3.1 MICROSTRUCTURAL CHARACTERIZATION
The thermodynamic calculations depicted in Fig. 2a-e illustrate the FCC single-phase field for the Ni50Pd50 Ni63.2V36.8, Cr33Co33Ni33, Cr30Co30Ni30Pd10, Cr30Co30Ni30V10 alloys. Subsequently, Fig. 3 displays the X-ray diffraction (XRD) patterns, for the recrystallized samples, confirming the exclusive presence of the FCC phase in all samples and the correspondent inverse pole figure (IPF-Z) maps from EBSD analysis showing the microstructure of the alloys.
For comparison, XRD for pure Ni is shown in Supplementary Fig. 1. The SEM-EDS elemental distribution maps showcased in Supplementary Figs. 2–6 revealed the compositional homogeneity of the recrystallized microstructure at the micro scale on each alloy.
3.2 INFLUENCE OF ATOMIC VOLUME
Analysis of the diffractograms in Fig. 3a-b shows that adding Pd and V to pure Ni causes a noticeable shift of all reflections to lower 2θ values compared to pure Ni (Supplementary Fig. 1). A similar trend is seen in Fig. 3d-e with the addition of Pd and V to the CrCoNi system (Fig. 3c). According to Bragg's law [32], these shifts to lower 2θ values are related to an increase in atomic radius (atomic volume).
Accurately determining atomic volume in solid solutions with multiple atoms is challenging due to variations in atomic radii depending on the environment of which a certain atom is inserted. For example, α-Fe (BCC) and γ-Fe (FCC) have different radii (1.239 Å and 1.287 Å, respectively), as do α and β Ti in different structures (1.475 Å and 1.432 Å, respectively) [26]. Pure Ni (FCC) has a radius of 1.243Å, but a radius of 1.257 Å in a binary FCC solution with Cr (30 at. %) [8]. To address this, we propose using the concept of apparent atomic volume (Aav), which relies on the principles that the unit cell represents the entire system and contains atoms in the same proportion as the solid solution.
Considering a system with a composition of Cr33Co33Ni33, forming an FCC solid solution, the atomic volume of the unit cell is crucial because it must include the three elements in the same stoichiometry as the system. Thus, each element—Cr, Co, and Ni—occupies approximately one third of the atomic volume in the cell. For an FCC cell, and assuming the simplified spherical geometry for the atoms, the following relationship is established:
$$\:0.74=\frac{QV}{{a}^{3}}$$
1
Where \(\:Q\) is the number of atoms in a FCC unit cell, \(\:V\) is the volume of one atom, and \(\:{a}^{3}\) is the unit cell volume. Thus, taking into account that an FCC cell is occupied by four atoms, the \(\:{A}_{av}\) might be expressed as:
$$\:{A}_{av}=\frac{\text{0,74}{a}^{3}}{4}$$
2
Therefore, for solid solutions, \(\:{A}_{av}\) provides an effective and functional measure of the volume occupied by atoms in a unit cell, regardless of the quantity of atoms composing the solid solution. It should be noted that atoms are not spheres and the volume occupied by a single atom might be the entire 1/4th of the FCC unit cell if the atom is now viewed in this way. However, the simplification of taking atoms as spheres is convenient to directly apply the results to many equations in the literature that consider the atomic radii’ as the intrinsic measurement for the atomic volume and either option, atomic radii or volume, will lead to the same outcome using these models.
Furthermore, a distorted lattice may result if a significant difference in atomic sizes is present. The distortion generated in the crystalline lattice (\(\:\delta\:\)) can be estimated by the model proposed by Zhang et al.[27]:
$$\:\delta\:=\:\sqrt{\sum\:_{i=1\:\:\:}^{n}{c}_{i}\:{\left(1-{r}_{i}/\stackrel{-}{r}\right)}^{2}}$$
3
Where n is the total number of elements in solid solution, \(\:{c}_{i}\) is the atomic fraction of the ith element, \(\:{r}_{i}\) is the atomic radius of element i and \(\:\stackrel{-}{r}\) (\(\:\sum\:_{i=1}^{n}{c}_{i}{r}_{i})\) is the average atomic radius. The \(\:\delta\:\) parameter is very common in the HEA literature to describe the difference in atomic radii in solid solutions.
Using the XRD results shown in Fig. 3, the apparent atomic volume (\(\:{A}_{av}\)) was calculated for the investigated materials, detailed in Tables 1 and 2. These tables also present the δ values for pure Ni and the CrCoNi system, including changes from adding Pd and V. For precise δ determination, we used ‘solution atomic radii’. These radii represent the size that an element would have in a face-centered cubic (FCC) system. These values were derived from binary FCC alloys of the element with Ni [8, 33], as detailed in Supplementary Table 2.
Table 1
– Lattice parameter (a), variation in the lattice parameter (Δa), apparent atomic volume (Aav), variation in the apparent atomic volume (ΔAav), and crystalline lattice distortion (δ) for pure Ni, Ni50Pd50, and Ni63.2V36.8 alloys.
Material | a (Å) | Δa* (%) | Aav (Å3) | ΔAav**(%) | δx100 |
Ni | 3.5246 | - | 8.1003 | - | - |
Ni50Pd50 | 3.7396 | 6.10 | 9.6747 | 19.44 | 4.97 |
Ni63,2V36,8 | 3.6094 | 2.40 | 8.6988 | 7.39 | 4.53 |
Table 2
Lattice parameter (a), variation in the lattice parameter (Δa), apparent atomic volume (Aav), variation in the apparent atomic volume (ΔAav), and crystalline lattice distortion (δ) for Cr33Co33Ni33, Cr30Co30Ni30Pd10, and Cr30Co30Ni30V10 alloys.
Material | a (Å) | Δa* (%) | Aav (Å3) | ΔAav**(%) | δx100 |
Cr33Co33Ni33 | 3.5621 | - | 8.3616 | 3.23 | 1.40 |
Cr30Co30Ni30Pd10 | 3.6091 | 1.32 | 8.6971 | 4.01 | 3.13 |
Cr30Co30Ni30V10 | 3.5841 | 0.62 | 8.5176 | 1.87 | 1.94 |
Table 1 reveals that adding Pd to pure Ni increases Aav by 19.44%, much more than the 7.39% increase from adding V. The δ values are also higher with Pd additions. Similarly, Table 2 shows that Pd increases Aav by 4.01% in the CrCoNi system, compared to 1.87% for V, with higher \(\:\delta\:\) values for Pd as well. This larger increase in Aav and δ when adding Pd to both Ni and CrCoNi is due to Pd's larger atomic size compared to V. Consequently, considering only atomic volume differences, alloys with Pd are predicted to exhibit stronger SSS due to greater lattice distortion.
Moreover, it is worth noting that these findings corroborate what has been shown before by other authors [34, 35], indicating that the presence of a greater number of elements in solid solution does not invariably induce larger distortions in the crystal lattice. This is exemplified by the higher values of Aav and δ observed for the binary alloys in comparison to their counterparts in the ternary and quaternary alloy systems.
3.3 INFLUENCE OF ELECTRONEGATIVITY
Based on the assumption that it's possible to predict configurational fluctuations of charge transfer and atomic-level pressure, the model introduced by Oh et al. [24] depends on the average charge transfer of each element in a solid solution, approximated by the local difference in electronegativity. The electronegativity difference between the constituent elements (\(\:\varDelta\:{\chi\:}\)) can be calculated as follows:
$$\:\varDelta\:{\chi\:}=\sqrt{\sum\:_{x}{c}_{x}{\left({{\chi\:}}_{x}-{⟨{\chi\:}⟩}_{element}\right)}^{2}}$$
4
Where \(\:{{\chi\:}}_{x}\) represents the electronegativity of element X, and \(\:{⟨{\chi\:}⟩}_{element}\) stands for the weighted average electronegativity across the element. Using the model proposed by Oh et al. [24], the \(\:\varDelta\:{\chi\:}\) values were calculated for the alloys under investigation and are presented in Table 3 and Table 4. In their model, the authors claim the Allen electronegativity scale should be the one to used. The values of electronegativity used for each element are listed in Supplementary Table 2.
Table 3
Electronegativity difference caused by the addition of Pd and V to pure Ni
Material | δχ |
Ni | - |
Ni50Pd50 | 0.150 |
Ni63,2V36,8 | 0.175 |
Table 4
Electronegativity difference caused by the addition of Pd and V to CrCoNi system
Material | δχ |
Cr33Co33Ni33 | 0.100 |
Cr30Co30Ni30Pd10 | 0.118 |
Cr30Co30Ni30V10 | 0.126 |
Tables 3 and 4 show that the introduction of V into pure Ni results in slightly higher \(\:\varDelta\:{\chi\:}\) values compared to the addition of Pd. Similarly, V addition to the CrCoNi system leads to a higher \(\:\varDelta\:{\chi\:}\) values than the addition of Pd. According to the Oh model [24], alloys with V should exhibit stronger SSS due to the greater \(\:\varDelta\:{\chi\:}\) values. It is important to note that the values presented in Table 4 may vary depending on the electronegativity scale used. For example, based on the electronegativity according to the Pauling scale, the Ni50Pd50 alloy would exhibit the greatest \(\:\varDelta\:{\chi\:}\) values (Supplementary Tables 2 and 3).
In light of the exposition thus far, it becomes clear that the models proposed by Varvenne and Oh contain a fundamental contradiction. According to the Varvenne model, alloys containing palladium (Ni50Pd50 and Cr30Co30Ni30Pd10) should exhibit the highest contribution of SSS. Conversely, the Oh model posits alloys with vanadium addition (Ni63.2V36.8 and Cr30Co30Ni30V10) should have the greatest SSS contribution. Therefore, the following section delves into both models and attempts to clarify the fundamental difference between them.
3.4 PREDICTION OF SSS SOLUTION USING THE EVALUATED MODELS
The Varvenne model evaluates the energy associated with the interaction between a dislocation and a solute atom. This energy calculation is subsequently included in a standard equation to adapt to thermally induced deformation. As a result, the determined interaction energy becomes part of an equation to handle thermally driven deformation. The resulting outcome is a model for the SSS component in the yield strength, which considers the impact of strain rate and temperature. This final formulation offers a direct method for calculating the activation energy needed for dislocation movement (\(\:{\varDelta\:\text{E}}_{\text{b}}\)) and the Peierls stress at absolute zero (\(\:{{\tau\:}}_{0}\)), as expressed in the following equations:
$$\:{{\tau\:}}_{0}=0.051{{\alpha\:}}^{-\raisebox{1ex}{$1$}\!\left/\:\!\raisebox{-1ex}{$3$}\right.}\text{G}{\left(\frac{1+{\nu\:}}{1-{\nu\:}}\right)}^{\raisebox{1ex}{$4$}\!\left/\:\!\raisebox{-1ex}{$3$}\right.}{\text{f}}_{1}\left({\text{W}}_{\text{c}}\right){\left(\sum\:_{\text{n}}\frac{{\text{x}}_{\text{n}}{\varDelta\:\stackrel{-}{V}}_{n}^{2}}{{\text{b}}^{6}}\right)}^{\raisebox{1ex}{$2$}\!\left/\:\!\raisebox{-1ex}{$2$}\right.}$$
5
\(\:{\varDelta\:\text{E}}_{\text{b}}=0.274{{\alpha\:}}^{\raisebox{1ex}{$1$}\!\left/\:\!\raisebox{-1ex}{$3$}\right.}\text{G}{\text{b}}^{3}{\left(\frac{1+{\nu\:}}{1-{\nu\:}}\right)}^{\raisebox{1ex}{$2$}\!\left/\:\!\raisebox{-1ex}{$3$}\right.}{\text{f}}_{2}\left({\text{W}}_{\text{c}}\right){\left(\sum\:_{\text{n}}\frac{{\text{x}}_{\text{n}}{\varDelta\:\stackrel{-}{V}}_{n}^{2}}{{\text{b}}^{6}}\right)}^{\raisebox{1ex}{$1$}\!\left/\:\!\raisebox{-1ex}{$3$}\right.}\) | (6) |
Where \(\:{\nu\:}\) is the Poisson’s ratio, \(\:{\alpha\:}\) (0.123 in this work) represents a constant related to the value of the line tension of the dislocation (Γ = \(\:{\alpha\:}\)G𝑏²), G denotes the shear modulus, b represents the Burger’s vector and Xn represents de fraction of nTh element in solid solution. The functions f\(\:\left({\text{W}}_{\text{c}}\right)\), denoted as \(\:{\text{f}}_{1}\left({\text{W}}_{\text{c}}\right)\) and \(\:{\text{f}}_{2}\left({\text{W}}_{\text{c}}\right)\), are termed to minimize dislocation core-related coefficients. They account for the curved nature of dislocations, which deviate to locate local energy minima; Varvenne employed values of 0.35 and 5.70 for \(\:{\text{f}}_{1}\left({\text{W}}_{\text{c}}\right)\) and \(\:{\text{f}}_{2}\left({\text{W}}_{\text{c}}\right)\), respectively.
In both equations, the pivotal term is \(\:\varDelta\:{\stackrel{-}{V}}_{n}\), denoting the average volumetric misfit per atom. It is computed as the difference between the volume of the nth atom and the average atomic volume within the mixture. The atomic volume is derived from the FCC unit cell's volume divided by four.
The resulting terms from equations 5 and 6 are then inserted into the subsequent equation, delineated as Eq. 7, for the computation of yield strength.
$$\:{\tau\:}_{y}\left(T,\dot{ϵ}\right)={\tau\:}_{0}exp\:\left(-\frac{1}{0.51}\frac{kT}{\varDelta\:{E}_{b}}ln\frac{\dot{{ϵ}_{0}}}{\dot{ϵ}}\right)$$
7
Where k is the Boltzmann constant and \(\:\dot{{ϵ}_{0}}\) is a reference term for the strain rate, which was 10− 3 s− 1 in this work. Hence, the Eq. 7 is a model for the SSS component to the critical resolved shear stress (converted into the respective contribution to the yield strength by the Taylor factor) that incorporates the dependence on strain rate and temperature.
On other hand, Oh model suggests that the SSS component (σss) in multicomponent alloys, comprising atoms from the 3d family, can be elucidated by the electronegativity difference between the constituent elements (Δχ), outlined as follows:
$$\:{\sigma\:}_{ss\:}\propto\:\varDelta\:{\chi\:}\:$$
8
More specifically, Oh model suggests that the SSS could be predicted by the following equation:
$$\:{\sigma\:}_{ss}=\left(4293\pm\:448\right)\varDelta\:\chi\:+\left(84\pm\:37\right)\:MPa\:$$
9
Therefore, taking into consideration Eq. 7, Fig. 4a presents the values of the SSS contribution alloys under study. The variation of AAv was used to express the difference in atomic volume in each alloy. Moreover, the σss value for pure Ni was considered to be 0. Moreover, based on Eq. 9, Fig. 4b illustrates the SSS contribution values as per Oh model.
Figure 4a shows that adding Pd to pure Ni results in higher SSS, compared to the addition of V, with Ni50Pd50 demonstrating the highest SSS. Similarly, in the CrCoNi system, Cr30Co30Ni30Pd10 exhibits higher SSS than with V addition. These results align with Varvenne model, which indicates that greater differences in AAv between constituent atoms lead to stronger SSS.
On other hand, Fig. 4b shows that adding V to pure Ni results in higher SSS than adding Pd, with Ni63.2Pd36.8 demonstrating the highest SSS. Similarly, in the CrCoNi system, Cr30Co30Ni30V10 alloy shows higher SSS compared to Pd addition. These findings support Oh’s model, which suggests that greater Δχ differences result in stronger SSS. However, the model proposed in Eq. 9 should be applied with caution, once the intrinsic dispersion associated with the model (indicated in Fig. 4b) could led to a widely distribution of σss values. Therefore, the inherent dispersion associated with the model and the choice of the electronegativity scale to be employed may lead to inconclusive results when applying Oh's model, potentially limiting its practicality.
3.5 EXCLUDING THE EFFECT OF GRAIN BOUNDARY STRENGTHENING
In fact, Oh et al. [24] support their perspective by examining the σy and σuts values of the Ni63.2V36.8 alloy, in contrast to the equiatomic CrCoNi and CrCoNiMnFe alloys (all with similar grain sizes). Notably, the Ni63.2V36.8 alloy demonstrates higher yield strength and ultimate tensile strength values than the other alloys.
However, in order to conduct an accurate analysis of the isolated contribution of the σss, it is essential to disregard the influence of grain refinement strengthening. This approach was adopted in the present study. According to the Hall-Petch model, the strengthening attributed to grain boundaries can be elucidated by Eq. 10.
$$\:\sigma\:={\sigma\:}_{0}\:+\:{K}_{hp}{d}^{-\text{0,5}}$$
10
In this context, σo represents the friction lattice stress, Khp is the Hall-Petch coefficient, and d denotes the grain size. Theoretically, σ0 includes various contributions to material strength, but since only substitutional Solid Solution is considered here (all alloys are single-phase), we focus solely on the SSS contribution to σ0. Additionally, Eq. 10 can analyze the influence of grain size on hardness (H) by replacing σo with H0.
Therefore, examining Eq. 10 reveals that the same d can produce different hardness and σy values across materials with different values for Khp. Consequently, assessing the σss contribution based on absolute σy and σuts for fine-grained materials may be misleading. Hence, to isolate the σss component, the grain boundary strengthening contribution Khp must be removed. This can be done using Hall-Petch plots, measuring hardness (or σy) for the same material at different grain sizes, as shown in Fig. 5 for the alloys assessed in this study. The σy values were extracted from several stress-strain curves for each alloy, shown in Supplementary Fig. 7.
Figures 5a-b present hardness and σy values, respectively, plotted against d− 0.5 for the alloys under study. Moreover, the dashed lines in both figures represent a linear regression (y = ax + b) based on the data for each material. Thus, by extrapolating the curve to an infinite grain size (d− 0.5→0), it is possible to estimate the intrinsic hardness (or σy) of each alloy and, consequently, the SSS.
Analyzing the data in Figs. 5a-b, it is evident that both Ho and σ0 values for the binary alloys are higher than those for pure Ni (Ho = 68.6 Hv and σ0 = 14 MPa ) [35, 36]. Additionally, the binary alloy with Pd shows higher Ho and σ0 values compared to those with V. This trend is also observed in the quaternary alloys, where those containing Pd and V exhibit higher Ho and σ0 values compared to the Cr33Co33Ni33 alloy. These results suggest that adding V and Pd enhances the SSS of both pure Ni and the CrCoNi system. However, while V-based alloys show higher hardness and σy, their H0 and σ0 values indicate that V contributes less to SSS compared to Pd.
Examining the Khp values (extracted from hardness and σy Hall-Petch plots) for the alloys it is revealed that Ni63.2V36.8 has a significantly higher Khp than Ni50Pd50. Similar trends are observed in Cr30Co30Ni30Pd10 and Cr30Co30Ni30V10 alloys, with the V-based alloy showing a higher Khp than the Pd-based alloy. This indicates that grain boundary strengthening is much more pronounced in V-based alloys. Also, Ni63.2V36.8 exhibited the highest grain boundary strengthening among the studied alloys. Thus, directly comparing absolute hardness and σy values across different alloys is not feasible due to the varying contributions of grain boundary strengthening.
Figures 6a-b show the projected hardness and σy of these alloys, assuming a uniform grain size of 10 µm, based on their respective Khp values. The Ni63.2V36.8 alloy has higher hardness and σy values at this grain size compared to other alloys. Despite Ni50P50 exhibiting the lowest Khp for σy (Fig. 5b), its σy value (Fig. 6b) exceeds those of the ternary and quaternary alloys due to its highest σ0 for a large range of grain sizes. This demonstrates that analyzing hardness or σy values alone can lead to incorrect conclusions regarding the SSS effect in different metallic alloys.
The Khp values should be approached with caution, as the same constant can exhibit different values depending on whether hardness or σy is being analyzed. As illustrated in Figs. 5a-b, except for the Ni50Pd50 alloy, the Khp values associated with hardness are consistently higher than those for σy. This discrepancy arises because hardness measurements are conducted in a localized region of the material, where the plastic deformation is restricted to only a few grains. Additionally, during hardness testing indentation, significant work-hardening effects occur, directly influencing the hardness values obtained. Specifically, the Ni50Pd50 alloy exhibits an unusually low Khp for hardness (i.e., the hardness remains unchanged with increasing grain size), potentially indicating a low work-hardening coefficient for this alloy. Consequently, the mechanical behavior of the Ni50Pd50 alloy requires more in-depth analysis.
Figure 7 presents a comparison between the experimental and calculated σ0 values using the Varvenne and Oh models, in Fig. 7a and 7b, respectively. The analysis indicates that, although both models fail to precisely predict the SSS, the theoretical σss values predicted by the Varvenne model exhibit a closer correlation to the experimental data.
Hence, the analysis of the data presented so far shows not only that the Varvenne model is more suitable for predicting SSS, but also that the atomic volume difference between species in a solid solution is the more significant factor in SSS of FCC alloys. However, the Varvenne model has not yet been able to accurately predict the SSS component for all the alloys in study. This suggests that additional factors should be considered when developing models to predict SSS. For example, the dislocation line constants might change from case to case or maybe the non-linear variations in the atomic volume of each element.
Interestingly, the results indicate that alloys exhibiting the highest δ values also demonstrated the highest SSS values, primarily observed in the binary alloy systems. (Supplementary Fig. 8). This indicates that, while solid solutions comprising a greater diversity of atoms may not always display the greatest lattice distortion, increased lattice distortions are likely responsible for the greater contribution to solid solution strengthening.
It was also noted that, contrary to current literature on HEAs [5, 11], SSS is not always the predominant strengthening mechanism, as shown in Fig. 8. The SSS contribution values, obtained using supplementary Equations 1 and 2, indicate that SSS becomes predominant (SSS > 50%) only above a critical grain size, while grain boundary strengthening (GBS) dominates below this size. This critical grain size varies for each alloy and is detailed in Supplementary Table 4. Notably, the Ni50Pd50 alloy, due to its relatively low Khp value, has an extremely small critical grain size (Supplementary Fig. 9), making SSS nearly the sole hardening mechanism. These findings highlight that much work has yet to be done to fully understand SSS for different alloys, and suggest the Ni50Pd50 alloy promising material for future fundamental metallurgical studies due to its minimal mechanical property sensitivity to grain growth.