Next Article in Journal
Effect of Temperature and Ceramization Atmosphere on the Structure and Microstructure of Boron-Modified SiBOC Materials
Previous Article in Journal
Influence of Alloying Elements on the Phase Structure, Stress–Strain Behavior, and Fracture Toughness of Ni3Sn: A First-Principles Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Sn Content on Wettability and Interfacial Structure of Cu–Sn–Cr/Graphite Systems: Experimental and First-Principles Investigations

State Key Laboratory of Advanced Processing and Recycling of Nonferrous Metals, Lanzhou University of Technology, No. 287 Langongping Road, Lanzhou 730050, China
*
Author to whom correspondence should be addressed.
Materials 2025, 18(8), 1793; https://doi.org/10.3390/ma18081793
Submission received: 11 March 2025 / Revised: 28 March 2025 / Accepted: 2 April 2025 / Published: 14 April 2025

Abstract

:
The co-addition of chromium (Cr) and tin (Sn) is known to enhance the wettability between copper (Cu) and graphite (Cgr), but the effect of Sn content remains poorly understood. This study aims to systematically investigate the influence of Sn content a (a = 0, 10, 20, 30, 40, 50, 80, 99 at. %) on the wettability, interfacial structure, surface/interface energy (σlv/σsl), and adhesion behavior of the Cu–aSn–1Cr/Cgr system at 1100 °C. The experimental results show that as the Sn content increases, the equilibrium contact angle (θe) of the metal droplet shows a non-monotonic trend; the thickness of the reaction product layer (RPL, consisting of Cr carbides (CrmCn)) gradually increases, accompanied by a decrease in the calculated adhesion work ( W ad cal ). A “sandwich” interface structure is observed, consisting of two interfaces: metal||CrmCn and CrmCn||Cgr. Sn content mainly affects the former. At metal||CrmCn, Sn exists in various forms (e.g., Cu–Sn solid solution, CuxSny compounds) in contact with CrmCn. To elucidate the wetting and bonding mechanisms of metal||CrmCn, simplified interfacial models are constructed and analyzed based on first-principles calculations of density functional theory (DFT). The trend of theoretically calculated results (σmetal and Wad) agrees with the experimental results (σlv and W ad cal ). Further analysis of the partial density of state (PDOS) and charge density difference (CDD) reveals that charge distribution and bonding characteristics vary with Sn content, providing the microscopic insight into the nature of wettability and interfacial bonding strength.

1. Introduction

Graphite (Cgr) has low density, high thermal conductivity, low thermal expansion coefficient, high-temperature resistance, and self-lubricating properties [1]. Copper (Cu) is commonly used to prepare composites with Cgr due to its economic feasibility, easy availability of resources, and excellent alloy solubility. With the continuous development of Cu/Cgr composites, their mechanical and functional properties have attracted increasing attention [2,3,4]. However, the applications of Cu matrices are limited by their inherent low hardness and strength, as well as poor creep and corrosion resistance. Recently, alloying copper matrix to prepare Cu/C composites has become a promising research hotspot. The addition of Sn to the Cu matrix significantly enhances its strength, wear resistance, and corrosion resistance through solid solution strengthening. Moreover, Sn improves the functional properties of both Cu/Cgr composites and brazed joints, such as electrical and thermal conductivity. These performance improvements are closely dependent on the Sn content [5,6,7,8,9]. Therefore, optimizing the Sn content is critical for enhancing the performance of Cu–Sn/Cgr composites and solder joints. It is expected to demonstrate significant potential in a wide range of advanced applications, including construction machinery, electronic packaging, sliding conductive components, and high-efficiency thermal management systems.
Good wettability and strong interfacial bonding are essential for brazed joints and Cu/C composites. However, the intrinsic wettability of Cu–Sn/Cgr is poor [10], and the bonding is weak due to van der Waals forces, making reliable connections difficult. Two approaches are commonly used to improve wettability: surface modification of carbon materials and copper-based alloying. The latter is more practical, as surface modification often struggles with coating uniformity and reliability. Copper-based alloying involves adding carbide-forming elements such as Ti, Cr, V, Mn, Nb, W, Mo, or Zr [11] to enhance interfacial reactivity. Among them, Cr is particularly effective. Most studies focus on Cu–Cr/C systems, with limited attention to the role of Sn. References [12,13] showed that adding a small amount of Cr significantly improves Cu/C interfacial wettability. However, Cr’s high reactivity causes rapid formation of dense carbides (CrmCn), which hinders infiltration. Lin et al. [14] found that adding Sn to Cu–Cr alloys enhances wettability on porous Cgr by modifying Cr reactivity and reducing droplet surface energy, resulting in deeper infiltration. Thus, Sn content can effectively regulate wetting behavior. Moreover, Sn facilitates low-temperature joining, expanding the processing window for Cu/C composite preparation via melt infiltration.
Hsieh et al. [15] showed that Sn addition in Cu–Ti alloys lower the melting point and affects alloy microstructure and properties. As alloy complexity increases, interfacial structures become more intricate, affecting the interfacial bonding properties (IBPs) (abbreviations in this paper can be found in Supplementary Materials). Optimizing wettability and IBPs typically requires extensive trial-and-error experiments, which are time-consuming and inefficient. Thermodynamic calculations offer qualitative insights, but the underlying mechanisms remain unclear. Moreover, key parameters related to IBPs are environment-sensitive and difficult to measure directly. First-principles calculations based on density functional theory (DFT) have been widely used to study wettability and bonding behavior without empirical inputs [16]. A commonly used approach is the solid/solid slab model, which approximates the liquid/solid interface by neglecting atomic thermal vibrations at high temperatures [17]. When the trends from theoretical and experimental results are consistent, correction methods can be applied to improve accuracy [18]. Despite its advantages, this method is limited by computational cost, as only small-scale systems (typically 100–200 atoms) can be modeled. DFT studies are typically limited to simple interfaces involving pure metals or compounds and cannot easily handle large supercell models or multicomponent alloys. Meanwhile, extensive research exists on non-reactive metal/carbide interfaces (e.g., Al/TiC [19], Co/WC [20], Ag/Ti (C, N) [21], etc.).
This study investigates the wettability and interfacial behavior of Cu–aSn–1Cr alloys on Cgr using a modified sessile drop method, which minimizes pre-reaction and oxidation during heating. The interfacial structure and elemental composition of the samples are characterized using a scanning electron microscope (SEM) equipped with energy-dispersive spectroscopy (EDS). Typical interfacial configurations observed experimentally are then selected for simplified DFT modeling, aiming to uncover the mechanisms by which alloying elements and their concentrations influence wettability and IBPs. The results offer both experimental and theoretical guidance for designing Cu–Sn/C composites and solder alloys, and provide a reference for incorporating other non-reactive, low-melting-point elements (e.g., gallium, germanium, silver, indium, lead) into the Cu matrix.

2. Experimental and Calculation

2.1. Experimental Materials and Methods

The Cgr substrate has a purity of >99.9%, a size of 20 mm × 20 mm × 5 mm, an ash content of <40 ppm, a density of ~1.85 g/cm3, a porosity of 13%, and a surface roughness (Ra) of approximately 30–50 nm. The metallic materials used include Cu and Sn foils with a purity of 99.99% and Cr powder with a particle size of 400 μm. The Cu–aSn–1Cr alloys (a = 0, 10, 20, 30, 40, 50, 80, 99 at. %) are prepared based on atomic ratios. High-purity Ar gas (99.999%) is repeatedly introduced into the KDH-300B mini vacuum arc furnace, equipped with Ti getters, to purify the furnace chamber. To ensure the uniformity of alloy melting, the wrapped metal is first melted into spheres in an Ar atmosphere before using a robotic arm to flip the alloy and continue the melting process. This procedure is repeated 3–5 times. After the melting, the sample is cooled to room temperature under flowing Ar at a rate of ~15 °C/min. Based on the modified sessile drop method, the metal is first stored in a stainless-steel tube outside the furnace. The furnace is evacuated to ~6–7 × 10⁴ Pa, and the chamber temperature is then heated to the preset experimental temperature of 1100 °C at a rate of 20 °C/min. After the dynamic vacuum is maintained at a stable level (~3–4 × 10−4 Pa), the metal is dropped onto the Cgr surface through an open alumina tube. Simultaneously, a high-resolution charge-coupled device (CCD) camera is used to record the dynamic spreading and wetting process of the molten metal, with a holding time of 30 min. The specific experimental setup and procedure are referenced in Ref. [22]. The contact angle (CA) in the photographs is measured using the Surface Meter software (SM, version 1.1.0.1, developed by NB Scientific Instruments Co., Ltd., Ningbo, China), and the variation curve of CA with time is obtained. The modified sessile drop method, compared to the traditional sessile drop method, effectively avoids the pre-reaction between the melt and the substrate during the experimental heating process. To prevent oxidation of the metal droplets, the droplets are first cooled to 600 °C under high vacuum, followed by a cooling process at a rate of 15 °C/min with flowing Ar until room temperature. The interface structure, microstructure, morphology of reaction products, and microregional elemental composition of the samples’ cross-section are analyzed and characterized using a Scanning Electron Microscopy (SEM, FEG 450, Thermo Fisher Scientific, Waltham, MA, USA) equipped with an Energy Dispersive Spectrometer (EDS, Oxford Instruments, Abingdon, UK) with a spot size of approximately 2 μm. The samples for high-resolution transmission electron microscope (HRTEM, JEOL, JEM–F200, Tokyo, Japan) are prepared using a focused ion beam (FIB; Helios G4 PFIB, Thermo Fisher Scientific, Waltham, MA, USA).

2.2. Calculation Methods

The calculations in this paper are performed using the Cambridge Sequential Total Energy Package (CASTEP, version 22.11, BIOVIA, San Diego, CA, USA) software, which is based on first-principles density functional theory [23]. The Perdew–Burke–Ernzerhof (PBE) function within the Generalized Gradient Approximation (GGA) is used to describe the exchange-correlation energy of Cu ([Ar] 3d104s1), Sn ([Kr] 4d105s25p2), Cr ([Ne] 3s23p63d54s1), and C ([He] 2s22p2), respectively. The OTFG ultrasoft pseudopotentials are employed to model the interaction between atomic nuclei and valence electrons [24], and the energy cutoff is set as 500 eV. The energy fluctuation threshold for static self-consistent calculations is set to 2 × 10−5 eV/atom. The Broyden–Fletcher–Goldfarb–Shanno (BFGS) algorithm is used for structure optimization, and the iterative self-consistent convergence criterion is that the energy fluctuation per atom is less than 1 × 10−5 eV. The interatomic force is less than 0.05 eV/Å. The maximum stress is below 0.05 GPa, and the atomic displacement fluctuation is less than 1.0 × 10−3 Å [25]. Using the density of 0.04 Å−1 Monkhorst-Pack k-point for Brillouin zone sampling for the bulk, surface, and interface. The bulk models are presented in Figure S1 in Supplementary Materials. In the interface calculations, the interactions among surface layer atoms play a crucial role, while fixing the topmost and bottom layers of atoms ensures the convergence of energy.

3. Results

3.1. Wetting Experiment

Eustathopoulos et al. [26] have systematically analyzed the equilibrium contact angle (CAeq) and adhesion work (Wa(Cu–XSn/Cv)) of Cu–XSn alloys on the vitreous carbon (Cv) surface as a function of Sn content, as shown in Figure 1a. The results show that CAeq increases first and then decreases with increasing Sn, while Wa(Cu–XSn/CV) exhibits an inverse trend. In this study, after adding 1 at. % Cr, the measured θe (Figure 1b) follows a different trend—decreasing first, then increasing, and finally decreasing again. When Sn reaches 20 at. %, θe drops to 12°. According to Dupre’s definition of the adhesion work W ad cal at solid/liquid interfaces, as shown in Equation (1) [27]:
W ad cal   = σ lv   +   σ sv     σ sl
where σlv represents the surface energy of the liquid, σsv represents the surface energy of the solid, and σsl represents the solid/liquid interface energy. However, the value of W ad cal , σsv, and σsl cannot yet be accurately measured through experiments. Therefore, we used experimentally measured wetting parameters to estimate the adhesion work W ad cal based on the well-known Young–Dupre Equation (2) [28]:
W ad cal   = σ lv   ( 1   +   cos   θ e )
where σlv represents the surface energy of the Cu–Sn metallic melt, and its values under different Sn contents are listed in Table 1, based on data reported in Ref. [29]. θe is expressed in radians. The measured values of θe are shown in Figure 1c–j. Substituting σlv and θe into Equation (1) yields the W ad cal , and the variation of W ad cal with Sn content is shown as the blue curve in Figure 1b. Compared with the Cu–XSn/Cv system, θe is generally smaller and Wa (0.8–2.4 J/mol) is significantly larger than that of the Cr-free case (0.1–0.3 J/mol). This suggests that Cr addition enhances wettability and interfacial bonding. Lin et al. [14] attributed this wetting behavior to the decrease surface energy (σlv) of the alloy melt with increasing Sn content and a shift in reaction products from Cr7C3 (major) + Cr3C2 (minor) to Cr7C3 (minor) + Cr3C2 (major). This indicates that Sn affects both melt properties and solid/liquid interface energy (σsl). Unfortunately, they did not provide an in-depth explanation for this. Beyond σlv, Sn content also influences the alloy’s density, viscosity, and mixing enthalpy (ΔHmixCu–Sn). According to References [30,31], both the density and viscosity of the Cu–Sn melt decrease with increasing Sn content, but do not explain the trend in θe. Interestingly, the variation of ΔHmixCu–Sn [32] aligns well with θe (blue curve, Figure 1b). There may be a relationship between ΔHmixCu–Sn and θe, as ΔHmixCu–Sn influences the σlv and σsl of the melt droplet. However, no relevant studies to date have demonstrated a relationship between the two. This may be related to the fact that the complex mechanisms of action within the alloy and at the interface between the alloy and the reaction products have not yet been solved. Therefore, this study will attempt to explain the relationship between alloying elements and their content variations with wettability and IBPs, and their potential mechanisms in Section 4.
Figure 1. Variation of wetting parameters: (a) the variation of CAeq and Wad(Cu–XSn/Cv) with Sn content (X at. %) of Cu–Sn alloy on the CV at 1150 °C [26]; (b) variation of θe, W ad cal , and ΔHmixCu–Sn [32] with Sn content; (cj) measured results of θe with Sn content.
Figure 1. Variation of wetting parameters: (a) the variation of CAeq and Wad(Cu–XSn/Cv) with Sn content (X at. %) of Cu–Sn alloy on the CV at 1150 °C [26]; (b) variation of θe, W ad cal , and ΔHmixCu–Sn [32] with Sn content; (cj) measured results of θe with Sn content.
Materials 18 01793 g001

3.2. Interfacial Structure

The solidified cross-section of the Cu–aSn–1Cr/Cgr system after wetting is examined using SEM (backscattered mode) and EDS. As shown in Figure 2(a-1–d-1), the interfacial transition zone exhibits a typical “sandwich” structure, composed of two sub-interfaces: metal||reaction product layer (RPL) and RPL||Cgr. EDS analysis near the interface (Figure 2(a-4–c-4)) confirms that the RPL mainly consists of Cr7C3 and Cr3C2. With increasing Sn content, the dominant phase in the RPL gradually shifts from Cr7C3 (metallic) to Cr3C2 (covalent). At 10 at. % Sn (Figure 2(a-1)), Cr precipitates are observed on the RPL (point 3, Figure 2(d-4)), due to Cr’s limited solubility (0.98 at. %) in Cu–10Sn alloy at 1100 °C, which is lower than the nominal 1 at.%. As a result, a portion of Cr becomes supersaturated dissolution and precipitates out. For detailed calculations, please refer to the References [34,35,36]. When Sn reaches 20 at. %, the solubility increases to 1.96 at. %, and the Cr precipitates disappear, indicating complete dissolution into the melt (Figure 2(b-1)). The RPL becomes thicker and more continuous, increasing from ~1.4 μm (Figure 2(a-3)) to ~3 μm (Figure 2(d-3)). The alloy side exhibits various phases in contact with the RPL, including Cu–Sn solid solutions, CuxSny intermetallic (e.g., Cu10Sn3, Cu3Sn, Cu6Sn5), and Sn/Cr segregation regions. EDS analyses in Figure 2(a-2,d-2) and points 1, 2, and 4–9 confirm that the Cu and Sn concentrations dominate, while Cr is nearly undetectable inside the alloy, justifying subsequent model simplifications. When the Sn content increases to 20 at. %, the alloy side transforms into a network structure (Figure 2(b-1)). Figure 2(b-2) shows an enlarged view near the interface of Figure 2(b-1), where the RPL thickness increases to 2.66 μm, and at the dark gray (point 4) and light gray (point 5) locations on the alloy side, Cu–Sn = 3.3:1, which may correspond to the Cu10Sn3 phase. To determine the alloy phase, FIB sampling is conducted at the location marked by the red rectangle in Figure 2(b-1). The HRTEM characterization confirms that the alloy phase is indeed Cu10Sn3, as shown in Figure 2(b-3). From Figure 2(c-1,c-2,d-1,d-3), it can be seen that Cu3Sn and Cu6Sn5 are the key intermetallic compounds precipitated. In addition, regions of Sn segregation may appear near the interface, as shown in Figure 2(c-3). These observations suggest multiple interfacial contact forms—Cu–Sn solid solution, CuxSny (Cu10Sn3, Cu3Sn, and Cu6Sn5) intermetallic, Cr precipitates, and Sn segregated zones—between the alloy and RPL (Figure 2(a-1–c-3)), corresponding to areas 1–6 in Figure 2(a-1,b-1,c-1–c-3), which likely influence the IBPs, as discussed in Section 4.
Figure 3 shows the top view SEM morphology (secondary electron mode) and EDS results of the interface RPL after etching the alloy droplets using FeCl3 solution (mass ratio of FeCl3–H2O = 3:1). At 10 at. % Sn, dendritic Cr-rich precipitates appear in the RPL (Figure 3(a-1,a-2)), confirmed by EDS (point 1, Figure 3(d-4)) as mainly Cr with traces of C, O, Fe, and Cl—the latter likely from residual etchant. Figure 3(a-3–d-3) show the morphology near the triple line (TL). EDS analysis in Figure 3(a-4,b-4) indicates that the RPL is primarily composed of Cr, with small amounts of Cu, Sn, and C. The precipitates on the RPL are mainly concentrated at the center of the interface, with fewer near the TL. As the Sn content increases to 20 at. %, the dendritic Cr-rich phases disappear, replaced by numerous polygonal CrmCn phases (Figure 3(b-1)). EDS analysis of point 3 suggests that the particles (point 3) in Figure 3(b-2) have a composition similar to that of the point 2 and may be the Cr7C3 phase. A precursor film enriched in Cr and C elements ahead of the RPL (Figure 3(b-3,b-4)), possibly due to Sn-induced Cr migration toward the TL. With 50 at. % Sn, no precipitates are seen on the RPL (Figure 3(c-1)). The RPL consists of nanoparticles in orthorhombic Cr3C2 (point 4) and hexagonal Cr7C3 (point 5) phases, as shown in Figure 3(c-1,c-2). At 80 at. % Sn, there are some cracks distributed in the continuous and dense CrmCn layer (Figure 3(d-2)), possibly caused by stress concentration during cooling. EDS of point 6 suggests central RPL regions are Cr7C3, while river-like structures at the TL front (point 7) show Cr/C atomic ratios close to Cr3C2. These results indicate that Sn content not only alters the RPL morphology but also affects Cr diffusion and interfacial phase distribution. Further phase confirmation can be found in the X-ray diffraction results in Ref. [14].

4. Discussion

From the above, the interface structure analysis indicates that the Cu–aSn–1Cr/Cgr reactive wetting system exhibits a “sandwich” structure, composed of two interfaces: metal||CrmCn and CrmCn||Cgr. For the latter, although this study does not directly investigate the bonding mechanism at the CrmCn||Cgr interface, previous work (Ref. [37]) has reported that bonding at the Cr3C2||diamond interface primarily arises from hybridization between C–2p states of diamond (001) and C–2p or Cr–3d states of Cr3C2 (001). Given the structural and electronic similarity between diamond and Cgr, it is reasonable to expect a similar bonding mechanism at the CrmCn||Cgr interface in our system. Therefore, detailed discussion of this interface is omitted here. In this work, we focus on the metal||CrmCn interface, as it plays a more dominant role in determining the overall wettability behavior observed in our experiments. As illustrated in Figure 4, Sn and Cr may interact with the RPL surface in multiple forms, including Cu–Sn solid solutions, CuxSny compounds, and Sn/Cr segregation.
From Section 3.2, it can be seen that when the Sn content is low, Cr undergoes supersaturation and precipitates out, and when the Sn content is high, according to the Cu–Sn binary phase diagram [38], Sn segregation occurs, and CuxSny varies with the Sn content. The contact phenomenon with the RPL surface may be caused by the strong interactions among Sn/Cr and CuxSny with CrmCn. The EDS analysis in Figure 2 of Section 3.2 shows that the main precipitate phases in contact with RPL are Cu10Sn3, Cu3Sn, and Cu6Sn5, which may result from reaction formulas Equations (3)–(10) occurring during the cooling process:
10Cu + 3Sn → Cu10Sn3
3Cu + Sn → Cu3Sn
Cu + 5Sn → Cu6Sn5
Cu10Sn3→ Cu + 3Cu3Sn
Cu6Sn5 + 9Cu → 5Cu3Sn
5Cu3Sn → 9Cu + Cu6Sn5
2Cu3Sn + 3Sn → Cu6Sn5
L + Cu3Sn → Cu6Sn5
Li et al. [39] reported that adding 1~2 wt. % Cr does not significantly affect the growth of CuxSny. In the initial reaction stage, Cu10Sn3, Cu3Sn, and Cu6Sn5 are formed (Equations (3)–(5)), where Cu10Sn3 is a metastable phase that readily transforms into Cu3Sn (Equation (6)). When there is less Sn, Cu6Sn5 may convert to Cu3Sn (Equation (7)). Conversely, when there is more Sn, Cu3Sn will transform into Cu6Sn5, and Cu3Sn will decompose to form Cu6Sn5 (Equation (8)). Sn atoms continuously diffuse into Cu3Sn and also react with Sn to produce Cu6Sn5 (Equation (9)). Furthermore, the outer layer of Cu3Sn can also transform into Cu6Sn5 through a peritectic reaction (Equation (10)) [40]. To investigate the effect of the aforementioned forms of metal presence on the interfacial properties and to illustrate the specific mechanism of alloying elements, this paper focuses on the metal||CrmCn interface. The electronic structure calculations of Cr7C3 and Cr3C2 indicate that the chemical bonds in these two compounds exhibit complex characteristics with a mixture of metallic, covalent, and ionic properties [41,42]. The difference between the two was primarily reflected in the stronger metallic of Cr7C3 than that of Cr3C2. Naidich [11] reported that the metallic properties of ceramics enhance the metal/ceramic adhesion. However, Cr3C2 is more stable than Cr7C3. Considering the computational cost (Cr7C3 requires more atoms for modeling than Cr3C2), the most stable Cr3C2 is chosen for this study. Based on the interface structure obtained from the experiments, this study approximates several typical configurations at the metal||Cr3C2 interface as follows: the effects of Cu–Cr and Cu–Sn solid solutions are simplified to the issue of Cr/Sn mono-/multi-atomic doping in the Cu matrix. The precipitation of Cr/Sn on the interface is simplified to the issue of contact between pure Cr/Sn and Cr3C2. The case with high Sn content is simplified to the influence of CuxSny (the effect of CuxSny on the interface also needs to be explored). Simplified interface models are established for pure metals (Cu/Sn/Cr), Cu-doped Cr/Sn single/multi-atom systems, and CuxSny (Cu10Sn3, Cu3Sn, Cu6Sn5) with Cr3C2.

4.1. DFT Calculations

4.1.1. Surface Properties

The lattice parameters of Cu, β–Sn, Cr, CuxSny (Cu10Sn3, Cu3Sn, Cu6Sn5), and Cr3C2 are shown in Table S1 in the Supplementary Materials. The optimized lattice constant values are compared with the calculated values from References [37,43,44,45,46], and there is a slight deviation of ± 2% in the obtained lattice parameters. The close-packed surfaces of Cu (111), β–Sn (100), Cr (110), and Cr3C2 (001) have stable atomic structures and the lowest surface energies [37,47,48,49], and the surface models are shown in Figure S2a–d. Here, the reason for selecting the surface with the lowest surface energy and high adhesion work for the calculation is explained below. The atomic radius difference between Cu and Sn atoms is large, and the lattice distortion is small when the doping quantity is low. Conversely, at higher doping quantities, the lattice distortion is larger. Therefore, to investigate the effect of high Sn content, this paper uses the CuxSny compounds shown in Figure 2(b-2,c-1,c-2) of Section 3.2 for comparison. The surface energy of CuxSny is investigated and found to be rarely reported in the literature, so the surface energy of CuxSny is calculated before building the interface model. According to Ref. [50], it is reported that the maximum Miller indices are 2 and 3 for noncubic and cubic crystals, respectively, when calculating the surface energy. Due to the limitation of the calculation conditions, in this paper, we only investigate the surface properties of (100), (010), and (001) of Cu10Sn3, Cu3Sn, and Cu6Sn5. Based on the symmetry and periodicity of the structures, surfaces with the same structure but different indices are combined for analysis, and the constructed surface structure models are presented in Supplementary Materials Figure S2e–j. σmetal is commonly defined as the energy required per unit area to form a new surface, which can be used to describe the stability of the surface, and is given by Equation (11) [51]:
σ metal = E surface n E bulk n i μ i + n j μ j 2 A surface
σmetal is the surface energy of the metal (J/m2); Esurface, n, Ebulk, and Asurface are the total energy of the metal surface, the number of unit cells, the energy of the unit cell, and the area of the surface, respectively. The doped surface model also needs to consider the number of dopant atoms and their chemical potential, where ni and nj are the number of dopant atoms and doped atoms, respectively. The μi and μj are the chemical potential of the dopant atoms and doped atoms, respectively. The σmetal values for Cu (111), Sn (100), Cr (110), Cu6Sn5 (100), and Cr3C2 (001) are close to the reported results in the References [37,45,52,53,54], as presented in Table S2 in Supplementary Materials. The doping cases are as follows: single Cr atom doped Cu (111) and Sn (100) surfaces, named Cu55Cr1 and Sn49Cr1, respectively. Multiple Sn atoms doped Cu (111) surface with up to four dopant atoms, named Cu56-kSnk (k = 1, 2, 3, 4), the doping position is shown in Figure 5(b-2). As shown in Table S2, the doping of a small number of Cr atoms increases the σmetal of Cu (111) but has little effect on the σmetal of Sn (100). When multiple Sn atoms are doped, the σ (Cu56-kSnk) decreases slightly with increasing Sn atoms, while the σmetal of CuxSny decreases significantly with the increase in Sn content. This trend is consistent with the changes in σlv of Cu–Sn alloys reported in Ref. [29]. Cu10Sn3 (100), Cu3Sn (001) type2, and Cu6Sn5 (001) type2 with lower σmetal values are selected for the study in the case of higher Sn content.

4.1.2. Interfacial Details

Before establishing the interface models of pure metals (Cu/Sn/Cr), Cu-doped single/multiple Cr/Sn atoms, and CuxSny with Cr3C2, the mismatch degree δ of the interfaces is calculated. Ref. [55] shows that when δ < 8%, the two surfaces can be considered well-matched, which allows the interface model to achieve good convergence. The formula for calculating δ is shown in Equation (12) [56]:
δ = a 1   -   a 2 0.5 ( a 1 + a 2 ) × 100 %
where a1 and a2 (Å) represent the lengths of the supercells that need to be matched in the same direction. a 1 a 2 is the difference between the matched lattice parameters, and 0.5   ( a 1 + a 2 ) is their average. The δ–values for the supercell interface models are all below 8%, which satisfies the matching criteria. The calculated δ–values are summarized in Table S3.
Based on this, interface models are constructed for Cu (111)||Cr3C2 (001), Sn (100)||Cr3C2 (001), Cr (110)||Cr3C2 (001), and the three compound interfaces Cu10Sn3 (100), Cu3Sn (001), and Cu6Sn5 (001) with Cr3C2 (001). These are denoted as Cu||Cr3C2, Sn||Cr3C2, Cr||Cr3C2, and CuxSny||Cr3C2 (CuxSny = Cu10Sn3, Cu3Sn, and Cu6Sn5), respectively, and are shown in Figure 5(a-1–a-3,c-1–c-3). A vacuum layer of 15 Å is added along the c-axis in all interface models to eliminate the effect of periodic boundary conditions.
To study the effect of atomic doping, 14 and 12 possible substitution positions are considered for single Cr or Sn atoms in Cu||Cr3C2 and Sn||Cr3C2 models, respectively, as shown in Figure 5(b-2). The doping difficulty at each site is evaluated by calculating the formation energy Esub [57] (details in Supplementary Materials). The configurations with the lowest Esub are selected and denoted as Cu55Cr1||Cr3C2 and Sn49Cr1||Cr3C2. For multiple Sn atom doping, four low Esub positions are selected, labeled as Cu56-kSnk||Cr3C2 (k = 1, 2, 3, 4).

4.1.3. Interface Properties

The adhesion, stability, and wettability in the interfacial properties are related to the adhesion work and the surface/interfacial energy of the metal and Cr3C2. Therefore, this section primarily analyzes the effects of typical interface structures on interfacial properties at the atomic level by calculating the theoretical adhesion work Wad and the interfacial energy γint. Wad is obtained by Equation (13) [55]:
W ad = E metal + E C r 3 C 2 E int A int
where E metal is the energy of the metal part, E C r 3 C 2 is the energy of Cr3C2, Eint is the total energy of the interface model, and Aint is the area of the interface. When Wad is larger, the interface is more stable and the interfacial bond strength is higher. The interfacial equilibrium distance of the interface is determined based on the maximum value of the Wad, as detailed in Figure S3 in Supplementary Materials. The γint is calculated as shown in Equation (14) [55]:
γ int = σ metal + σ Cr 3 C 2 W ad
where σCr3C2 is the surface energy of the Cr3C2. A lower γint indicates a more stable interface, as wetting is preferred to be carried out at interfaces with lower values of γint, which is often used as a basis for determining model parameters. The minimum value of σCr3C2 (001) is 3.99 J/m2. σmetal and Wad vary with the composition of the metal. For a stable interface to form between σmetal and σCr3C2, γint must be positive. If γint is negative, the interface is unstable. The lower the γint value (closer to zero), the more stable the interface. To ensure the value of γint is the minimum, the surface/interface models with the smallest difference of (σmetalWad) are chosen for calculation as much as possible. However, selecting a surface/interface model with a minimum (σmetalWad), γint > 0, and low δ requires extensive calculations. Therefore, the calculations in this paper utilize a maximum Wad and a relatively low σmetal.
(1)
Single Sn/Cr atom and multiple Sn atoms doped Cu||Cr3C2 interface
Figure 6 shows the variation of Wad (Figure 6a) and γint (Figure 6b) for a single Sn/Cr atom doping at different positions of Cu||Cr3C2. From Figure 6a, it can be seen that the magnitude of Wad is in the order of Sn||Cr3C2 (2.60 J/m2) < Cu||Cr3C2 (3.05 J/m2) < Cr||Cr3C2 (6.20 J/m2). This suggests that the precipitation of Cr on RPL, as shown in Figure 3(a-2) of Section 3.2, plays a role in enhancing the interface bonding. The Wad of Cu||Cr3C2 is slightly higher than the value of Wad (2.16 J/m2) in Ref [33]. Comparing the Wad of Cu55Cr1||Cr3C2 and Cu55Sn1||Cr3C2, it is clear that the influence of Sn/Cr on the Wad at positions 5–12 of the Cu||Cr3C2 is much greater than that at positions 1–4. Thus, the further the doping site is from the Cr3C2 surface, the smaller the effect. As shown in Figure 6a, except for positions 1–4, Sn doping at other sites decreases the Wad of the interface. From Figure 6b, it can be seen that the Cr||Cr3C2 is more stable and possesses the lower γint (0.97 J/m2) compared to Cu||Cr3C2(2.31 J/m2) and Sn||Cr3C2(1.83 J/m2). Different Sn doping sites have less effect on the γint of Cu||Cr3C2, whereas Cr doping significantly reduces the γint of Cu||Cr3C2 and enhances the stability of the interface. The Wad and γint at sites 13 and 14 in Figure 6 correspond to Sn atoms occupying the Cr and C sites in Cr3C2. It can be seen that the doping at these two sites significantly decreases the Wad of the Cu||Cr3C2 and increases the γint, resulting in the interface being unstable. Therefore, the possibility of Sn occupying sites in Cr3C2 is relatively low. In addition, the Wad of Sn49Cr1||Cr3C2 (3.99 J/m2) is higher than that of Sn||Cr3C2 (2.60 J/m2), and the γint of Sn49Cr1||Cr3C2 (0.84 J/m2) is lower than that of Sn||Cr3C2 (1.83 J/m2), indicating a significant improvement in interface stability.
(2)
CuxSny||Cr3C2 interface
After being compared to Cu||Cr3C2, Sn||Cr3C2, Cr||Cr3C2, and Cu56-kSnk||Cr3C2, the Cu10Sn3||Cr3C2 (Wad = 1.28 J/m2, γint = 3.58 J/m2), Cu3Sn||Cr3C2 (Wad = 1.09 J/m2, γint = 3.93 J/m2), and Cu6Sn5||Cr3C2 (Wad = 0.96 J/m2, γint = 3.79 J/m2) interface possesses lower Wad and higher γint. The Wad gradually decreases, with increasing Sn content in CuxSny increases. Although the γint fluctuates, they are overall relatively high, which indicates that the interface between CuxSny and Cr3C2 is unstable. However, once CuxSny interacts with Cr3C2 to form a CuxSny||Cr3C2 interface structure, it will be unfavorable for interfacial bonding. Yu et al. [10] studied the interfacial properties of CuxSny with diamond and the results indicated that Cu6Sn5 (100) with diamond (111) had a stronger hybridization, lower γint, and better wettability compared to Cu (100) and Cu3Sn (100). CuxSny had different effects on the CuxSny||diamond and CuxSny||Cr3C2 interfaces. Therefore, when considering the adjustment of Sn content to improve the wettability of Cu–aSn–1Cr/Cgr or enhance the mechanical and functional properties of the Cu matrix, the effects of CuxSny on Wad and γint should also be taken into account. This provides theoretical guidance for the design and property optimization of compositions in Cu–aSn–1Cr/Cgr and similar systems.
(3)
Comparison of theoretical and calculated values
Table 1 has the W ad cal obtained in Section 3.1 from experimental values of θe, σlv [29] and solid–liquid interfacial energy σsl; σsl can be obtained from Equation (15) [58,59]:
cos θ e = σ sv σ sl σ lv
The value of σsv of Cr3C2 is 4.13 J/m2 [37]. The a (at. %) values are obtained by converting the Sn content in the metal surface model according to the atomic ratio. Similarly, the σmetal, Wad, and γint obtained from the theoretical calculations are displayed in Table 2. Comparing Table 1 and Table 2, although there is a certain deviation between the theoretically calculated values and the experimentally calculated values, there is a trend of decreasing σmetal and σlv, as well as Wad and W ad cal with the increase in a. The values of σmetal and σlv are close to each other. However, some theoretically calculated values (Wad and γint) are larger than the experimentally calculated values ( W ad cal and σsl), such as the Cu||Cr3C2 and Sn||Cr3C2 interfaces, which results in an inability to back extrapolate θe and thus hinders wettability predictions. The reason may be that Wad and γint are based on first-principles calculations at 0 K, where the atoms are in equilibrium positions, whereas the experimentally calculated values are obtained from measurement and calculation after heating to 1100 °C. The theoretically calculated values of Wad are smaller than W ad cal , and γint is larger than σsl for Cu56-kSnk||Cr3C2 and CuxSny||Cr3C2, probably because the composition of the RPL in the experiment includes not only Cr3C2 but also Cr7C3, which affects the experimental values and hence the larger values.

4.1.4. Interfacial Electronic Structure

The bonding of interface atoms depends on the configuration of the metal||CrmCn interface. To better explore the interactions between atoms at the metal||CrmCn interface, this section calculates the changes in the partial density of states (PDOS), charge density difference (CDD), and Mulliken population of the interface after relaxation and reveals the variation mechanism of the IBPs from an electronic perspective (charge transfer and distribution).
(1)
Partial density of states
Figure 7 shows the PDOS of the above interfacial structures near the Fermi energy level (EF). The PDOS of the electronic orbitals of the relaxation state system can be used to determine the origin of the electronic orbitals that contribute to interatomic interactions in each system. In Figure 7, the upper and lower parts of each subplot represent the PDOS of the metal and the Cr3C2, respectively. The PDOS is divided into s, p, and d orbitals based on the orbitals in which the valence electrons are located, and the total is the sum of all orbitals of the metal||Cr3C2 part. Figure 7a–c correspond to the PDOS for the interfaces Cu||Cr3C2, Sn||Cr3C2, and Cr||Cr3C2, respectively. In the Cu||Cr3C2 interface (Figure 7a), the main contributions stem from the electronic interactions of the Cr–d and Cu–d orbitals. The density of states (DOS) of the d orbital electrons of Cu and Cr3C2 are mainly located in the range of −6eV~−0.9eV and −7.5eV~5.5eV. The range of DOS distribution of the d orbital electrons of Cr3C2 is generally consistent with that reported in Ref. [37]. Similarly, for the Cr||Cr3C2 interface, the d orbital contributes to the main electronic state (Figure 7c), where the Cr–d orbital of Cr hybridizes with the Cr–d orbital of Cr3C2. The metallic bonds formed at the interface of Cu||Cr3C2 and Cr||Cr3C2 provide the main role for adhesion at the interface. The DOS of Sn is mainly contributed by the Sn–p orbital electrons (Figure 7b), which are mainly distributed in the range of −5eV~9eV, with a small contribution from the d orbital electrons. The Sn||Cr3C2 interface is mainly contributed by metallic bonds formed by Sn–p orbital and Cr–d orbital electrons. The bonding strength Cr||Cr3C2 > Cu||Cr3C2 > Sn||Cr3C2 is due to the metallic bonds formed by pd orbital hybridization being weaker than those formed by dd orbital hybridization, and thus Sn||Cr3C2 is lower. Comparing Figure 7a,c, the DOS peak for Cr||Cr3C2 shows dd orbital hybridization over a wider energy range, which has more homomorphic electrons compared to Cu||Cr3C2, and thus Cr||Cr3C2 is higher. Comparing Figure 7a,d, it is observed that the DOS near the EF for the Cu55Cr1||Cr3C2 interface is higher than that of Cu||Cr3C2. This is because Cr doping effectively increases the d orbital electronics near the EF at the Cu||Cr3C2 interface, thereby enhancing the bonding strength between the interface atoms. Comparing Figure 7a,d, it is observed that the DOS of d orbital electronics below the EF for Cu55Cr1 is higher than for Cu. Similarly, Figure 7b,e show that Sn49Cr1 increases the DOS of d orbital electronics below the EF for Sn. Therefore, Cr doping can effectively enhance the interface bonding strength. Figure 7f–i show the effect of doping with different numbers of Sn atoms on the PDOS near the EF at the interface. Comparing Figure 7a,f, it is evident that Sn doping causes the peak value of the DOS for the d orbital electrons of Cu to decrease from 28 eV/states to 25 eV/states. With the increase in Sn doping number k (k = 1, 2, 3, 4), the DOS peak of the d orbital electrons below the EF in the Cu56-kSnk decreases gradually, which indicates that Sn doping weakens the interfacial bonding strength. Figure 7j–l shows the effect of CuxSny on PDOS near the EF at the interface. With the increase in Sn content, the DOS peak widths of the d orbital electrons below the EF of CuxSny are in the order of Cu10Sn3 > Cu3Sn > Cu6Sn5. Compared with Figure 7f–i, the DOS peak widths of the d orbital electrons below the EF of CuxSny are narrower than those of Cu56-kSnk, which indicates that higher Sn content results in the number of d orbital electrons of CuxSny being less than that of Cu56-kSnk, so the bonding strength of the interface is gradually weakened.
(2)
Charge density difference analysis
To further explore the interfacial bonding mechanisms at the atomic scale, charge density difference (CDD) analyses are performed for typical metal||Cr3C2 interfaces with different alloying configurations. The CDD (Δρ) is used to indicate the charge density redistribution and the degree of electron transfer at the interface. The Δρ can be expressed as Equation (16) [52]:
Δ ρ   =   ρ int ρ metal   ρ Cr 3 C 2
where ρint is the total charge density of the metal||Cr3C2 interface, ρmetal, and ρCr3C2 are the charge densities of isolated metal slab and Cr3C2 slab, respectively. The red regions represent charge accumulation (electron gain), while the blue regions indicate charge dissipation (electron loss). The intensity of the color corresponds to the degree of charge accumulation or dissipation. The CDD results (Figure 8) show significant charge accumulation and dissipation at the Cr||Cr3C2 interface compared to Cu||Cr3C2 and Sn||Cr3C2, indicating strong interfacial charge transfer and bond formation, which is consistent with its high adhesion work (Wad). In addition, the Cr-doped interfaces (Cu55Cr1||Cr3C2 and Sn49Cr1||Cr3C2) exhibit enhanced charge redistribution, confirming that Cr doping strengthens interfacial bonding. As the Sn doping content increases (Cu56kSnk||Cr3C2, k = 1–4), the charge accumulation near the interface gradually weakens, and dissipation becomes more prominent, as shown in Figure 8f–i. This trend continues in the CuxSny||Cr3C2 interfaces, as shown in Figure 8j–l, particularly at Cu6Sn5||Cr3C2, where charge dissipation is strongest, corresponding to the lowest Wad observed.

5. Conclusions

In the reactive wetting experiments of the Cu–aSn–1Cr/Cgr system, variations in the Sn content can affect the physical and chemical properties of the alloy, the dissolution/precipitation of Cr elements, the microstructure of the alloy, and the phase composition and morphology of the RPL. These changes further impact the interface configurations and characteristics. Based on the first-principles study, several typical interfacial structures are selected, and simplified interfacial models are constructed. By calculating the Wad, σmetal, and γint of these models at the atomic level, the effects of these factors on the interface characteristics are elucidated. Additionally, from the electronic level, the nature of the effects of interatomic charge distribution on interfacial properties is explored using the PDOS and CDD. The conclusions from the experiments and calculations are as follows:
(1)
With the increase in Sn content, the σlv and W ad cal of the alloy show a decreasing trend, and the RPL thickness increases. The overall interface presents a “sandwich” structure, consisting of the metal||CrmCn (Cr3C2 and Cr7C3) and CrmCn||Cgr interfaces. The influence of varying Sn content is primarily reflected in the metal||CrmCn interface. The phase composition and morphology of CrmCn change with the Sn content.
(2)
The various forms (Cu–Sn solid solutions, CuxSny compounds, and Sn/Cr element segregation) exist on the alloy side in contact with the CrmCn surface. The doping and segregation of Cr can significantly increase the Wad and stability of the metal||Cr3C2 interface. The effect of Sn is opposite to that of Cr and becomes more significant with the increase in Sn content. This is primarily because the metallic bond formed by Sn–p and Cr–d is weaker than the dd metallic bonds (Cu–d and Cr–d, Cr–d and Cr–d). Sn increases the dissipation (or accumulation) of the interface’s CDD. The increase in Sn content reduces the number of d orbital electrons in the metal portion, leading to a decrease in the PDOS of electronics below the EF.
(3)
The σmetal of the Cu–Sn alloy obtained from theoretical calculations is in good agreement with the σlv reported in the literature. The trend in the theoretical calculation of Wad agrees with that of W ad cal calculated in the experiment.
(4)
The systematic experiments and simplified interfacial models based on the first-principles DFT calculations help elucidate the interaction among alloying elements and the effects of Sn content changes; reveal how typical interfacial configurations affect wettability, adhesion, and stability; and reduce the need for extensive experimental screening to avoid the influence of complex experimental conditions. This combined approach provides useful insights for designing interfaces in metal/carbon composites. In future work, advanced characterization techniques coupled with multiscale and dynamic interface simulations can be employed to refine the understanding of atomic-scale wetting behavior.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma18081793/s1, Figure S1: Structures of Cu, Sn, Cr, CuxSny (Cu10Sn3, Cu3Sn, Cu6Sn5), and Cr3C2: (a) Cu, (b) Sn, (c) Cr, (d) Cu10Sn3, (e) Cu3Sn, (f) Cu6Sn5, (g) Cr3C2; Table S1: Space groups and optimized lattice parameters (literature values in parentheses) of the bulk phases. Figure S2: Surface model: (a) Cu(111), (b) Sn(100), (c) Cr(110), (d)Cr3C2(001), (e) Cu10Sn3(100)/(010), (f) Cu10Sn3(001) type 1–3, (g) Cu3Sn(100) type 1, 2, (h) Cu6Sn5(100) type 1–3, (i) Cu6Sn5(010), (j) Cu6Sn5(001) type 1–3; Table S2: Surface energy calculation results; Figure S3: The UBER curves of Wad with interfacial equilibrium distance; Table S3: δ–values of supercell interfaces; Equation (S1): Formula for calculating formation energy (Esub) of a Cr or Sn atom doped at the Cu||Cr3C2 interface. References [37,38,43,44,45,46,52,53,54,57,60,61,62] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, W.C.; Methodology, Q.L. and X.L.; Validation, Y.S. and L.Y.; Formal analysis, W.C.; Investigation, W.C. and W.W.; Resources, Q.L. and W.W.; Data curation, Q.L. and X.L.; Writing—original draft, W.C.; Writing—review & editing, W.C., Q.L. and L.Y.; Visualization, W.W.; Supervision, Q.L., X.L., Y.S. and L.Y.; Funding acquisition, Q.L., X.L., Y.S. and L.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the central government’s special funds for guiding local science and technology development (grant number 24ZYQA054, 23ZYQA308), the major science and technology projects-industry (grant number 23ZDGA010), the National Natural Science Foundation of China (grant number 52461009, 52165004, 52305370), the Natural Science Foundation of Gansu Province (grant number 24JRRA199).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Materials. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Inagaki, M.; Kang, F. Carbon Materials Science and Engineering; Tsinghua University Press: Beijing, China, 2006. [Google Scholar]
  2. Hong, Q.; Li, M.; Wei, J.; Song, H.; Xue, F.; Sun, M. New carbon–copper composite material applied in rail-type launching system. IEEE Trans. Magn. 2007, 43, 137–140. [Google Scholar] [CrossRef]
  3. Sohn, Y.; Han, T.; Han, J.H. Effects of shape and alignment of reinforcing graphite phases on the thermal conductivity and the coefficient of thermal expansion of graphite/copper composites. Carbon 2019, 149, 152–164. [Google Scholar] [CrossRef]
  4. Silvain, J.F.; Heintz, J.M.; Veillere, A.; Constantin, L.; Lu, Y.F. A review of processing of Cu/C base plate composites for interfacial control and improved properties. Int. J. Extrem. Manuf. 2020, 2, 012002. [Google Scholar] [CrossRef]
  5. Karthik, J.M.; Abhinav, K.V. Shankar, Morphological and mechanical behavior of Cu–Sn alloys—A review. Met. Mater. Int. 2021, 27, 1915–1946. [Google Scholar] [CrossRef]
  6. Li, H.; Li, K.; Fan, Y.; Liu, C.; Wang, C. Influence of brazing temperature on interfacial reaction layer characteristics of Cu–Sn–Ti/diamond composites. Diam. Relat. Mater. 2022, 128, 109276. [Google Scholar] [CrossRef]
  7. Dong, R.; Cui, Z.; Zhu, S.; Xu, X.; Yang, X. Preparation, characterization and mechanical properties of Cu–Sn alloy/graphite composites. Metall. Mater. Trans. A 2014, 45, 5194–5200. [Google Scholar] [CrossRef]
  8. Naidich, Y.; Zhuravlev, V.; Gab, I.; Kostyuk, B.; Krasovskyy, V.; Adamovskyy, A.; Taranets, N. Liquid metal wettability, and advanced ceramic brazing. J. Eur. Ceram. Soc. 2008, 28, 717–728. [Google Scholar] [CrossRef]
  9. Asnavandi, M.; Ghorbani, M.; Kahram, M. Production of Cu–Sn–graphite–SiC composite coatings by electrodeposition. Surf. Coat. Technol. 2013, 216, 207–214. [Google Scholar] [CrossRef]
  10. Yu, J.; Huang, L.; Wang, J.; Hu, Z.; Hai, Y. Effects of Cu, Cu3Sn, Cu6Sn5 on interfacial properties between Cu–Sn alloys and diamond. Phys. Status Solidi (RRL) Rapid Res. Lett. 2022, 16, 2200013. [Google Scholar] [CrossRef]
  11. Naidich, J.V. The wettability of solids by liquid metals. Prog. Surf. Membr. Sci. 1981, 14, 353–484. [Google Scholar] [CrossRef]
  12. Deng, S. Interfacial Phenomena in Carbon-Copper Alloy System and Preparation of Carbon Fiber Fabric Reinforced Cu Matrix Composites. Ph.D. Thesis, Taiyuan University of Technology, Taiyuan, China, 2016. (In Chinese) [Google Scholar] [CrossRef]
  13. Sobczak, N.; Sobczak, J.; Rohatgi, P.; Ksiazek, M.; Radziwill, W.; Morgiel, J. Interaction between Ti or Cr containing copper alloys and porous graphite substrate. In Proceedings of the 2nd International Conference High Temperature Capillarity, Cracow, Poland, 29 June–2 July 1997; pp. 145–152. [Google Scholar]
  14. Lin, Q.; Yang, F.; Yang, H.; Sui, R.; Shi, Y.; Wang, J. Wetting of graphite by molten Cu–xSn–yCr ternary alloys at 1373 K. Carbon 2020, 159, 561–569. [Google Scholar] [CrossRef]
  15. Hsieh, Y.-C.; Lin, S.-T. Microstructural development of Cu–Sn–Ti alloys on graphite. J. Alloys Compd. 2008, 466, 126–132. [Google Scholar] [CrossRef]
  16. Dong, E. A-First-Principles Study of Matal-Ceramic Interfacial Bonding. Master’s Thesis, Jilin University, Changchun, China, 2013. (In Chinese). [Google Scholar]
  17. David, S.; Sholl, A.; Steckel, S. Density Functional Theory: A Practical Introduction; John Wiley & Sons: Hoboken, NJ, USA, 2009. [Google Scholar]
  18. Hashibon, A.; Elsässer, C.; Mishin, Y.; Gumbsch, P. First-principles study of thermodynamical and mechanical stabilities of thin copper film on tantalum. Phys. Rev. B 2007, 76, 245434. [Google Scholar] [CrossRef]
  19. Wang, Z.; Liu, S.; Qiu, Z.; Sun, H.; Liu, W. First-principles calculations on the interface of the Al/TiC aluminum matrix composites. Appl. Surf. Sci. 2020, 505, 144502. [Google Scholar] [CrossRef]
  20. Christensen, M.; Dudiy, S.; Wahnström, G. First-principles simulations of metal-ceramic interface adhesion: Co/WC versus Co/TiC. Phys. Rev. B 2002, 65, 045408. [Google Scholar] [CrossRef]
  21. Dudiy, S.V.; Lundqvist, B.I. Wetting of TiC and TiN by metals. Phys. Rev. B 2004, 69, 125421. [Google Scholar] [CrossRef]
  22. Lin, Q.; Shen, P.; Yang, L.; Jin, S.; Jiang, Q. Wetting of TiC by molten Al at 1123–1323K. Acta. Mater. 2011, 59, 1898–1911. [Google Scholar] [CrossRef]
  23. Clark, S.J.; Segall, M.D.; Pickard, C.J.; Hasnip, P.J.; Probert, M.I.J.; Refson, K.; Payne, M.C. First principles methods using CASTEP. Z. Kristallogr. Cryst. Mater. 2005, 220, 567–570. [Google Scholar] [CrossRef]
  24. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  25. Head, J.D.; Zerner, M.C. A Broyden–Fletcher–Goldfarb–Shanno optimization procedure for molecular geometries. Chem. Phys. Lett. 1985, 122, 264–270. [Google Scholar] [CrossRef]
  26. Eustathopoulos, N.; Nicholas, M.G.; Drevet, B. Wettability at High Temperatures; Elsevier Science: Kidlington, UK, 1999. [Google Scholar]
  27. Davies, J.T.; Rideal, E.K. Interfacial Phenomena, 2nd ed.; Academic Press: New York, NY, USA, 1963. [Google Scholar]
  28. Schrader, M.E. Young-Dupre revisited. Langmuir 1995, 11, 3585–3589. [Google Scholar] [CrossRef]
  29. Zhang, C.; Fan, T.; Cao, W.; Zhang, D. A model for work of solid–liquid adhesion in multicomponent melts. Acta. Mater. 2009, 57, 3623–3632. [Google Scholar] [CrossRef]
  30. Gasior, W.; Moser, Z.; Adam, D.; Janusz, P. SURDAT–Database of lead-free soldering materials. J. Min. Metall. Sect. B Metall. 2007, 43, 125–130. [Google Scholar]
  31. Zhao, Y.; Hou, X. A direct correlation between viscosity and liquid structure in Cu–Sn alloys. Adv. Condens. Matter Phys. 2017, 1, 7536853. [Google Scholar] [CrossRef]
  32. Hultgren, R.R.; Desai, P.D.; Hawkins, D.T.; Gleiser, M.; Kelley, K.K. Selected Values of The Thermodynamic Properties of Binary Alloys; American Society for Metals: Detroit, MI, USA, 1973; p. 795. Available online: https://cir.nii.ac.jp/crid/1130282270182197376 (accessed on 1 January 2025).
  33. Li, J. Adhesion between nonreactive liquid metals and transition metal carbides. Rare Met. 1992, 11, 177–184. [Google Scholar]
  34. Doi, T. Studies on Cu alloys containing Cr (lst Report). On the Cu side phase diagram. J. Jpn. Inst. Met. 1957, 21, 337–340. [Google Scholar] [CrossRef]
  35. Venkatraman, M.; Neumann, J.P. The Cr−Sn (Chromium–Tin) system. J. Phase Equilibria 1988, 9, 159–162. [Google Scholar] [CrossRef]
  36. Ci, W.; Shi, Y.; Lin, Q.; Gu, Y.; Li, G. Effects of temperature on reaction path, spreading and wetting mechanism of Cu–20Sn–2Cr/graphite. Ceram. Int. 2023, 49, 36907–36919. [Google Scholar] [CrossRef]
  37. Gu, K.; Pang, M.; Zhan, Y. Insight into interfacial structure and bonding nature of diamond(001)/Cr3C2(001) interface. J. Alloys Compd. 2019, 770, 82–89. [Google Scholar] [CrossRef]
  38. Fürtauer, S.; Li, D.; Cupid, D.; Flandorfer, H. Flandorfer, The Cu–Sn phase diagram, Part I: New experimental results. Intermetallics 2013, 34, 142–147. [Google Scholar] [CrossRef]
  39. Li, J.; Mannan, S.; Clode, M.; Whalley, D.; Hutt, D. Interfacial reactions between molten Sn–Bi–X solders and Cu substrates for liquid solder interconnects. Acta. Mater. 2006, 54, 2907–2922. [Google Scholar] [CrossRef]
  40. Mookam, N.; Kanlayasiri, K. Evolution of Intermetallic compounds between Sn-0.3Ag-0.7Cu low-silver lead-free solder and Cu substrate during thermal aging. J. Mater. Sci. Technol. 2012, 28, 53–59. [Google Scholar] [CrossRef]
  41. Li, Y.; Gao, Y.; Xiao, B.; Min, T.; Yang, Y.; Ma, S.; Yi, D. The electronic, mechanical properties and theoretical hardness of chromium carbides by first-principles calculations. J. Alloys Compd. 2011, 509, 5242–5249. [Google Scholar] [CrossRef]
  42. Jiang, C. First-principles study of structural, elastic, and electronic properties of chromium carbides. Appl. Phys. Lett. 2008, 92, 041909. [Google Scholar] [CrossRef]
  43. Ellner, M.; Kolatschek, K.; Predel, B. On the partial atomic volume and the partial molar enthalpy of aluminum in some phases with Cu and Cu3Au structures. J. Less-Common Met. 1991, 170, 171–184. [Google Scholar] [CrossRef]
  44. Tu, K.N. Cu/Sn interfacial reactions: Thin-film case versus bulk case. Mater. Chem. Phys. 1996, 46, 217–223. [Google Scholar] [CrossRef]
  45. Bolef, D.I.; de Klerk, J. Anomalies in the elastic constants and thermal expansion of chromium single crystals. Phys. Rev. 1963, 129, 1063–1067. [Google Scholar] [CrossRef]
  46. Larsson, A.-K.; Stenberg, L.; Lidin, S. The superstructure of domain-twinned η’–Cu6Sn5. Acta Crystallogr. Sect. B Struct. Sci. 1994, 50, 636–643. [Google Scholar] [CrossRef]
  47. Lei, C.; Xue, H.; Tang, F.; Luo, X. The effect of solute segregation on stability and strength of Cu symmetrical tilt grain boundaries from the first-principles study. J. Mater. Res. Technol. 2022, 21, 3274–3284. [Google Scholar] [CrossRef]
  48. Sellers, M.S.; Schultz, A.J.; Basaran, C.; Kofke, D.A. Atomistic modeling of β–Sn surface energies and adatom diffusivity. Appl. Surf. Sci. 2010, 256, 4402–4407. [Google Scholar] [CrossRef]
  49. Xue, H.; Lei, C.; Tang, F.; Li, X.; Luo, Y.; Ren, J.; Lu, X. Structure-composition-property correlations of symmetrical tilt grain boundaries in copper-based binary alloys. J. Phys. Chem. Solids 2021, 154, 110082. [Google Scholar] [CrossRef]
  50. Tran, R.; Xu, Z.; Radhakrishnan, B.; Winston, D.; Sun, W.; Persson, K.A.; Ong, S.P. Surface energies of elemental crystals. Sci. Data. 2016, 3, 160080. [Google Scholar] [CrossRef] [PubMed]
  51. Rapcewicz, K.; Chen, B.; Yakobson, B.; Bernholc, J. Consistent methodology for calculating surface and interface energies. Phys. Rev. B 1998, 57, 7281–7291. [Google Scholar] [CrossRef]
  52. Wu, Z.; Pang, M.; Zhan, Y.; Shu, S.; Xiong, L.; Li, Z. The bonding characteristics of the Cu(111)/WC(0001) interface: An insight from first-principle calculations. Vacuum 2021, 191, 110218. [Google Scholar] [CrossRef]
  53. Ma, Z.; Li, H.; Zhang, M.; Xu, Z. The first principles calculation of the interfacial properties of β–Sn and 4H–SiC. Ordnance. Mater. Sci. Eng. 2019, 42, 17–20. [Google Scholar] [CrossRef]
  54. Chen, Y.; Dai, W.; Liu, Y.; Chen, C.; Tu, K.; Chen, G. Surface protrusion induced by inter-diffusion on Cu–Sn micro-pillars. Mater. Des. 2022, 224, 111318. [Google Scholar] [CrossRef]
  55. Hu, J.; Jian, X.; Yang, T.; Peng, X. Investigation on the interface characteristic between WC(001) and diamond(111) by first-principles calculation. Diam. Relat. Mater. 2022, 123, 108864. [Google Scholar] [CrossRef]
  56. Wu, Y.; Zhang, W.; Guo, J.; Hou, J.; Li, X.; Huang, R.; Ma, X.; Zhang, Q. The first-principles study on the occupation behavior and the ductility mechanism of Zr in Ni–Ni3Al system with lattice misfit. J. Mater. Sci. Technol. 2014, 30, 517–522. [Google Scholar] [CrossRef]
  57. Huang, J.; Li, M.; Liu, Y.; Chen, J.; Lai, Z.; Hu, J.; Zhou, F.; Zhu, J. A first-principles study on the doping stability and micromechanical properties of alloying atoms in aluminum matrix. Vacuum 2023, 207, 111596. [Google Scholar] [CrossRef]
  58. Eustathopoulos, N. Dynamics of wetting in reactive metal/ceramic systems. Acta. Mater. 1998, 46, 2319–2327. [Google Scholar] [CrossRef]
  59. Morra, M.; Occhiello, E.; Garbassi, F. Knowledge about polymer surfaces from contact angle measurements. Adv. Colloid Interface Sci. 1990, 32, 79–116. [Google Scholar] [CrossRef]
  60. Materials Project: An Open Database of Materials Properties. Available online: https://next-gen.materialsproject.org (accessed on 10 April 2024).
  61. Leineweber, A. The Cu–Sn System: A Comprehensive Review of the Crystal Structures of its Stable and Metastable Phases. J. Phase Equilibria Diffus. 2023, 44, 343–393. [Google Scholar] [CrossRef]
  62. Xie, H.; Zhao, N.; Shi, C.; He, C.; Liu, E. Effects of active elements on adhesion of the Al2O3/Fe interface: A first principles calculation. Comput. Mater. Sci. 2021, 188, 110226. [Google Scholar] [CrossRef]
Figure 2. Microstructure near the cross-sectional interface of Cu–aSn–1Cr/Cgr (backscattered mode): (a-1a-4) 89Cu–10Sn–1Cr/Cgr, (b-1b-4) 79Cu–20Sn–1Cr/Cgr, (c-1c-4) 49Cu–50Sn–1Cr/Cgr, (d-1d-4) 19Cu–80Sn–1Cr/Cgr, (a-1,d-1) at the center of the interface, (a-2,d-2) corresponds to the EDS analysis results near the interface in (a-1,d-1,a-3,b-2,c-3,d-3) correspond to the magnified maps near the interface, (b-3) HRTEM of the red rectangle area in (b-1,a-4c-4) correspond to the EDS analysis of lines 1–3 in (a-3,b-1,c-3,d-4) EDS analysis of points 1–9, areas 1–6 correspond to the Cu–Sn solid solution, Cr precipitates, Cu10Sn3, Cu3Sn, Cu6Sn5, and Sn segregation regions in contact with the RPL, the blue dotted line in (a-3) indicates the boundary between CuxSny intermetallic compound and Cu-Sn solid solution.
Figure 2. Microstructure near the cross-sectional interface of Cu–aSn–1Cr/Cgr (backscattered mode): (a-1a-4) 89Cu–10Sn–1Cr/Cgr, (b-1b-4) 79Cu–20Sn–1Cr/Cgr, (c-1c-4) 49Cu–50Sn–1Cr/Cgr, (d-1d-4) 19Cu–80Sn–1Cr/Cgr, (a-1,d-1) at the center of the interface, (a-2,d-2) corresponds to the EDS analysis results near the interface in (a-1,d-1,a-3,b-2,c-3,d-3) correspond to the magnified maps near the interface, (b-3) HRTEM of the red rectangle area in (b-1,a-4c-4) correspond to the EDS analysis of lines 1–3 in (a-3,b-1,c-3,d-4) EDS analysis of points 1–9, areas 1–6 correspond to the Cu–Sn solid solution, Cr precipitates, Cu10Sn3, Cu3Sn, Cu6Sn5, and Sn segregation regions in contact with the RPL, the blue dotted line in (a-3) indicates the boundary between CuxSny intermetallic compound and Cu-Sn solid solution.
Materials 18 01793 g002
Figure 3. The top view SEM morphology of RPL after etching the alloy droplets (secondary electron mode): (a-1a-4) 89Cu10Sn1Cr/Cgr, (b-1b-4) 79Cu–20Sn–1Cr/Cgr, (c-1c-4) 49Cu50Sn1Cr/Cgr, and (d-1d-4) 20Cu79Sn1Cr/Cgr, (a-1d-1) the center of the interface, (a-2d-2) correspond to the local enlarged views of the interfaces in (a-1d-1,a-3d-3) local enlarged views near the TL, (a-4b-4) correspond to the EDS analysis results of (a-3,b-3,d-4) EDS analysis results of points 1–7.
Figure 3. The top view SEM morphology of RPL after etching the alloy droplets (secondary electron mode): (a-1a-4) 89Cu10Sn1Cr/Cgr, (b-1b-4) 79Cu–20Sn–1Cr/Cgr, (c-1c-4) 49Cu50Sn1Cr/Cgr, and (d-1d-4) 20Cu79Sn1Cr/Cgr, (a-1d-1) the center of the interface, (a-2d-2) correspond to the local enlarged views of the interfaces in (a-1d-1,a-3d-3) local enlarged views near the TL, (a-4b-4) correspond to the EDS analysis results of (a-3,b-3,d-4) EDS analysis results of points 1–7.
Materials 18 01793 g003
Figure 4. Schematic diagram of the interface structure: (a) Cr precipitation on the CrmCn, surface, (b) Sn segregation on the CrmCn surface, (c) Cu10Sn3 phase in contact with the CrmCn, (d) Cu3Sn and Cu6Sn5 in contact with the CrmCn.
Figure 4. Schematic diagram of the interface structure: (a) Cr precipitation on the CrmCn, surface, (b) Sn segregation on the CrmCn surface, (c) Cu10Sn3 phase in contact with the CrmCn, (d) Cu3Sn and Cu6Sn5 in contact with the CrmCn.
Materials 18 01793 g004
Figure 5. Interface model: (a-1a-3) pure metal Cu/Sn/Cr with Cr3C2: (a-1) Cu||Cr3C2, (a-2) Sn||Cr3C2, (a-3) Cr||Cr3C2, (b-1) Cr/Sn atom doped regions, (b-2) doping positions and their numbers, Numbers 1–12 indicate the atomic positions for Cr atom doping in the Cu||Cr3C2 interface, Numbers 1–14 represent the positions for Sn atom doping in the same interface. (c-1c-3) CuxSny||Cr3C2: (c-1) Cu10Sn3||Cr3C2, (c-2) Cu3Sn||Cr3C2, (c-3) Cu6Sn5||Cr3C2.
Figure 5. Interface model: (a-1a-3) pure metal Cu/Sn/Cr with Cr3C2: (a-1) Cu||Cr3C2, (a-2) Sn||Cr3C2, (a-3) Cr||Cr3C2, (b-1) Cr/Sn atom doped regions, (b-2) doping positions and their numbers, Numbers 1–12 indicate the atomic positions for Cr atom doping in the Cu||Cr3C2 interface, Numbers 1–14 represent the positions for Sn atom doping in the same interface. (c-1c-3) CuxSny||Cr3C2: (c-1) Cu10Sn3||Cr3C2, (c-2) Cu3Sn||Cr3C2, (c-3) Cu6Sn5||Cr3C2.
Materials 18 01793 g005
Figure 6. The variation of Wad and γint when a single Sn/Cr atom and multiple Sn atoms are doped at interface of the Cu||Cr3C2 interface and the magnitude of Esub: (a) the Wad of Sn/Cr atom dope into different sites at the Cu||Cr3C2 interface; (b) γint, the red, green, and blue horizontal lines correspond to Cu||Cr3C2, Sn||Cr3C2 and Cr||Cr3C2, respectively; (c) Wad and γint for the Cu56-kSnk||Cr3C2 (k = 0, 1, 2, 3, 4) interface; (d) the magnitude of Esub.
Figure 6. The variation of Wad and γint when a single Sn/Cr atom and multiple Sn atoms are doped at interface of the Cu||Cr3C2 interface and the magnitude of Esub: (a) the Wad of Sn/Cr atom dope into different sites at the Cu||Cr3C2 interface; (b) γint, the red, green, and blue horizontal lines correspond to Cu||Cr3C2, Sn||Cr3C2 and Cr||Cr3C2, respectively; (c) Wad and γint for the Cu56-kSnk||Cr3C2 (k = 0, 1, 2, 3, 4) interface; (d) the magnitude of Esub.
Materials 18 01793 g006
Figure 7. Partial density of states (PDOS) at interfaces: (a) Cu||Cr3C2, (b) Sn||Cr3C2, (c) Cr||Cr3C2, (d) Cu54Cr1||Cr3C2, (e) Sn49Cr1||Cr3C2, (fi) Cu56-kSnk||Cr3C2: (f) k = 1, (g) k = 2, (h) k = 3, (i) k = 4, (jl) CuxSny||Cr3C2: (j) Cu10Sn3||Cr3C2, (k) Cu3Sn||Cr3C2, (l) Cu6Sn5||Cr3C2.
Figure 7. Partial density of states (PDOS) at interfaces: (a) Cu||Cr3C2, (b) Sn||Cr3C2, (c) Cr||Cr3C2, (d) Cu54Cr1||Cr3C2, (e) Sn49Cr1||Cr3C2, (fi) Cu56-kSnk||Cr3C2: (f) k = 1, (g) k = 2, (h) k = 3, (i) k = 4, (jl) CuxSny||Cr3C2: (j) Cu10Sn3||Cr3C2, (k) Cu3Sn||Cr3C2, (l) Cu6Sn5||Cr3C2.
Materials 18 01793 g007
Figure 8. Charge density difference (CDD) maps of selected metal||Cr3C2 interfaces; red and blue regions represent charge accumulation and dissipation, respectively: (a) Cu||Cr3C2, (b) Sn||Cr3C2, (c) Cr||Cr3C2, (d) Cu54Cr1||Cr3C2, (e) Sn49Cr1||Cr3C2, (fi) Cu56-kSnk|| Cr3C2: (f) k = 1, (g) k = 2, (h) k = 3, (i) k = 4, (jl) CuxSny||Cr3C2: (j) Cu10Sn3||Cr3C2, (k) Cu3Sn||Cr3C2, (l) Cu6Sn5||Cr3C2.
Figure 8. Charge density difference (CDD) maps of selected metal||Cr3C2 interfaces; red and blue regions represent charge accumulation and dissipation, respectively: (a) Cu||Cr3C2, (b) Sn||Cr3C2, (c) Cr||Cr3C2, (d) Cu54Cr1||Cr3C2, (e) Sn49Cr1||Cr3C2, (fi) Cu56-kSnk|| Cr3C2: (f) k = 1, (g) k = 2, (h) k = 3, (i) k = 4, (jl) CuxSny||Cr3C2: (j) Cu10Sn3||Cr3C2, (k) Cu3Sn||Cr3C2, (l) Cu6Sn5||Cr3C2.
Materials 18 01793 g008
Table 1. Experimental measurements and calculated values.
Table 1. Experimental measurements and calculated values.
a (at. %)010203040507999
θe (°)43
43 [33]
14121324313931
σlv (J/m2) [29]1.31.00.860.760.690.640.540.49
σsl (J/m2)3.053.023.163.263.363.463.593.58
W ad cal (J/m2)2.25
2.16 [33]
1.981.701.501.331.180.950.91
Table 2. Theoretical calculated values.
Table 2. Theoretical calculated values.
a (at. %)01.793.575.367.1423.102545.45100
σmetal (J/m2)1.321.221.241.161.130.870.850.760.43
γint (J/m2)2.312.302.863.063.203.583.933.791.83
Wad (J/m2)3.052.912.762.672.681.281.090.962.6
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ci, W.; Lin, Q.; Lu, X.; Shi, Y.; Yang, L.; Wang, W. Effect of Sn Content on Wettability and Interfacial Structure of Cu–Sn–Cr/Graphite Systems: Experimental and First-Principles Investigations. Materials 2025, 18, 1793. https://doi.org/10.3390/ma18081793

AMA Style

Ci W, Lin Q, Lu X, Shi Y, Yang L, Wang W. Effect of Sn Content on Wettability and Interfacial Structure of Cu–Sn–Cr/Graphite Systems: Experimental and First-Principles Investigations. Materials. 2025; 18(8):1793. https://doi.org/10.3390/ma18081793

Chicago/Turabian Style

Ci, Wenjuan, Qiaoli Lin, Xuefeng Lu, Yu Shi, Likai Yang, and Wenkai Wang. 2025. "Effect of Sn Content on Wettability and Interfacial Structure of Cu–Sn–Cr/Graphite Systems: Experimental and First-Principles Investigations" Materials 18, no. 8: 1793. https://doi.org/10.3390/ma18081793

APA Style

Ci, W., Lin, Q., Lu, X., Shi, Y., Yang, L., & Wang, W. (2025). Effect of Sn Content on Wettability and Interfacial Structure of Cu–Sn–Cr/Graphite Systems: Experimental and First-Principles Investigations. Materials, 18(8), 1793. https://doi.org/10.3390/ma18081793

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop