3.1. Microstructure
Figure 1 illustrates the evolution of the microstructure in Al-5Ni alloys with different levels of samarium addition. The base alloy (
Figure 1a) exhibits characteristic dendritic α-Al grains, while incremental Sm incorporation induces progressive microstructural refinement. As can be seen from
Figure 1, a distinct trend of secondary dendrite arm spacing (SDAS) reduction can be observed. The secondary dendrite arm spacing (SDAS) of Al-5Ni-xSm alloys was measured by using a digital image analysis software (Nano measurement) [
13], is presented in
Figure 2. As depicted in this figure, the secondary dendrite arm spacing (SDAS) measures 45.7 ± 8.2 μm in the absence of Sm. When the additional amount of samarium is 0.1 wt.%, the SDAS decreases markedly to 23.4 ± 4.4 μm. Upon increasing the Sm content to 0.2 wt.%, the SDAS shows a slight increment to 25.4 ± 3.8 μm, followed by no discernible change at 0.3 wt.% Sm. An increase in SDAS to 27.5 ± 3.9 μm is observed when the Sm content reaches 0.5 wt.%. Results indicate that the SDAS reaches the minimum value when the addition amount of Sm reaches 0.1 wt.%. However, this effect gradually diminishes as the Sm content continues to increase. This microstructural refinement, accompanied by homogenization of α-Al and eutectic structure (α-Al + Al
3Ni) morphology, provides favorable conditions for mechanical property enhancement. The observed microstructural modification indicates that Sm plays a distinct role as a phase morphology modifier in the Al-Ni alloy system.
X-ray diffraction analysis of Al-5Ni-xSm alloys (
Figure 3) revealed consistent α-Al and Al
3Ni phase formation across all compositions, with Sm-containing variants exhibiting additional Al
3Sm intermetallic phases (
Figure 3a). High-resolution examination of Al
3Ni’s strongest diffraction peaks (
Figure 3b) showed invariant peak positions regardless of Sm content, indicating preserved lattice parameters. This crystallographic stability suggests Sm primarily interacts through surface segregation at Al
3Ni interfaces rather than bulk solid solution formation. The combined XRD evidence supports a microstructure evolution mechanism involving independent Al
3Sm precipitation alongside unmodified Al
3Ni phase formation.
3.2. Crystallization
Figure 4 presents the cooling curves of Al-5Ni alloys with varying Sm concentrations, accompanied by their corresponding first and second derivative curves. The solidification parameters are tabulated in
Table 1. The base Al-5Ni alloy exhibited α-Al nucleation temperature (
TN) at 656.8 °C, eutectic growth temperature (
TEG) at 640.4 °C, and solidus temperature at 621.6 °C. Sm-modified alloys displayed significant solidification characteristic parameter modifications, including a secondary exothermic peak following the (α-Al + Al
3Ni) eutectic reaction (
Figure 4b,c). This appearing peak, attributed to the formation of the Al
3Sm phase as confirmed by XRD analysis and the Al-Sm phase diagram [
19], shows that the nucleation temperatures of the Al
3Sm phase are 615.2 °C and 606.3 °C for the 0.1 wt.% and 0.5 wt.% Sm-containing alloys, respectively. The progressive intensity enhancement of this peak with increasing Sm content reflects elevated Al
3Sm phase fraction.
Analysis of solidification parameters (
Table 1) reveals a non-monotonic dependence of α-Al nucleation temperature (
TN) on Sm concentration in Al-5Ni-xSm alloys.
TN decreases from 656.8 °C (0 wt.% Sm) to 643.5 °C (0.1 wt.% Sm), followed by a marginal increase to 649.7 °C at higher Sm levels. And the nucleation temperature of the eutectic structure (
TEG) exhibits the same trend. The decrease in the initial nucleation temperature implies a greater degree of undercooling to promote the nucleation, thereby refining the grain size.
The most interesting results obtained from the first-derivative curves show that recalescence occurs in both the Al-5Ni alloy and Al-5Ni-0.1Sm alloy (see
Figure 4a,b for details). Recalescence is a specific thermal effect observed during the solidification of metals or alloys, manifesting as a temporary temperature rise on the cooling curve [
24]. Fundamentally, it results from the latent heat released during solidification, which causes a transient local temperature increase. The presence of recalescence is indicated by the region where the first-order derivative of the cooling curve is greater than zero. In the Al-5Ni alloy, recalescence occurs at the onset of the (α-Al + Al
3Ni) eutectic reaction. This indicates that during the initial stage of the eutectic reaction, nuclei of the eutectic structure form and grow rapidly, releasing substantial latent heat of crystallization. When the latent heat release rate exceeds the system’s heat dissipation rate, a localized temporary temperature rise creates a “recalescence peak”. During recalescence, the undercooling decreases as temperature rebounds, reducing the nucleation driving force and slowing the crystallization rate, thereby forming a large amount of coarse and brittle eutectic structures.
The addition of Sm elements to the Al-5Ni melt can eliminate the recalescence phenomenon during the eutectic reaction. This elimination helps hinder the development of (α-Al + Al3Ni) eutectic structures, thereby leading to the refinement of eutectic morphology. During solidification, Sm undergoes solute redistribution, accumulating preferentially at solid–liquid interfaces. This interfacial segregation enhances liquid-phase compositional gradients, increasing composition undercooling that promotes eutectic nucleation while simultaneously restricting eutectic growth through solute redistribute effects. The coupled mechanisms of nucleation site multiplication and diffusion-limited growth kinetics facilitate the development of refined, uniformly distributed eutectic microstructures.
Meanwhile, a pronounced recalescence phenomenon of α-Al was also observed in the Al-5Ni-0.1Sm alloy. Due to the addition of 0.1% Sm, the initial nucleation temperature (
TN) decreases from 656.8 °C to 643.5 °C, generating a higher degree of undercooling. When the undercooling reaches the critical value, massive nucleation and growth of α-Al occur, releasing substantial latent heat of crystallization. However, compared with the Al-5Ni alloy, the growth duration of α-Al is significantly shortened, leading to finer and more uniform α-Al in the Al-5Ni-0.1Sm alloy. As shown in
Figure 4c, the recalescence phenomenon of both α-Al and eutectic structures disappears in the Al-5Ni-0.5Sm alloy. Compared with the Al-5Ni-0.1Sm alloy, the growth duration of α-Al increases and its nucleation temperature rises, which leads to the coarsening of α-Al.
The difference (
ΔT) between the liquidus temperature (
TN) and solidus temperature (
TS) is the non-equilibrium solidification range of the alloy melt. As shown in
Table 1, the addition of 0.1wt% Sm to the Al-5Ni alloy reduces the non-equilibrium solidification range of the melt from 35.2 °C to 28.3 °C, indicating improved fluidity of the alloy. However, when the Sm content increases to 0.5 wt.%, the non-equilibrium solidification range expands to 43.4 °C, leading to deteriorated fluidity of the melt.
3.3. Effect of Sm Addition on Mechanical Properties
Figure 5 displays the tensile stress–strain profiles and mechanical property statistics for Al-5Ni alloys with varying Sm concentrations. The base alloy (0 wt.% Sm) demonstrates yield strength (YS) of 93.2 ± 1.5 MPa, ultimate tensile strength (UTS) of 130.5 ± 2.2 MPa, and elongation (EL) of 19.5 ± 0.5%. Sm-modified alloys exhibit significant mechanical enhancement, with optimal performance achieved at 0.1 wt.% Sm: YS = 125.5 ± 1.4 MPa (+34.7%), UTS = 172.4 ± 2.5 (+32.1%), and EL = 20.0 ± 0.5% (+2.5%). Further Sm addition to 0.5 wt.% induces progressive property reduction (YS = 112.5 ± 1.3 MPa, UTS = 158.4 ± 1.9 MPa, EL = 18.5 ± 0.5%), demonstrating a concentration-dependent strengthening mechanism. This non-monotonic response suggests an optimal Sm threshold for mechanical performance enhancement in the Al-Ni system. Combining the microstructural observations in
Figure 1 and the cooling curve thermal analysis, the addition of Sm reduces the nucleation temperatures of α-Al and eutectic structures, increases the degree of undercooling, and promotes nucleation. Meanwhile, the solute redistribution effect hinders the growth of eutectic structures, which is corroborated by the variation trend of SDAS in
Figure 2. Sm addition facilitates the refinement of dendritic α-Al grains and optimization of the size, morphology, and distribution of eutectic structures, thereby significantly enhancing the material strength. Excessive incorporation of Sm induces a poisoning effect, suppressing the effect of lowering the nucleation temperature. This leads to the coarsening of eutectic structures (as illustrated in
Figure 1e), thereby causing a gradual decline in mechanical properties.
This research determines the Q value, a holistic quality metric, to evaluate the strength and elongation of Al-5Ni-xSm alloys. The Q value is denoted in the form of flow:
For aluminum alloy, the value of
a is 150 MPa [
25,
26]. For Al-5Ni alloys containing Sm at concentrations of 0%, 0.1%, 0.2%, 0.3%, and 0.5%, the respective Q values are 323.5 MPa, 367.2 MPa, 360.5 MPa, 356.8 MPa, and 348.1 MPa. Clearly, the Q value for the Al-5Ni-0.1Sm alloy stands at 367.2 MPa, marking it as the highest in the sample set. The figure is approximately 13.5% higher compared to the Sm-free alloy. Therefore, incorporating 0.1% Sm may lead to an impressive combination of strength and elongation performance.
Figure 6 illustrates the tensile fracture morphologies of Al-5Ni-xSm alloys. Examining the structure of tensile fractures allows for a clearer understanding of how these alloys break. In the sample without Sm (as clearly shown in
Figure 6a), distinct quasi-cleavages and tear ridges are observable. These features are characteristic of brittle fracture. Moreover, the rod-like eutectic Al
3Ni phases are readily identifiable on the fracture surface. The experimental findings align well with the relatively inferior tensile performance of the Sm-free alloy. For the Al-5Ni-0.1Sm alloy,
Figure 6b vividly illustrates how its fracture surface displays multiple small, uniformly distributed dimples. This morphological feature indicates a transition in the fracture mechanism from brittle to ductile. As the Sm concentration escalates, the dimples grow rougher. Additionally, the fracture surfaces display visible tear ridges, as illustrated in
Figure 6c,d, which suggests a mixed fracture morphology characterized by quasi-cleavage and micropore aggregation. The findings from the analysis of the tensile fracture morphology align well with the results of the tensile experiments. Of all the alloys examined, the Al-5Ni-0.1Sm alloy exhibits the most advantageous tensile characteristics.
The mechanical performance of near-eutectic Al-5Ni-xSm alloys is governed by the morphological characteristics and spatial distribution of the (α-Al + Al
3Ni) eutectic architecture. In Sm-free alloys (
Figure 1a), the eutectic structure exhibits a coarse, irregular morphology with uneven dispersion, creating stress concentration sites that detrimentally impact mechanical integrity. This interconnected network of brittle Al
3Ni phases facilitates preferential crack propagation pathways under applied loads, as evidenced by the progressive elongation reduction at elevated Sm concentrations. Sm addition modifies the eutectic structure through morphological refinement and homogenized dispersion (
Figure 1b–d), enhancing load transfer efficiency while mitigating stress concentrations. The resultant microstructure optimization, achieved through Sm-induced eutectic modification, demonstrates a strengthening–toughening synergy that significantly improves tensile performance despite minor ductility trade-offs at higher Sm levels. This microstructural engineering approach effectively decouples conventional strength–ductility conflicts in near-eutectic Al-Ni systems through controlled interface design.
3.4. Effect of Sm Addition on the Thermal Conductivity
Figure 7a presents the thermophysical properties of Al-5Ni-xSm alloys, illustrating a linear decrease in specific heat capacity from 0.878 J/(g·K) (0 wt.% Sm) to 0.874 J/(g·K) (0.5 wt.% Sm), attributed to the comparatively low heat capacity of Sm J/(g·K). Concurrently, density increased from 3.009 g/cm
3 to 3.033 g/cm
3 due to Sm’s higher atomic mass.
Figure 7b depicts the properties of thermal diffusivity, electrical conductivity, and thermal conductivity. The thermal conductivities are calculated based on thermal diffusivity, heat capacity, and density, in accordance with Equation (1). It is observable that the changes in electrical and thermal conductivity of Al-5Ni-
xSm alloys, when varying in Sm content, are similar to the fluctuations in thermal diffusivities. Initially, there is an increase in the electrical and thermal conductivities, alongside a rise in thermal diffusivity, followed by a gradual decline. The alloy without Sm exhibits the minimal levels of thermal diffusivity and electrical and thermal conductivity, recorded at 82.55 ± 0.38 mm
2/s, 31.33 ± 0.15 MS/m, and 218.1 ± 1.0 W/(m·K), respectively. The alloy of Al-5Ni-0.1Sm exhibits maximum thermal diffusivity, electrical and thermal conductivity, achieving 82.93 ± 0.29 mm
2/s, 31.50 ± 0.11 MS/m, and 219.3 ± 0.8 W/(m·K). The data indicates a rough rise of 0.5%, 0.5%, and 0.6% in comparison to the original Al-5Ni alloy. With an increase in Sm content to 0.5%, there is a gradual reduction in thermal diffusivity, electrical conductivity, and thermal conductivity, dropping to 81.56 ± 0.39 mm
2/s, 31.08 ± 0.15 MS/m, and 216.3 ± 1.0 W/(m·K).
Research on Al-5Ni-xSm alloys indicates that their increased heat transfer rate is due to the poor solubility of Ni and Sm within the α-Al matrix. Sm exhibits moderate capability for microstructural modification, and the variation in thermal conductivity is intricately correlated with the evolution of the alloy’s microstructure. The trace additions of Sm refine (α-Al + Al
3Ni) eutectic structures, concomitantly increasing electron conduction pathways. However, excessive Sm incorporation (0.5 wt.%) gives rise to eutectic coarsening (as shown in
Figure 1e), thereby deteriorating thermal conductivity. Previous studies [
27] confirm that rare earth elements purify the α-Al matrix through impurity segregation effects. However, detailed interfacial chemistry analysis (to be presented subsequently) is required to fully elucidate these conductivity enhancement mechanisms.
3.5. Promotion Mechanism of Sm Addition on Al-5Ni Alloys
The SEM results, as provided in
Figure 8, disclose the effect of samarium addition on the compositions and morphologies of secondary phases in the Al-5Ni alloy. The morphologies of secondary phases in the Al-5Ni alloy are displayed in
Figure 8a. As seen, two kinds of morphologies of secondary phases can be found in the alloy. The secondary phase near the boundary of the α-Al grain, marked as phase ‘A’ and termed as ‘Eutectic boundary phase’, is flake-like. The needle-like phases, labeled as phase ‘B’ and termed as ‘Eutectic internal phase’, are far from the boundary of the α-Al grain. These secondary phases with different morphologies are extensively generated in the Al-Ni hypoeutectic alloy with different preparation methods [
28,
29].
The morphological difference between the eutectic internal phase and eutectic boundary phase is closely connected to the solidification process. Previous studies [
30,
31] establish 27.6 vol.% as the critical secondary phase fraction for morphological transitions. For this study, the theoretical volume fraction of the Al
3Ni phase, about 8.3 vol.%, is computed by the thermodynamic software JmatPro 7.0. During the early phase of solidification, the Ni element’s concentration remains fairly consistent, indicating that the Al
3Ni phase’s phase volume fraction is less than 27.6 vol.%. Observing the rod or needle-shaped Al
3Ni phase within the eutectic grain is straightforward. During the last phase of solidification, the residual liquid exhibits a high Ni element concentration due to the redistribution of solutes and non-equilibrium solidification. This occurrence reveals that the Al
3Ni phase’s related phase volume fraction exceeds 27.6 vol.%. Hence, the eutectic boundary phase manifests a morphology resembling flakes.
Introducing the Sm element might lead to the creation of a flake-like eutectic boundary phase due to adsorption poisoning. Highly active Sm element will be enriched at the boundary of the liquid/solid interface, inhibiting the migration and diffusion of Ni atoms. Therefore, there are more remaining liquids with high Ni concentration, which facilitates the formation of flake-like eutectic boundary phase. The relative line scanning results for the secondary phase in the Al-5Ni-0.1Sm alloy are provided in
Figure 8b. The experimental results demonstrate that the Sm element mainly distributes alongside the flake-like eutectic boundary phase. Combined with the XRD analysis, the Sm-rich layer may be the Al
3Sm phase. In other words, the flake-like Al
3Ni phases are coated with Al
3Sm phases.
Figure 8c displays the morphology of the Al-Sm phase. By referring to the XRD patterns, the Al-Sm phase can be determined as Al
3Sm, indicating that the excessive addition of samarium will lead to the precipitation of Al
3Sm in the Al-5Ni matrix.
Figure 9 presents EPMA elemental mapping of the Al-5Ni alloy. Illustrated in
Figure 9c,d, the minimal presence of Fe and Si impurities renders these elements almost undetectable in the initial alloy, a fact supported by the experimental findings.
Figure 10 depicts the EPMA mapping outcomes for the Al-5Ni-0.1Sm alloy. Importantly,
Figure 10c,d distinctly show clusters of Fe and Si located at the edges of grains. The finding suggests that introducing a minor quantity of the Sm element markedly reduces the solubility of Fe and Si impurities in the α-Al matrix. Furthermore, an examination of
Figure 10e makes it evident that Sm predominantly accumulates in the region surrounding the eutectic Al
3Ni phase.
In metallic systems, thermal transport arises from coupled electron and phonon contributions, with total thermal conductivity (
λ) comprising electronic (
λe) and lattice (
λg) components as per established models [
22]. Among these, the electronic thermal conductivity plays a more dominant role compared to the lattice component. According to the Wiedemann–Franz rule [
32,
33], a proportional correlation is observed between electronic thermal and electrical conductivity. Concurrently, the Mattissen–Flemming rule [
34,
35] suggests that the main contributors to electrical resistance are solid solute atoms, contaminants, and imperfections. The associated expression is shown as follows:
The value of
is 2.665 × 10
−8 Ω·m, as reported in reference [
34].
represents the electrical resistivity caused by precipitates. The negligible effect of minor Sm additions on precipitate volume fraction permits the exclusion of precipitation-mediated resistivity contributions. The coefficients (
,
,
, and
), represent the individual contributions of grain boundaries, dislocations, vacancies, and solid solute atoms to the electrical resistance.
is the grain boundary fraction per unit volume,
is the dislocation density,
is the vacancy concentration and
is the concentration of solid solute atom
i. According to Miyajama’s study [
36], for aluminum,
is approximately 2.6 × 10
−15 Ω·m
2,
is around 2.7 × 10
−25 Ω·m
3, and
is about 2.6 × 10
−8 Ω·m/at.%. Considering the minimal magnitudes of
,
, and
, the impact of grain boundaries, dislocations, and vacancies on electrical resistance is trivial. As a result, the electrical resistance is greatly influenced by the levels of atoms in solid solutions. Mulazimoglu [
37] and Cooper [
38] quantified solute-specific resistivity coefficients through systematic alloy characterization, deriving empirical relationships between aluminum’s electrical resistivity and solute concentrations. These correlations are mathematically expressed as
According to the aforementioned formulas, each 1 wt.% rise in Si or Fe content results in a roughly 26.4% or 63.4% reduction in the electrical resistance of pure Al, respectively. In reality, trace amounts of Fe and Si elements are invariably present in commercially pure aluminum. Moreover, as Jiang et al. [
8] and Zhang et al. [
27] emphasized that rare earth elements present in molten aluminum possess pronounced interfacial adsorption capabilities. Such a trait grants them exceptional purification capabilities for the α-Al matrix. In this study, the same observation (as illustrated in
Figure 9c,d) reveals that Sm can also extract Fe and Si elements from the Al-5Ni melt. This extraction process effectively reduces the solid solubility of these impurities within the alloy. Moreover, as illustrated in
Figure 1, the incorporation of Sm results in notable refinement of both α-Al grains and the eutectic (α-Al + Al
3Ni) structure. This microstructural refinement contributes significantly to improved thermal conductivity. The mechanism driving this phenomenon has been thoroughly detailed in our earlier work [
13]. The reduction in grain size and optimization of the eutectic arrangement increases intergranular contact surface, thereby expanding pathways for electron flow. As a result, Sm-induced refinement directly promotes higher thermal conductivity in the near-eutectic Al-5Ni alloy. However, the excessive addition of samarium in the Al-5Ni alloy results in the precipitation of Al
3Sm within the Al matrix (as illustrated in
Figure 8c). As an intermetallic phase in the Al-5Ni alloy, Al
3Sm acts as a scattering source that hinders the transmission of free electrons. Consequently, as illustrated in
Figure 7b, the TC property of the Al-5Ni alloy decreases with increasing samarium content.