches 0 K. At an ultra-low temperature, the pyrochlores with a small cationic radius ratio will not be completely amorphous and even does not appear amorphous at a larger irradiation dose with low-energy ions. For example, no amorphization occurred for Y2Sn2O7 (rA/rB = 1.48), Er2Sn2O7 (rA/rB = 1.46), Gd2Zr2O7 (rA/rB = 1.46), and Gd2Hf2O7 (rA/rB = 1.48), under 1 MeV Kr ion irradiation at 25 K at a dose of ~5.18, ~5.94,~5, and ~4.54 dpa, respectively [194,197]. Moreover, under 1.5MeV Xe ion-irradiation at 25 K, only disordered fluorite transform occurs for Gd2Zr2O7 (rA/rB = 1.46), Sm2Zr2O7 (rA/rB = 1.50), Nd2Zr2O7 (rA/rB = 1.54) at a dose of ~36, ~7.3, and ~7 dpa [91]. Ce2Zr2O7 (rA/rB = 1.58) is a special one; there is no amorphous for it under MeV Kr ion irradiation at 25 K at a dose of ~7.08 dpa [65], which also suggests that zirconated pyrochlores have better irradiation resistance.
What is more, except for rA/rB, the composition of B-site also influences the critical temperature. The critical amorphization temperatures of stannate pyrochlores are much higher than those of titanate, hafnate, and zirconated pyrochlores having similar ionic radius ratios, TC(A2Sn2O7) > TC(A2Ti2O7) > TC(A2Hf2O7) > TC(A2Zr2O7), even stannate pyrochlores have a smaller rA/rB and higher TC than those of titanate pyrochlores. For example, the rA/rB of La2Sn2O7 and Y2Ti2O7 are both 1.68, however, the critical amorphization temperatures of La2Sn2O7 and Y2Ti2O7 with 1 MeV Kr irradiations are 960 and 780 K, respectively [194,195]. Additionally, the TC of La2Sn2O7 is higher than that of A2Ti2O7 (A = Ho, Er, Yb, Lu) whose rA/rB are from 1.677 to 1.615 and TC are from 850 K to 480 K [194,195]. The critical amorphization temperature of Nd2Sn2O7 is 700 K is significantly higher than that of Lu2Ti2O7 (480 K) and Y2Ti1.2Sn0.8O7 (335 K), although the ionic radius ratio of Nd2Sn2O7 (rA/rB = 1.607) is slightly less than that of Lu2Ti2O7 (rA/rB = 1.61) and is slightly larger than that of Y2Ti1.2Sn0.8O7 (rA/rB = 1.595) [55,194,195]. In addition, although the rA/rB of Gd2Sn2O7 (1.526) is similar to that of Gd2(Zr0.75Ti0.25)2O7 (1.523), complete amorphization cannot be approached in Gd2(Zr0.75Ti0.25)2O7 with 1 MeV Kr irradiation at 25 K while the critical amorphization doses are ~0.81 dpa for Gd2Sn2O7 [139,197]. Furthermore, La2Zr2O7 has a larger cation radius ratio (rA/rB = 1.61) than Gd2Sn2O7 (rA/rB = 1.526), the amorphization does (DC) for La2Zr2O7 irradiated by 1 MeV Kr at 25 K and room temperature are ~1.19 dpa and ~3.42 dpa, respectively, whereas DC for Gd2Sn2O7 at 25 K and room temperature are ~0.81 dpa and ~3.4 dpa, respectively [65,194,197]. The rA/rB of La2Hf2O7 (1.633) is slight higher to that of Yb2Ti2O7 (1.628), while the TC of La2Hf2O7 is 563 K, which is significantly less than that of Yb2Ti2O7 (594 K (poly-crystal) or 611 K (single-crystal)) [195,196], and the similar rule for La1.6Y0.4Hf2O7 (rA/rB = 1.594, TC = 473 K) and Nd2Sn2O7 (rA/rB = 1.607, TC = 700 K). It is of note that although Y3+ (0.1019 nm) is slightly larger than Ho3+ (0.1015 nm), Y2Sn2O7 and Y2Ti2O7 have a higher irradiation resistance (i.e., a higher critical dose for amorphization or lower critical amorphous temperature) to ion beam induced amorphization than Ho2Sn2O7 and Ho2Ti2O7, respectively. This deviation for yttrium pyrochlores is consistent with the mass of yttrium (88.906 g/mol) which almost half of that of holmium (164.93 g/mol). Which will lead to low density, small energy deposition (both electronic and nuclear), the deeper ion range and more easily transform to fluorite structure than holmium pyrochlores [194,195].
This more complex behavior of different types of pyrochlores may reflect different cation electronic configurations in bond type. The stronger covalent character of the <Sn–O> bond of stannate pyrochlores implies a lesser degree of distortion of the SnO6 coordination octahedron which result in a structure more close to the ordered pyrochlore superstructure [198]. For example, the <Sn–O> bond length in Gd2Sn2O7 is approximately 0.2048 nm, which is smaller than the sum of the effective ionic radii of O2− (0.138 nm for CN = 4) and Sn4+ (0.069 nm for CN = 6), implying the more covalent for the <Sn–O> bond. Thus, the position parameter of O48f (x48f) reported for Gd2Sn2O7 (0.3348) is slightly lower than Nd2Zr2O7 (0.335) although that the cation radius ratio of Nd2Zr2O7 (1.540) is greater than that of Gd2Sn2O7 (1.526) [35]. Additionally, it lead to a greater susceptibility of stannate pyrochlore to ion irradiation induced amophization, comparing with the titanate, hafnate, and zirconate pyrochlores with similar rA/rB.
First-principle calculations based density-functional theory (DFT) point that a conspicuous covalency for the <Sn–O> bond and mainly ionic character for the <Ti–O>, <Hf–O> and <Zr–O> bonds [117,196,199]. Depending on the Mulliken overlap population analysis [196], the covalency of <B–O> bonds increases in the order Sn > Ti > Hf > Zr and this trend roughly follows that of Pauling’s electronegativity. Additionally, the disordering energies (the energy difference between the fully ordered pyrochlore and the fully disordered defect-fluorite structures) of A2B2O7 were calculated and result implies that for zirconate, hafante, and stannate pyrochlores, disordering energy approximately linear increases with increasing A3+ size [37]. However, the titanate pyrochlore series have a maximum disordering energy at Gd2Ti2O7 [37] which has the highest critical amorphization temperature among all titanate pyrochlores [195]. Other theoretical calculation based empirical potentials [38,47,48] or DFT [200], showing agreement with these results. Disordering energies of stannate pyrochlores are significantly higher than corresponding titanate pyrochlores, although the ionic radius of Sn4+ (0.069 nm) is much larger than that of Ti4+ (0.0605 nm). Furthermore, it is obvious that the disordering energy of a given hafnate pyrochlore is always larger than that of its corresponding zirconate substitution, despite the very similar ionic radius of Zr4+ (0.072 nm) and Hf4+ (0.071 nm) [37]. Such phenomena also can be revealed that the covalency of <B–O> increases in order of Sn > Ti > Hf > Zr, which cannot be simply explained by the rA/rB. The phase transform temperature of the order-disorder transition also proves above conclusion [201,202].

5.3.2. Recover Process

Re-crystallization process of fully amorphous pyrochlores has also been investigated. Firstly, amorphous materials crystallize to form a disorder fluorite structure; then, the stable pyrochlore phase is recovered. Temperatures of forming defective fluorite structure and pyrochlore structure are related to the irradiation resistance. For example, for Nd2Zr2O7, the formation of fluorite structure occurred at 598 K [62], while the temperature of crystallization of Gd2TiZrO7 is 948 K [153], indicating that the amorphous phase is less stable in Nd2Zr2O7 than in Gd2TiZrO7, and, as we all knew, Nd2Zr2O7 has better irradiation resistance than Gd2TiZrO7. The annealing process is different for Gd2Ti2O7, which shows that the recrystallization of amorphization towards its pyrochlore structure occurs directly at 1073 K [153]. It is due to the extortionate formation energy of cation antisite defects to no thermal cation disordering could occur [39,203]. Park et al. [204] pointed that the recrystallization process depends on the energy difference between the amorphous and possible crystalline structures, because of the different driven force of the atomic rearrangements to recover the structures. La2Ti2O7 has a higher structural freedom degree, fewer structural constraints, than Gd2Ti2O7, hence, a lower energetic barrier to rearrangement is, which lead to a lower onset crystallization temperature than that of Gd2Ti2O7 [204]. The formation enthalpies of those compounds could be another piece of evidence. Gd2Ti2O7 has a formation enthalpy of −3822.5 kJ/mol which is 33 kJ/mol higher than that of La2Ti2O7 [205]. Chung et al. [41,42] suggest that use amorphization enthalpy accounting for all the factors that affect the amorphization resistance as a more effective parameter to predict the irradiation toleration of a given chemical composition pyrochlore. Additionally, this method can also be used for other potential waste form material research [206]. Combing high temperature differential scanning calorimetry (DSC) and neutron total scattering experiments, the damage recovery process of pyrochlore is shown a complex multistep processes with decoupled long-range recrystallization and short-range defect recovery operation at different temperature and time scales [41,42]. The disorderd and the amorphous pyrochlores have a similar local atomic arrangements described as weberite-like [36,44,46], which explains the onset temperatures for the second event of Dy2Ti2O7 [42], Dy2Sn2O7 [41], Er2Ti2O7 [45], and Gd2Ti2O7 [204] are all close to 1148 K (the thermal energies required to trigger the effective atomic movements for the local structural re-arrangement for different compositions are nearly identical). Yet the initial long-range recrystallization temperatures are compositionally dependent, which is conformity to the experimental results mentioned above. Although the re-crystallization processes are multifaceted, the initial crystalline temperature increase with the increasing of rA/rB, globally [41,42,45,62,153,204,207]. Atom-scale simulations based on density functional theory have revealed that differences between the formation enthalpies or disordering enthalpies of weberite-type and defect fluorite for A2Sn2O7 and A2Ti2O7 are nearly constants, respectively, while the formation enthalpy of pyrochlore is composition dependent [208,209]. Additionally, those DFT simulations also suggest that the weberite-type structure is more energetically stable than the fully ordered pyrochlore structure for most zirconate and hafnate pyrochlores while almost the whole titanate and stannate pyrochlores prefer ordered pyrochlore structure [208,209]. Janish et al. [105] show the re-crystallization process of Gd2Ti2O7 amorphized by He ion irradiation. The results show that at lower temperature (1098 K), the initial recrystallization starts at the interface between the irradiated and unirradiated material, where the primary pyrochlore acts as a center of nucleation, but the new crystal grain grown is quickly hindered by the porous network of the He implanted region. At higher temperature (1148 K), the recrystallization front occurs at the top of the porous zone, which the He bubbles act as the nucleation center [105].

5.4. Size Dependence

It is worth mentioning that the critical amorphous dose of material and critical amorphous temperature of the material are also affected by the grain size. Zhang et al. [210] have systematically studied the Kr irradiation induced amorphous response to Gd2(Ti0.65Zr0.35)2O7 with different processing temperature and corresponding crystal size. Gd2(Ti0.65Zr0.35)2O7 powders maintain nanocrystal (~18 nm) annealing at less than 1000 °C, and are not completely amorphous under 1MeV Kr irradiation at room temperature (partially amorphous with ~1.72 dpa for 1000 °C), whereas the critical amorphous dose of larger-grained crystals (more than 100 nm) is ~0.73 dpa [210], which is similar to Wang’s result (DC = 0.66 dpa) [139]. Wen et al. [93] have reported that the grain nanomatization improves the critical amorphous dose at room temperature and the critical amorphous temperature of titanate pyrochlores. On the one hand, the defect formation energy is reduced as the defect is formed closer to the surface, which indicates that, at the same defect concentration, the larger specific surface areas, the lower average defect formation energy is [93]. Which also explains the variation of the degree of disordering as grain size, e.g., occupancy of Gd in 16c changes from 0.57 to 0.87 as the crystal size changing 17 to 380 nm, although most variation occurs at annealing temperatures during 1273 K (~18 nm, occupancy 0.57) and 1473 K (~100 nm, occupancy 0.8) [210,211]. Furthermore, susceptibility of pyrochlores to amorphization has been proved to be related to the energetics of disordering [39]. Thus, nanomaterials can able to tolerate disorder defects in the structure than coarse crystal materials, that is, higher radiation-induced amorphous resistance. Similarly, enhanced irradiation resistance has been shown in other nanomaterials including alloys and ceramics, e.g., Au (~23 nm) [212], Pd (<30 nm) [213,214], TiNi alloy (23–31 nm) [215], MgGa2O4 (4–12 nm) [216], CePO4 (20 nm or 40 nm) [134] and ZrO2 (<15 nm) [213,214]. On the other hand, MD simulations have shown that the interfaces especially, grain boundary, act a sink to effectively annihilate defects, facilitating the self-healing of the radiation-induced damage [217,218]. The cell volume expansion, the amorphization fraction and the threshold He or Xe concentrations of bubble formation are size dependent for Gd2Zr2O7 [94,170]. For example, as the average grain size ranges from 221 to 55 nm, the He concentration thresholds for helium bubble nucleation along grain boundaries, formation of helium bubble chains within the grain, and intragranular ribbon helium bubbles are 1.7% to 2.8%, 5.9% to 15.2%, and 19.2% to more than 25.8%, respectively [170]. However, nano-crystalline materials are not intrinsically radiation-tolerant, and an optimized size regime may exist that displays the greatest radiation tolerance against amorphization [134,210]. Meldrum et al. [219] found that nano-crystalline ZrO2 (~3 nm) embedded in an amorphous SiO2 matrix is completely amorphous with a dose of 0.9 dpa at ambient temperature, while bulk pure ZrO2 cannot be amorphized even at very high ion does of 680 dpa [220]. Similarly, for monazite CePO4, nanoparticles show lower critical temperature and higher critical dose than their bulk counterpart (~450 K), whereas, ~20 nm size CePO4 nanoparticles shows a higher critical temperature (~435 K) and lower critical dose (~0.35 dpa) than the 40 nm sized monazite (~413 K, ~0.47 dpa) [134,221]. This phenomenon has been attributed to the delicate balance between the bulk free energy and surface free energy of polymorphs and the defect free energy introduced by energetic ions [219]. A small grain size has two opposing effects on the energy of an irradiated material: (i) a smaller grain size hinders the accumulation of vacancies in the grain interior, which decreases the free energy resulting from irradiation-induced defects; and (ii) a smaller grain size increases the free energy resulting from the grain boundary [222]. The higher density defect and the higher surface free energy of ultrafine materials provide additional energy for the amorphous process. Thus, the grain size must be optimized.

6. Conclusions

6.1. Summary

During irradiation, pyrochlore ceramics lose their long-range order gradually, transform to defect fluorite structure, and eventually convert into amorphizaiton. For swift ion irradiation, the morphology of ion track is core (amorphous) + shell (defect fluorite) ion track in pyrochlore, while for low energy implanted ion, loss their crystal structure through defect accumulation without obvious ion track. In addition, the cation radii ratio rA/rB is a key parameter of pyrochlore A2B2O7, which is related to the lattice parameter, phase structure, properties, especially radiation resistance. Generally, cation radii ratio rA/rB plays a dominant role of the radiation response of different pyrochlore compositions, but the B site ionic species also shows significant influence of the radiation tolerance. The irradiation resistances of zirconate and hafnate pyrochlore are much stronger than that of titanate and stannate pyrochlore, which is mainly related to disordering energy. The weak covalency of <Zr–O> in zirconate pyrochlore enables to accommodate more defects and has lower disordering energy, which is the fundamental reason for the excellent irradiation resistance performance of zirconate pyrochlore. It is of great significance for the design of waste form for different waste streams.

6.2. Outlook

Although the research so far on pyrochlore ceramics has demonstrated their excellent radiation resistance and great potential as a candidate waste form for radioactive waste, further effort on some fundamental and processing aspects is still necessary. The areas identified for further investigation include but are not limited to the following.
Matrices with Zr, being the dominant B site metal hosted, are preferred as these materials offer great irradiation resistance. Doped Ti or other element on B site should also be investigated to further decreased leach rates or improve synthesis process on an industrial scale to find a balance between cost and performance. In recent years, high entropy materials have attracted a lot of attention. High entropy alloys show excellent properties in anti-irradiation applications [223,224,225,226,227,228], while high entropy ceramics as potential waste forms are rarely studied [77,107,229,230] and need to be comprehensive and in-depth research. The maximum amount of radionuclide solidification needs to be accurately studied, which is essential for industrial applications and the development of safety guidelines. Additional methods such as glass-ceramic should be investigated. This would result in further increased stability but also limit the load density of radionuclides.
The ion irradiation experiment is an accelerated simulation test of radionuclides self-irradiation in waste forms. There is a distinct lack of literature comparing the ion irradiation test to actual radionuclides self-irradiation. Additionally, for long term geological disposition, a long-term ion irradiation test, and an industrial scale synthesis and irradiation study are necessary. This review mainly focuses on the irradiation damage on pyrochlores, but the leaching studies of pyrochlores are equally important, especially the long- term leaching research and irradiation damage coupled leaching study which are heavily lacking.

Author Contributions

Conceptualization, Y.W. and J.-H.O.; validation, Y.-Z.Z., Y.W. and T.W.; formal analysis, Y.W. and C.J.; investigation, Y.-J.W., C.J., J.-H.O. and Y.-M.W.; resources, J.-H.O.; data curation, Y.W, and C.J.; writing—original draft preparation, Y.W. and C.J.; writing—review and editing, Y.W., J.-H.O., Y.-J.W. and Z.-Y.D.; visualization, Y.W.; supervision, Z.-G.L.; project administration, Z.-G.L.; funding acquisition, J.-H.O. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to thank the financial support from the National Science and Technology Major Project of China, grant number 2017-VI-0020-0093.

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.

Appendix A

The ion irradiation experiences mentioned and other reported in literature are summarized in Table A1, including composition, irradiation ion, the range of flux or fluence, critical amorphization dose or fluence and corresponding reference.
Table A1. Summary of publications of pyrochlore ion-irradiation.
Table A1. Summary of publications of pyrochlore ion-irradiation.
CompositionIrradiation IonRang of Flux or FluenceDC or FCRef.
Gd2Ti2O73 MeV Ar5 × 1013–3 × 1016 ions/cm2 [231]
Gd2Ti2O7
Er2Ti2O7
Lu2Ti2O7
CaZrTi2O7
1.5 MeV Xe,
1 MeV Kr,
0.6 MeV Ar
20–1073K
8.5 × 1011 ions/cm2/s0.23 dpa (RT)
0.25 dpa (RT)
0.51 dpa (RT)
[102,133]
Gd2Ti2O7
Sm2Ti2O7
Eu2Ti2O7
Y2Ti2O7
1 MeV Kr
25–1023 K
8.5 × 1011 ions/cm2/s~0.5 dpa (RT)
0.8 dpa (RT)
[190]
Gd2(Ti2−xZrx)O7
(x = 0, 0.5, 1, 1.5, 2)
1 MeV Kr
20–1100 K
0.2 dpa (320–500 K) for x = 0[139]
Gd2(Ti2−xZrx)O7
(x = 0, 0.5, 1, 1.5, 2)
2 MeV Au RT5 × 1014 ions/cm23 × 1014 ions/cm2 for x = 0[138]
Gd2(Ti0.65Zr0.35)O71 MeV Kr2.5 × 1015 ions/cm20.73 dpa (FA, coarse-grained, >100 nm)
1.75 dpa (PA, nanocrystal, <20 nm)
[210]
A2Ti2O7
(A = Y, Sm, Gd, Lu)
2 MeV Au RT 0.18 ± 0.01 dpa (RT)[191]
Sm2Ti2O71 MeV Au
(170, 300, 700 K)
1 × 1012–8 × 1013 ions/cm20.14, 0.14, 0.22 dpa for 170, 300, 700 K[184]
A2Zr2O7
(A = La, Nd, Sm, Gd)
1.5 MeV Xe
25–300 K
8.6 × 1011 ions/cm2/s5.5 dpa (RT) and 0.65 dpa (25 K) for La2Zr2O7[91]
Gd2Zr2O7200 keV Ti RT1017 ions/cm2
Single crystal Er2Ti2O7
Single crystal Er2Zr2O7
350 KeV Xe++ 120 K1 × 1015–5 × 1016 ions/cm21 × 1015 ions/cm2[39,232]
(La1−xCex)2Zr2O7
(x = 0, 0.1, 0.2, 1)
1 MeV Kr
25 K and RT
1.25 × 1012 ions/cm2/s3.42 dpa (RT) and 1.19 dpa (25 K) for x = 0; 3.55, 5.20, 7.08 dpa (25 K) for x = 0.1, 0.2, 1[65]
An2Zr2O7
An = Am, Cm, Bk, Cf
Self-irradiation [233,234]
Single crystal A2Ti2O7
(A = Sm, Eu, Gd,
Dy and Er)
1 MeV Kr (RT)5 × 1014 ions/cm2
(1.74 × 1014 ions/cm2 for Er)
0.11, 0.138, 0.143, 0.160, 0.229 dpa (RT) respectively[103]
Single crystal A2Ti2O7
(A = Sm–Lu, Y)
1 MeV Kr
(RT–1073 K)
6.25 × 1011 ions/cm2/s0.25, 0.23, 0.18, 0.11, 0.25, 0.27, 0.25, 0.45, 0.55, 0.37 dpa (RT) for A = Sm–Lu, Y[195]
A2Sn2O7
(A = La–Lu, Y)
1 MeV Kr
(25–973 K)
6.25 × 1011 ions/cm2/s0.32, 0.61, 1.41, 3.40 dpa (RT) for A = La, Nd, Eu, Gd;
0.37, 0.81, 2.57, 4.36 dpa (25 K) for A = Nd, Gd, Dy, Ho.
[194]
Gd2Sn2O7
Gd2Hf2O7
1 MeV Kr
(25–300 K)
1.25 × 1012 ions/cm2/s3.4 dpa (RT)
4.54 dpa (RT), NA
[197]
Ca1.04Ce0.97Ti1.99O6.96, Ca0.62Gd0.97U0.23(Zr0.84Ti1.34)O6.90
Ca0.47Gd0.95Th0.40Zr1.29Ti0.89O7.05,
(Ca0.44GdTh0.42)Zr2.13O7.05,
Ca0.91Th0.84Zr2.25O7.09,
1 MeV Kr
(293–1173 K)
2.0 × 1014 ions/cm2 (RT)
3.75 × 1014 ions/cm2 (RT)
5.62 × 1014 ions/cm2 (RT)
NA
NA
[66]
Nd2Zr2O7
Gd2Zr2O7
Gd2Ti2O7
120MeV Au
90MeV I
70MeV Ni
1 × 1012–1 × 1014 ions/cm2
3 × 1010 ions/cm2/s
[67]
Gd2Ti2O7
Gd2ZrTiO7
Gd2Zr2O7
4 MeV Au
1.5 GeV Xe
2.6 GeV U
1 × 1012–1 × 1016 ions/cm20.4 dpa
0.8 dpa
[54]
1 × 1011–5 × 1012 ions/cm2
Gd2Zr2−xTixO7
(x = 0, 0.5, 1, 1.5, 2)
1.43 GeV Xe;
4 GeV Xe
7 GeV U
(10, 20, 40 GPa)
1 × 1010–1 × 1013 ions/cm2
~3 × 1012 ions/cm2
[140,141]
Y2Ti2−xSnxO7
(x = 0, 0.4, 0.8, 1.2, 1.6, 0)
La2Hf2O7, La1.6Y0.4Hf2O7
Nd2Zr2O7, Nd2Ti0.8Zr1.2O7
1 MeV Kr
(50–650 K)
6.25 × 1011 ions/cm2/s6.4, 23.4, 26.6, 5.3 × 1014 ions/cm2 (RT) for x = 0.4, 0.8, and La1.6Y0.4Hf2O7, Nd2Ti0.8Zr1.2O7[55]
Am2Zr2O7Self-irradiation
(α-decay)
40 days (0.02 dpa)
370 days (0.21 dpa)
[189]
100-oriented single crystals YSZ940 MeV Pb
850 MeV Xe
2.6 GeV U
1 × 1010–1 × 1013 ions/cm2
1 × 108 ions/cm2/s
[152]
REE2(U0.06Sn1.94)O7 REE2(U0.06SnZr.94)O7
REE2(U0.06Zr1.94)O7
REE = 0.12La + 0.24Ce + 0.10Pr + 0.38Nd + 0.08Sm + 0.05Eu + 0.03Gd
1 MeV Kr1 × 1014 ions/cm2/s [235]
Gd2Zr2−xTixO7
(x = 0, 1, 2)
870 MeV Xe
780 MeV Kr
(RT)
1 × 1010–1 × 1013 ions/cm2
1 × 108 ions/cm2/s
[60]
Gd2Zr2−xTixO7 (x = 1, 2)870 MeV Xe2 × 1011–1 × 1013 ions/cm2 [236]
Single crystal Ho2Ti2O71.0 MeV Au
(300 K, 7° off <001> direction)
2 × 1012–3.1 × 1013 ions/cm20.12 dpa[104]
Lu2(Ti2−xLux)O7−x/2
(x = 0, 0.4, 0.67)
400 keV Ne
(77 K)
1 × 1015–1 × 1016 ions/cm2
1 × 1012 ions/cm2/s
[56]
Y2Ti2O7
Yb2Ti2O7
Y2TiO7
YbYTi2O7
Yb2Ti2O7
1 MeV Kr
(50–650 K)
6.25 × 1011 ions/cm2/s [68]
Gd2TiO5, Gd2Ti2O7, Gd2TiZrO7, Gd2Zr2O71.1 GeV Ru
2.2 GeV Au
(8 K or 300 K)
5 × 1010–1 × 1011 ions/cm2
1 × 108 ions/cm2/s
[137]
REE2(Ti1−xZrx)2O7, x = 0–1
REE = 0.24La + 0.46Ce + 0.21Pr + 0.75Nd + 0.15Sm + 0.02Eu + 0.03Gd + 0.14Y
1 MeV Kr
(RT)
0.3 dpa for REE2Ti2O7[237,238]
Er2Ti2O7400 keV Ne
(77 K)
2 × 1018–2 × 1019 ions/cm2
1 × 1016 ions/cm2/s
[71]
Y2Ti2O7, Gd2Ti2O7, Gd2Hr2O7400 keV Ne
(77 K)
5 × 1014–6 × 1016 ions/cm2
1 × 1012 ions/cm2/s
[72,239]
Gd2Ti2O7, Gd2ZrTiO7, Gd2Zr2O7, Eu2Zr2O7, Sm2Zr2O7, Nd2Zr2O7120MeV U RT2 × 1011–4 × 1013 ions/cm2
2 × 109 ions/cm2/s
[142]
Gd2Ti2O7, Gd2Zr2O7, Eu2Zr2O7, Sm2Zr2O7, Nd2Zr2O7119MeV U RT2 × 1011–4 × 1013 ions/cm2
1 × 108 ions/cm2/s
[61]
A2TiO5
A = La, Nd, Sm, Gd
1.47 GeV Xe
(RT)
[240]
Nd2TiO5
Gd2TiO5
Yb2TiO5
2.2 GeV Au
(RT)
−3 × 1013 ions/cm2~7.5 × 1012 ions/cm2
~1.5 × 1013 ions/cm2
~3 × 1013 ions/cm2
[207]
Nd2Zr2O74 MeV Au1 × 1012–1 × 1016 ions/cm2 [62]
74 Mev Kr
92 MeV Xe
119 MeV U
900 MeV U
940 MeV Pb
2 × 1011–2 × 1014 ions/cm2
2 × 109 ions/cm2/s
Gd2Zr2−xTixO7
(x = 0, 0.5, 1, 1.5, 2)
58Ni, 101Ru, 129Xe, 181Ta, 197Au, 208Pb, 238U 11.1MeV/u RT5 × 109–2 × 1014 ions/cm2
1 × 108 ions/cm2/s
[135]
Gd2Ti2O7, La2Ti2O72.0 GeV Ta (RT)1 × 1010–1 × 1013 ions/cm2 [204]
A2Ti2O7
A = Gd, Y
92 MeV Xe (RT)
109 MeV U (RT)
2 × 1011–2 × 1016 ions/cm2
5 × 1012 ions/cm2
5, 10 × 1012 ions/cm2 for A2Ti2O7[63]
Lu2Ti2O7600 keV Kr
(RT, 723 K)
4 × 1015 ions/cm2
1 × 1012 ions/cm2/s
[241]
Lu2(Ti2−xLux)O7−x/2
x = 0–0.667
2.7 MeV Ar2 × 1014–8 × 1014 ions/cm2
[171]
Gd2Zr2O712 MeV C60
2.6 GeV U
1 × 1011 ions/cm2
5 × 1011 ions/cm2
[242]
Gd2(TixZr1−x)2O7
(x = 0, 1, 2)
30 MeV C60
1.5 GeV U
5 × 1010–1 × 1011 ions/cm2 [177]
Lu2Ti2O7200 keV He2 × 1015–2 × 1017 ions/cm2 [166]
Gd2Zr2O7120MeV Au RT3 × 1011–1 × 1014 ions/cm2 [69]
Gd2Zr2O7320 MeV Ar2 × 1014–1 × 1016 ions/cm2 [243]
A2Ti2O7
(A = Yb, Er, Y, Gd, Sm)
2.2GeV Au5 × 1010–3 × 1013 ions/cm2 [244]
Er2Sn2O7
Dy2Sn2O7
Dy2Ti2O7
2.2 GeV 197Au RT (2.1GeV, after Al foil)8 × 1012 ions/cm2 [41,42,44,46]
Er2Ti2O72.2 GeV Au RT5 × 1012 ions/cm2 [45]
Y2Ti2O793 MeV Xe2 × 1011–1 × 1013 ions/cm2 [188]
Gd2Zr2O77 MeV Au +
200 keV He
or 65 keV He
2.2 × 1015
2 × 1015–2× 1016
2 × 1017 ions/cm2
[167]
Nano-A2Ti2O7
(A = Gd, Ho, Lu)
1 MeV Kr
(RT–600 K)
6.25 × 1011 ions/cm2/s0.36, 0.53, 0.81 dpa (RT), respectively[93]
Nd2Zr2O70.5 MeV He1 × 1014–1 × 1017 ions/cm2 [245]
(Gd1−4xU2x)2(Zr1−xUx)2O7
(x = 0, 0.1, 0.14)
1.5 MeV Xe1 × 1012–1 × 1015 ions/cm2
5 × 1011 ions/cm2/s
[246,247]
Lu2−xCexTi2O7
x = 0–0.7
400 keV Ne1 × 1014–2 × 1015 ions/cm2 [248]
Gd1.891244Cm0.091Pu0.013Sn2O7Self-irradiation 1.52 dpa (RT)[249]
La2Zr2O71 MeV Zr3.98 × 1013–7.96 × 1014 ions/cm2 [250]
Gd2Zr2O7+xTRPO
x = 0, 25, 40wt%
0.5 MeV He1 × 1014–1 × 1017 ions/cm2/s [168]
Lu2Ti2O70.4 MeV Ne (77 K)
2.7 MeV Ar,
6.5 MeV Xe (RT)
1 × 1012 ions/cm2/s [57]
(Gd1−xNdx)2(Zr1−yCey)2O7
(0 ≤ x, y ≤ 1)
0.5MeV He1 × 1015–1 × 1017 ions/cm2/s [169,251]
Gd2Zr2O7, Nd2Zr2O7,
Nd2Ce2O7
2.0 MeV Xe4.08 × 1013–5.28 × 1016 ions/cm2/s [252,253]
Nd2Zr2O7, Sm2Zr2O7, Er2Zr2O72.2 GeV Au5 × 1010–3 × 1013 ions/cm2 [254]
Gd2Ti2O7,Sm2Ti2O7, Yb2Ti2O7, Gd2ZrTiO755 MeV I5 × 1010 ions/cm2 [255]
Nd2Zr2O7
Nd1.9U0.1Zr2O7
2 MeV Kr6 × 1015, 2 × 1016 ions/cm2 [256]
Gd2Ti2O7, Gd2Zr2O7400 keV Kr5.36 × 1016 ions/cm2 [58]
Gd2Zr2O7190 keV He5 × 1017 ions/cm2 [170]
Nd2Zr2O7100 MeV Ag
(RT, 1000 K)
3 × 1011–2 × 1013 ions/cm2 [257]
Gd2Zr2O7 +35wt%TRPO0.5 MeV He1 × 1015–1 × 1017 ions/cm2 [258]
Single-crystal Gd2Ti2O7190 keV He1 × 1017 ions/cm2/sRe-crystallizaton[105]
Gd2Zr2O75 MeV Xe5 × 1014–4 × 1015 ions/cm2 [94]
Y2−xAxTi2Oδ (A = U, Ce; x = 0.05, 0.075, 0.10, 0.20, 0.30)60Co
(γ-source.)
1000 kGy
5.8 kGy/h
[95]
La2Zr2O7500 keV Kr
1 MeV Xe
(88 K or 300 K)
1 × 1013–1 × 1014 ions/cm2 [259,260]
Gd2(Ti2-yZry)O7,
(y = 0, 0.4, 0.8, 1.2, 1.6, 2)
120MeV Au RT1 × 1013–1 × 1014 ions/cm2 [64]
Single crystal (Yb0.2Tm0.2Lu0.2Ho0.2Er0.2)2Ti2O74 MeV Au RT
(4° off normal surface)
5 × 1012–6 × 1014 ions/cm2
4.2 × 1011 ions/cm2/s
0.13 dpa[107]
(La0.2Ce0.2Nd0.2Sm0.2Gd0.2)2Zr2O79 MeV Au RT4.5 × 1015 ions/cm2/s [77]
(Sm0.2Eu0.2Gd0.2Y0.2Lu0.2)2Ti2O7800 keV Kr RT 0.27 dpa[230]
Note, NA, PA and FA means no amorphization, partial amorphization and fully amorphization, respectively.

References

  1. Ojovan, M.I.; Lee, W.E. An Introduction to Nuclear Waste Immobilisation, 2nd ed.; Elsevier: Oxford, UK, 2014. [Google Scholar]
  2. Feiveson, H.; Mian, Z.; Ramana, M.V.; von Hippel, F. (Eds.) Managing Spent Fuel from Nuclear Power Reactors: Experience and Lessons from Around the World; International Panel on Fissile Material: Bedfordshire, UK, 2011. [Google Scholar]
  3. Yudintsev, S.V. Isolation of Separated Waste of Nuclear Industry. Radiochemistry 2021, 63, 527–555. [Google Scholar] [CrossRef]
  4. Ewing, R.C. Long-term storage of spent nuclear fuel. Nat. Mater. 2015, 14, 252–257. [Google Scholar] [CrossRef] [PubMed]
  5. Ewing, R.C.; Weber, W.J.; Lian, J. Nuclear waste disposal—Pyrochlore (A2B2O7): Nuclear waste form for the immobilization of plutonium and “minor” actinides. J. Appl. Phys. 2004, 95, 5949–5971. [Google Scholar] [CrossRef] [Green Version]
  6. McMaster, S.A.; Ram, R.; Faris, N.; Pownceby, M.I. Radionuclide disposal using the pyrochlore supergroup of minerals as a host matrix-A review. J. Hazard. Mater. 2018, 360, 257–269. [Google Scholar] [CrossRef] [PubMed]
  7. Ewing, R.C.; Weber, W.J. Actinide Waste Forms and Radiation Effects. In The Chemistry of the Actinide and Transactinide Elements; Springer: Dordrecht, The Netherlands, 2010; Volume 1. [Google Scholar]
  8. Weber, W.J.; Ewing, R.C.; Angell, C.A.; Arnold, G.W.; Cormack, A.N.; Delaye, J.M.; Griscom, D.L.; Hobbs, L.W.; Navrotsky, A.; Price, D.L.; et al. Radiation Effects in Glasses Used for Immobilization of High-level Waste and Plutonium Disposition. J. Mater. Res. 1997, 12, 1948–1978. [Google Scholar] [CrossRef] [Green Version]
  9. Donald, I.W.; Metcalfe, B.L.; Taylor, R.N.J. The immobilization of high level radioactive wastes using ceramics and glasses. J. Mater. Sci. 1997, 32, 5851–5887. [Google Scholar] [CrossRef]
  10. Lopez, C.; Deschanels, X.; Bart, J.M.; Boubals, J.M.; Den Auwer, C.; Simoni, E. Solubility of actinide surrogates in nuclear glasses. J. Nucl. Mater. 2003, 312, 76–80. [Google Scholar] [CrossRef]
  11. Donald, I.W. Waste Immobilization in Glass and Ceramic Based Hosts; John Wiley & Sons Ltd.: Chichester, UK, 2010. [Google Scholar]
  12. Gin, S.; Abdelouas, A.; Criscenti, L.J.; Ebert, W.L.; Ferrand, K.; Geisler, T.; Harrison, M.T.; Inagaki, Y.; Mitsui, S.; Mueller, K.T.; et al. An international initiative on long-term behavior of high-level nuclear waste glass. Mater. Today 2013, 16, 243–248. [Google Scholar] [CrossRef]
  13. Cassingham, N.; Corkhill, C.L.; Backhouse, D.J.; Hand, R.J.; Ryan, J.V.; Vienna, J.D.; Hyatt, N.C. The initial dissolution rates of simulated UK Magnox–ThORP blend nuclear waste glass as a function of pH, temperature and waste loading. Mineral. Mag. 2018, 79, 1529–1542. [Google Scholar] [CrossRef] [Green Version]
  14. Bohre, A.; Avasthi, K.; Pet’kov, V.I. Vitreous and crystalline phosphate high level waste matrices: Present status and future challenges. J. Ind. Eng. Chem. 2017, 50, 1–14. [Google Scholar] [CrossRef]
  15. Kong, L.; Karatchevtseva, I.; Zhang, Y. A new method for production of glass-Ln2Ti2O7 pyrochlore (Ln = Gd, Tb, Er, Yb). J. Eur. Ceram. Soc. 2017, 37, 4963–4972. [Google Scholar] [CrossRef]
  16. Bhuiyan, A.; Wong, V.; Abraham, J.L.; Aughterson, R.D.; Kong, L.; Farzana, R.; Gregg, D.J.; Sorrell, C.C.; Zhang, Y.; Koshy, P. Phase assemblage and microstructures of Gd2Ti2-xZrxO7 (x = 0.1–0.3) pyrochlore glass-ceramics as potential waste forms for actinide immobilization. Mater. Chem. Phys. 2021, 273, 125058. [Google Scholar] [CrossRef]
  17. Zhang, Y.; Kong, L.; Ionescu, M.; Gregg, D.J. Current advances on titanate glass-ceramic composite materials as waste forms for actinide immobilization: A technical review. J. Eur. Ceram. Soc. 2022, 42, 1852–1876. [Google Scholar] [CrossRef]
  18. Ringwood, A.E.; Kesson, S.E.; Ware, N.G.; Hibberson, W.; Major, A. Immobilisation of high level nuclear reactor wastes in SYNROC. Nature 1979, 278, 219–223. [Google Scholar] [CrossRef]
  19. Lutze, W.; Ewing, R.C. Radioactive Waste Forms for the Future; OSTI.GOV: Oak Ridge, TN, USA, 1988.
  20. Begg, B.D.; Vance, E.R.; Conradson, S.D. The incorporation of plutonium and neptunium in zirconolite and perovskite. J. Alloys Compd. 1998, 271–273, 221–226. [Google Scholar] [CrossRef]
  21. Wang, L.; Liang, T. Ceramics for high level radioactive waste solidification. J. Adv. Ceram. 2012, 1, 194–203. [Google Scholar] [CrossRef] [Green Version]
  22. Gregg, D.J.; Farzana, R.; Dayal, P.; Holmes, R.; Triani, G. Synroc technology: Perspectives and current status (Review). J. Am. Ceram. Soc. 2020, 103, 5424–5441. [Google Scholar] [CrossRef]
  23. Jostsons, A.; Vance, E.R.; Mercer, D.J.; Oversby, V.M. Synroc for Immobilising Excess Weapons Plutonium. In Proceedings of the Symposium—Scientific Basis for Nuclear Waste Management XVIII, Kyoto, Japan, 23–27 October 1994; Volume 353. [Google Scholar] [CrossRef]
  24. Vance, E.R.; Jostsons, A.; Day, R.A.; Ball, C.J.; Begg, B.D.; Angel, P.J. Excess Pu Disposition in Zirconolite-Rich synroc. In Proceedings of the Symposium V—Scientific Basis for Nuclear Waste Management XIX, Boston, MA, USA, 27 November–1 December 1995; Volume 412. [Google Scholar] [CrossRef]
  25. Weber, W.J.; Wald, J.W.; Matzke, H. Self-radiation damage in Gd2Ti2O7. Mater. Lett. 1985, 3, 173–180. [Google Scholar] [CrossRef]
  26. Lumpkin, G.; Ewing, R. Alpha-decay damage in minerals of the pyrochlore group. Phys. Chem. Miner. 1988, 16, 2–20. [Google Scholar] [CrossRef]
  27. Vance, E.R.; Begg, B.D.; Day, R.A.; Ball, C.J. Zirconolite-Rich Ceramics for Actinide Wastes. In Proceedings of the Symposium—Scientific Basis for Nuclear Waste Management XVIII, Kyoto, Japan, 23–27 October 1994; p. 353. [Google Scholar] [CrossRef]
  28. Gong, W.L.; Lutze, W.; Ewing, R.C. Zirconia ceramics for excess weapons plutonium waste. J. Nucl. Mater. 2000, 277, 239–249. [Google Scholar] [CrossRef]
  29. Weber, W.J.; Ewing, R.C.; Catlow, C.R.A.; de la Rubia, T.D.; Hobbs, L.W.; Kinoshita, C.; Matzke, H.; Motta, A.T.; Nastasi, M.; Salje, E.K.H.; et al. Radiation effects in crystalline ceramics for the immobilization of high-level nuclear waste and plutonium. J. Mater. Res. 1998, 13, 1434–1484. [Google Scholar] [CrossRef] [Green Version]
  30. Lumpkin, G.R. Alpha-decay damage and aqueous durability of actinide host phases in natural systems. J. Nucl. Mater. 2001, 289, 136–166. [Google Scholar] [CrossRef]
  31. Ewing, R.C.; Weber, W.J.; Clinard, F.W. Radiation effects in nuclear waste forms for high-level radioactive waste. Prog. Nucl. Energy 1995, 29, 63–127. [Google Scholar] [CrossRef]
  32. Clinard, F.W.; Hobbs, L.W.; Land, C.C.; Peterson, D.E.; Rohr, D.L.; Roof, R.B. Alpha decay self-irradiation damage in 238Pu-substituted zirconolite. J. Nucl. Mater. 1982, 105, 248–256. [Google Scholar] [CrossRef]
  33. Zhang, Y.-B.; Wang, J.; Wang, J.-X.; Huang, Y.; Luo, P.; Liang, X.-F.; Tan, H.-B. Phase evolution, microstructure and chemical stability of Ca1-xZr1-xGd2xTi2O7 (0.0 ≤ x ≤ 1.0) system for immobilizing nuclear waste. Ceram. Int. 2018, 44, 13572–13579. [Google Scholar] [CrossRef]
  34. Chakoumakos, B.C. Systematics of the pyrochlore structure type, ideal A2B2X6Y. J. Solid State Chem. 1984, 53, 120–129. [Google Scholar] [CrossRef]
  35. Subramanian, M.A.; Aravamudan, G.; Subba Rao, G.V. Oxide pyrochlores—A review. Prog. Solid State Chem. 1983, 15, 55–143. [Google Scholar] [CrossRef]
  36. Subramanian, M.A.; Marshall, W.J.; Hoffmann, R.D.; Sleight, A.W. Synthesis and Structure of Some MII/MIII Mixed Fluorides with Pyrochlore and Weberite Related Structures. Z. Für Nat. B 2006, 61, 808–812. [Google Scholar] [CrossRef]
  37. Jiang, C.; Stanek, C.R.; Sickafus, K.E.; Uberuaga, B.P. First-principles prediction of disordering tendencies in pyrochlore oxides. Phys. Rev. B 2009, 79, 104203. [Google Scholar] [CrossRef]
  38. Minervini, L.; Grimes, R.W.; Sickafus, K.E. Disorder in Pyrochlore Oxides. J. Am. Ceram. Soc. 2000, 83, 1873–1878. [Google Scholar] [CrossRef]
  39. Sickafus, K.E.; Minervini, L.; Grimes, R.W.; Valdez, J.A.; Ishimaru, M.; Li, F.; McClellan, K.J.; Hartmann, T. Radiation tolerance of complex oxides. Science 2000, 289, 748–751. [Google Scholar] [CrossRef]
  40. Wang, Y.; Ouyang, J.; Wei, T.; Cao, G.; Liu, Z.; Ding, Z.; Wang, Y.; Wang, Y. Mechanical properties, thermal conductivity and defect formation energies of samarium immobilization in Gd2Zr2O7: First-principles study and irradiation experiment. J. Rare Earths 2022, in press. [Google Scholar] [CrossRef]
  41. Chung, C.-K.; Lang, M.; Xu, H.; Navrotsky, A. Thermodynamics of radiation induced amorphization and thermal annealing of Dy2Sn2O7 pyrochlore. Acta Mater. 2018, 155, 386–392. [Google Scholar] [CrossRef]
  42. Chung, C.-K.; Shamblin, J.; O’Quinn, E.C.; Shelyug, A.; Gussev, I.; Lang, M.; Navrotsky, A. Thermodynamic and structural evolution of Dy2Ti2O7 pyrochlore after swift heavy ion irradiation. Acta Mater. 2018, 145, 227–234. [Google Scholar] [CrossRef]
  43. Lang, M.; O’Quinn, E.C.; Shamblin, J.; Neuefeind, J. Advanced Experimental Technique for Radiation Damage Effects in Nuclear Waste Forms: Neutron Total Scattering Analysis. MRS Adv. 2018, 3, 1735–1747. [Google Scholar] [CrossRef]
  44. Shamblin, J.; Tracy, C.L.; Palomares, R.I.; O’Quinn, E.C.; Ewing, R.C.; Neuefeind, J.; Feygenson, M.; Behrens, J.; Trautmann, C.; Lang, M. Similar local order in disordered fluorite and aperiodic pyrochlore structures. Acta Mater. 2018, 144, 60–67. [Google Scholar] [CrossRef]
  45. Chung, C.-K.; O’Quinn, E.C.; Neuefeind, J.C.; Fuentes, A.F.; Xu, H.; Lang, M.; Navrotsky, A. Thermodynamic and structural evolution of mechanically milled and swift heavy ion irradiated Er2Ti2O7 pyrochlore. Acta Mater. 2019, 181, 309–317. [Google Scholar] [CrossRef]
  46. Shamblin, J.; Feygenson, M.; Neuefeind, J.; Tracy, C.L.; Zhang, F.; Finkeldei, S.; Bosbach, D.; Zhou, H.; Ewing, R.C.; Lang, M. Probing disorder in isometric pyrochlore and related complex oxides. Nat Mater 2016, 15, 507–511. [Google Scholar] [CrossRef]
  47. Minervini, L.; Grimes, R.W.; Tabira, Y.; Withers, R.L.; Sickafus, K.E. The oxygen positional parameter in pyrochlores and its dependence on disorder. Philos. Mag. A 2002, 82, 123–135. [Google Scholar] [CrossRef]
  48. Stanek, C.R.; Minervini, L.; Grimes, R.W. Nonstoichiometry in A2B2O7 Pyrochlores. J. Am. Ceram. Soc. 2002, 85, 2792–2798. [Google Scholar] [CrossRef]
  49. Livshits, T.S.; Zhang, J.; Yudintsev, S.V.; Stefanovsky, S.V. New titanate matrices for immobilization of REE–actinide high-level waste. J. Radioanal. Nucl. Chem. 2014, 304, 47–52. [Google Scholar] [CrossRef]
  50. Guo, Y.; Feng, S.; Yang, Y.; Zheng, R.; Zhang, Y.; Fu, J.; Wang, H.; Li, J. High-entropy titanate pyrochlore as newly low-thermal conductivity ceramics. J. Eur. Ceram. Soc. 2022, 42, 6614–6623. [Google Scholar] [CrossRef]
  51. Zhang, K.; Wen, G.; Jia, Z.; Teng, Y.; Zhang, H. Self-Propagating High-Temperature Synthesis of Zirconolite Using CuO and MoO3 as the Oxidants. Int. J. Appl. Ceram. Technol. 2015, 12, E111–E120. [Google Scholar] [CrossRef]
  52. Zhang, K.; Wen, G.; Zhang, H.; Teng, Y. Self-propagating high-temperature synthesis of Y2Ti2O7 pyrochlore and its aqueous durability. J. Nucl. Mater. 2015, 465, 1–5. [Google Scholar] [CrossRef]
  53. Zhang, K.; He, Z.; Peng, L.; Zhang, H.; Lu, X. Self-propagating synthesis of Y2−xNdxTi2O7 pyrochlore and its aqueous durability as nuclear waste form. Scr. Mater. 2018, 146, 300–303. [Google Scholar] [CrossRef]
  54. Sattonnay, G.; Moll, S.; Thomé, L.; Legros, C.; Herbst-Ghysel, M.; Garrido, F.; Costantini, J.M.; Trautmann, C. Heavy-ion irradiation of pyrochlore oxides: Comparison between low and high energy regimes. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2008, 266, 3043–3047. [Google Scholar] [CrossRef]
  55. Lumpkin, G.R.; Smith, K.L.; Blackford, M.G.; Whittle, K.R.; Harvey, E.J.; Redfern, S.A.T.; Zaluzec, N.J. Ion Irradiation of Ternary Pyrochlore Oxides. Chem. Mater. 2009, 21, 2746–2754. [Google Scholar] [CrossRef]
  56. Li, Y.H.; Wang, Y.Q.; Zhou, M.; Xu, C.P.; Valdez, J.A.; Sickafus, K.E. Light ion irradiation effects on stuffed Lu2(Ti2−xLux)O7−x/2 (x=0, 0.4 and 0.67) structures. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2011, 269, 2001–2005. [Google Scholar] [CrossRef]
  57. Yang, D.; Xia, Y.; Wen, J.; Liang, J.; Mu, P.; Wang, Z.; Li, Y.; Wang, Y. Role of ion species in radiation effects of Lu2Ti2O7 pyrochlore. J. Alloys Compd. 2017, 693, 565–572. [Google Scholar] [CrossRef] [Green Version]
  58. Holesinger, T.G.; Valdez, J.A.; Janish, M.T.; Wang, Y.; Uberuaga, B.P. Potential benefit of amorphization in the retention of gaseous species in irradiated pyrochlores. Acta Mater. 2019, 164, 250–260. [Google Scholar] [CrossRef]
  59. Mandal, B.P.; Krishna, P.S.R.; Tyagi, A.K. Order–disorder transition in the Nd2−yYyZr2O7 system: Probed by X-ray diffraction and Raman spectroscopy. J. Solid State Chem. 2010, 183, 41–45. [Google Scholar] [CrossRef]
  60. Moll, S.; Sattonnay, G.; Thomé, L.; Jagielski, J.; Legros, C.; Monnet, I. Swift heavy ion irradiation of pyrochlore oxides: Electronic energy loss threshold for latent track formation. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2010, 268, 2933–2936. [Google Scholar] [CrossRef]
  61. Sattonnay, G.; Moll, S.; Thomé, L.; Legros, C.; Calvo, A.; Herbst-Ghysel, M.; Decorse, C.; Monnet, I. Effect of composition on the behavior of pyrochlores irradiated with swift heavy ions. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2012, 272, 261–265. [Google Scholar] [CrossRef]
  62. Sattonnay, G.; Sellami, N.; Thomé, L.; Legros, C.; Grygiel, C.; Monnet, I.; Jagielski, J.; Jozwik-Biala, I.; Simon, P. Structural stability of Nd2Zr2O7 pyrochlore ion-irradiated in a broad energy range. Acta Mater. 2013, 61, 6492–6505. [Google Scholar] [CrossRef]
  63. Sellami, N.; Sattonnay, G.; Grygiel, C.; Monnet, I.; Debelle, A.; Legros, C.; Menut, D.; Miro, S.; Simon, P.; Bechade, J.L.; et al. Modifications of structural and physical properties induced by swift heavy ions in Gd2Ti2O7 and Y2Ti2O7 pyrochlores. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2015, 365, 371–375. [Google Scholar] [CrossRef]
  64. Kumar, A.; Sharma, S.K.; Grover, V.; Singh, Y.; Kumar, V.; Shukla, V.K.; Kulriya, P.K. Probing the Short-Range Ordering of Ion Irradiated Gd2Ti2-yZryO7 (0.0 ≤ y ≤ 2.0) Pyrochlore under Electronic Stopping Regime. J. Nucl. Mater. 2022, 564, 153682. [Google Scholar] [CrossRef]
  65. Lian, J.; Wang, L.M.; Haire, R.G.; Helean, K.B.; Ewing, R.C. Ion beam irradiation in La2Zr2O7–Ce2Zr2O7 pyrochlore. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2004, 218, 236–243. [Google Scholar] [CrossRef]
  66. Lian, J.; Yudintsev, S.V.; Stefanovsky, S.V.; Wang, L.M.; Ewing, R.C. Ion beam irradiation of U-, Th- and Ce-doped pyrochlores. J. Alloys Compd. 2007, 444–445, 429–433. [Google Scholar] [CrossRef]
  67. Patel, M.K.; Vijayakumar, V.; Avasthi, D.K.; Kailas, S.; Pivin, J.C.; Grover, V.; Mandal, B.P.; Tyagi, A.K. Effect of swift heavy ion irradiation in pyrochlores. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2008, 266, 2898–2901. [Google Scholar] [CrossRef]
  68. Whittle, K.R.; Blackford, M.G.; Aughterson, R.D.; Lumpkin, G.R.; Zaluzec, N.J. Ion irradiation of novel yttrium/ytterbium-based pyrochlores: The effect of disorder. Acta Mater. 2011, 59, 7530–7537. [Google Scholar] [CrossRef]
  69. Kumari, R.; Kulriya, P.K.; Grover, V.; Shukla, R.; Saravanan, K.; Mohapatra, S.; Tyagi, A.K.; Avasthi, D.K. Radiation stability of Gd2Zr2O7: Effect of stoichiometry and structure. Ceram. Int. 2016, 42, 103–109. [Google Scholar] [CrossRef]
  70. Chen, S.; Liu, X.; Shu, X.; Yang, J.; Wang, L.; Zhang, K.; Wang, X.; Zhang, H.; Lu, X. Rapid synthesis and chemical durability of Gd2Zr2-xCexO7 via SPS for nuclear waste forms. Ceram. Int. 2018, 44, 20306–20310. [Google Scholar] [CrossRef]
  71. Li, Y.H.; Wang, Y.Q.; Xu, C.P.; Valdez, J.A.; Tang, M.; Sickafus, K.E. Microstructural evolution of the pyrochlore compound Er2Ti2O7 induced by light ion irradiations. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2012, 286, 218–222. [Google Scholar] [CrossRef]
  72. Li, Y.H.; Wen, J.; Wang, Y.Q.; Wang, Z.G.; Tang, M.; Valdez, J.A.; Sickafus, K.E. The irradiation effects of Gd2Hf2O7 and Gd2Ti2O7. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2012, 287, 130–134. [Google Scholar] [CrossRef]
  73. Brykała, U.; Diduszko, R.; Jach, K.; Jagielski, J. Hot pressing of gadolinium zirconate pyrochlore. Ceram. Int. 2015, 41, 2015–2021. [Google Scholar] [CrossRef]
  74. Teng, Z.; Zhu, L.; Tan, Y.; Zeng, S.; Xia, Y.; Wang, Y.; Zhang, H. Synthesis and structures of high-entropy pyrochlore oxides. J. Eur. Ceram. Soc. 2020, 40, 1639–1643. [Google Scholar] [CrossRef]
  75. Wang, Y.; Jin, Y.-J.; Ding, Z.-Y.; Cao, G.; Liu, Z.-G.; Wei, T.; Ouyang, J.-H.; Wang, Y.-M.; Wang, Y.-J. Microstructure and electrical properties of new high-entropy rare-earth zirconates. J. Alloys Compd. 2022, 906, 164331. [Google Scholar] [CrossRef]
  76. Wang, Y.; Jin, Y.-J.; Wei, T.; Wang, Z.-G.; Cao, G.; Ding, Z.-Y.; Liu, Z.-G.; Ouyang, J.-H.; Wang, Y.-J.; Wang, Y.-M. Size disorder: A descriptor for predicting the single- or dual-phase formation in multi-component rare earth zirconates. J. Alloys Compd. 2022, 918, 165636. [Google Scholar] [CrossRef]
  77. Xu, L.; Niu, M.; Wang, H.; Su, L.; Gao, H.; Zhuang, L. Response of structure and mechanical properties of high entropy pyrochlore to heavy ion irradiation. J. Eur. Ceram. Soc. 2022, 42, 6624–6632. [Google Scholar] [CrossRef]
  78. Han, W.; Ye, Y.; Lu, K.; Wu, Y.; Wang, H.; Huang, Z.; Qi, J.; Lu, T. High-entropy transparent (Y0.2La0.2Gd0.2Yb0.2Dy0.2)2Zr2O7 ceramics as novel phosphor materials with multi-wavelength excitation and emission properties. J. Eur. Ceram. Soc. 2022, 43, 143–149. [Google Scholar] [CrossRef]
  79. Li, S.; Ma, K.; Ye, Y.; Han, W.; Li, S.; Shi, Y.; Tang, Y.; Cheng, G.; Zhu, X.; Qi, J. Effect of annealing temperature on optical properties of Erbium doped Y2Zr2O7 transparent ceramics. Ceram. Int. 2022, 48, 37180–37185. [Google Scholar] [CrossRef]
  80. Li, Y.; Ye, Y.; Quan, W.; Qin, Y.; Han, W.; Chen, Y.; Liao, W.; Wu, Y.; Wang, J.; Qi, J. Investigation of up/down-converted luminescence and related mechanisms in Ho3+/Yb3+ co-doped Y2Zr2O7 highly transparent ceramics as phosphor material. J. Lumin. 2022, 251, 119255. [Google Scholar] [CrossRef]
  81. Lu, K.; Ye, Y.; Chen, R.; Li, Y.; Qi, J.; Lu, T. Simultaneous coloration erasure and significant photoluminescence enhancement via vacuum re-annealing in lanthanide-doped transparent ceramics. Scr. Mater. 2022, 215, 114695. [Google Scholar] [CrossRef]
  82. Wang, J.; Lu, K.; Ye, Y.; Qin, Y.; Han, W.; Qi, J. Samarium doped yttrium zirconate transparent ceramics for highly transparent reddish-orange light emitting applications. Ceram. Int. 2022, 48, 35050–35055. [Google Scholar] [CrossRef]
  83. Liu, Z.-G.; Ouyang, J.-H.; Zhou, Y. Heat capacities and derived thermodynamic functions of neodymium–gadolinium zirconates from 298.15 to 1050 K. J. Alloys Compd. 2009, 475, 21–24. [Google Scholar] [CrossRef]
  84. Liu, Z.-G.; Ouyang, J.-H.; Wang, B.-H.; Zhou, Y.; Li, J. Thermal expansion and thermal conductivity of SmxZr1−xO2−x/2 (0.1≤x≤0.5) ceramics. Ceram. Int. 2009, 35, 791–796. [Google Scholar] [CrossRef]
  85. Liu, Z.-G.; Ouyang, J.-H.; Zhou, Y. Structural evolution and thermophysical properties of (SmxGd1−x)2Zr2O7 (0 ≤ x ≤ 1.0) ceramics. J. Alloys Compd. 2009, 472, 319–324. [Google Scholar] [CrossRef]
  86. Liu, Z.-G.; Ouyang, J.-H.; Zhou, Y.; Li, J.; Xia, X.-L. Densification, Structure, and Thermophysical Properties of Ytterbium-Gadolinium Zirconate Ceramics. Int. J. Appl. Ceram. Technol. 2009, 6, 485–491. [Google Scholar] [CrossRef]
  87. Liu, Z.-G.; Ouyang, J.-H.; Zhou, Y. Preparation and thermophysical properties of (NdxGd1–x)2Zr2O7 ceramics. J. Mater. Sci. 2008, 43, 3596–3603. [Google Scholar] [CrossRef]
  88. Liu, Z.-G.; Ouyang, J.-H.; Zhou, Y.; Li, J.; Xia, X.-L. Influence of ytterbium- and samarium-oxides codoping on structure and thermal conductivity of zirconate ceramics. J. Eur. Ceram. Soc. 2009, 29, 647–652. [Google Scholar] [CrossRef]
  89. Liu, Z.-G.; Ouyang, J.-H.; Zhou, Y.; Xia, X.-L. Electrical conductivity of samarium–ytterbium zirconate ceramics. Electrochim. Acta 2009, 54, 3968–3971. [Google Scholar] [CrossRef]
  90. Liu, Z.-G.; Ouyang, J.-H.; Zhou, Y.; Xia, X.-L. Preparation and microstructural characterization of (Nd1−xGdx)2(Ce1−xZrx)2O7 solid solutions. Ceram. Int. 2009, 35, 2387–2392. [Google Scholar] [CrossRef]
  91. Lian, J.; Zu, X.T.; Kutty, K.V.G.; Chen, J.; Wang, L.M.; Ewing, R.C. Ion-irradiation-induced amorphization of La2Zr2O7 pyrochlore. Physical Review B 2002, 66. [Google Scholar] [CrossRef]
  92. Wang, X.; Jiang, K.; Zhou, L. Characterization and phase stability of pyrochlore (Nd1−xCex)2Zr2O7+y (x= 0–1). J. Nucl. Mater. 2015, 458, 156–161. [Google Scholar] [CrossRef]
  93. Wen, J.; Sun, C.; Dholabhai, P.P.; Xia, Y.; Tang, M.; Chen, D.; Yang, D.Y.; Li, Y.H.; Uberuaga, B.P.; Wang, Y.Q. Temperature dependence of the radiation tolerance of nanocrystalline pyrochlores A2Ti2O7 (A = Gd, Ho and Lu). Acta Mater. 2016, 110, 175–184. [Google Scholar] [CrossRef] [Green Version]
  94. Liu, K.; Zhang, K.; Deng, T.; Luo, B.; Zhang, H. Heavy-ion irradiation effects of Gd2Zr2O7 nanocrystalline ceramics as nuclear waste immobilization matrix. J. Nucl. Mater. 2020, 538, 152236. [Google Scholar] [CrossRef]
  95. Gumber, N.; Pai, R.V.; Phatak, R.; Adiraju, B.; Sahu, M.; Jagannath, J.; Sudarshan, K. Synthesis, Characterization and Crystal Chemistry of Uranium and Cerium Doped Yttrium Titanate Pyrochlore: A Potential Waste Immobilization Matrix. J. Nucl. Mater. 2021, 556, 153191. [Google Scholar] [CrossRef]
  96. Li, B.-r.; Gao, H.; Liu, J.-p.; Yang, Z.-w. Synthesis of Lu2Ti2O7 nano-powders by salt modified sol–gel process. Ceram. Int. 2012, 38, 5453–5459. [Google Scholar] [CrossRef]
  97. Mao, Y.; Li, G.; Sun, Y.; Feng, S. Hydrothermal Synthesis and Characterization of Bi2Pb2O7 with Pyrochlore Structure. J. Solid State Chem. 2000, 149, 314–319. [Google Scholar] [CrossRef]
  98. Zeng, J.; Wang, H.; Zhang, Y.; Zhu, M.K.; Yan, H. Hydrothermal Synthesis and Photocatalytic Properties of Pyrochlore La2Sn2O7 Nanocubes. J. Phys. Chem. C 2007, 111, 11879–11887. [Google Scholar] [CrossRef]
  99. Trujillano, R.; Martín, J.A.; Rives, V. Hydrothermal synthesis of Sm2Sn2O7 pyrochlore accelerated by microwave irradiation. A comparison with the solid state synthesis method. Ceram. Int. 2016, 42, 15950–15954. [Google Scholar] [CrossRef]
  100. Wang, Q.; Cheng, X.; Li, J.; Jin, H. Hydrothermal synthesis and photocatalytic properties of pyrochlore Sm2Zr2O7 nanoparticles. J. Photochem. Photobiol. A Chem. 2016, 321, 48–54. [Google Scholar] [CrossRef]
  101. Palpandi, K.; Raman, N. Hydrothermally synthesized a pyrochlore-type bismuth stannate (Bi2Sn2O7): Efficient electrochemical detection of nitroxoline. J. Electroanal. Chem. 2021, 899, 115679. [Google Scholar] [CrossRef]
  102. Lian, J.; Wang, L.; Chen, J.; Sun, K.; Ewing, R.C.; Matt Farmer, J.; Boatner, L.A. The order–disorder transition in ion-irradiated pyrochlore. Acta Mater. 2003, 51, 1493–1502. [Google Scholar] [CrossRef]
  103. Lian, J.; Wang, L.M.; Ewing, R.C.; Boatner, L.A. Ion beam implantation and cross-sectional TEM studies of lanthanide titanate pyrochlore single crystals. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2005, 241, 365–371. [Google Scholar] [CrossRef]
  104. Zhang, Y.; Jagielski, J.; Bae, I.-T.; Xiang, X.; Thomé, L.; Balakrishnan, G.; Paul, D.M.; Weber, W.J. Damage evolution in Au-implanted Ho2Ti2O7 titanate pyrochlore. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2010, 268, 3009–3013. [Google Scholar] [CrossRef]
  105. Janish, M.T.; Schneider, M.M.; Valdez, J.A.; McClellan, K.J.; Byler, D.D.; Wang, Y.; Chen, D.; Holesinger, T.G.; Uberuaga, B.P. In-situ re-crystallization of heavily-irradiated Gd2Ti2O7. Acta Mater. 2020, 194, 403–411. [Google Scholar] [CrossRef]
  106. Moll, S.; Sattonnay, G.; Thomé, L.; Jagielski, J.; Decorse, C.; Simon, P.; Monnet, I.; Weber, W.J. Irradiation damage in Gd2Ti2O7 single crystals: Ballistic versus ionization processes. Phys. Rev. B 2011, 84, 064115. [Google Scholar] [CrossRef]
  107. Kinsler-Fedon, C.; Nuckols, L.; Nelson, C.T.; Qi, Z.H.; Huang, Q.; Mandrus, D.; Zhang, Y.W.; Weber, W.J.; Keppens, V. Effects of Au2+ irradiation induced damage in a high-entropy pyrochlore oxide single crystal. Scr. Mater. 2022, 220, 114916. [Google Scholar] [CrossRef]
  108. Li, B.; Gao, H.; Guo, Y.; Hou, S. Synthesis and thermal properties of Lu2Ti2O7 nano-wires. CrystEngComm 2012, 14, 4168–4172. [Google Scholar] [CrossRef]
  109. Li, B.-r.; Chang, H.-b.; Gao, H.; Hou, S.-x. Synthesis of Lu2Ti2O7 nano-rods from molten salt with two-step calcinations. Mater. Lett. 2012, 79, 219–221. [Google Scholar] [CrossRef]
  110. Li, B.; Yang, Y.; Hu, Z.; Guo, Y.; Zhang, N.; Cheng, Y.; Gao, H. Formation mechanism and its photo-luminescent behavior of Lu2Ti2O7 nano-plates composed of single crystal grains. CrystEngComm 2013, 15, 5956–5960. [Google Scholar] [CrossRef]
  111. Yu, H.; Gao, X.; Deng, Y.; Xu, S.; Chen, B.; Zhang, Y.; Zhao, H. Electrospinning preparation and electrochemical properties of Lu2Ti2O7 nanowires, nanotubes and wire-in-tube architectures for their potential lithium-ion battery applications. J. Phys. Chem. Solids 2022, 169, 110889. [Google Scholar] [CrossRef]
  112. Feng, J.; Xiao, B.; Wan, C.L.; Qu, Z.X.; Huang, Z.C.; Chen, J.C.; Zhou, R.; Pan, W. Electronic structure, mechanical properties and thermal conductivity of Ln2Zr2O7 (Ln=La, Pr, Nd, Sm, Eu and Gd) pyrochlore. Acta Mater. 2011, 59, 1742–1760. [Google Scholar] [CrossRef]
  113. Zhao, F.A.; Xiao, H.Y.; Liu, Z.J.; Li, S.; Zu, X.T. A DFT study of mechanical properties, thermal conductivity and electronic structures of Th-doped Gd2Zr2O7. Acta Mater. 2016, 121, 299–309. [Google Scholar] [CrossRef]
  114. Toulemonde, M.; Dufour, C.; Paumier, E. Transient thermal process after a high-energy heavy-ion irradiation of amorphous metals and semiconductors. Phys. Rev. B Condens. Matter. 1992, 46, 14362–14369. [Google Scholar] [CrossRef] [PubMed]
  115. Lan, G.; Ouyang, B.; Song, J. The role of low-lying optical phonons in lattice thermal conductance of rare-earth pyrochlores: A first-principle study. Acta Mater. 2015, 91, 304–317. [Google Scholar] [CrossRef]
  116. Zhao, F.A.; Xiao, H.Y.; Bai, X.M.; Liu, Z.J.; Zu, X.T. Effects of doping Yb3+, La3+, Ti4+, Hf4+, Ce4+ cations on the mechanical properties, thermal conductivity, and electronic structures of Gd2Zr2O7. J. Alloys Compd. 2019, 776, 306–318. [Google Scholar] [CrossRef]
  117. Yang, J.; Shahid, M.; Zhao, M.; Feng, J.; Wan, C.; Pan, W. Physical properties of La2B2O7(B Zr, Sn, Hf and Ge) pyrochlore: First-principles calculations. J. Alloys Compd. 2016, 663, 834–841. [Google Scholar] [CrossRef] [Green Version]
  118. Li, F.; Zhou, L.; Liu, J.-X.; Liang, Y.; Zhang, G.-J. High-entropy pyrochlores with low thermal conductivity for thermal barrier coating materials. J. Adv. Ceram. 2019, 8, 576–582. [Google Scholar] [CrossRef] [Green Version]
  119. Wright, A.J.; Wang, Q.; Ko, S.-T.; Chung, K.M.; Chen, R.; Luo, J. Size disorder as a descriptor for predicting reduced thermal conductivity in medium- and high-entropy pyrochlore oxides. Scr. Mater. 2020, 181, 76–81. [Google Scholar] [CrossRef]
  120. 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]
  121. Xia, X.-L.; Liu, Z.-G.; Ouyang, J.-H.; Zheng, Y. Influence of trivalent Gd and Dy codoping on the structure and electrical conductivity of pyrochlore-type Sm2Zr2O7. J. Power Sources 2012, 217, 316–321. [Google Scholar] [CrossRef]
  122. Xia, X.-L.; Liu, Z.-G.; Ouyang, J.-H.; Zheng, Y. Preparation, Structural Characterization, and Enhanced Electrical Conductivity of Pyrochlore-type (Sm1–xEux)2Zr2O7 Ceramics. Fuel Cells 2012, 12, 624–632. [Google Scholar] [CrossRef]
  123. Liu, Z.-G.; Ouyang, J.-H.; Sun, K.-N.; Zhou, Y.; Xiang, J. Preparation, structure and electrical conductivity of pyrochlore-type Gd1–xEu2xSm1–xZr2O7 ceramics with a constant lattice parameter. Electrochim. Acta 2011, 56, 7045–7050. [Google Scholar] [CrossRef]
  124. Xia, X.-L.; Liu, Z.-G.; Ouyang, J.-H. Order–disorder transformation and enhanced oxide-ionic conductivity of (Sm1−xDyx)2Zr2O7 ceramics. J. Power Sources 2011, 196, 1840–1846. [Google Scholar] [CrossRef]
  125. Jin, Y.-J.; Liu, Z.-G.; Cao, G.; Feng, W.; Chen, L.; Wang, Y.-J.; Ouyang, J.-H. Order-disorder transition and electrical performance of (Nd1–xYx)2Zr2O7 solid solutions. J. Alloys Compd. 2019, 811, 151974. [Google Scholar] [CrossRef]
  126. Van Dijk, M.P.; de Vries, K.J.; Burggraaf, A.J. Oxygen ion and mixed conductivity in compounds with the fluorite and pyrochlore structure. Solid State Ion. 1983, 9–10, 913–919. [Google Scholar] [CrossRef] [Green Version]
  127. Liu, Z.-G.; Gao, S.; Ouyang, J.-H.; Xia, X.-L. Influence of MoO3 doping on structure and electrical conductivity of defect fluorite-type Gd2Zr2O7. J. Alloys Compd. 2010, 506, 868–871. [Google Scholar] [CrossRef]
  128. Zhang, R.; Xu, Q.; Pan, W.; Wan, C.L.; Qi, L.H.; Miao, H.Z. Structure and Ionic Conductivity of Ln2Zr2O7 -Type Rare Earth Zirconates. Key Eng. Mater. 2007, 336–338, 420–423. [Google Scholar] [CrossRef]
  129. Matzke, H. Radiation damage in crystalline insulators, oxides and ceramic nuclear fuels. Radiat. Eff. 1984, 64, 3–33. [Google Scholar] [CrossRef]
  130. Konings, R.J.M.; Wiss, T.; Guéneau, C. Nuclear Fuels. In The Chemistry of the Actinide and Transactinide Elements, 4th ed.; Morss, L.R., Edelstein, N.M., Fuger, J., Eds.; Springer: Dordrecht, The Netherlands, 2010; Volume 1. [Google Scholar]
  131. Firestone, R.B. Table of Isotopes, 8th ed.; Baglin, C.M., Chu, S.Y.F., Eds.; Wiley: New York, NY, USA, 1999. [Google Scholar]
  132. Ojovan, M.I.; Lee, W.E. Safety and Performance Assessments. In An Introduction to Nuclear Waste Immobilisation, 2nd ed.; Ojovan, M.I., Lee, W.E., Eds.; Elsevier: Oxford, UK, 2014; pp. 337–362. [Google Scholar]
  133. Wang, S.X.; Wang, L.M.; Ewing, R.C.; Was, G.S.; Lumpkin, G.R. Ion irradiation-induced phase transformation of pyrochlore and zirconolite. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 1999, 148, 704–709. [Google Scholar] [CrossRef]
  134. Lu, F.; Shen, Y.; Sun, X.; Dong, Z.; Ewing, R.C.; Lian, J. Size dependence of radiation-induced amorphization and recrystallization of synthetic nanostructured CePO4 monazite. Acta Mater. 2013, 61, 2984–2992. [Google Scholar] [CrossRef]
  135. Lang, M.; Toulemonde, M.; Zhang, J.; Zhang, F.; Tracy, C.L.; Lian, J.; Wang, Z.; Weber, W.J.; Severin, D.; Bender, M.; et al. Swift heavy ion track formation in Gd2Zr2−xTixO7 pyrochlore: Effect of electronic energy loss. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2014, 336, 102–115. [Google Scholar] [CrossRef] [Green Version]
  136. Ziegler, J.F.; Ziegler, M.D.; Biersack, J.P. SRIM—The stopping and range of ions in matter (2010). Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2010, 268, 1818–1823. [Google Scholar] [CrossRef] [Green Version]
  137. Zhang, J.; Lang, M.; Ewing, R.C.; Devanathan, R.; Weber, W.J.; Toulemonde, M. Nanoscale phase transitions under extreme conditions within an ion track. J. Mater. Res. 2011, 25, 1344–1351. [Google Scholar] [CrossRef]
  138. Begg, B.D.; Hess, N.J.; McCready, D.E.; Thevuthasan, S.; Weber, W.J. Heavy-ion irradiation effects in Gd2(Ti2−xZrx)O7 pyrochlores. J. Nucl. Mater. 2001, 289, 188–193. [Google Scholar] [CrossRef]
  139. Wang, S.X.; Begg, B.D.; Wang, L.M.; Ewing, R.C.; Weber, W.J.; Kutty, K.V.G. Radiation stability of gadolinium zirconate: A waste form for plutonium disposition. J. Mater. Res. 1999, 14, 4470–4473. [Google Scholar] [CrossRef]
  140. Lang, M.; Zhang, F.; Zhang, J.; Wang, J.; Schuster, B.; Trautmann, C.; Neumann, R.; Becker, U.; Ewing, R.C. Nanoscale manipulation of the properties of solids at high pressure with relativistic heavy ions. Nat. Mater. 2009, 8, 793–797. [Google Scholar] [CrossRef]
  141. Lang, M.; Zhang, F.; Zhang, J.; Wang, J.; Lian, J.; Weber, W.J.; Schuster, B.; Trautmann, C.; Neumann, R.; Ewing, R.C. Review of A2B2O7 pyrochlore response to irradiation and pressure. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2010, 268, 2951–2959. [Google Scholar] [CrossRef]
  142. Sattonnay, G.; Grygiel, C.; Monnet, I.; Legros, C.; Herbst-Ghysel, M.; Thomé, L. Phenomenological model for the formation of heterogeneous tracks in pyrochlores irradiated with swift heavy ions. Acta Mater. 2012, 60, 22–34. [Google Scholar] [CrossRef]
  143. Jozwik-Biala, I.; Jagielski, J.; Arey, B.; Kovarik, L.; Sattonnay, G.; Debelle, A.; Mylonas, S.; Monnet, I.; Thomé, L. Effect of combined local variations in elastic and inelastic energy losses on the morphology of tracks in ion-irradiated materials. Acta Mater. 2013, 61, 4669–4675. [Google Scholar] [CrossRef]
  144. Weber, W.J. Models and mechanisms of irradiation-induced amorphization in ceramics. Nuclear Instruments and Methods. Phys. Res. Sect. B Beam Interact. Mater. At. 2000, 166–167, 98–106. [Google Scholar] [CrossRef]
  145. Szenes, G. General features of latent track formation in magnetic insulators irradiated with swift heavy ions. Phys Rev B Condens Matter. 1995, 51, 8026–8029. [Google Scholar] [CrossRef]
  146. Schiwietz, G.; Czerski, K.; Roth, M.; Staufenbiel, F.; Grande, P.L. Femtosecond dynamics—Snapshots of the early ion-track evolution. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2004, 226, 683–704. [Google Scholar] [CrossRef]
  147. Gibbons, J.F. Ion implantation in semiconductors—Part II: Damage production and annealing. Proc. IEEE 1972, 60, 1062–1096. [Google Scholar] [CrossRef]
  148. Avasthi, D.K.; Ghosh, S.; Srivastava, S.K.; Assmann, W. Existence of transient temperature spike induced by SHI: Evidence by ion beam analysis. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2004, 219–220, 206–214. [Google Scholar] [CrossRef]
  149. Izui, K. Fission Fragment Damage in Semiconductors and Ionic Crystals. J. Phys. Soc. Jpn. 1965, 20, 915–932. [Google Scholar] [CrossRef]
  150. Fujimoto, J.G.; Liu, J.M.; Ippen, E.P.; Bloembergen, N. Femtosecond Laser Interaction with Metallic Tungsten and Nonequilibrium Electron and Lattice Temperatures. Phys. Rev. Lett. 1984, 53, 1837–1840. [Google Scholar] [CrossRef]
  151. Toulemonde, M.; Dufour, C.; Meftah, A.; Paumier, E. Transient thermal processes in heavy ion irradiation of crystalline inorganic insulators. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2000, 166–167, 903–912. [Google Scholar] [CrossRef]
  152. Moll, S.; Thomé, L.; Vincent, L.; Garrido, F.; Sattonnay, G.; Thomé, T.; Jagielski, J.; Costantini, J.M. Damage induced by electronic excitation in ion-irradiated yttria-stabilized zirconia. J. Appl. Phys. 2009, 105, 023512. [Google Scholar] [CrossRef]
  153. Sattonnay, G.; Moll, S.; Thomé, L.; Decorse, C.; Legros, C.; Simon, P.; Jagielski, J.; Jozwik, I.; Monnet, I. Phase transformations induced by high electronic excitation in ion-irradiated Gd2(ZrxTi1−x)2O7 pyrochlores. Journal of Applied Physics 2010, 108. [Google Scholar] [CrossRef]
  154. Nordlund, K.; Zinkle, S.J.; Sand, A.E.; Granberg, F.; Averback, R.S.; Stoller, R.E.; Suzudo, T.; Malerba, L.; Banhart, F.; Weber, W.J.; et al. Primary radiation damage: A review of current understanding and models. J. Nucl. Mater. 2018, 512, 450–479. [Google Scholar] [CrossRef]
  155. Chartier, A.; Meis, C.; Crocombette, J.-P.; Corrales, L.R.; Weber, W.J. Atomistic modeling of displacement cascades in La2Zr2O7 pyrochlore. Phys. Rev. B 2003, 67, 174102. [Google Scholar] [CrossRef]
  156. Devanathan, R.; Weber, W.J.; Gale, J.D. Radiation tolerance of ceramics—Insights from atomistic simulation of damage accumulation in pyrochlores. Energy Environ. Sci. 2010, 3, 1551–1559. [Google Scholar] [CrossRef] [Green Version]
  157. Dong, L.; Li, Y.; Devanathan, R.; Gao, F. Molecular dynamics simulation of the structural, elastic, and thermal properties of pyrochlores. RSC Adv. 2016, 6, 41410–41419. [Google Scholar] [CrossRef]
  158. Dong, L.; Li, Y.; Devanathan, R.; Setyawan, W.; Gao, F. Low energy ion-solid interactions and chemistry effects in a series of pyrochlores. J. Am. Ceram. Soc. 2017, 100, 3132–3144. [Google Scholar] [CrossRef]
  159. Dong, L.Y.; Setyawan, W.; Li, Y.H.; Devanathan, R.; Gao, F. Molecular dynamics simulation of low-energy recoil events in titanate pyrochlores. Rsc Adv. 2017, 7, 35403–35410. [Google Scholar] [CrossRef] [Green Version]
  160. Xiao, H.Y.; Zhang, Y.; Weber, W.J. Ab initiomolecular dynamics simulations of low-energy recoil events in ThO2, CeO2, and ZrO2. Phys. Rev. B 2012, 86, 054109. [Google Scholar] [CrossRef]
  161. Liu, B.; Xiao, H.Y.; Zhang, Y.; Weber, W.J. Ab initio molecular dynamics simulations of overlapping recoil events in ThO2. J. Phys. Condens. Matter 2013, 25, 395004. [Google Scholar] [CrossRef]
  162. Wang, X.J.; Xiao, H.Y.; Zu, X.T.; Zhang, Y.; Weber, W.J. Ab initio molecular dynamics simulations of ion–solid interactions in Gd2Zr2O7 and Gd2Ti2O7. J. Mater. Chem. C 2013, 1, 1665–1673. [Google Scholar] [CrossRef]
  163. Xiao, H.Y.; Weber, W.J.; Zhang, Y.; Zu, X.T. Ab initio molecular dynamics simulations of ion–solid interactions in zirconate pyrochlores. Acta Mater. 2015, 87, 273–282. [Google Scholar] [CrossRef] [Green Version]
  164. Xiao, H.Y.; Gao, F.; Weber, W.J. Threshold displacement energies and defect formation energies in Y2Ti2O7. J. Phys. Condens. Matter. 2010, 22, 415801. [Google Scholar] [CrossRef]
  165. Ojovan, M.I.; Lee, W.E. Short-Lived Waste Radionuclides. In An Introduction to Nuclear Waste Immobilisation, 2nd ed.; Ojovan, M.I., Lee, W.E., Eds.; Elsevier: Oxford, UK, 2014; pp. 99–106. [Google Scholar]
  166. Zhang, J.; Wang, Y.Q.; Tang, M.; Sun, C.; Yin, D.M.; Li, N. Helium irradiation induced micro-swelling and phase separation in pyrochlore Lu2Ti2O7. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2015, 342, 179–183. [Google Scholar] [CrossRef] [Green Version]
  167. Taylor, C.A.; Patel, M.K.; Aguiar, J.A.; Zhang, Y.; Crespillo, M.L.; Wen, J.; Xue, H.; Wang, Y.; Weber, W.J. Bubble formation and lattice parameter changes resulting from He irradiation of defect-fluorite Gd2Zr2O7. Acta Mater. 2016, 115, 115–122. [Google Scholar] [CrossRef] [Green Version]
  168. Shi, K.; Shu, X.; Shao, D.; Zhang, H.; Liu, Y.; Peng, G.; Lu, X. Helium ions’ irradiation effects on Gd2Zr2O7 ceramics holding complex simulated radionuclides. J. Radioanal. Nucl. Chem. 2017, 314, 2113–2122. [Google Scholar] [CrossRef]
  169. Hu, Q.; Zeng, J.; Wang, L.; Shu, X.; Shao, D.; Zhang, H.; Lu, X. Helium ion irradiation effects on neodymium and cerium co-doped Gd2Zr2O7 pyrochlore ceramic. J. Rare Earths 2018, 36, 398–403. [Google Scholar] [CrossRef]
  170. Huang, Z.; Ma, N.; Qi, J.; Guo, X.; Yang, M.; Tang, Z.; Zhang, Y.; Han, Y.; Wu, D.; Lu, T. Defect-fluorite Gd2Zr2O7 ceramics under helium irradiation: Amorphization, cell volume expansion, and multi-stage bubble formation. J. Am. Ceram. Soc. 2019, 102, 4911–4918. [Google Scholar] [CrossRef] [Green Version]
  171. Yang, D.Y.; Liu, C.G.; Zhang, K.Q.; Xia, Y.; Chen, L.J.; Liu, H.; Li, Y.H. 2.7 MeV Ar11+ ion irradiation induced structural evolution in Lu2(Ti2−xLux)O7−x/2 pyrochlores. J. Nucl. Mater. 2015, 466, 496–501. [Google Scholar] [CrossRef]
  172. Li, Y.H.; Uberuaga, B.P.; Jiang, C.; Choudhury, S.; Valdez, J.A.; Patel, M.K.; Won, J.; Wang, Y.Q.; Tang, M.; Safarik, D.J.; et al. Role of antisite disorder on preamorphization swelling in titanate pyrochlores. Phys. Rev. Lett. 2012, 108, 195504. [Google Scholar] [CrossRef]
  173. Seitz, F.; Koehler, J.S. Displacement of Atoms during Irradiation. In Solid State Physics; Seitz, F., Turnbull, D., Eds.; Academic Press: New York, NY, USA, 1956; Volume 2, p. 305. [Google Scholar]
  174. Toulemonde, M.; Assmann, W.; Dufour, C.; Meftah, A.; Trautmann, C. Nanometric transformation of the matter by short and intense electronic excitation: Experimental data versus inelastic thermal spike model. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2012, 277, 28–39. [Google Scholar] [CrossRef]
  175. Wilde, P. Defects and diffusion in pyrochlore structured oxides. Solid State Ion. 1998, 112, 173–183. [Google Scholar] [CrossRef]
  176. Jozwik-Biala, I.; Jagielski, J.; Thomé, L.; Arey, B.; Kovarik, L.; Sattonnay, G.; Debelle, A.; Monnet, I. HRTEM study of track evolution in 120-MeV U irradiated Gd2Ti2O7. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2012, 286, 258–261. [Google Scholar] [CrossRef]
  177. Zhang, J.; Toulemonde, M.; Lang, M.; Costantini, J.M.; Della-Negra, S.; Ewing, R.C. C60 and U ion irradiation of Gd2TixZr2−xO7 pyrochlore. J. Mater. Res. 2015, 30, 2456–2466. [Google Scholar] [CrossRef]
  178. Radha, A.V.; Ushakov, S.V.; Navrotsky, A. Thermochemistry of lanthanum zirconate pyrochlore. J. Mater. Res. 2011, 24, 3350–3357. [Google Scholar] [CrossRef]
  179. Hecking, N.; Heidemann, K.F.; Te Kaat, E. Model of temperature dependent defect interaction and amorphization in crystalline silicon during ion irradiation. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 1986, 15, 760–764. [Google Scholar] [CrossRef]
  180. Lindner, J.K.N.; Zuschlag, R.; te Kaat, E.H. Radiation damage and amorphization of silicon by 2 MeV nitrogen ion implantation. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 1992, 62, 314–318. [Google Scholar] [CrossRef]
  181. Lindner, J.K.N. Radiation damage of 2 MeV Si ions in Si0.75Ge0.25: Optical measurements and damage modelling. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 1996, 112, 316–320. [Google Scholar] [CrossRef]
  182. Wendler, E.; Heft, A.; Wesch, W. Ion-beam induced damage and annealing behaviour in SiC. Nucl. Instrum. Methods Phys. Res. Sect. B: Beam Interact. Mater. At. 1998, 141, 105–117. [Google Scholar] [CrossRef]
  183. Lindner, J.K.N.; Frohnwieser, A.; Rauschenbach, B.; Stritzker, B. keV- and MeV- Ion Beam Synthesis of Buried SiC Layers in Silicon. In Proceedings of the Symposium A—Beam Solid Interactions for Materials Synthesis and Characterization, Boston, MA, USA, 28 November–2 December 1994; Volume 354. [Google Scholar] [CrossRef]
  184. Zhang, Y.; Weber, W.J.; Shutthanandan, V.; Devanathan, R.; Thevuthasan, S.; Balakrishnan, G.; Paul, D.M. Damage evolution on Sm and O sublattices in Au-implanted samarium titanate pyrochlore. J. Appl. Phys. 2004, 95, 2866–2872. [Google Scholar] [CrossRef] [Green Version]
  185. Jagielski, J.; Thomé, L. Multi-step damage accumulation in irradiated crystals. Appl. Phys. A 2009, 97, 147–155. [Google Scholar] [CrossRef]
  186. Thomé, L.; Jagielski, J.; Binet, C.; Garrido, F. Structural properties of fission-product doped ZrO2 and MgAl2O4 single crystals. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2000, 166–167, 258–262. [Google Scholar] [CrossRef]
  187. Jagielski, J.; Thomé, L. Accumulation of radiation damage in spinel crystals: Quantitative analysis and interpretation of RBS/C data. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2007, 261, 1155–1158. [Google Scholar] [CrossRef]
  188. Soulié, A.; Menut, D.; Crocombette, J.-P.; Chartier, A.; Sellami, N.; Sattonnay, G.; Monnet, I.; Béchade, J.-L. X-ray diffraction study of the Y2Ti2O7 pyrochlore disordering sequence under irradiation. J. Nucl. Mater. 2016, 480, 314–322. [Google Scholar] [CrossRef]
  189. Martin, P.M.; Belin, R.C.; Valenza, P.J.; Scheinost, A.C. EXAFS study of the structural phase transition in the americium zirconate pyrochlore. J. Nucl. Mater. 2009, 385, 126–130. [Google Scholar] [CrossRef]
  190. Wang, S.X.; Wang, L.M.; Ewing, R.C.; Govindan Kutty, K.V. Ion irradiation of rare-earth- and yttrium-titanate-pyrochlores. Nucl. Instrum. Methods Phys. Res. Sect. B: Beam Interact. Mater. At. 2000, 169, 135–140. [Google Scholar] [CrossRef]
  191. Begg, B.D.; Hess, N.J.; Weber, W.J.; Devanathan, R.; Icenhower, J.P.; Thevuthasan, S.; McGrail, B.P. Heavy-ion irradiation effects on structures and acid dissolution of pyrochlores. J. Nucl. Mater. 2001, 288, 208–216. [Google Scholar] [CrossRef]
  192. Wang, S.X.; Wang, L.M.; Ewing, R.C.; Govindan Kutty, K.V. Ion Irradiation Effects for Two Pyrochlore Compositions: Gd2Ti2O7 and Gd2Zr2O7. MRS Online Proc. Libr. 1998, 540, 355–360. [Google Scholar] [CrossRef]
  193. Wang, S.X.; Wang, L.M.; Ewing, R.C. Irradiation-induced amorphization: Effects of temperature, ion mass, cascade size, and dose rate. Phys. Rev. B 2001, 63, 024105. [Google Scholar] [CrossRef]
  194. Lian, J.; Helean, K.B.; Kennedy, B.J.; Wang, L.M.; Navrotsky, A.; Ewing, R.C. Effect of structure and thermodynamic stability on the response of lanthanide stannate pyrochlores to ion beam irradiation. J. Phys. Chem. B 2006, 110, 2343–2350. [Google Scholar] [CrossRef]
  195. Lian, J.; Chen, J.; Wang, L.M.; Ewing, R.C.; Farmer, J.M.; Boatner, L.A.; Helean, K.B. Radiation-induced amorphization of rare-earth titanate pyrochlores. Phys. Rev. B 2003, 68, 134107. [Google Scholar] [CrossRef]
  196. Lumpkin, G.R.; Pruneda, M.; Rios, S.; Smith, K.L.; Trachenko, K.; Whittle, K.R.; Zaluzec, N.J. Nature of the chemical bond and prediction of radiation tolerance in pyrochlore and defect fluorite compounds. J. Solid State Chem. 2007, 180, 1512–1518. [Google Scholar] [CrossRef]
  197. Lian, J.; Ewing, R.C.; Wang, L.M.; Helean, K.B. Ion-beam irradiation of Gd2Sn2O7 and Gd2Hf2O7 pyrochlore: Bond-type effect. J. Mater. Res. 2004, 19, 1575–1580. [Google Scholar] [CrossRef]
  198. Kennedy, B.J.; Hunter, B.A.; Howard, C.J. Structural and Bonding Trends in Tin Pyrochlore Oxides. J. Solid State Chem. 1997, 130, 58–65. [Google Scholar] [CrossRef]
  199. Panero, W.R.; Stixrude, L.; Ewing, R.C. First-principles calculation of defect-formation energies in theY2(Ti,Sn,Zr)2O7pyrochlore. Phys. Rev. B 2004, 70, 054110. [Google Scholar] [CrossRef]
  200. Li, Y.; Kowalski, P.M. Energetics of defects formation and oxygen migration in pyrochlore compounds from first principles calculations. J. Nucl. Mater. 2018, 505, 255–261. [Google Scholar] [CrossRef]
  201. Andrievskaya, E.R. Phase equilibria in the refractory oxide systems of zirconia, hafnia and yttria with rare-earth oxides. J. Eur. Ceram. Soc. 2008, 28, 2363–2388. [Google Scholar] [CrossRef]
  202. Shlyakhtina, A.V. Morphotropy, isomorphism, and polymorphism of Ln2M2O7-based (Ln = La-Lu, Y, Sc; M = Ti, Zr, Hf, Sn) oxides. Crystallogr. Rep. 2013, 58, 548–562. [Google Scholar] [CrossRef]
  203. Li, Y.; Kowalski, P.M.; Beridze, G.; Birnie, A.R.; Finkeldei, S.; Bosbach, D. Defect formation energies in A2B2O7 pyrochlores. Scr. Mater. 2015, 107, 18–21. [Google Scholar] [CrossRef]
  204. Park, S.; Lang, M.; Tracy, C.L.; Zhang, J.; Zhang, F.; Trautmann, C.; Rodriguez, M.D.; Kluth, P.; Ewing, R.C. Response of Gd2Ti2O7 and La2Ti2O7 to swift-heavy ion irradiation and annealing. Acta Mater. 2015, 93, 1–11. [Google Scholar] [CrossRef] [Green Version]
  205. Helean, K.B.; Ushakov, S.V.; Brown, C.E.; Navrotsky, A.; Lian, J.; Ewing, R.C.; Farmer, J.M.; Boatner, L.A. Formation enthalpies of rare earth titanate pyrochlores. J. Solid State Chem. 2004, 177, 1858–1866. [Google Scholar] [CrossRef]
  206. Zhao, M.; O’Quinn, E.; Birkner, N.; Xu, Y.; Lang, M.; Brinkman, K. Radiation damage and thermal annealing in tunnel structured hollandite materials. Acta Mater. 2021, 206, 116598. [Google Scholar] [CrossRef]
  207. Park, S.; Tracy, C.L.; Zhang, F.; Palomares, R.I.; Park, C.; Trautmann, C.; Lang, M.; Mao, W.L.; Ewing, R.C. Swift-heavy ion irradiation response and annealing behavior of A2TiO5 (A = Nd, Gd, and Yb). J. Solid State Chem. 2018, 258, 108–116. [Google Scholar] [CrossRef]
  208. Kowalski, P.M. Formation enthalpy of Ln2B2O7-type (B=Ti, Sn, Hf, Zr) compounds. Scr. Mater. 2020, 189, 7–10. [Google Scholar] [CrossRef]
  209. Matsumoto, U.; Ogawa, T.; Kitaoka, S.; Moriwake, H.; Tanaka, I. First-Principles Study on the Stability of Weberite-Type, Pyrochlore, and Defect-Fluorite Structures of A23+B24+O7 (A = Lu3+–La3+, B = Zr4+, Hf4+, Sn4+, and Ti4+). J. Phys. Chem. C 2020, 124, 20555–20562. [Google Scholar] [CrossRef]
  210. Zhang, J.; Lian, J.; Fuentes, A.F.; Zhang, F.; Lang, M.; Lu, F.; Ewing, R.C. Enhanced radiation resistance of nanocrystalline pyrochlore Gd2(Ti0.65Zr0.35)2O7. Appl. Phys. Lett. 2009, 94, 243110. [Google Scholar] [CrossRef]
  211. Moreno, K.J.; Fuentes, A.F.; Maczka, M.; Hanuza, J.; Amador, U.; Santamaría, J.; León, C. Influence of thermally induced oxygen order on mobile ion dynamics in Gd2(Ti0.65Zr0.35)2O7. Phys. Rev. B 2007, 75, 184303. [Google Scholar] [CrossRef] [Green Version]
  212. Chimi, Y.; Iwase, A.; Ishikawa, N.; Kobiyama, M.; Inami, T.; Okuda, S. Accumulation and recovery of defects in ion-irradiated nanocrystalline gold. J. Nucl. Mater. 2001, 297, 355–357. [Google Scholar] [CrossRef]
  213. Rose, M.; Balogh, A.G.; Hahn, H. Instability of irradiation induced defects in nanostructured materials. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 1997, 127–128, 119–122. [Google Scholar] [CrossRef]
  214. Rose, M.; Gorzawski, G.; Miehe, G.; Balogh, A.G.; Hahn, H. Phase stability of nanostructured materials under heavy ion irradiation. Nanostruct. Mater. 1995, 6, 731–734. [Google Scholar] [CrossRef]
  215. Kilmametov, A.R.; Gunderov, D.V.; Valiev, R.Z.; Balogh, A.G.; Hahn, H. Enhanced ion irradiation resistance of bulk nanocrystalline TiNi alloy. Scr. Mater. 2008, 59, 1027–1030. [Google Scholar] [CrossRef]
  216. Shen, T.D.; Feng, S.; Tang, M.; Valdez, J.A.; Wang, Y.; Sickafus, K.E. Enhanced radiation tolerance in nanocrystalline MgGa2O4. Applied Physics Letters 2007, 90. [Google Scholar] [CrossRef]
  217. Bai, X.M.; Voter, A.F.; Hoagland, R.G.; Nastasi, M.; Uberuaga, B.P. Efficient annealing of radiation damage near grain boundaries via interstitial emission. Science 2010, 327, 1631–1634. [Google Scholar] [CrossRef]
  218. Chen, D.; Wang, J.; Chen, T.; Shao, L. Defect annihilation at grain boundaries in alpha-Fe. Sci. Rep. 2013, 3, 1450. [Google Scholar] [CrossRef] [Green Version]
  219. Meldrum, A.; Boatner, L.A.; Ewing, R.C. Nanocrystalline zirconia can be amorphized by ion irradiation. Phys. Rev. Lett. 2002, 88, 025503. [Google Scholar] [CrossRef]
  220. Sickafus, K.E.; Matzke, H.; Hartmann, T.; Yasuda, K.; Valdez, J.A.; Chodak Iii, P.; Nastasi, M.; Verrall, R.A. Radiation damage effects in zirconia. J. Nucl. Mater. 1999, 274, 66–77. [Google Scholar] [CrossRef]
  221. Meldrum, A.; Boatner, L.A.; Ewing, R.C. Displacive radiation effects in the monazite- and zircon-structure orthophosphates. Phys. Rev. B 1997, 56, 13805–13814. [Google Scholar] [CrossRef]
  222. Shen, T.D. Radiation tolerance in a nanostructure: Is smaller better? Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2008, 266, 921–925. [Google Scholar] [CrossRef]
  223. Jin, K.; Lu, C.; Wang, L.M.; Qu, J.; Weber, W.J.; Zhang, Y.; Bei, H. Effects of compositional complexity on the ion-irradiation induced swelling and hardening in Ni-containing equiatomic alloys. Scr. Mater. 2016, 119, 65–70. [Google Scholar] [CrossRef] [Green Version]
  224. Lu, C.; Niu, L.; Chen, N.; Jin, K.; Yang, T.; Xiu, P.; Zhang, Y.; Gao, F.; Bei, H.; Shi, S.; et al. Enhancing radiation tolerance by controlling defect mobility and migration pathways in multicomponent single-phase alloys. Nat. Commun. 2016, 7, 13564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Lu, C.; Yang, T.; Jin, K.; Gao, N.; Xiu, P.; Zhang, Y.; Gao, F.; Bei, H.; Weber, W.J.; Sun, K.; et al. Radiation-induced segregation on defect clusters in single-phase concentrated solid-solution alloys. Acta Mater. 2017, 127, 98–107. [Google Scholar] [CrossRef]
  226. Barr, C.M.; Nathaniel, J.E.; Unocic, K.A.; Liu, J.; Zhang, Y.; Wang, Y.; Taheri, M.L. Exploring radiation induced segregation mechanisms at grain boundaries in equiatomic CoCrFeNiMn high entropy alloy under heavy ion irradiation. Scr. Mater. 2018, 156, 80–84. [Google Scholar] [CrossRef]
  227. Kombaiah, B.; Jin, K.; Bei, H.; Edmondson, P.D.; Zhang, Y. Phase stability of single phase Al0.12CrNiFeCo high entropy alloy upon irradiation. Mater. Des. 2018, 160, 1208–1216. [Google Scholar] [CrossRef]
  228. Lin, Y.; Yang, T.; Lang, L.; Shan, C.; Deng, H.; Hu, W.; Gao, F. Enhanced radiation tolerance of the Ni-Co-Cr-Fe high-entropy alloy as revealed from primary damage. Acta Mater. 2020, 196, 133–143. [Google Scholar] [CrossRef]
  229. Zhou, L.; Li, F.; Liu, J.X.; Sun, S.K.; Liang, Y.; Zhang, G.J. High-entropy A2B2O7-type oxide ceramics: A potential immobilising matrix for high-level radioactive waste. J. Hazard. Mater. 2021, 415, 125596. [Google Scholar] [CrossRef]
  230. Wang, H.; Zhu, C.; Wang, Z.; Wang, M.; Li, Y. Mechanical property and irradiation resistance of high entropy pyrochlore (Sm0.2Eu0.2Gd0.2Y0.2Lu0.2)2Ti2O7. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2022, 533, 17–22. [Google Scholar] [CrossRef]
  231. Weber, W.J.; Hess, N.J. Ion beam modification of Gd2Ti2O7. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 1993, 80–81, 1245–1248. [Google Scholar] [CrossRef]
  232. Sickafus, K.E.; Minervini, L.; Grimes, R.W.; Valdez, J.A.; Hartmann, T. A comparison between radiation damage accumulation in oxides with pyrochlore and fluorite structures. Radiat. Eff. Defects Solids 2001, 155, 133–137. [Google Scholar] [CrossRef]
  233. Sykora, R.E.; Raison, P.E.; Haire, R.G. Self-irradiation induced structural changes in the transplutonium pyrochlores An2Zr2O7 (An=Am, Cf). J. Solid State Chem. 2005, 178, 578–583. [Google Scholar] [CrossRef]
  234. Haire, R.G.; Raison, P.E.; Assefa, Z. Systematic Studies of the Fundamental Chemistry of Pyrochlore oxides: An2Zr2O7 [An = Pu, Am, Cm, Bk and Cf]. J. Nucl. Sci. Technol. 2002, 39, 616–619. [Google Scholar] [CrossRef]
  235. Laverov, N.P.; Yudintsev, S.V.; Livshits, T.S.; Stefanovsky, S.V.; Lukinykh, A.N.; Ewing, R.C. Synthetic minerals with the pyrochlore and garnet structures for immobilization of actinide-containing wastes. Geochem. Int. 2010, 48, 1–14. [Google Scholar] [CrossRef]
  236. Sattonnay, G.; Moll, S.; Desbrosses, V.; Menvie Bekale, V.; Legros, C.; Thomé, L.; Monnet, I. Mechanical properties of fluorite-related oxides subjected to swift ion irradiation: Pyrochlore and zirconia. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2010, 268, 3040–3043. [Google Scholar] [CrossRef]
  237. Laverov, N.P.; Yudintsev, S.V.; Stefanovsky, S.V.; Ewing, R.C. Phase composition and radiation stability of matrices for isolation of REE-actinide waste. Dokl. Earth Sci. 2012, 443, 526–531. [Google Scholar] [CrossRef]
  238. Yudintsev, S.V.; Lizin, A.A.; Livshits, T.S.; Stefanovsky, S.V.; Tomilin, S.V.; Ewing, R.C. Ion-beam irradiation and 244Cm-doping investigations of the radiation response of actinide-bearing crystalline waste forms. J. Mater. Res. 2015, 30, 1516–1528. [Google Scholar] [CrossRef]
  239. Li, Y.H.; Wang, Y.Q.; Valdez, J.A.; Tang, M.; Sickafus, K.E. Swelling effects in Y2Ti2O7 pyrochlore irradiated with 400keV Ne2+ ions. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2012, 274, 182–187. [Google Scholar] [CrossRef]
  240. Tracy, C.L.; Lang, M.; Zhang, J.; Zhang, F.; Wang, Z.; Ewing, R.C. Structural response of A2TiO5 (A=La, Nd, Sm, Gd) to swift heavy ion irradiation. Acta Mater. 2012, 60, 4477–4486. [Google Scholar] [CrossRef]
  241. Xie, Q.R.; Zhang, J.; Dong, X.N.; Guo, Q.X.; Li, N. Heavy ion irradiation-induced microstructural evolution in pyrochlore Lu2Ti2O7 at room temperature and 723 K. J. Solid State Chem. 2015, 231, 159–162. [Google Scholar] [CrossRef]
  242. Zhang, J.; Lang, M.; Lian, J.; Liu, J.; Trautmann, C.; Della-Negra, S.; Toulemonde, M.; Ewing, R.C. Liquid-like phase formation in Gd2Zr2O7 by extremely ionizing irradiation. J. Appl. Phys. 2009, 105, 113510. [Google Scholar] [CrossRef]
  243. Kurpaska, L.; Jagielski, J. Mechanical properties of irradiated Gd2Zr2O7 pyrochlores as studied by nanoindentation technique—Effect of grains and grain boundaries. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2016, 379, 107–111. [Google Scholar] [CrossRef]
  244. Shamblin, J.; Tracy, C.L.; Ewing, R.C.; Zhang, F.; Li, W.; Trautmann, C.; Lang, M. Structural response of titanate pyrochlores to swift heavy ion irradiation. Acta Mater. 2016, 117, 207–215. [Google Scholar] [CrossRef] [Green Version]
  245. Chen, S.; Shu, X.; Wang, L.; Chi, F.; Shao, D.; Ding, Y.; Yuan, X.; Qing, Q.; Lu, X. Effects of alpha irradiation on Nd2Zr2O7 matrix for nuclear waste forms. J. Aust. Ceram. Soc. 2017, 54, 33–38. [Google Scholar] [CrossRef]
  246. Shu, X.; Lu, X.; Fan, L.; Yang, R.; Ding, Y.; Pan, S.; Zhou, P.; Wu, Y. Design and fabrication of Gd2Zr2O7-based waste forms for U3O8 immobilization in high capacity. J. Mater. Sci. 2016, 51, 5281–5289. [Google Scholar] [CrossRef]
  247. Lu, X.; Shu, X.; Chen, S.; Zhang, K.; Chi, F.; Zhang, H.; Shao, D.; Mao, X. Heavy-ion irradiation effects on U3O8 incorporated Gd2Zr2O7 waste forms. J. Hazard. Mater. 2018, 357, 424–430. [Google Scholar] [CrossRef]
  248. Xia, Y.; Liu, C.G.; Yang, D.Y.; Wen, J.; Liu, H.; Mu, P.C.; Li, Y.H. Synthesis and radiation tolerance of Lu2−xCexTi2O7 pyrochlores. J. Nucl. Mater. 2016, 480, 182–188. [Google Scholar] [CrossRef]
  249. Yudintsev, S.V.; Tomilin, S.V.; Livshits, T.S.; Lizin, A.A.; Goryatchev, I.A. Curium-doped stannate pyrochlore: Durability under radiation and leaching in water. Dokl. Earth Sci. 2016, 469, 732–736. [Google Scholar] [CrossRef]
  250. Kaspar, T.C.; Gigax, J.G.; Shao, L.; Bowden, M.E.; Varga, T.; Shutthanandan, V.; Spurgeon, S.R.; Yan, P.; Wang, C.; Ramuhalli, P.; et al. Damage evolution of ion irradiated defected-fluorite La2Zr2O7 epitaxial thin films. Acta Mater. 2017, 130, 111–120. [Google Scholar] [CrossRef]
  251. Shu, X.; Chen, S.; Wang, L.; Li, B.; Xie, Y.; Yang, Y.; Li, L.; Lu, X. Simulated self-irradiation effects of Gd2Ce2O7 nuclear waste form. J. Radioanal. Nucl. Chem. 2020, 324, 271–276. [Google Scholar] [CrossRef]
  252. Lu, X.; Shu, X.; Shao, D.; Chen, S.; Zhang, H.; Yuan, X.; Chi, F. Radiation stability of Gd2Zr2O7 and Nd2Ce2O7 ceramics as nuclear waste forms. Ceram. Int. 2018, 44, 760–765. [Google Scholar] [CrossRef]
  253. Qing, Q.; Shu, X.; Shao, D.; Zhang, H.; Chi, F.; Lu, X. Irradiation response of Nd2Zr2O7 under heavy ions irradiation. J. Eur. Ceram. Soc. 2018, 38, 2068–2073. [Google Scholar] [CrossRef]
  254. Park, S.; Tracy, C.L.; Zhang, F.; Park, C.; Trautmann, C.; Tkachev, S.N.; Lang, M.; Mao, W.L.; Ewing, R.C. Radiation-induced disorder in compressed lanthanide zirconates. Phys. Chem. Chem. Phys. 2018, 20, 6187–6197. [Google Scholar] [CrossRef]
  255. Sachan, R.; Chisholm, M.F.; Ou, X.; Zhang, Y.; Weber, W.J. Energetic Ion Irradiation-Induced Disordered Nanochannels for Fast Ion Conduction. JOM 2018, 71, 103–108. [Google Scholar] [CrossRef]
  256. Wang, L.L.; Li, J.B.; Xie, H.; Deng, S.H.; Zhang, K.X.; Yi, F.C. Solubility and ion-irradiation effects of uranium in Nd2Zr2O7 pyrochlore. Acta Phys. Sin. 2018, 67, 192801. [Google Scholar] [CrossRef]
  257. Sharma, S.K.; Grover, V.; Tyagi, A.K.; Avasthi, D.K.; Singh, U.B.; Kulriya, P.K. Probing the temperature effects in the radiation stability of Nd2Zr2O7 pyrochlore under swift ion irradiation. Materialia 2019, 6, 100317. [Google Scholar] [CrossRef]
  258. Zhou, M.; Hou, C.; Xie, Y.; Wang, L.; Shu, X.; Shao, D.; Lu, X. Alpha-radiation effects of Gd2Zr2O7 bearing simulated multi-nuclides. J. Aust. Ceram. Soc. 2019, 55, 831–836. [Google Scholar] [CrossRef]
  259. Panghal, A.; Kumar, Y.; Kulriya, P.K.; Shirage, P.M.; Singh, N.L. Atomic order-disorder engineering in the La2Zr2O7 pyrochlore under low energy ion irradiation. Ceram. Int. 2021, 47, 20248–20259. [Google Scholar] [CrossRef]
  260. Panghal, A.; Kumar, Y.; Kulriya, P.K.; Shirage, P.M.; Singh, N.L. Structural assessment and irradiation response of La2Zr2O7 pyrochlore: Impact of irradiation temperature and ion fluence. J. Alloys Compd. 2021, 862, 158556. [Google Scholar] [CrossRef]
Figure 1. Pyrochlore structures described based on (a) the polyhedral network and (b) the derivative of fluorite structure, which is one eighth of the conventional unit cell of pyrochlore. Corresponding fluorite unit cells (c) are included for comparison. (df) Structural relations between the (d) pyrochlore, (e) weberite-type, and (f) disordered fluorite. Both the pyrochlore and weberite-type structures are supercells of disordered fluorite with unit cell parameters of aP = aWB = 2aF and bWBcWB ≈ 21/2aF (the subscripts refer to pyrochlore, weberite-type and disordered fluorite respectively). Red spheres represent oxygen sites that are fully occupied in the pyrochlore and weberite-type structures, and grey spheres represent oxygen vacancy. Purple and green spheres represent A-site (CN = 8) and B-site (CN = 6) cations, respectively. Half-purple and half-green spheres designate cation site mixing (CN = 7).
Figure 1. Pyrochlore structures described based on (a) the polyhedral network and (b) the derivative of fluorite structure, which is one eighth of the conventional unit cell of pyrochlore. Corresponding fluorite unit cells (c) are included for comparison. (df) Structural relations between the (d) pyrochlore, (e) weberite-type, and (f) disordered fluorite. Both the pyrochlore and weberite-type structures are supercells of disordered fluorite with unit cell parameters of aP = aWB = 2aF and bWBcWB ≈ 21/2aF (the subscripts refer to pyrochlore, weberite-type and disordered fluorite respectively). Red spheres represent oxygen sites that are fully occupied in the pyrochlore and weberite-type structures, and grey spheres represent oxygen vacancy. Purple and green spheres represent A-site (CN = 8) and B-site (CN = 6) cations, respectively. Half-purple and half-green spheres designate cation site mixing (CN = 7).
Crystals 13 00143 g001
Figure 2. (a) The total conductivity of pyrochlores at 1073 K is a function of rA/rB. (b) Comparison of the total conductivity between high entropy rare-earth zirconates and other low/medium entropy rare-earth zirconates at 1073 K. Note, HEC from Ref. [75], (Sm1−xYbx)2Zr2O7 from Ref. [89], (Sm1−xEux)2Zr2O7 from Ref. [122], Sm1−xEu2xGd1−xZr2O7 from Ref. [123], SmGd1−xDyxZr2O7 from [121], (Sm1−xDyx)2Zr2O7 from Ref. [124], (Nd1−xYx)2Zr2O7 from Ref. [125], Gd2Zr2O7 from Ref. [127], and Er2Zr2O7 from Ref. [128]. * in (a) means the data of SmGd1−xDyxZr2O7 is grain conductivity.
Figure 2. (a) The total conductivity of pyrochlores at 1073 K is a function of rA/rB. (b) Comparison of the total conductivity between high entropy rare-earth zirconates and other low/medium entropy rare-earth zirconates at 1073 K. Note, HEC from Ref. [75], (Sm1−xYbx)2Zr2O7 from Ref. [89], (Sm1−xEux)2Zr2O7 from Ref. [122], Sm1−xEu2xGd1−xZr2O7 from Ref. [123], SmGd1−xDyxZr2O7 from [121], (Sm1−xDyx)2Zr2O7 from Ref. [124], (Nd1−xYx)2Zr2O7 from Ref. [125], Gd2Zr2O7 from Ref. [127], and Er2Zr2O7 from Ref. [128]. * in (a) means the data of SmGd1−xDyxZr2O7 is grain conductivity.
Crystals 13 00143 g002
Figure 3. Cumulative Dose as a function of waste storage time. (a) β-decay and α-decay events per gram for a multiphase ceramics or glass-ceramics containing DOE high-level tank waste. Additionally shown is the cumulative number of α-decay events per gram for a ceramic containing 10 wt.% 239Pu. (b) Cumulative α-decay dose as a function of time for different 239Pu concentrations. Additionally, the dose-time ranges for available data from Pu/Cm containing glasses and ceramics, as well as for natural mineral analogues are also shown. (a) Reproduced with permission from Reference [29], Copyright © Springer Nature 2011; (b) Reproduced with permission from Reference [7], Copyright © Springer Nature 2010.
Figure 3. Cumulative Dose as a function of waste storage time. (a) β-decay and α-decay events per gram for a multiphase ceramics or glass-ceramics containing DOE high-level tank waste. Additionally shown is the cumulative number of α-decay events per gram for a ceramic containing 10 wt.% 239Pu. (b) Cumulative α-decay dose as a function of time for different 239Pu concentrations. Additionally, the dose-time ranges for available data from Pu/Cm containing glasses and ceramics, as well as for natural mineral analogues are also shown. (a) Reproduced with permission from Reference [29], Copyright © Springer Nature 2011; (b) Reproduced with permission from Reference [7], Copyright © Springer Nature 2010.
Crystals 13 00143 g003
Figure 4. Schematic diagram of the relationship between energy loss and bombarding particle energy.
Figure 4. Schematic diagram of the relationship between energy loss and bombarding particle energy.
Crystals 13 00143 g004
Figure 5. (a) High-resolution TEM images of individual tracks in Gd2Ti2O7 pyrochlore produced by 1.1 GeV Ru ions (21.7 keV/nm), 1.4 GeV Xe ions (28.5 keV/nm), 2.0 GeV Ta ions (37.0 keV/nm), and 2.2 GeV Au ions (40.1 keV/nm). The track diameters increase systematically with dE/dx as shown from left to right. (b) Schematic of track morphology showing the amorphous core (red) surrounded by a defect-fluorite structured shell (blue) embedded in the pyrochlore matrix. Fast Fourier transform (FFT) analysis of TEM images confirm the crystalline phase of matrix (pyrochlore, p) and track shells (defect-fluorite, df), displayed in (a) for the Ru ion track. The loss of the spots in the inner circles corresponds to the loss of the long-range ordered pyrochlore superstructure leading to the defect-fluorite structure. (c) The amorphous fraction within an ion track (normalized core-to-shell ratio) increases significantly as a function of dE/dx. The error bars represent the dE/dx variation over the entire sample thickness and the uncertainties of determining the area of amorphous core and entire track, respectively. The values in % represent the amorphous fraction, and the numbers in parenthesis are the total track diameters in nm (core + shell). The scale bar applies to all TEM images. Reproduced with permission from Reference [135], Copyright © Elsevier 2014.
Figure 5. (a) High-resolution TEM images of individual tracks in Gd2Ti2O7 pyrochlore produced by 1.1 GeV Ru ions (21.7 keV/nm), 1.4 GeV Xe ions (28.5 keV/nm), 2.0 GeV Ta ions (37.0 keV/nm), and 2.2 GeV Au ions (40.1 keV/nm). The track diameters increase systematically with dE/dx as shown from left to right. (b) Schematic of track morphology showing the amorphous core (red) surrounded by a defect-fluorite structured shell (blue) embedded in the pyrochlore matrix. Fast Fourier transform (FFT) analysis of TEM images confirm the crystalline phase of matrix (pyrochlore, p) and track shells (defect-fluorite, df), displayed in (a) for the Ru ion track. The loss of the spots in the inner circles corresponds to the loss of the long-range ordered pyrochlore superstructure leading to the defect-fluorite structure. (c) The amorphous fraction within an ion track (normalized core-to-shell ratio) increases significantly as a function of dE/dx. The error bars represent the dE/dx variation over the entire sample thickness and the uncertainties of determining the area of amorphous core and entire track, respectively. The values in % represent the amorphous fraction, and the numbers in parenthesis are the total track diameters in nm (core + shell). The scale bar applies to all TEM images. Reproduced with permission from Reference [135], Copyright © Elsevier 2014.
Crystals 13 00143 g005
Figure 6. Schematic diagram of decay series for (a) 90Sr, (b) 137Cs.
Figure 6. Schematic diagram of decay series for (a) 90Sr, (b) 137Cs.
Crystals 13 00143 g006
Figure 7. Amorphous fraction as a function of normalized ion fluence for direct-impact model, single-cascade-overlap, direct-impact/defect-stimulated (D-I/D-S), multiple step damage accumulation (MSDA), and heterogeneous track overlap model (HTOM). Reproduced with permission from Reference [142], Copyright © Elsevier 2012.
Figure 7. Amorphous fraction as a function of normalized ion fluence for direct-impact model, single-cascade-overlap, direct-impact/defect-stimulated (D-I/D-S), multiple step damage accumulation (MSDA), and heterogeneous track overlap model (HTOM). Reproduced with permission from Reference [142], Copyright © Elsevier 2012.
Crystals 13 00143 g007
Figure 8. (a) Electron-phonon mean free path (λ) deduced from thermal spike fits versus optical band gap energy for different insulators [174]. (b) Inelastic thermal spike calculations of the evolution of the energy density deposited in Gd2Ti2O7 vs. time for different diameters of the track cylinder (given in nm) for Au ion (7 MeV/u). Horizontal lines indicate the required energy density to obtain track diameters in agreement with experimental data from high-resolution TEM [135]. (c) Experimental track diameters for different ion beams in Gd2Ti2O7 from high-resolution TEM measurements as a function of dE/dx: core-shell tracks (red circles) and amorphous track cores (blue squares). The solid red and dashed black curves are the results of thermal spike calculations based on Et = 0.90 eV/at and Et = 0.56 eV/at, respectively [135]. (a) Reproduced with permission from Reference [174], Copyright © Springer Nature 2009; (b,c) Reproduced with permission from Reference [135], Copyright © Elsevier 2014.
Figure 8. (a) Electron-phonon mean free path (λ) deduced from thermal spike fits versus optical band gap energy for different insulators [174]. (b) Inelastic thermal spike calculations of the evolution of the energy density deposited in Gd2Ti2O7 vs. time for different diameters of the track cylinder (given in nm) for Au ion (7 MeV/u). Horizontal lines indicate the required energy density to obtain track diameters in agreement with experimental data from high-resolution TEM [135]. (c) Experimental track diameters for different ion beams in Gd2Ti2O7 from high-resolution TEM measurements as a function of dE/dx: core-shell tracks (red circles) and amorphous track cores (blue squares). The solid red and dashed black curves are the results of thermal spike calculations based on Et = 0.90 eV/at and Et = 0.56 eV/at, respectively [135]. (a) Reproduced with permission from Reference [174], Copyright © Springer Nature 2009; (b,c) Reproduced with permission from Reference [135], Copyright © Elsevier 2014.
Crystals 13 00143 g008
Figure 9. Accumulated damage (fd) in YSZ and MAS crystals irradiated at RT with low-energy (160 keV) Cs ions. Solid lines are fits to RBS/C data using the MSDA model with m = 3 (three-step process). The conversion factors of fluence (expressed in 1015 cm−2) to dpa scales are: 2.45 for MAS and 3.46 for YSZ crystals. Reproduced with permission from Reference [185], Copyright © Springer Nature 2009.
Figure 9. Accumulated damage (fd) in YSZ and MAS crystals irradiated at RT with low-energy (160 keV) Cs ions. Solid lines are fits to RBS/C data using the MSDA model with m = 3 (three-step process). The conversion factors of fluence (expressed in 1015 cm−2) to dpa scales are: 2.45 for MAS and 3.46 for YSZ crystals. Reproduced with permission from Reference [185], Copyright © Springer Nature 2009.
Crystals 13 00143 g009
Figure 10. Schematic representations of the physical processes underlying the HTOM model with the cross sections of cylinders corresponding to swift heavy ion tracks: (a), first impact, (b) second impact. σPA and σPF are the areas directly converted by each ion impact into amorphous or disordered regions, and σFA is the area related to the transformation of a defective fluorite region into an amorphous one owing to a second ion impact.
Figure 10. Schematic representations of the physical processes underlying the HTOM model with the cross sections of cylinders corresponding to swift heavy ion tracks: (a), first impact, (b) second impact. σPA and σPF are the areas directly converted by each ion impact into amorphous or disordered regions, and σFA is the area related to the transformation of a defective fluorite region into an amorphous one owing to a second ion impact.
Crystals 13 00143 g010
Figure 11. Variation of the fraction of (a) pyrochlore, (b) fluorite, and (c) amorphous phases vs the ion fluence for pyrochlores irradiated with 120 MeV U ions. Solid lines are fits to the data with HTOM (Equation (20a–c)). Reproduced with permission from Reference [142], Copyright © Elsevier 2012.
Figure 11. Variation of the fraction of (a) pyrochlore, (b) fluorite, and (c) amorphous phases vs the ion fluence for pyrochlores irradiated with 120 MeV U ions. Solid lines are fits to the data with HTOM (Equation (20a–c)). Reproduced with permission from Reference [142], Copyright © Elsevier 2012.
Crystals 13 00143 g011
Figure 12. Variation in of (a) cross-section σi, (b) σi/σT, (c) the total track diameter (dT), and the amorphous track core diameter (dA) as a function of the cation radius ratio for pyrochlore ceramics. The solid marks are Gd2Ti2O7, Gd2TiZrO7, Gd2Zr2O7, Sm2Zr2O7, Eu2Zr2O7, and Nd2Zr2O7 irradiated with 119 MeV U ions [61,62,142], while the hollow marks represent Nd2Zr2O7 and Y2Ti2O7 irradiated with 93 MeV Xe ions from Refs. [62,188]. The lines are guides for the eye.
Figure 12. Variation in of (a) cross-section σi, (b) σi/σT, (c) the total track diameter (dT), and the amorphous track core diameter (dA) as a function of the cation radius ratio for pyrochlore ceramics. The solid marks are Gd2Ti2O7, Gd2TiZrO7, Gd2Zr2O7, Sm2Zr2O7, Eu2Zr2O7, and Nd2Zr2O7 irradiated with 119 MeV U ions [61,62,142], while the hollow marks represent Nd2Zr2O7 and Y2Ti2O7 irradiated with 93 MeV Xe ions from Refs. [62,188]. The lines are guides for the eye.
Crystals 13 00143 g012
Figure 13. The critical amorphization temperature TC of pyrochlores irradiated by ions as a function of the cation ionic radius ratio rA/rB. The triangle, square, circle, and rhombus represent irradiated ions 1 MeV Kr, 1.5 MeV Xe, 0.6 MeV Ar, and 0.6 MeV Bi, respectively. Grey, red, green, purple, yellow, orange, and blue represent actinide doped pyrochlores [66], titanate pyrochlores (A2Ti2O7) [68,93,133,190,191,192,195], Zr-doped titanate pyrochlores (A2(Ti1−xZrx)2O7) [55,139], stannate pyrochlores (A2Sn2O7) [55,194], Sn-doped yttrium titanate pyrochlores (Y2(Ti1−xSnx)2O7) [55], hafnate pyrochlores (A2Hf2O7) [55], and lanthanide zirconate pyrochlores (La2Zr2O7) [91], respectively. Additionally, the triangles with black edges represent data [55] calculated from the model proposed by Lumpkin et al. (Equation (2) in Ref. [196]). The red solid line is obtained by linear fitting all data points irradiated by 1 MeV Kr ion, except the three points with black arrows (Y2TiO5, Yb2TiO5 and YbYTiO5) and the data marked by black-edged triangles.
Figure 13. The critical amorphization temperature TC of pyrochlores irradiated by ions as a function of the cation ionic radius ratio rA/rB. The triangle, square, circle, and rhombus represent irradiated ions 1 MeV Kr, 1.5 MeV Xe, 0.6 MeV Ar, and 0.6 MeV Bi, respectively. Grey, red, green, purple, yellow, orange, and blue represent actinide doped pyrochlores [66], titanate pyrochlores (A2Ti2O7) [68,93,133,190,191,192,195], Zr-doped titanate pyrochlores (A2(Ti1−xZrx)2O7) [55,139], stannate pyrochlores (A2Sn2O7) [55,194], Sn-doped yttrium titanate pyrochlores (Y2(Ti1−xSnx)2O7) [55], hafnate pyrochlores (A2Hf2O7) [55], and lanthanide zirconate pyrochlores (La2Zr2O7) [91], respectively. Additionally, the triangles with black edges represent data [55] calculated from the model proposed by Lumpkin et al. (Equation (2) in Ref. [196]). The red solid line is obtained by linear fitting all data points irradiated by 1 MeV Kr ion, except the three points with black arrows (Y2TiO5, Yb2TiO5 and YbYTiO5) and the data marked by black-edged triangles.
Crystals 13 00143 g013
Table 1. Ion substitute eligibility for the various sites of the pyrochlore supergroup [6,34,35,36].
Table 1. Ion substitute eligibility for the various sites of the pyrochlore supergroup [6,34,35,36].
SiteIon
ARare-earth, U, Th, Na, Ca, Ag, Mn, Sr, Ba, Fe2+, Sn2+, Sb3+, , …
BTi, Sn4+, Zr, Hf, Ce, Ta, Nb, Sb5+, Mo, Cr, U, Pu, W, V, Fe3+, Mg, Al, Si, …
XO, OH, F
YO, F, OH, □, H2O, K, Cs, Rb, …
Note: □ denotes a site vacancy. Bold indicates the most common substitutes.
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, Y.; Jing, C.; Ding, Z.-Y.; Zhang, Y.-Z.; Wei, T.; Ouyang, J.-H.; Liu, Z.-G.; Wang, Y.-J.; Wang, Y.-M. The Structure, Property, and Ion Irradiation Effects of Pyrochlores: A Comprehensive Review. Crystals 2023, 13, 143. https://doi.org/10.3390/cryst13010143

AMA Style

Wang Y, Jing C, Ding Z-Y, Zhang Y-Z, Wei T, Ouyang J-H, Liu Z-G, Wang Y-J, Wang Y-M. The Structure, Property, and Ion Irradiation Effects of Pyrochlores: A Comprehensive Review. Crystals. 2023; 13(1):143. https://doi.org/10.3390/cryst13010143

Chicago/Turabian Style

Wang, Yuhao, Chong Jing, Zhao-Ying Ding, Yun-Zhuo Zhang, Tao Wei, Jia-Hu Ouyang, Zhan-Guo Liu, Yu-Jin Wang, and Ya-Ming Wang. 2023. "The Structure, Property, and Ion Irradiation Effects of Pyrochlores: A Comprehensive Review" Crystals 13, no. 1: 143. https://doi.org/10.3390/cryst13010143

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