Downloaded from http://rspa.royalsocietypublishing.org/ on June 18, 2017 rspa.royalsocietypublishing.org Research Cite this article: Das NK, Rigby K, de Leeuw NH. 2014 The effect of helium nano-bubbles on the structures stability and electronic properties of palladium tritides: a density functional theory study. Proc. R. Soc. A 470: 20140357. http://dx.doi.org/10.1098/rspa.2014.0357 Received: 2 May 2014 Accepted: 19 August 2014 Subject Areas: computational chemistry Keywords: density functional theory simulations, Pd defects, helium cluster, diffusion, migration Author for correspondence: N. K. Das e-mail: [email protected] The effect of helium nano-bubbles on the structures stability and electronic properties of palladium tritides: a density functional theory study N. K. Das, K. Rigby and N. H. de Leeuw Department of Chemistry, University College London, 20 Gordon St., London WC1H 0AJ, UK Density functional theory calculations have been used to study the incorporation of helium in perfect and defect-containing palladium tritides, where we have calculated the energetics of incorporation and the migration behaviour. Helium atoms preferably occupy the octahedral interstitial and substitutional sites in the perfect and Pd vacancy-containing tritides, respectively. The energetics reveal that helium clusters can form in the lattice, which displace the Pd metal atoms. The defective lattice shows less expansion compared with the perfect lattice, which can accommodate the helium less easily. The path from octahedral–tetrahedral–octahedral sites is the lowest energy pathway for helium diffusion, and the energetics indicate that the helium generated from tritium decay can accumulate in or near the octahedral sites. Density of states analyses shows the hybridization between palladium d and tritium s orbitals and repulsion between palladium d and helium s orbitals, which can distort the lattice as a result of generating localized stress. 1. Introduction One contribution to a Special feature ‘New developments in the chemistry and physics of defects in solids’. Tritium is important in the nuclear industry and its storage has been suggested in the form of metal tritides [1]. Tritium is radioactive, has a half-life of about 12.3 years and decays to 3 He. During tritium decay, helium atoms agglomerate in metals owing to their extremely low solubility [2]. The tritium decay products do not immediately diffuse out of the lattice, but they can 2014 The Author(s) Published by the Royal Society. All rights reserved. Downloaded from http://rspa.royalsocietypublishing.org/ on June 18, 2017 DFT calculations were performed using the Vienna ab initio simulation package [20,21]. The generalized gradient approximation (GGA) and Perdew–Burke–Ernzerhof (PBE) pseudopotentials [22] were used to describe the electronic structure of palladium, hydrogen and helium atoms, with the palladium 4d and 4s electrons treated as valence electrons. The electronic structures of hydrogen and tritium are almost the same, and therefore it is reasonable to use hydrogen to simulate the electronic behaviour of tritium. Standard hydrogen and helium PBE potentials were used, with the hydrogen and helium 1s electrons treated as valence electrons and the hydrogen atom was manually modified to have an atomic mass of 3 instead of 1. The number of Monkhorst–Pack k-points was determined by performing individual DFT calculations on a PdT system with an increasing number of k-points. An appropriate number of k-points was deemed to have been reached when a total energy convergence of 1 meV per atom was observed, and as result a 6 × 6 × 6 k-point grid was used for 2 × 2 × 2 supercell calculations. The calculations ................................................... 2. Computational details 2 rspa.royalsocietypublishing.org Proc. R. Soc. A 470: 20140357 be strongly trapped by interstitial sites and defects containing a free volume, particularly vacancies and grain boundaries, where they can subsequently form helium clusters or nanobubbles [3–5]. The presence of helium nano-bubbles plays an important role in the microstructural evolution of storage materials, and the bubbles can significantly degrade the mechanical properties of materials. The behaviour of helium in metal and metal tritides has been investigated using experimental approaches and theoretical methods. For example, a nuclear magnetic resonance study has suggested 3 He bubble formation in LiT, TiT1.9 and UT3 and has further indicated an increase in the bubble sizes as the metal tritides age [6]. Weaver et al. [7] have shown that 3 He atoms were initially trapped interstitially in titanium tritide, whereas older samples were found to contain helium gas bubbles. The stability of metal tritides decreases with the increased generation of helium and in titanium tritide, helium was shown to migrate between two neighbour interstitial sites [8]. The effects of helium bubbles on metal tritides are clearly of considerable interest and it is important to gain understanding of the atomic-level interactions of helium atoms and clusters with metal tritides. Palladium is an outstanding tritium storage material, which is of particular interest owing to its high resistance to oxidation and poisoning and its ability to easily absorb, as well as release, tritium and retain the helium generated in the matrix [9]. Several experimental studies have been carried out on the helium behaviour in palladium tritide [10–12]. A transmission electron microscopy study found bubble formation in an aged Pd-T. However, the study revealed that grain boundaries and other extended defects do not appear to play a major role in bubble formation [10]. Other studies have focused on helium release from aged palladium tritide, with X-ray diffraction analysis results showing that palladium tritides retain more than 95% of helium at room temperature for at least 9 years of ageing [13]. Additionally, density functional theory (DFT) and molecular dynamics methods have been used to investigate the effect of helium and helium clusters in Pd, Ti, W, Fe and Mo [14–19], which provide an atomistic understanding of helium in these metals. Zeng et al. [14] have studied the properties of small helium-vacancy clusters in palladium and they reported that vacancies in Pd are the preferred sites to capture helium atoms. The formation energies of helium vacancies in metals have also been calculated using molecular dynamics and semi-empirical methods [15], but, to date, no computational studies have been carried out on helium in tritides. A suitable tritium storage material would need to have a large capacity for helium and the ability to expand and withstand mechanical stress, while still retaining its overall macrostructure. As the interaction between helium clusters and palladium tritides plays an important role in the material’s integrity, in this study, we have employed DFT methods to investigate the structural and electronic properties of helium clusters in perfect and defect-containing palladium tritide lattices and calculate the helium migration energies. Downloaded from http://rspa.royalsocietypublishing.org/ on June 18, 2017 3 Figure 1. The PdT0.22 structure. The dark blue and white spheres indicate the palladium and tritium atoms, respectively. (Online version in colour.) were performed using a high planewave cut-off energy of 660 eV. Total energy convergence was achieved for each calculation when all forces on the system were lower than 0.001 eV per Å and the atomic positions, shape and volume of the supercells were optimized. The tritium atoms were placed in interstitial sites, and the most stable structure was selected for further calculations. It has been observed experimentally that 60% of the octahedral interstitial sites in palladium metal can accommodate tritium atoms and still retain the metal’s fcc crystal structure [10]. However, here, we have considered a relatively low concentration of tritium to retain more vacant sites for helium incorporation. Seven tritium atoms were placed in octahedral interstitial sites of the 2 × 2 × 2 palladium supercell giving a composition of PdT0.22 , as shown in figure 1. Although this is not a realistic concentration at room temperature, where the material would phase separate into low concentration α and high concentration β domains, the simulation cell is convenient to study explicitly all interactions that we are interested in, not only between helium and the tritium atoms, but also between tritium atoms themselves, and it provides a number of sites for helium incorporation and diffusion. The unrealistic tritium concentration overall is, in reality, an artefact of the periodic boundary conditions, which repeats this very local concentration throughout the lattice, although the cell is big enough that there are no unrealistic interactions between the mirror images. For the defect-containing lattice, a Pd vacancy was created through the removal of a Pd atom in the perfect lattice. For the calculation of the interaction between helium clusters and Pd vacancies, helium atoms were added to the interstitial and vacant sites one by one. The formation energy of a helium cluster composed of n helium atoms in an interstitial site and in a Pd vacancy are given, respectively, by Ef (Hen ) = Etot (PdT0.22 Hen ) − nEHe − Etot (PdT0.22 ) (2.1) Ef (Hen V) = Etot (PdT0.22 Hen V) − nEHe − E(PdT0.22 V), (2.2) and where Etot (PdT0.22 Hen ) is the total energy of a system containing a Hen cluster, Etot (PdT0.22 Hen V) is the total energy of a system containing a Hen and Pd vacancy, Etot (PdT0.22 ) is the energy of the perfect tritide, Etot (PdT0.22 V) is the energy of a Pd vacancy-containing system without helium and EHe is the energy of an isolated helium atom. The binding energy of a helium atom to Hen clusters is calculated for the perfect and defective PdT0.22 lattices as follows Eb (He) = Ef (He) + Ef (Hen−1 ) − Ef (Hen ) (2.3) Eb (He) = Ef (He) + Ef (Hen−1 V) − Ef (Hen V), (2.4) and which means that a positive binding energy is an energetically preferred process. ................................................... X Y rspa.royalsocietypublishing.org Proc. R. Soc. A 470: 20140357 Z Downloaded from http://rspa.royalsocietypublishing.org/ on June 18, 2017 (a) Bulk properties The calculated lattice constant, cohesive energy and bulk modulus of the Pd bulk are 3.96 Å, 3.76 eV and 190 GPa, respectively, in good agreement with previously obtained results [14] and experimental values of 3.89 Å, 3.89 eV and 180 GPa [24], although the GGA–PBE functional used leads to a small overestimation of the lattice constant and a consequent underestimation of the cohesive energy. The Pd bulk lattice was affected by the insertion of helium and tritium atoms into the interstitial sites. Insertion of a single tritium atom in the octahedral site resulted in a small (less than 0.5%) expansion of the palladium lattice volume, whereas a helium atom expanded the structure by about 1.5%. When both the tritium and helium were present in two different octahedral sites, the palladium lattice expanded by 2.5%. Increasing the tritium concentration in the interstitial sites naturally increased the size of the palladium lattice. For example, the PdT0.22 structure showed a 3.5% lattice volume expansion compared with Pd metal, which is consistent with a previous study [25]. The palladium tritide lattice was further expanded by 5.7% compared with the pure Pd metal owing to insertion of one helium atom in the PdT0.22 lattice. (b) Tritium distribution in Pd lattice When hydrogen is occluded in palladium, initially, the α phase is formed, which converts to the β phase at high hydrogen concentrations. Both are fcc structures, but a volume expansion occurs on going from the α to the β phase [26]. We have first studied the distribution of tritium in the bulk palladium structure, where a 2 × 2 × 2 supercell of PdT0.6875 was modelled, containing 32 palladium atoms and 22 tritium atoms. PdT0.6875 was chosen for this part of our study as it has a similar molecular formula to the experimental β-phase of palladium tritide (PdT0.7 ). The PdT0.6875 composition gives rise to 53 unique configurations of the Pd and T in the 2 × 2 × 2 cell, where we have calculated the total energy for each configuration. It was found that each configuration had a similar total energy, where the largest energy difference between the highest and lowest energy configurations was only 0.05 eV. As such, it can be assumed that the probability of each configuration existing experimentally is relatively equal. This also supports the theory that tritium atoms diffuse quickly through palladium and do not remain in the same octahedral interstitial site for an extended time. However, it was found that the configurations with the lowest energies had octahedral interstitial vacancies evenly distributed throughout the material, whereas, in the highest energy configurations, the octahedral interstitial vacancies were clustered together. This suggests that configurations where vacancies are spread across the material are most stable. A converse system can also be considered where there are few tritium atoms and many octahedral interstitial vacancies, such as in PdT0.22 . Although for tritium concentrations up to x = 0.6, there is a phase miscibility gap at equilibrium between the α- and β-phases, it has been suggested that at low temperatures (less than 50 K) an ordered α phase can exist [27], whereas the β-phase can be maintained in hydride samples through a continuous concentration gradient from 0.03 to 0.60 [26], whereas materials with a tritium concentration of x = 0.22 might also be expected to occur in small regions at interfaces between α- and β-phases within the miscibility gap. As already outlined in §2, we have chosen to study the PdT0.22 composition as a model for ................................................... 3. Results and discussion 4 rspa.royalsocietypublishing.org Proc. R. Soc. A 470: 20140357 The dimer method was used to investigate the migration mechanism and the energy barriers of helium diffusion in the lattice, which calculations were performed in a supercell with and without a fixed volume condition for comparison [23]. The dimer method was developed to identify saddle points in the potential energy surface within a solid, without knowledge of the final state of transition, and without the use of second derivatives of the potential. The final converged transition state in each case was verified to have only one imaginary frequency. Downloaded from http://rspa.royalsocietypublishing.org/ on June 18, 2017 (a) (b) 11 3.0 2.8 Ef /n (eV) 7 6 2.6 2.4 2.2 5 2.0 4 1.8 3 1.6 1 2 3 4 no. He atoms (n) 5 6 1 2 3 4 no. He atoms (n) 5 6 Figure 2. Formation energy of small helium clusters versus number of helium atoms in PdT0.22 (a) the total formation energy and (b) average formation energy per helium atom. (Online version in colour.) the system with a relatively low concentration of tritium in the lattice, which allows for explicit consideration of T–T as well as He–He and He–T interactions. Again, we have studied all the unique configurations of PdT0.22 , where the largest energy difference between the highest and lowest energy configurations was 0.04 eV. The most stable configurations of PdT0.22 have tritium atoms evenly distributed throughout the material and this structure was therefore selected for further study of helium incorporation and diffusion. (c) Clustering of helium atoms in perfect PdT0.22 Initially, we considered tetrahedral and octahedral interstitial sites for the incorporation of a single helium atom in both fcc Pd and PdT0.22 systems corresponding to a helium concentration of 0.03 per Pd, which we henceforth call PdHe0.03 and PdT0.22 He0.03 , respectively. In both systems, the octahedral interstitial site is more stable than the tetrahedral sites for helium insertion, although the energy differences between octahedral and tetrahedral sites are only 0.12 and 0.16 eV for Pd and PdT0.22 , respectively. Considering the favourable octahedral interstitial site, the stabilities of helium clusters in PdT0.22 are investigated, with a variation of cluster size from one to six atoms, corresponding to helium concentrations of 0.03–0.19 per Pd, respectively. We next inserted helium atoms one by one into the octahedral site of the PdT0.22 system and the energetics of the formation of these interstitial helium clusters into the system are shown in figure 2a,b. The formation energy per helium atom is decreasing as the number of helium atoms increases, indicating that a helium atom in an octahedral site can attract more helium atoms and form a cluster of at least up to six helium atoms, as calculated here. The He2 forms a pair in the octahedral site with a He–He distance of 1.73 Å, which is very close to the value of 1.704 Å in the hcp Ti structure [16]. For He3 , the most stable structure formed is a scalene triangle in the octahedral site with He–He distances of 1.87, 1.88 and 1.886 Å. The ground-state structure of the He4 cluster was a planar rhombus with an average He–He distance of 1.90 Å, whereas the He5 cluster was found to be a trigonal bipyramidal structure. In the case of He6 , the lowest energy structure was the octahedron structure indicating the onset of three-dimensional structures. The average He–He distance again was 1.90 Å. The average He–He distances of the helium clusters therefore vary between 1.73 and 1.90 Å, i.e. significantly less than that of gas phase helium at 4.50 Å. The bigger clusters displayed longer He–He distances, acquiring more space in the interstitial site. Large clusters could therefore generate stress in the tritide lattice. The helium binding energies in PdT0.22 have also been calculated, i.e. the energy required to bind a helium to a cluster. The binding energy for the smallest cluster of He2 is 0.58 eV and that energy is increasing as more helium atoms are added to the cluster, suggesting that an interstitial ................................................... 9 rspa.royalsocietypublishing.org Proc. R. Soc. A 470: 20140357 3.2 8 Ef (eV) 5 3.4 10 Downloaded from http://rspa.royalsocietypublishing.org/ on June 18, 2017 It is important to understand the migration behaviour of helium in palladium tritide to assess the mechanism of helium bubble formation. The diffusion pathways are not intuitive, but we have investigated all major migration pathways for helium diffusion, that can reasonably be expected in the PdT0.22 lattice, and calculated the lowest energy paths with respect to the tritium positions in the system. In order to evaluate if the size of the simulation cell could affect the diffusion activation energies, we have first calculated the cohesive energies and energies for helium incorporation in the octahedral and tetrahedral interstitial sites in both (2 × 2 × 2) and (3 × 3 × 3) simulation supercells. However, these energies do not deviate significantly and as such the activation energies for helium diffusion are unlikely to be affected by cell size, and we have therefore carried out all diffusion calculations in a (2 × 2 × 2) cell. From an octahedral site, the helium can migrate to another available octahedral site following different paths, such as directly from octahedral to octahedral site (parallel and perpendicular) or octahedral–tetrahedral– octahedral (O–T–O), as shown in figure 3. The migration energy of helium for the path O–T–O was lower than that of O1–O2 and O1–O3 diffusion. Previous studies have shown that the O–T–O path is also the lowest energy path for helium diffusion in Al and Pd metals [14]. The migration of an interstitial helium via O–T–O was determined to require about 0.48 eV to overcome the energy barrier, with the T site being 0.44 eV higher in energy than the O sites. Figure 3, on the other hand, shows that the diffusion along path O1–O2 for an interstitial helium requires an energy of 1.05 eV, and 1.14 eV for path O1–O3, both of which are higher than the O–T–O migration path, and thus would only be likely to occur at high temperatures. A previous experimental study found a value of 0.35 eV for the helium activation energy in Ni metal, whereas a hybrid molecular dynamics method obtained a value of 0.43 eV for the same system [31,32]. Moreover, the ErT system found a value of 0.74 eV experimentally [33], which is larger than that of the present study on PdT0.22 . Helium bubble growth rate is slow when activation energies are high, and helium bubble formation is thus more difficult in PdT0.22 compared with the pure Ni metal. However, this study shows that an interstitial helium in PdT0.22 is less likely to jump directly from one O site to another without passing through an intermediate tetrahedral site, owing to the much ................................................... (d) Helium diffusion in PdT0.22 6 rspa.royalsocietypublishing.org Proc. R. Soc. A 470: 20140357 helium atom in PdT0.22 could be easily trapped by another. The largest binding energy in PdT0.22 was obtained for helium addition to a He4 cluster at 2.78 eV but this decreased to 1.93 eV for the addition to the He5 cluster to make a He6 cluster. Previous DFT and experimental studies have found that the octahedral interstitial site is more stable than the tetrahedral interstitial site in the Pd lattice [14,28,29], suggesting that the helium from tritium decay would preferably occupy octahedral interstitial sites. Zeng et al. [14] have shown that the palladium metal can accommodate up to four helium atoms in an interstitial site, but there are no reports about the interaction between helium clusters and palladium tritides. The calculated formation energy in the PdT0.22 system is gradually increasing with increasing cluster size, whereas the formation energy per helium atom is decreasing as the number of helium atoms increases, consistent with previous studies in Pd and Ti metal [14,16]. The values for the PdT0.22 system are lower than those for helium insertion in palladium metal, which are reported at 3.70– 11 eV for He1 –He4 clusters [14], and the presence of tritium in the octahedral sites thus clearly affects the formation energies. When we compare helium binding energies with the previous studies, we see that tritium occlusion in palladium has an effect on helium cluster formation. Helium binding energies for the PdT0.22 system are very close to a number of metals, e.g. 0.65 eV for Pd, 0.66 eV for Ti and 0.43 eV for Fe [4,14,16]. The helium binding energy in nickel, on the other hand, is found to be near constant at about 2 eV for He6 and large clusters [30] and in an fcc Pd lattice, a value of 1.33 eV was found for a He4 cluster [14]. Such differences in binding energies are indicative of the effect of interstitial tritium. Therefore, helium from tritium decay can accumulate in an interstitial site to form a bubble in the bulk PdT0.22 lattice. The binding energies reveal that the cluster is formed preferentially over dispersed atoms and the helium would be trapped until the system was exposed to very high temperatures [30]. Downloaded from http://rspa.royalsocietypublishing.org/ on June 18, 2017 (a) 7 (ii) O1 O3 (iii) (b) 1.2 1.0 0.8 0.6 0.4 0.2 0 O1–O2 O1–T–O2 O1–O3 Figure 3. (a) Illustration of helium migration paths. The dashed teal line indicates direct octahedral to octahedral, the red line shows octahedral to octahedral via tetrahedral site and the dashed green line indicates helium motion from octahedral to octahedral between two layers. The dark blue atoms are palladium, tritium is not shown. The red and green spheres indicate the octahedral and tetrahedral interstitial sites, respectively, and (b) schematic illustration of the helium activation energy. (Online version in colour.) higher migration energy. Moreover, the migration of helium in the tritides will also depend on the amount of tritium decay, because the number and distribution of octahedral vacancies will affect the diffusion. (e) Clustering of helium atoms in a defect-containing lattice The calculated monovacancy formation energy in a (2 × 2 × 2) bulk cell of Pd metal was 1.26 eV, which is 0.28 eV lower than that found by thermal diffusivity experiments [34]. Tritium atoms were again added in the octahedral interstitial sites of the vacancy-containing Pd lattice and the structure optimized. In the vacancy-containing PdT0.22 lattice, we placed the helium atom in all possible octahedral and tetrahedral interstitial sites as well as the vacancy site, in effect replacing ................................................... (i) rspa.royalsocietypublishing.org Proc. R. Soc. A 470: 20140357 T O2 O1 Downloaded from http://rspa.royalsocietypublishing.org/ on June 18, 2017 (a) (b) 11 10 8 2.8 Ef /n (eV) Ef (eV) 7 6 2.2 2.0 5 1.8 4 1.6 3 1 2 3 4 no. He atoms (n) 5 6 1 2 3 4 no. He atoms (n) 5 6 Figure 4. Formation energy of small He-vacancy clusters versus number of helium atoms (a) the total formation energy and (b) average formation energy per helium atom. (Online version in colour.) 20 change of lattice volume (%) 18 16 PdT0.22 perfect lattice PdT0.22 lattice with a Pd defect 14 12 10 8 6 4 2 0 no. He atoms Figure 5. The change in local PdT0.22 lattice volume caused by the presence of helium clusters. (Online version in colour.) a Pd by a helium atom, to calculate the lowest energy site. The preferred site for one helium atom was in the vacancy site, where the energy of the helium incorporation was 2.30 eV lower than in the octahedral interstitial site. A molecular dynamics study found that the formation energy of helium in a vacancy site in fcc Ni was 2.70 eV lower than that of the interstitial site [30], in accordance with our PdT0.22 study. Thus, the interstitial helium will be trapped by the vacancy. The addition of helium atoms one by one into the vacant site shows that the cluster formation energy is again gradually increasing with increasing cluster size, as shown in figure 4a, but all values are lower than in the perfect PdT0.22 He0.13 lattice. The cluster formation energy per atom was gradually decreasing with increasing cluster size up to our calculated He6 cluster. The perfect PdT0.22 lattice showed the same trend, predicting the clustering of a number of helium atoms could be trapped. The helium binding trend indicates that one monovacancy can accommodate at least six helium atoms and that Pd vacancies could therefore be nucleation sites for helium bubble formation. The helium cluster geometries are very similar to those in the perfect PdT0.22 lattice, but the He–He distances are different, i.e. 1.62 Å for the dimer with the average He–He distances varying from 1.70 to 1.82 Å for the clusters He3 –He6 . With less than seven helium atoms in the monovacancy, He–He distances in vanadium are in the range of 1.75–1.85 Å, suggesting that the ................................................... 2.4 8 rspa.royalsocietypublishing.org Proc. R. Soc. A 470: 20140357 2.6 9 Downloaded from http://rspa.royalsocietypublishing.org/ on June 18, 2017 (a) 140 DOS (states eV–1) ................................................... 100 s of Pd rspa.royalsocietypublishing.org Proc. R. Soc. A 470: 20140357 120 9 total p of Pd d of Pd 80 60 40 20 0 –10.0 –7.5 –5.0 –2.5 energy (eV) 0 2.5 20.0 total s of Pd p of Pd d of Pd s of T DOS (states eV–1) 17.5 (b) 140 total 120 DOS (states eV–1) 15.0 12.5 enlarged view 10.0 7.5 5.0 2.5 s of Pd 0 –10.0 p of Pd 100 5.0 d of Pd –7.5 –5.0 –2.5 0 energy (eV) 2.5 5.0 s of T 80 60 40 20 0 –10.0 –7.5 –5.0 –2.5 0 2.5 5.0 energy (eV) Figure 6. The calculated total density of states of (a) bulk palladium, (b) PdT0.22 , (c) PdT0.22 He0.03 and (d) PdT0.22 He0.19 . (Online version in colour.) interstitial clusters are in a compressed state in the cluster compared with a free cluster [35]. Helium clusters in the vacancy significantly displace the metal atoms away from their original positions and the lattice expands as a result. (f) The effect of helium on the structural parameters of palladium tritides Interstitial helium atoms have a significant effect on the palladium tritide structures. If we compare the PdT0.22 with PdT0.22 He0.03 , incorporation of a single helium atom leads to an increase in the lattice vector by about 1.9%. Increasing the number of helium atoms in the PdT0.22 lattice gradually expands the structure, as shown in figure 5. It is interesting to note that expansion of the perfect PdT0.22 lattice is very small from He5 to He6 clusters, perhaps owing to the structure of the Downloaded from http://rspa.royalsocietypublishing.org/ on June 18, 2017 (c) 10 20.0 DOS (states eV–1) 100 80 12.5 enlarged view 10.0 7.5 5.0 140 120 15.0 total s of Pd p of Pd d of Pd s of T s of He 2.5 0 –10.0 –7.5 –5.0 –2.5 energy (eV) 0 2.5 5.0 60 40 20 0 –10.0 –7.5 –5.0 –2.5 energy (eV) 0 2.5 5.0 (d) 20.0 DOS (states eV–1) 17.5 DOS (states eV–1) 100 80 12.5 enlarged view 10.0 7.5 5.0 140 120 15.0 total s of Pd p of Pd d of Pd s of T s of He total s of Pd p of Pd d of Pd s of T s of He 2.5 0 –10.0 –7.5 –5.0 –2.5 energy (eV) 0 2.5 5.0 60 40 20 0 –10.0 –7.5 –5.0 –2.5 energy (eV) 0 2.5 5.0 Figure 6. (Continued.) helium clusters. The structure of the tetramer is two-dimensional, but the pentamer and hexamer clusters are three-dimensional. This result is qualitatively consistent with X-ray studies, where it was found that after several months of ageing the metal tritide was distorted in an extremely inhomogeneous fashion to the point that a unique lattice parameter could not be defined [36]. ................................................... DOS (states eV–1) 17.5 rspa.royalsocietypublishing.org Proc. R. Soc. A 470: 20140357 total s of Pd p of Pd d of Pd s of T s of He Downloaded from http://rspa.royalsocietypublishing.org/ on June 18, 2017 To gain more insight into the bonding and electronic structure of the systems, we have plotted the density of states (DOS) of Pd, PdT0.22 , PdT0.22 He0.03 and PdT0.22 He0.19 , as shown in figure 6. Figure 6a clearly shows that Pd d orbitals are highly localized at the Fermi level (EF) and the total orbitals spread up to 7.0 eV below the Fermi energy. The PdT0.22 DOS in figure 6b shows a new peak at around −7.5 eV, which indicates that the tritium s orbital is hybridizing with the Pd d and s orbitals. Previously, Gupta & Gupta [39] stated that a peak below approximately −5.0 eV results from the bonding Pd-d–T-s interactions. The insertion of one helium in an octahedral site led to only a very small change in the electronic properties, as shown in figure 6c. With an increasing number of helium atoms in the octahedral site, a significant contribution from the helium s orbitals has modified the peak from −6.25 to −7.5 eV (figure 6b). This peak is mainly owing to the admixture of helium s and Pd d orbitals with some contribution from tritium s orbitals. The Pd d states have been pushed upward through an interaction with helium atoms, indicating the repulsive character of the atom in the matrix [25]. 4. Conclusion This study shows that helium from tritium decay can become trapped in interstitial sites and form clusters in the bulk PdT0.22 lattice. When Pd vacancies are present in the lattice, helium atoms will accumulate preferentially in the vacancy site. The calculated energetics show that clusters are thermodynamically preferred over helium dispersed in the lattice, where it should be retained until the system is exposed to high temperatures. The activation energy of migration of an interstitial helium is 0.48 eV for an O–T–O path, which is lower than the O–O paths. Of course, the migration of helium in the metal tritides may have to follow a path between octahedral sites in different layers, depending on the distribution of octahedral vacancies. The average He–He distances in the helium clusters vary from 1.62 to 1.90 Å, significantly less than that of gas phase helium. Even so, increasing cluster sizes require more space in the interstitial site and the process leads to an expansion of the PdT0.22 lattice, which the DOS analysis shows may be due to the repulsion between helium s and Pd d orbitals. Thus, the accumulation of helium atoms in the interstitial site in the lattice induces local stress around the cluster, as shown by the calculated bulk moduli for the pure and He-containing systems, which may cause material failure. ................................................... (g) Electronic structures 11 rspa.royalsocietypublishing.org Proc. R. Soc. A 470: 20140357 The accumulation of helium atoms in a vacancy showed the same trend but the percentage of lattice expansion of the defective PdT0.22 was much smaller than that of the perfect lattice. The vacant site has extra space that can accommodate more helium before significant expansion of the lattice becomes necessary. An ab initio study of the effect of helium on the peak tensile strength of tungsten metal showed that the presence of interstitial helium led to the highest increase in the tensile strength [37], which indicates that interstitial helium atoms are more harmful than substitutional helium, which is consistent with our observation. The optimized PdT0.22 and PdT0.22 Hen configurations were used to calculate the bulk moduli, which is a material property that relates the change in volume with a change in pressure. The bulk moduli are calculated to be 150 and 129.2 GPa for the PdT0.22 and PdT0.22 He0.03 systems, respectively, whereas further increases in helium cluster size in the PdT0.22 system lead to a gradual increase of the bulk modulus, for example, the bulk modulus varies from 180 to 257 GPa for the PdT0.22 He0.06 to PdT0.22 He0.19 systems, respectively. A previous study has shown that the bulk modulus for some compounds in inversely proportional to the lattice volume and proportional to the lattice pressure [38]. However, here we observe that the accumulation of helium atoms in an interstitial site increases the lattice volume, as well as the bulk modulus of PdT0.22 structures. This means that the larger helium clusters create higher pressures in the PdT0.22 systems as a result of more deformation of the crystal structures, which may eventually lead to microstructural failure. Downloaded from http://rspa.royalsocietypublishing.org/ on June 18, 2017 Acknowledgements. This work made use of the facilities of HECToR, the UK’s national high-performance 1. Lässer R. 1989 Tritium and helium-3 in metals. Berlin, Germany: Springer. 2. Rothaut J, Schroeder H, Ullmaier H. 1983 The growth of helium bubbles in stainless steel at high temperatures. Phil. Mag. A 47, 781–795. (doi:10.1080/01418618308245265) 3. Gao N, Victoria M, Chen J, Van Swygenhoven H. 2011 Helium-vacancy cluster in a single bcc iron crystal lattice. J. Phys.: Condensed Matter 23, 245403. (doi:10.1088/0953-8984/ 23/24/245403) 4. Fu CC, Willaime F. 2005 Ab initio study of helium in α-Fe: dissolution, migration, and clustering with vacancies. Phys. Rev. B 72, 064117. (doi:10.1103/PhysRevB.72.064117) 5. Singh BN, Leffers T, Green WV, Victoria M. 1984 Nucleation of helium bubbles on dislocations, dislocation networks and dislocations in grain boundaries during 600 MeV proton irradiation of aluminium. J. Nucl. Mater. 125, 287–297. (doi:10.1016/0022-3115(84)90556-7) 6. Brown Jr RC. 1978 Distribution of helium in metal tritides. Nature 271, 531–532. (doi:10.1038/271531a0) 7. Weaver HT, Camp WJ. 1975 Detrapping of interstitial helium in metal tritides-NMR studies. Phys. Rev. B 12, 3054. (doi:10.1103/PhysRevB.12.3054) 8. Liang JH, Dai YY, Yang L, Peng SM, Fan KM, Long XG, Zhou XS, Zu XT, Gao F. 2013 Ab initio study of helium behavior in titanium tritides. Comput. Mater. Sci. 69, 107–112. (doi:10.1016/j.commatsci.2012.11.033) 9. Emig JA, Garza RG, Christensen LD, Coronado PR, Souers PC. 1992 Helium release from 19-year-old palladium tritide. J. Nucl. Mater. 187, 209–214. (doi:10.1016/0022-3115(92)90499-B) 10. Thomas GJ, Mintz JM. 1983 Helium bubbles in palladium tritide. J. Nucl. Mater. 116, 336–338. (doi:10.1016/0022-3115(83)90124-1) 11. Shanahan KL, Hölder JS, Bell DR, Wermer JR. 2003 Tritium aging effects in LaNi4.25 Al0.75 . J. Alloys Comp. 356–357, 382–385. (doi:10.1016/S0925-8388(03)00139-7) 12. Thiébaut S, Demoment J, Limacher B, Paul-Boncour V, Décamps B, Percheron-Guégan A, Prem M, Krexner G. 2003 Aging effects on palladium pressure-composition isotherms during tritium storage. J. Alloys Comp. 356–357, 36–40. (doi:10.1016/S0925-8388(03)00097-5) 13. Thiébaut S, Douilly M, Contreras S, Limacher B, Paul-Boncour V, Décamps B, PercheronGuégan A. 2007 3 He retention in LaNi5 and Pd tritides: dependence on stoichiometry, 3 He distribution and aging effects. J. Alloys Comp. 446–447, 660–669. (doi:10.1016/j.jallcom. 2007.01.041) 14. Zeng X, Deng H, Hu W. 2009 First-principles approach to the properties of point defects and small helium-vacancy clusters in palladium. Nucl. Instrum. Methods Phys. Res. B 267, 3037– 3040. (doi:10.1016/j.nimb.2009.06.050) 15. Yang J, Ao B, Hu W, Wang X. 2006 The formation energies and binding energies of helium vacancy cluster: comparative study in Ni and Pd. J. Phys. Conf. Ser. 29, 190–193. (doi:10.1088/1742-6596/29/1/037) 16. Wang Y-L, Liu S, Rong L-J, Wang Y-M. 2010 Atomistic properties of helium in hcp titanium: a first-principles study. J. Nucl. Mater. 402, 55–59. (doi:10.1016/j.jnucmat.2010.04.022) 17. Heinisch HL, Gao F, Kurtz RJ. 2010 Atomic-scale modelling of interactions of helium, vacancies and helium-vacancy clusters with screw dislocations in alpha-iron. Phil. Mag. 90, 885–895. (doi:10.1080/14786430903294932) 18. You Y-W, Kong X-S, Wu X-B, Fang QF, Chen J-L, Luo G-N, Liu CS. 2013 Effect of vacancy on the dissolution properties of hydrogen and helium in molybdenum. J. Nucl. Mater. 433, 167–173. (doi:10.1016/j.jnucmat.2012.09.025) 19. You Y-W, Kong X-S, Fang QF, Liu CS, Chen JL, Luo G-N, Dai Y. 2013 A first-principles study on hydrogen behavior in helium-implanted tungsten and molybdenum. J. Nucl. Mater. 450, 64–68. (doi:10.1016/j.jnucmat.2013.04.026) 20. Kresse G, Hafner J. 1993 Ab initio molecular dynamics for open-shell transition metals. Phys. Rev. B 48, 13115. (doi:10.1103/PhysRevB.48.13115) ................................................... References 12 rspa.royalsocietypublishing.org Proc. R. Soc. A 470: 20140357 computing service, made available to us through the EPSRC-supported (EP/L000202) Materials Chemistry Consortium. The authors acknowledge the use of the IRIDIS High-Performance Computing Facility, and associated support services at the University of Southampton, in the completion of this work. Funding statement. NHdL acknowledges the Royal Society for an Industry Fellowship and thanks AWE for funding. Downloaded from http://rspa.royalsocietypublishing.org/ on June 18, 2017 13 ................................................... rspa.royalsocietypublishing.org Proc. R. Soc. A 470: 20140357 21. Kresse G, Furthmüller J. 1996 Efficient iterative schemes for Ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 54, 11 169–11 186. (doi:10.1103/PhysRevB.54.11169) 22. Perdew JP, Burke K, Ernzerof M. 1996 Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868. (doi:10.1103/PhysRevLett.77.3865) 23. Henkelman G, Jónsson H. 1999 A dimer method for finding saddle points on high dimensional potential surfaces using only first derivatives. J. Chem. Phys. 111, 7010. (doi:10.1063/1.480097) 24. Kittel C. 2005 Introduction to solid state physics, 8th edn. London, UK: John Wiley and Sons. 25. Gupta M. 2000 Electronic structure of hydrogen storage materials. Int. J. Quant. Chem. 77, 982–990. (doi:10.1002/(SICI)1097-461X(2000)77:6<982::AID-QUA6>3.0.CO;2-#) 26. Goods SH, Guthrie SE. 1992 Mechanical properties of palladium and palladium hydride. Scr. Metall. Mater. 26, 561–565. (doi:10.1016/0956-716X(92)90284-L) 27. Flanagan TB, Oates WA. 1991 The palladium–hydrogen system. Annu. Rev. Mater. Sci. 21, 269–304. (doi:10.1146/annurev.ms.21.080191.001413) 28. Sugimoto H, Fukai Y. 1982 Energy and wave functions of interstitial hydrogen in fcc and bcc metals. J. Phys. Soc. Jpn 51, 2554–2561. (doi:10.1143/JPSJ.51.2554) 29. Klamt A, Teichler H. 1986 Quantum-mechanical calculations for hydrogen in niobium and tantalum. I. Model potential and properties of self-trapped states. Phys. Status Solidi (b) 134, 103–114. (doi:10.1002/pssb.2221340113) 30. Wilson WD, Bisson CL, Baskes MI. 1981 Self-trapping of helium in metals. Phys. Rev. B 24, 5616. (doi:10.1103/PhysRevB.24.5616) 31. Thomas GJ, Swansiger WA, Baskes MI. 1979 Low-temperature helium release in nickel. J. Appl. Phys. 50, 6942. (doi:10.1063/1.325848) 32. Melius CF, Bisson CL, Wilson WD. 1978 Quantum-chemical and lattice-defect hybrid approach to the calculation of defects in metals. Phys. Rev. B 18, 1647. (doi:10.1103/ PhysRevB.18.1647) 33. Kass WJ. 1977 Thermal desorption measurements of helium ion-implanted erbium tritide. J. Vac. Sci. Technol. 14, 518–522. (doi:10.1116/1.569293) 34. Kraftmakher Y. 1998 Equilibrium vacancies and thermophysical properties of metals. Phys. Rep. 299, 79–188. (doi:10.1016/S0370-1573(97)00082-3) 35. Zhang P, Li R, Zhao J, Wen B. 2013 Synergetic effect of H and He with vacancy in vanadium solid from first-principles simulations. Nucl. Instrum. Methods Phys. Res. B 303, 74–79. (doi:10.1016/j.nimb.2012.10.043) 36. Camp WJ. 1977 Helium detrapping and release from metal tritides. J. Vac. Sci. Technol. 14, 514–517. (doi:10.1116/1.569292) 37. Han YS, Tomar V. 2013 An ab initio study of the peak tensile strength of tungsten with an account of helium point defects. Int. J. Plast. 48, 54–71. (doi:10.1016/j.ijplas.2013.02.005) 38. Lam PK, Cohen ML, Martinez G. 1987 Analytical relation between bulk moduli and lattice constants. Phys. Rev. B 35, 9190. (doi:10.1103/PhysRevB.35.9190) 39. Gupta RP, Gupta M. 2002 Consequences of helium production from the radioactive decay of tritium on the properties of palladium tritide. Phys. Rev. B 66, 014105. (doi:10.1103/ PhysRevB.66.014105)
© Copyright 2025 Paperzz