This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Brought to you by:
Paper

A first-principles model for anomalous segregation in dilute ternary tungsten-rhenium-vacancy alloys

, , and

Published 3 March 2017 © Culham Centre for Fusion Energy
, , Citation J S Wróbel et al 2017 J. Phys.: Condens. Matter 29 145403 DOI 10.1088/1361-648X/aa5f37

0953-8984/29/14/145403

Abstract

The occurrence of segregation in dilute alloys under irradiation is a highly unusual phenomenon that has recently attracted attention, stimulated by the interest in the fundamental properties of alloys as well as by their applications. The fact that solute atoms segregate in alloys that, according to equilibrium thermodynamics, should exhibit full solubility, has significant practical implications, as the formation of precipitates strongly affects physical and mechanical properties of alloys. A lattice Hamiltonian, generalizing the so-called 'ABV' Ising model and including collective many-body inter-atomic interactions, has been developed to treat rhenium solute atoms and vacancies in tungsten as components of a ternary alloy. The phase stability of W–Re-vacancy alloys is assessed using a combination of density functional theory (DFT) calculations and cluster expansion (CE) simulations. The accuracy of CE parametrization is evaluated against the DFT data, and the cross-validation error is found to be less than 4.2 meV/atom. The free energy of W–Re-vacancy ternary alloys is computed as a function of temperature using quasi-canonical Monte Carlo simulations, using effective two, three and four-body interactions. In the low rhenium concentration range (<5 at.$ \% $ Re), solute segregation is found to occur in the form of voids decorated by Re atoms. These vacancy-rhenium clusters remain stable over a broad temperature range from 800 K to 1600 K. At lower temperatures, simulations predict the formation of Re-rich rhenium–vacancy clusters taking the form of sponge-like configurations that contain from 30 to 50 at.$ \% $ Re. The anomalous vacancy-mediated segregation of Re atoms in W can be rationalized by analyzing binding energy dependence as a function of Re to vacancy ratio as well as chemical Re–W and Re-vacancy interactions and short-range order parameters. DFT calculations show that rhenium–vacancy binding energies can be as high as 1.5 eV if the rhenium/vacancy ratio is in the range from 2.4 to 6.6. The predicted Re clustering agrees with experimental observations of precipitation in self-ion irradiated W-2$ \% $ Re alloys and neutron-irradiated alloys containing 1.4 at.$ \% $ Re.

Export citation and abstract BibTeX RIS

1. Introduction

Tungsten (W) is a candidate material for various plasma-facing components because of its high melting temperature, high thermal conductivity and low sputtering and erosion yields [1, 2]. In tungsten, in addition to radiation defects, neutrons initiate nuclear transmutation reactions [3], where rhenium (Re) is the main solid transmutation element forming in W under neutron irradiation. It has long been recognized that room-temperature brittleness of tungsten can be alleviated by alloying it with rhenium [4]. Microstructural evolution in neutron irradiated W, W–Re and W–Re–Os alloys was investigated in [510]. It was found that nucleation and growth of voids dominate microstructure of pure W, but the addition of rhenium suppresses void formation. Hardening is thought to be caused by radiation-induced precipitation of σ phase (WRe) and χ phase (WRe3), which form in W-5%Re and W-10%Re alloys after irradiation to 0.5–0.7 dpa at 600–1500 °C, see [8]. This anomalous segregation of Re occurs despite the fact that the thermodynamic solubility limit of Re in W is high, close to 30 at.$ \% $ [11]. Furthermore, the precipitation of α-Mn-type phase in W-25%Re alloy, resulting from exposure to the neutron dose of several dpa, was reported in [12]. Platelet-like precipitates in neuron-irradiated W-10%Re and W-25%Re alloys were found using a combination of field-ion microscopy (FIM) and atom probe tomography (APT) in [13, 14]. A recent TEM investigation [15] of neutron-irradiated polycrystalline and single crystal tungsten exposed to the neutron dose of 1.6 dpa at 900 °C, where transmutation reactions resulted in the accumulation of 1.4 at.$ \% $ Re, shows the formation of rhenium clusters as well as voids decorated by rhenium.

Ion irradiation provides means for investigating the dynamics of microstructural evolution in alloys, and help circumvent the problem of sample activation [16, 17]. Experimental examination of W-2$ \% $ Re and W-1$ \% $ Re–1$ \% $ Os alloys, irradiated with 2 MeV W+ self-ions [18, 19], by atom probe tomography (APT) and nano-hardness measurements, provides evidence for the occurrence of radiation-induced segregation in under-saturated solid solutions [20]. For example, experimental observations show the formation of Re-rich clusters, approximately 3 nm in diameter, in atom maps of Re solutes in a highly dilute 2 at.$ \% $ binary alloy irradiated at 773 K to the dose of 33 dpa. In W-1$ \% $ Re–1$ \% $ Os alloy, just one percent rhenium concentration also appears to be sufficient for the formation of Re-rich clusters under irradiation. The presence of Os suppresses the formation of Re precipitates, with Os-rich clusters forming instead. Segregation in under-saturated solid solutions under irradiation was also found in Ni–Si, Ni–Ge and Al–Zn alloys [2124]. Segregation of Cr was discovered in Fe-5$ \% $ Cr alloys irradiated with Fe ions to a relatively high dose at 400 °C [25]. The latter finding is particularly striking and significant since binary Fe–Cr alloys are a well known model alloy family [26], the thermodynamic properties of which resemble those of ferritic/martensitic steels [27]. Hence, we have extensive experimental evidence [15, 20, 25] suggesting that irradiation stimulates the formation of secondary-phase precipitates in under-saturated iron and tungsten alloys that otherwise exhibit no precipitation under thermodynamic equilibrium conditions.

Interpretation of experimental observations of solute segregation effects in the context of theoretical models linking solute segregation to binding between defects and solute atoms, and/or different rates of diffusion of solute and solvent atoms, has so far primarily focused on the kinetic aspects of phase stability [21, 2830]. The treatments of solute fluctuations in solid solution under irradiation in combination with density functional theory (DFT) calculations [3134] provide insight into the highly complex phenomena of solute-defect trapping, solute segregation, point-defect recombination, dislocation interactions, nucleation and growth of voids. Similar models have recently been applied to W-based alloys [3544], and iron alloys and steels [4550].

In this work, we adopt a conceptually different approach to modelling radiation-stimulated precipitation of solutes in irradiated alloys, where point defects are represented as extra 'chemical' components within a concentration-constrained thermodynamic treatment of solid solutions. This makes it possible to interpret radiation-induced precipitation using free energy minimization, which is applied to an alloy containing defects produced by irradiation. The treatment overcomes a major difficulty encountered in the context of a kinetic approach, associated with the fact that a kinetic model requires information not only about energies of configurations but also about the transition rates and defect mobilities. In this paper, as a proof of concept, we treat vacancies as a component of a binary W–Re alloy, mapping the alloy onto a ternary alloy system containing tungsten and rhenium atoms as well as vacancies.

The formation energies of self-interstitial atom defects, including Re–Re and mixed W–Re dumbbells in bcc-W, are about three times the vacancy formation energy [39, 43, 44, 51]. Binding energies that characterize interaction between self-interstitial atom defects and solutes in W–Re alloys predicted by DFT, are also high [39, 40, 52] in comparison with energies describing binding between vacancies and rhenium solute atoms computed in this study (see figure 4 and table S3 in supplementary material (SM) (stacks.iop.org/JPhysCM/29/145403/mmedia)). Migration energies of self-interstitial atom (SIA) and mixed interstitial defects are very small [37, 44, 52]. Putting this in the context of experimental observations of elongated precipitates in neutron-irradiated W-10$ \% $ Re [13], W-25$ \% $ Re [14] and W-26$ \% $ Re [7, 9], we conclude that interstitial defects may in principle play a part in the kinetics of phase decomposition of irradiated W–Re alloys in the limit where the concentration of Re is high.

Here, we investigate radiation-induced precipitation (RIS) occurring in the solid solution region of the W–Re binary phase diagram, where the concentration of Re is low. The main difference between the two (interstitial and vacancy mediated) RIS models can be understood by comparing the mixing enthalpies of formation calculated separately for the defective self-interstitial and vacancy dominated W–Re alloys. Results of DFT calculations illustrated in figure 5(c) of [52] show that in self-interstitial-dominated defective alloy structures, the mixing energies are positive and large, and the convex hull line is maximum at 300 meV/atom at the rhenium concentration of approximately 15$ \% $ Re. Comparing this with the much lower mixing energies calculated in this work for the vacancy-dominated alloy structures, where the corresponding convex hull line exhibits the maximum value of only 30 meV/atom, we conclude that the lower energy configurations of defective W–Re alloys in the concentration range below 15$ \% $ Re are mostly of Re-vacancy type. Hence, from the constrained thermodynamic point of view, it is natural to adopt a starting approximation where we consider only vacancies when evaluating the free energy of defective W–Re solid solutions exposed to irradiation. Earlier models for the free energy of W–Re alloys, also taking into account point defects, were based either on the relatively simple effective pair interaction model of Bragg and Williams [53, 54] or on the higher-order approximations that take into account multisite correlations, including for example the cluster variation method (CVM) [55, 56].

In this paper, we investigate the anomalous segregation of Re in dilute W–Re alloys using a combination of first-principles calculations and statistical mechanics simulations based on a generalized Ising alloy model, known as cluster expansion (CE) [57, 58]. CE makes it possible to explore the phase stability of magnetic fcc and bcc Fe–Cr–Ni ternary alloys [59], and even high-entropy alloys [60]. A ternary CE model, also based on DFT calculations, was recently applied to the investigation of vacancies in fcc Cu–Ni at equilibrium [61]. Microstructural evolution associated with the formation and growth of voids in supersaturated vacancy solutions was investigated using classical phase-field models [62] and also first-principles models for heterogeneous nucleation, where vacancy content was close to 0.1 at.$ \% $ [63].

A CE model that treats vacancies as an alloy component, is applied to W–Re–Vac ternary alloys in section 2. In section 3, DFT data for binding energies of Re-vacancy clusters in tungsten are analyzed as functions of the Re to vacancy concentration ratio. Section 4 focuses on finite-temperature effects, where rhenium clustering is investigated as a function of Re and vacancy solute concentrations using quasi-canonical Monte Carlo simulations. In section 5, we discuss the origin of anomalous segregation in dilute W–Re–Vac alloys, and compare theoretical predictions with experimental APT and TEM observations.

2. Computational methodology

2.1. Cluster expansion formalism for ternary W–Re–Vac alloys

We describe the state of a solid solute under irradiation by three concentrations: xA, xB, xC that are, respectively, the solvent, solute and vacancy concentrations expressed in terms of the number per lattice site units. In our model, the three concentrations are subject to a constraint

Equation (1)

In cluster expansion (CE), the configuration enthalpy of mixing of a multi-component alloy is defined as [58]

Equation (2)

where summation is performed over all the clusters ω that are distinct under group symmetry operations applied to the underlying lattice, $m_{\omega}^{\text{lat}}$ are the multiplicities indicating the number of clusters equivalent to ω by symmetry, divided by the number of lattice sites, and $\langle {{\Gamma}_{{{\omega}^{\prime}}}}(\vec{\sigma})\rangle $ are the cluster functions defined as products of functions of occupation variables on a specific cluster ω averaged over all the clusters ${{\omega}^{\prime}}$ that are equivalent by symmetry to cluster ω. ${{J}_{\omega}}$ are the concentration-independent effective cluster interaction (ECI) parameters, derived from a set of ab initio calculations using the structure inversion method (SIM) [64]. A cluster ω is defined by its size (i.e. the number of lattice points) $|\omega |$ and relative positions of points. For clarity, each cluster ω is described by two parameters $(|\omega |,n)$ , where $|\omega |$ is the cluster size and n is a label, see table S1 from the supplementary information (SI) for bcc lattice.

In CE developed for a K-component system, a cluster function is not a simple product of occupation variables, $\{{{\sigma}_{i}}\}$ . Instead, it is defined as a product of orthogonal point functions ${{\gamma}_{{{j}_{i}},K}}({{\sigma}_{i}})$ ,

Equation (3)

where sequence $(s)=(\,{{j}_{1}}{{j}_{2}}\ldots ~{{j}_{|\omega |}})$ is the decoration [65] of a cluster by point functions. We use the following definition of point functions for a K-component system

Equation (4)

where ${{\sigma}_{i}}=0,1,2,\ldots,(K-1)$ , j is the index of point functions ($j=0,1,2,\ldots,(K-1)$ ). In a ternary alloy, index K equals 3 and occupation variables are defined as $\sigma =0,1,2$ , referring to the constituent components of the alloy A, B and C, which here correspond to W, Re and Vac.

According to the derivation given in [59], the configuration enthalpy of mixing of a ternary alloy can be expressed analytically in terms of three-body interaction parameters as a function of concentrations, xi, and the average pair and 3-body probabilities, $y_{n}^{ij}$ and $y_{n}^{ijk}$ as

Equation (5)

It is worth mentioning here that equation (5) with C  =  V (vacancy) represents a generalized Hamiltonian of the so-called 'ABV' Ising model [66], the conventional form of which contains only the first and second nearest-neighbor effective pair interactions. The 'ABV' models have been applied to the investigation of non-equilibrium phenomena such as enhanced diffusion and segregation in irradiated materials [67]. We now apply the more accurate cluster-expansion based ABV model to the treatment of radiation-induced precipitation in bcc W-rich alloys, where anomalous segregation of Re atoms was observed experimentally [15, 20]. The enthalpy of mixing of a W–Re–Vac configuration is found by subtracting the total enthalpies corresponding to pure end terms as:

Equation (6)

where the reference enthalpy of a vacancy is assumed to be zero, ${{E}_{\text{tot}}}(\text{Vac})=0$ .

The values of ECIs ($J_{|\omega |,n}^{(s)}$ ) for ternary W–Re–Vac bcc alloys are derived by mapping DFT energies onto the CE energies computed for 224 structures, which are described in the next sections of this paper. According to the definition of cluster functions (equation (3)) and point functions (equation (4)) for ternary alloys, each cluster can be decorated in various ways for each nearest-neighbour shell of interactions. Here, we have used a set of 15 two-body, 12 three-body, and 6 four-body clusters on bcc lattice. Values of all the optimized ECIs for ternary W–Re–Vac alloys are shown in figure 1 and also in the last column of table S1 in the SI. Section 3 shows that the inclusion of five nearest neighbour shells in two-body interactions is essential for understanding the origin of rhenium–vacancy binding energy trends characterizing under-saturated solid solutions.

Figure 1.

Figure 1. Effective cluster interactions in ternary W–Re–Vac alloy.

Standard image High-resolution image

To arrive at a suitable choice of Hamiltonian (equation (2)), the fitness of a cluster expansion can be, in general, quantified by means of leave-many-out cross-validation (LMO-CV) [68]. The set of structures $\vec{\sigma}$ for which we carry out first-principles (DFT) energy calculations, $ \Delta {{E}_{\text{DFT}}}({{\vec{\sigma}}_{\text{input}}})$ is subdivided into a fitting set ${{\{\vec{\sigma}\}}_{\text{fit}}}$ and a prediction set ${{\{\vec{\sigma}\}}_{\text{pred}}}$ . We determine ECIs ($J_{|\omega |,n}^{(s)}$ ) by fitting to $ \Delta {{E}_{\text{DFT}}}(\vec{\sigma})$ in ${{\{\vec{\sigma}\}}_{\text{fit}}}$ . The predicted values of $ \Delta {{E}_{\text{CE}}}(\vec{\sigma})$ for each structure ${{\{\vec{\sigma}\}}_{\text{fit}}}$ are then compared with their known first-principles counterparts $ \Delta {{E}_{\text{DFT}}}(\vec{\sigma})$ . This process can be iterated M times for various subdivisions of ${{\{\vec{\sigma}\}}_{\text{input}}}$ into fitting and prediction sets ${{\{{{\vec{\sigma}}_{\text{fit}}}\}}^{(m)}}$ and ${{\{{{\vec{\sigma}}_{\text{pred}}}\}}^{(m)}}$ (m  =  1... N). The values of the LMO-CV score is defined as the average prediction error of this iterative procedure,

Equation (7)

In this work, the CE Hamiltonian defined by equation (2) is applied to a defective vacancy-rich W–Re–Vac solid solution where, as opposed to a conventional ternary non-defective alloy [59], there are no further ground-state stable structures that can be found in comparison with the ones predicted from binary W–Re alloy calculations. Therefore, the LMO-CV score can be reduced to the leave-one-out cross validation (LOO-CV) score with M  =  1 in equation (7). The LOO-CV score characterizing the quality of agreement between DFT and CE has been assessed using 224 structures representing interactions in W–Re–Vac system. The value of the LOO-CV error is 4.12 meV/atom, confirming the very high accuracy of the set of ECIs used in the simulations.

2.2. Computational details

DFT calculations were performed using Vienna ab initio simulation package (VASP) with the interaction between ions and electrons described using the projector augmented waves (PAW) method [69, 70]. Exchange and correlation were treated in the generalized gradient approximation GGA-PBE [71], with PAW potentials containing semi-core p electron contributions. Supercell calculations were performed considering vacancy clusters interacting with Re atoms in bcc W lattice under constant pressure conditions, with structures optimized by relaxing both atomic positions as well as the shape and volume of the supercell. To treat clusters containing from 2 to 47 sites with Re atoms and vacancies, orthogonal supercells containing 128 and 250 atoms were used. Total energies were evaluated using the Monkhorst-Pack mesh [72] of k-points in the Brillouin zone, with the k-mesh spacing of 0.15 ${{{\mathring{\text{A}}}}^{-1}}$ . This corresponds to $4\times 4\times 4$ or $3\times 3\times 3$ k-point meshes for a bcc supercell of $4\times 4\times 4$ or $5\times 5\times 5$ bcc structural units, respectively. The plane wave cut-off energy was 400 eV. The total energy convergence criterion was set to 10−6 eV/cell, and force components were relaxed to 10−3 eV ${{{\mathring{\text{A}}}}^{-1}}$ . Vacancy and self-interstitial-atom (crowdion) formation energies evaluated using the above conditions for pure bcc W were 3.307 eV and 10.917 eV, respectively.

Mapping of DFT energies to CE was performed using the ATAT package [73]. For binary bcc alloys we used 58 structures from [74], and the corresponding results for the enthalpy of mixing in the binary W–Re are shown in figure 2 and table S2 in the SI. We find that negative values of the enthalpy of mixing characterizing bcc W–Re alloys were smaller than those obtained using similar DFT/CE calculations performed for binary W–Ta and W–V alloys [39]. The most stable configurations predicted for W–Re binary alloys remain the same as in the extended treatment of ternary W–Re–Vac alloys as well as in the simplified CE scheme discussed in the last section of the paper (see table S4 in the SI), where in the first nearest neighbour (1NN) shell the effective chemical interaction between W and Re is attractive but is more than 5 times smaller than that between Re and a vacancy. In the second nearest neighbour (2NN) coordination shell the effective W–Re interaction is weak and repulsive whereas Re–Vac interaction is strong and attractive. Hence we conclude that clustering of Re in bcc W is driven primarily by binding between Re solute atoms and the vacancies.

Figure 2.

Figure 2. Enthalpy of mixing of W–Re structures calculated using DFT and CE.

Standard image High-resolution image

The DFT database of structures used in this study uses data not only for the W–Re binary system but also data on vacancy clusters in W (W–Vac) [39], as well as data on Re-vacancy cluster interactions in W.

Quasi-canonical MC simulations were carried out using the ATAT package [73]. All the simulations described in section 4 used a $30\times 30\times 30$ bcc supercell containing 54 000 lattice sites. To evaluate the formation free energy for steady state configurations using thermodynamic integration, simulations were performed starting from a disordered high-temperature state, corresponding to T  =  3000 K. Configurations were then cooled down with the temperature step of $ \Delta T=10$ K, with 3000 MC steps per atom at both thermalization and accumulation stages. It was found that it was not necessary to use a higher number of MC steps to achieve convergence.

3. Re-vacancy binding energies at 0 K

Binding energy of a cluster containing n vacancies and m Re atoms in W matrix is the energy required to dissolve the cluster into n individual separate vacancies and m separate Re solute atoms in W matrix. It can be defined as follows [75]:

Equation (8)

where ${{E}_{\text{tot}}}(n\cdot \text{Vac},m\cdot \text{Re})$ is the total energy of a supercell of W atoms containing a cluster with n vacancies and m Re atoms, ${{E}_{\text{tot}}}(1\cdot \text{Vac})$ is the total energy of a supercell with one vacancy, ${{E}_{\text{tot}}}(1\cdot \text{Re})$ is the energy of a supercell with one Re atom, and ${{E}_{\text{tot}}}(\text{W})$ is the total energy of a supercell containing only W atoms. All the supercells contain the same number of sites.

3.1. Interaction of a mono-vacancy with Re

Calculations of binding energies of Re atoms with a mono-vacancy at zero Kelvin as a function of the ratio of the number of Re atoms to vacancies were performed by replacing W atoms by Re atoms in the first and then in the second coordination shells of a vacancy, see blue and pink spheres in figure 3(a). Results are given in figure 4 and table S3 in the SI. The binding energy of a vacancy with a Re atom is 0.183 eV, which is slightly smaller than the value of approximately 0.2 eV found earlier [40]. Since the enthalpy of mixing of W and Re atoms is small, see figure 2, the strength of binding in vacancy-Re clusters depends primarily on the number of Re atoms in the neighbourhood of a vacancy, and does not depend significantly on the specific configuration of Re atoms in the first or the second shells around a vacancy. As a result, the binding energy increases almost linearly as a function of Re atoms around a vacancy, up to the value of 1.304 eV corresponding to the case where Re atoms occupy eight positions in the first nearest neighbour coordination shell. Binding energy decreases once Re atoms start occupying positions in the second nearest neighbour coordination shell, see figure 4.

Figure 3.

Figure 3. Schematic representation of atoms in the neighbourhood of (a) a vacancy, (b) a di-vacancy in the third nearest neighbourhood, a four-vacancy cluster (figure 5(h)) and (d) a void (figure 5(h)). Colours indicate the number of vacancies in the first (1NN) and second (2NN) neighbourhood of a given atom: black—four 1NN and zero 2NN; brown—two 1NN and two 2NN; red—two 1NN and zero 2NN; orange—one 1NN and one 2NN; blue—one 1NN and zero 2NN; yellow—zero 1NN and three 2NN; green—zero 1NN and two 2NN; pink—zero 1NN and one 2NN.

Standard image High-resolution image
Figure 4.

Figure 4. Binding energy between Re atoms and vacancy clusters shown as a function of Re to vacancy ratio for di-vacancy (a), tri-vacancy (b), quarto-vacancy (c) clusters as well as for voids, the most compact vacancy clusters (d). Explanation of abbreviations and the schematic representations of vacancy clusters are given in figure 5. Vac x8(MC) are results referring to the configuration of 8 vacancies obtained from MC simulations with 2% at.Re and 0.05% at. vacancies (see table 1).

Standard image High-resolution image

3.2. Interaction of vacancy clusters with Re

Binding energies of di-vacancies as functions of Re/vacancy ratio for di-vacancies in the first, second, third, fourth and fifth nearest neighbourhood indicated as 1NN, 2NN, 3NN, 4NN and 5NN, respectively, are shown in figure 4(a) and table S3 in the SI. In pure W matrix, the binding energies of all the configurations apart from 4NN di-vacancy configuration, are negative, meaning that vacancies repel each other. The most negative is the binding energy of a 2NN di-vacancy configuration that is equal to  −0.177 eV per vacancy, which is slightly less negative than  −0.190 eV per vacancy reported earlier [36, 39]. The binding energy of the 1NN di-vacancy configuration equals to  −0.012 eV per vacancy, and is small but still negative.

In order to investigate the effect of Re atoms on the binding energies of various di-vacancy configurations, W atoms in the 1NN and 2NN positions were systematically replaced by Re atoms. Since the binding energy primarily depends on the number of Re atoms in the 1NN and 2NN positions with respect to vacancies, the largest contribution to the binding energy of a cluster comes from Re atoms that have two vacancies in their first nearest-neighbour coordination shells. Hence, a systematic investigation of the effect of Re atoms on binding energies of the 3NN di-vacancy configuration starts with a replacement of W atoms by Re atoms in the first nearest neighbourhood of two vacancies, see red spheres in figure 3(b). Afterwards the replacement of W atoms by Re atoms was performed in the following sequence: atoms that have one vacancy in the 1NN shell, two vacancies in the 2NN shell and one vacancy in the 2NN shell indicated by blue, green and pink spheres in figure 3(b), respectively. The same scheme was applied to the replacement of W atoms by Re atoms in other di-vacancy and multi-vacancy configurations, see figures 3(c) and (d) for the case of four-vacancy cluster and a 15-vacancy void, respectively.

Binding energies of all the di-vacancy configurations apart from the 2NN di-vacancy configuration (vacancies ordered in the (1 0 0) direction) are positive, meaning that vacancies in the neighbourhood of a Re atom attract each other. In other words, a Re atom provides a centre of nucleation for vacancies. The largest binding energy that is equal to 0.157 eV per vacancy characterizes the 3NN di-vacancy configuration (vacancies ordered in the (1 1 0) direction) bound by a Re atom, despite the fact that the binding energy of a 3NN di-vacancy configuration without Re atoms is only slightly more negative that that of the 1NN and 4NN di-vacancy configurations. The 3NN di-vacancy configuration is the most preferable energetically up to the Re/vacancy ratio approaching 7. For the ratios above 7, all the configurations of vacancies apart from the 1NN di-vacancy configuration, have quite similar binding energies. Moreover, they usually have smaller binding energies than those characterizing Re atoms and a mono-vacancy, see the black line in figure 4(a). It means that for the Re/vacancy ratios higher than 7, vacancies may prefer to occupy configurations other than the third nearest neighbour sites. The binding energies of tri-vacancy configurations are shown in figure 4(b) and table S3 in the SI as functions of the Re/vacancy ratio. In a pure W matrix the most stable tri-vacancy configuration is the one with vacancies forming the most compact cluster, see figure 5(f). This configuration remains the most stable in the presence of one Re atom and two Re atoms. For higher Re/vacancy ratios the most stable configuration is the one where vacancies are in the 3NN coordination shell, see figure 5(h).

Figure 5.

Figure 5. Schematic representation of vacancy clusters; tri-vacancy configurations: (a) Vac x3(a), (b) Vac x3(b), (c)Vac x3(c), (d) Vac x3(d), (e) Vac x3(e), (f) Vac x3(f); quarto-vacancy configurations: (g) Vac x4(g), (h) Vac x4(h), (i) Vac x4(i); void configurations: (i) with 4 vacancies—Vac x4:Void(i), (j) with 8 vacancies—Vac x8:Void(j), (k) with 15 vacancies—Vac x15:Void(k).

Standard image High-resolution image

Binding energies of quarto-vacancy configurations, forming a (1 0 0) surface, see figure 5(g), and with vacancies in the 3NN coordination shell to each other, see figure 5(h), are shown in figure 4(c) whereas those corresponding to the most compact vacancy cluster (void) configurations, see figure 5(i), are shown in figure 4(c). The values of binding energies are given in table S3 in the SI.

3.3. Voids and vacancy clusters

Void configurations, starting from three-vacancy clusters in the 1NN and 2NN configurations, to a 15-vacancy cluster, are stable in a pure W matrix, according to previous ab initio calculations [39, 76]. When the Re/vacancy ratio increases, configurations with vacancies in the 3NN coordination shell become more stable, similarly to how it was found in the tri-vacancy case with the same relative positions of vacancies. Similarly to the case of a tri-vacancy configuration, where the Re/vacancy ratio is higher than 7, binding between Re atoms and a mono-vacancy is stronger than in quarto-vacancy configurations.

Figure 4(d) and table S3 in the SI show that the binding energy of the most compact vacancy clusters (voids) without Re atoms increases as a function of the number of vacancies in a void. This agrees with the earlier work [39]. Results obtained with W atoms replaced by Re atoms show that binding energies increase to over 1.5 eV when the voids are surrounded by Re atoms. Overall, the Re-void configurations containing a larger number of vacancies are more stable than those with smaller number of vacancies over the entire range of Re/vacancy ratios.

In order to compare the relative stability of the most compact vacancy configurations (small voids) with other vacancy cluster configurations, a structure containing 8 vacancies, derived from MC simulations of W–Re alloys with 2% at.Re and 0.05% at. vacancies quenched down from high temperatures (see the section 4.2), is also included in table S3 of SI. The binding energy characterizing this structure, where each vacancy is in the third nearest-neighbour position with respect to others, is smaller than those of the 8-vacancy void configuration for the low ratio of Re/vacancy, but becomes larger when this ratio is $\geqslant $ 4. In particular, we find that the highest value (1.566 eV) of binding energy between Re atoms and 8-vacancy clusters corresponds to the Re/vacancy ratio of 6.63.

Concluding this section, we note that binding energies at T  =  0 K for various vacancy cluster configurations show consistent trends in terms of their stability as functions of Re concentrations. In the limit where the Re/vacancy ratio is small, the more compact configurations are preferred. As this ratio increases, configurations with vacancies being further apart become more stable. Configurations with vacancies being in the 3NN coordination shell with respect to each other have particularly high stability over a broad interval of Re/vacancy ratios.

4. Clustering of Re atoms at finite temperatures

4.1. Steady-state alloy configurations from constrained thermodynamic integration

Within the framework of a constrained model for phase stability of alloys containing high concentration of vacancies, the thermodynamic properties of W–Re–Vac system were analyzed using quasi-canonical MC simulations and ECIs derived from DFT calculations. The dependence of the configuration entropy on temperature is given by equation

Equation (9)

where the configurational contribution to the specific heat Cconf is related to fluctuations of the enthalpy of mixing at a given temperature [77, 78] through

Equation (10)

where $\langle {{H}_{\text{mix}}}({{x}_{\text{W}}},{{x}_{\text{Re}}},{{x}_{\text{Vac}}},T)\rangle $ and $\langle {{({{H}_{\text{mix}}}({{x}_{\text{W}}},{{x}_{\text{Re}}},{{x}_{\text{Vac}}},T))}^{2}}\rangle $ are the mean and mean square average enthalpies of mixing, respectively, computed by averaging over all the MC steps at the accumulation stage for a given temperature and for a given composition of the ternary W–Re–Vac system.

The accuracy of evaluation of configuration entropy depends on the temperature integration step in equation (9) and on the number of MC steps performed at the accumulation stage. Test simulations show that choosing a sufficiently small temperature integration step is particularly significant. Calculations of configuration entropy for all the alloy compositions considered below were performed with 3000 MC steps per atom at thermalization and accumulation stages, and with the temperature step of $ \Delta T=10$ K. Figure 6 shows the dependence of configuration entropy as a function of temperature for W-2%Re and W-5%Re with 0.1$ \% $ of vacancy concentration. Transitions from ideal random solid solutions into the Re-precipitation regime can be seen clearly as steps in the configuration entropy at T  =  1700 K and T  =  1950 K for the alloys containing 5 and 2 at.$ \% $ Re, respectively.

Figure 6.

Figure 6. Entropy of W-2%Re (black line) and W-5%Re (red line) alloys with 0.1% concentration of vacancies as a function of temperature.

Standard image High-resolution image

Table 1 describes structures of W-Re-vacancy alloys at 100 K containing 54 000 sites derived from MC simulations by quenching down from 2500 K with the temperature step of 100 K and shown as functions of Re concentrations of 1, 2, 5 and 10% and vacancy concentrations of 0.05, 0.1 and 0.2%. The high fraction of vacancy defects considered in this study is consistent with experimental observations of radiation induced precipitates in W–Re alloys forming under self-ion irradiation [20, 79]. The most stable configuration of Re-vacancy clusters at 100 K depends on the concentration of Re atoms. For each vacancy concentration, including 0.05%, 0.1% and 0.2%, for low concentrations of Re atoms, 1% or 2% at.Re, the most stable configurations are the voids surrounded by Re atoms. For higher Re concentrations, 5% or 10% at.Re, vacancies prefer to be further apart from each other, and the most stable configurations are sponge-like Re-vacancy clusters. Those results are in agreement with the analysis of binding energies of Re-vacancy configurations at 0 K treated as a function of Re/vacancy ratio described in section 4.

Table 1. Monte Carlo results as functions of Re and vacancy concentration for alloys quenched down from high temperatures. MC simulations were performed starting from 2500 K. Alloys were cooled down with the temperature step of 100 K to the temperature of 100 K with 3000 MC steps per atom performed both at thermalization and accumulation stages.

  1% at. Re 2% at. Re 5% at. Re 10% at. Re
0.05% at. Vac
0.1% at. Vac
0.2% at. Vac

4.2. Meta-stable finite temperature configurations

In this section, to understand the anomalous segregation of rhenium occurring in W–Re alloys under irradiation condition, where alloys are not quenched from high temperatures but are irradiated at a constant temperature, we performed MC simulations at various temperatures assuming a certain concentration of vacancies and Re atoms.

Table 2 shows structures of W–Re alloys with 2% at.Re obtained from MC simulations performed assuming T  =  300 K, 800 K, 1600 K and 2500 K and four different concentrations of vacancies: 0.05%, 0.1%, 0.2% and 0.5%, which can be related to various values of irradiation doses. At the lowest concentration of vacancies of 0.05%, clustering of vacancies in the form of sponge-like Re-vacancy clusters is observed only at 800 K. At 300 K Re atoms are bound to vacancies but do not aggregate. Formation of Re-vacancy clusters becomes significantly more pronounced as the concentration of vacancies increases. If it exceeds 0.1%, at 1600 K vacancies form voids, whereas at 300 K they aggregate into sponge-like Re-vacancy clusters. In W-2%Re alloy at 800 K we observe either the formation of voids surrounded by Re atoms or sponge-like Re-vacancy clusters. The former ones are observed in alloys where vacancy concentration is equal to 0.2% and 0.5%, whereas the latter ones form in alloys where the concentration of vacancies is lower, in the range from 0.05% to 0.1%. This can be explained by the fact that a high concentration of vacancies corresponds to low Re/vacancy ratio, which is the key parameter controlling the stability of Re-vacancy configurations, as described in section 3. Another important conclusion is that at the border between regions of stability of voids and sponge-like Re-vacancy clusters, it is possible to find both types of configurations. At 2500 K both vacancies and Re atoms are fully dissolved.

Table 2. Results of Monte Carlo simulations as functions of vacancy concentration and temperature. Concentration of Re atoms is fixed and equal to 2%. MC simulations were performed at various temperatures and screen shots were taken after 20 000 MC steps per atom.

  0.05% at. Vac 0.1% at. Vac 0.2% at. Vac 0.5% at. Vac
300 K
800 K
1600 K
2500 K

Table 3 shows structures of W–Re alloys with vacancy concentration of 0.1% obtained from MC simulations performed for temperatures of 300 K, 800 K, 1600 K and 2500 K for four different concentrations of Re atoms: 1%, 2%, 5% and 10% at.Re. Similarly to the case of W-2%Re alloys with various concentrations of vacancies, clustering of vacancies for each concentration of Re atoms is only observed at 800 K. At 2500 K all the vacancies and Re atoms are fully dissolved. For Re concentrations of 5% and 10%, vacancies do not aggregate at 300 K and 1600 K. In W–Re alloys containing 1% and 2% Re, sponge-like Re-vacancy clusters form at 300 K and voids form at 1600 K. Stability of Re-vacancy configurations at 800 K also depends on the Re/vacancy ratio. At low Re concentration close to 1% the Re/vacancy ratio is small enough to favour the formation of a void surrounded by Re atoms. At higher Re concentrations and higher values of the Re/vacancy ratio, sponge-like Re-vacancy clusters are more stable.

Table 3. Results of Monte Carlo simulations shown as functions of Re concentration and temperature. Concentration of vacancies is fixed and equal to 0.1%. MC simulations were performed at fixed temperatures and screen shots were taken after 20 000 MC steps per atom.

  1% at. Re 2% at. Re 5% at. Re 10% at. Re
300 K
800 K
1600 K
2500 K

5. Discussion and conclusion

To rationalize DFT and MC simulations of ternary W–Re–Vac alloys and their application to the interpretation of observations of radiation-induced precipitation of Re solute atoms in irradiated binary W–Re alloys, we analyzed results of simulations in terms of chemical pairwise interactions $V_{n}^{ij}$ derived from ECIs. Mapping DFT data to CE was performed by retaining only the effective interactions between different pairs of atoms within the 5NN range, corresponding to 15 two-body interactions. The DFT-CE cross-validation error is now 7.61 eV meV/atom, and although this value is higher than that obtained for figure 1 and table S1 in the SI, it is sufficient for interpreting the Re clustering phenomenon. With equation (5) expressed in terms of chemical pairwise interactions, the configuration enthalpy of mixing of a ternary alloy is [59]

Equation (11)

while in the binary alloy case, chemical pairwise interactions have a simple meaning: $V_{n}^{AB}>0$ corresponds to attraction and $V_{n}^{AB}<0$ to repulsion between atoms A and B, in the ternary case, equation (11), they depend not only on $V_{n}^{AB}$ , $V_{n}^{AC}$ , $V_{n}^{BC}$ but also on the chemical short-range order values associated with the pair probability functions $y_{n}^{AB}$ , $y_{n}^{AC}$ , $y_{n}^{BC}$ .

Values of effective pair interactions involving W–Re, W–Vac and Re–Vac at distances up to the radius of the 5NN shell in W–Re–Vac alloy are summarized in table S4 in the SI. Because of small mixing enthalpy values for W–Re pairs (see figure 2 and table S2 in the SI), interaction between W and Re atoms have a negligible effect on the order-disorder phase transformation in ternary W–Re–Vac alloys. Comparison between W–Vac and Re–Vac effective pair interactions is given in figure 7. In the 1NN shell, W–Vac interaction is repulsive whereas Re–Vac interaction is attractive, in agreement with the analysis of binding energies given in section 3. ECIs corresponding to the 2NN configurations both have positive sign, showing that Re and W atoms can occupy 2NN positions around a vacancy. For the 3NN and 4NN shells, the ECIs have negative sign, meaning that neither Re or W atoms are energetically favorable in the third and fourth NN coordination around a vacancy. As a result, this favors vacancies occupying positions in the third and fourth NN position, as illustrated in figure 3(b) for the di-vacancy 3NN configuration.

Figure 7.

Figure 7. Effective interactions of vacancies with W and Re atoms in the nth nearest-neighbour coordination shell in ternary W–Re–Vac alloys, in meV units.

Standard image High-resolution image

Figure 8(a) shows a representative structure of W–Re–Vac cluster along a $(1\,1\,1)$ direction, obtained from MC simulations for W-2%Re–0.1%Vac alloy at T  =  300 K (see table 3). In all the clusters, vacancies tend to occupy the 3NN positions, whereas Re atoms are usually in the 1NN shell with respect to vacancies, with W atoms usually located between the Re atoms. The estimated Re concentration within the cluster shown in figure 8(a) can be as high as 41% with vacancies surrounded by Re atoms. At T  =  800 K, Re concentration inside the cluster in a W-2%Re–0.1%Vac alloy is close to 29$ \% $ , which is many times higher than the average concentration of Re in the alloy. We note that both the atomic structure and Re concentration range associated with this cluster agree with the atomistic mechanism of phase transformation into the σ phase [80] in the W–Re binary phase diagram [11]. Figure 8(b) shows a typical configuration of a precipitate in the form of a void decorated by Re atoms, obtained from MC simulations for W-2%Re–0.1%Vac alloy at T  =  1600 K (see table 2). The concentration of Re in the first nearest-neighbour configuration around the void is found to be 19%. Finally, the phase transformation from binary W–Re alloys with negative enthalpy of mixing into the anomalous segregation with low Re and high vacancy concentrations can be cross-checked by analyzing the change of short-range order (SRO) parameters for the ternary W–Re–Vac system (see equation (21) from [59]). Indeed, it is confirmed that for W-2%Re the value of the SRO parameter in W–Re is negative (−0.0162) for the binary alloy and then increases and acquires positive values of 0.0066, 0.0512 and 0.1586 for ternary alloys containing 0.05%, 0.1% and 0.2% of vacancies, respectively.

Figure 8.

Figure 8. (a) Re–Vac cluster obtained from MC simulations for W-2%Re alloy with 0.1% concentration of vacancies at 300 K. (b) Void decorated by Re atoms predicted by MC simulations for W-2%Re alloy with 0.2% vacancy concentration at 1600 K. Green, red and blue spheres indicate W atoms, Re atoms and vacancies, respectively.

Standard image High-resolution image

In comparison with experimental APT observations [20], showing Re clustering in W-2%Re alloys induced by self-ion irradiation at 773 K, our MC simulations describing W-2%Re–0.1%Vac alloy at 800 K exhibit that the diameter of Re precipitates is close to 2.5 nm. This predictions correlates well with the observation of rhenium clusters of  ∼3 nm diameter found experimentally. The fact that Re clusters are coherent with the tungsten matrix [20] strongly supports the model proposed above, consistent with the view that clustering is associated with high, many orders of magnitude above equilibrium, local concentration of vacancies, and relatively small elastic distortion of the lattice. Another prediction that precipitates form in under-saturated Re–W alloys at high temperatures (up to 1600 K), in the form of voids decorated by rhenium atoms, is not only consistent with TEM observations of a neutron-irradiated W-1.4Re alloy [15] but it also agrees with the observation of small (∼2 nm) voids forming above 1200 °C in ion-irradiated tungsten [81].

6. Summary

We have developed a first-principles model for segregation and phase decomposition occurring in under-saturated binary W–Re alloys under irradiation. The model is based on the treatment of binary alloys as effective ternary alloys (A, B, vacancy), simulated under the constraint of a fixed vacancy concentration. Using an extensive DFT database combined with CE and quasi-canonical Monte-Carlo simulations, we investigated Re precipitation in irradiated bcc Re–W alloys. Depending on the Re/vacancy ratio, we found that precipitates formed at high temperature and low Re concentrations have the form of small voids decorated with Re atoms. At higher Re and low vacancy concentration, precipitates adopt a sponge-like structure, containing a high amount of rhenium approaching 40 at$ \% $ Re. Predictions derived from the model agree well with experimental observations of precipitation occurring in both neutron and self-ion irradiated tungsten–rhenium alloys.

It would be desirable to generalize the present first-principles CE approach by taking into account self-interstitial atom defects as well as mixed W–Re and Re–Re dumbbells. This may allow the study of alloy evolution under irradiation by extending the 'ABV' to the 'ABVI' Ising model, as was recently proposed in [82]. Preliminary assessment of values of the SRO parameter in the low Re concentration range of W–Re alloys based on the ABVI model, however, shows that the main result of our current investigation, namely the prediction that vacancies rather than self-interstitial atom defects are primarily responsible for the occurrence of anomalous segregation of solutes in dilute alloys under irradiation, remains both quantitatively and qualitatively correct.

Acknowledgments

This project has received funding from the European Unions Horizon 2020 research and innovation programme under grant agreement number 633053, and from the RCUK Energy Programme (Grant Number EP/I501045). JSW was funded by the Accelerated Metallurgy Project, which is co-funded by the European Commission in the 7th Framework Programme (Contract NMP4-LA-2011-263206), by the European Space Agency and by the individual partner organizations. To obtain further information on the data and models underlying this paper please contact PublicationsManager@ukaea.uk. The views and opinions expressed herein do not necessarily reflect those of the European Commission. DNM would like to acknowledge the International Fusion Energy Research Centre (IFERC) for providing access to a supercomputer (Helios) at Computational Simulation Centre (CSC) at Rokkasho (Japan) and would like to thank Jaime Marian for fruitful discussions. The authors are grateful to S G Roberts and G D W Smith for the discussions that stimulated this work.

Please wait… references are loading.
10.1088/1361-648X/aa5f37