Zeyun
Wang
a,
Zhenyu
Zhang
*b,
Pengfei
Hu
c,
Ganggang
Liu
d,
Jianjun
Hu
*e,
Jianan
Xu
*a,
Huaxiang
Cai
f,
Zehong
Pang
f and
Peng
Ding
f
aCollege of Mechanical and Electrical Engineering, Harbin Engineering University, Harbin 150001, China. E-mail: xujianan@hrbeu.edu.cn
bState Key Laboratory of High-performance Precision Manufacturing, Dalian University of Technology, Dalian 116024, China. E-mail: zzy@dlut.edu.cn
cGuizhou Equipment Manufacturing Polytechnic, Guiyang 551400, China
dGuizhou Zhenhua Fengguang Semiconductor Co., Ltd, Guiyang 550018, China
eSchool of Optoelectronic Science and Engineering, Soochow University, Suzhou 215021, China. E-mail: jianjunhu@suda.edu.cn
fGuizhou Aerospace Linquan Motor Co., Ltd, Guiyang 550081, China
First published on 7th August 2025
An aluminum (Al) alloy is a soft, plastic-like metal that is prone to embedding abrasives, scratches, corrosion pits, and deformation. Achieving an atomically smooth surface on an Al alloy presents a significant challenge. This study introduces a novel green chemical mechanical polishing (CMP) technique using hydrogen peroxide, tyrosine, sodium carbonate, and hybrid abrasives composed of silica, yttria, and ceria. The method produces a surface roughness (Sa) of 0.187 nm over a 50 × 50 μm2 scanning area, with a material removal rate of 17.23 μm h−1. Transmission electron microscopy (TEM) analysis shows a damaged layer thickness of 3.6 nm. To the best of our knowledge, this work reports the lowest Sa and damaged layer thickness for Al alloys to date. Molecular dynamics simulations are used to elucidate the mechanism of dynamic material removal during nanoscratching. As the cutting depth increases from 2.5 to 3 nm, the thickness of the damaged layer varies from 3.5 to 4.3 nm, aligning well with the TEM findings. The maximum von Mises stress recorded is 7.98 GPa, with the appearance of dislocation loops, vacancy defects, stacking faults, and an amorphous phase. The zeta potential measurements are −22.78, −9.46, −34.03, and −41.55 mV for silica in water, silica and yttria in water, silica at a pH of 10, and silica and yttria at a pH of 10, respectively. These correspond to polydispersity indices of 0.281, 0.412, 0.231, and 0.185. These data indicate that the hybrid abrasives of silica and yttria in a sodium carbonate solution exhibit the best stability and dispersion among the tested solutions. Characterization techniques, including X-ray photoelectron spectroscopy, Fourier transform infrared spectroscopy, and Raman spectroscopy, demonstrate the role of hydrogen peroxide in oxidizing the Al alloy surface to form aluminum hydroxide, alumina, and silica. Proposed complexing formulas suggest that tyrosine forms complexes with Al3+ ions, while ceria reacts with silica. The removal of oxides and complexes is facilitated by hybrid abrasives. The proposed novel green CMP method offers a new pathway for achieving atomic-level surface smoothness on Al alloys, potentially enhancing their application in high-performance devices.
Chemical components are crucial for the effectiveness of CMP, particularly for reactive metals like aluminum.14,15 Previous studies have explored various oxidants, including potassium permanganate,16 ferric nitrate,17 and hydrogen peroxide (H2O2).18 These oxidants aim to create passivated oxide layers that allow for controlled material removal. However, some oxidants can be unsafe or lack selectivity towards aluminum, resulting in uneven polishing and surface defects. In recent studies, complexing agents like amino acids have been introduced to enhance the complexation and dissolution of aluminum oxide.19,20 Nevertheless, most of these agents, such as glycine21 and serine,22 have limited solubility or weak interactions with oxidized aluminum, which restricts their effectiveness in alkaline environments. In this work, a novel chemical system is presented that combines hydrogen peroxide (H2O2) and tyrosine.23 In this system, H2O2 serves as a mild oxidant that forms a soft oxide layer, while tyrosine acts as an effective complexing agent for aluminum ions. The synergistic interaction between these two agents ensures a controlled balance between oxidation and dissolution, leading to improved surface smoothness and reduced damage.
The selection and dispersion stability of abrasive particles are crucial for the effectiveness of the CMP technique. Most commonly used abrasives in CMP include silica (SiO2),24 alumina (Al2O3),25 and ceria (CeO2)26, each chosen for their tunable hardness and surface reactivity. SiO2 is preferred for its relatively gentle mechanical action, while CeO2 is known for its redox-assisted reactivity with silicon (Si)-containing phases.27 However, conventional abrasive systems often face challenges such as poor dispersion stability and particle agglomeration, which can lead to uneven surfaces and random scratches.28 While several dispersants have been suggested, few provide long-term colloidal stability without compromising the slurry's reactivity.29,30 Furthermore, many studies neglect the importance of interface compatibility between abrasives and multiphase alloys, which is essential for effectively polishing composites like AlSi10Mg. To address these limitations, yttria (Y2O3) is proposed as a co-abrasive. Under alkaline conditions, Y2O3 enhances the dispersion of SiO2 through electrostatic and steric stabilization.31 The inclusion of CeO2 also provides targeted chemical reactivity with Si-rich phases. This combination aims to achieve uniform material removal while minimizing polishing defects.
Subsurface damage (SSD) is another critical factor influencing the performance of Al–Si alloy components, particularly in precision optical applications.32,33 During CMP and other material removal processes, localized stress concentrations can lead to plastic deformation, dislocation generation, and in severe cases, amorphization beneath the polished surface.34,35 Previous studies have investigated SSD through both experimental techniques, such as transmission electron microscopy (TEM), and computational methods, including molecular dynamics (MD) simulations.36,37 These investigations have provided valuable insights into dislocation nucleation, fault formation, and the transition from crystalline to amorphous states under mechanical loading.38 However, much of the previous research has concentrated on monolithic or Si-dominant systems,39 with limited attention given to the complex, multiphase nature of Al–Si alloys and their heterogeneous mechanical responses. Moreover, the coupling between abrasive-induced stress fields and phase-specific damage mechanisms remains poorly understood. A more comprehensive understanding of SSD evolution at different removal depths and across phase interfaces is essential for optimizing slurry design and minimizing degradation in optical performance.
This work developed a novel CMP slurry system tailored for the surface finishing of the AlSi10Mg alloy. This innovative slurry comprises H2O2, tyrosine, sodium carbonate (Na2CO3), SiO2, Y2O3, CeO2, and deionized water. It is chemically benign and compositionally optimized to promote stable dispersion, controlled chemical reactivity, and mechanical compatibility with multiphase Al–Si microstructures. In this study, a series of advanced testing and surface characterization techniques was employed to elucidate the material removal mechanisms and subsurface responses of AlSi10Mg. This integrated approach provides a comprehensive framework for understanding and optimizing the chemo-mechanical interactions during CMP, ultimately aiming to achieve defect-free, ultra-smooth surfaces.
| Element | Si | Mg | Fe | Mn | Al |
|---|---|---|---|---|---|
| Content (wt%) | 9–11 | 0.2–0.45 | ≤0.55 | ≤0.45 | Bal. |
As illustrated in Table 2, eight experiments with different parameters were designed to investigate the influence of the components of the CMP slurry on polishing performance. Among them, slurries S1–S4 explored how additives enhanced the dispersion of SiO2 abrasives, thereby reducing scratches and pits on the polished surface. Slurries S5–S8 investigated the chemical reaction mechanisms of the additives, aiming to achieve an excellent Sa and material removal rate (MRR).
| Group | SiO2 mass fraction (wt%) | Y2O3 mass fraction (wt%) | CeO2 mass fraction (wt%) | Na2CO3 pH | H2O2 mass fraction (wt%) | Tyrosine mass fraction (wt%) |
|---|---|---|---|---|---|---|
| S1 | 4.0 | 0 | 0 | 0 | 0 | 0 |
| S2 | 4.0 | 1.0 | 0 | 0 | 0 | 0 |
| S3 | 4.0 | 0 | 0 | 10.0 | 0 | 0 |
| S4 | 4.0 | 1.0 | 0 | 10.0 | 0 | 0 |
| S5 | 4.0 | 1.0 | 0 | 10.0 | 1.25 | 0 |
| S6 | 4.0 | 1.0 | 0 | 10.0 | 0 | 1.5 |
| S7 | 4.0 | 1.0 | 0 | 10.0 | 1.25 | 1.5 |
| S8 | 4.0 | 1.0 | 1.0 | 10.0 | 1.25 | 1.5 |
MRR is an important metric for assessing the polishing performance of the CMP slurry. The mass of the specimens was measured before and after polishing using a precision electronic balance (ME204E, Mettler-Toledo), with each measurement being repeated three times for accuracy. The formula used to calculate the MRR is as follows:40
![]() | (1) |
Molecular dynamics (MD) simulations of the constructed model were conducted using the Large-scale Atomic/Molecular Massively Parallel Simulator (LAMMPS).42 The initial structure was first relaxed using the conjugate gradient method to reduce the potential energy, thereby eliminating structural artifacts and relieving internal stress. To ensure adequate system equilibration, a relaxation process was first performed using the canonical ensemble (NVT) at 300 K for 50 ps. This was followed by a second relaxation phase under the isothermal-isobaric ensemble (NPT), also maintained at 300 K and 0.1 bar, for an additional 50 ps. Upon completion of these two sequential steps, the system achieved equilibrium, thereby allowing for the commencement of subsequent cutting simulations.
The particle was modeled using a repulsive force potential, as described by eqn (2).43 This approach has been extensively applied in MD simulations involving nanoindentation and abrasive particle cutting.
![]() | (2) |
The interatomic interactions in the simulation model were described using the Modified Embedded Atom Method (MEAM) potential. The general form of the MEAM potential is given by eqn (3).44
![]() | (3) |
The properties of the slurry were evaluated using a nanoparticle size and zeta potential analyzer (Zetasizer Pro, Malvern, UK) to determine the zeta potential and polydispersity index (PI). The electrochemical behavior was examined through Tafel polarization and electrochemical impedance spectroscopy (EIS), conducted on an electrochemical workstation (CHI760E, Shanghai Chenhua Instrument Co., Ltd). The measurements utilized a three-electrode configuration, consisting of a saturated calomel electrode as the reference electrode, a platinum counter electrode, and the sample acting as the working electrode.
The surface elemental composition was examined using X-ray photoelectron spectroscopy (XPS; K-Alpha, Thermo Scientific, USA), with the carbon (C) 1s peak at 284.8 eV serving as the reference for charge correction. Changes in functional groups were analyzed through Fourier-transform infrared spectroscopy (FTIR; Nicolet iS20, Thermo Scientific, USA), while structural alterations were characterized by Raman spectroscopy (DXR, Thermo Fisher, USA).
![]() | ||
| Fig. 1 The impact of different components on the Sa and MRR. (a) SiO2 concentration. (b) Slurry pH. (c) H2O2 concentration. (d) Tyrosine concentration. | ||
To enhance the dispersion and stability of nano-SiO2 abrasives, SEM characterization was performed on SiO2 slurries containing various additives. The zeta potential and PI values were also measured and analyzed.45 The measurement and characterization results for slurries S1–S4 are presented in Fig. 2. The SEM images of the abrasives in slurries S1–S4 are presented in Fig. 2(a–d), highlighting variations in particle size and aggregation degree, which directly impact the performance of the CMP process. It is evident from Fig. 2(a) that pure nano-silica abrasives exhibited significant aggregation and uneven dispersion, leading to larger agglomerates that caused non-uniform surface quality during polishing. However, the addition of Y2O3 abrasives exacerbated this situation, as shown in Fig. 2(b). In contrast, Fig. 2(c) illustrates that the dispersion of SiO2 abrasives is enhanced under alkaline conditions. Interestingly, in Fig. 2(d), the presence of Y2O3 abrasives further boosted the dispersion of SiO2 abrasives under these alkaline conditions. Overall, both the alkaline environment and the synergistic effect of Y2O3 contributed to increased electrostatic repulsion, significantly improving dispersion stability. This enhancement is essential for achieving high-quality polishing performance.
![]() | ||
| Fig. 2 SEM images and property characterization of the abrasives in slurries S1–S4. (a) S1 slurry. (b) S2 slurry. (c) S3 slurry. (d) S4 slurry. (e) PI value. (f) Zeta potential. | ||
Fig. 2(e) illustrates that the PI value of slurry S4 was the lowest, indicating that the dispersion of the SiO2 abrasives, sodium carbonate, and the Y2O3 abrasive slurry was relatively effective. Additionally, the absolute zeta potential of slurry S1 increased from 22.78 mV to 41.55 mV upon the addition of sodium carbonate and Y2O3 abrasives. This implies that these additives enhanced the dispersion and stability of the SiO2 abrasives, as observed from Fig. 2(f). A comprehensive analysis of both the PI value and zeta potential measurements revealed that combining Y2O3 abrasives and sodium carbonate solution with SiO2 abrasives enhanced the dispersion and stability of pure SiO2 abrasives. This improvement can be attributed to the transportation of a negative charge on the surface of Y2O3 abrasives under alkaline conditions, which collectively increased the electrostatic repulsion between the similarly charged SiO2 abrasives. These findings underscore the importance of optimizing both the chemical composition and dispersion properties of CMP slurries for the effective polishing of AlSi10Mg alloys.
The polishing efficiency of various slurry formulations, as shown in Table 2, was systematically investigated using AlSi10Mg substrates through controlled experimental trials. The resulting surface morphology and Sa outcomes are presented in Fig. 3. The Sa value was precisely measured over an area of 50 × 50 μm2. Notably, slurry S8 achieved the optimal Sa and effectively improved the polishing efficiency, with Sa reaching an impressive 0.187 nm and the MRR achieving an exceptional 17.23 μm h−1.
After polishing with slurries S1–S4, the value of Sa continuously decreased, and the number of scratches on the surface significantly reduced. This indicated that improving the dispersion of the abrasives enhanced polishing performance. Both Fig. 3(e) and (f) demonstrate that the addition of H2O2 and tyrosine improved Sa and accelerated the removal of surface protrusions. This phenomenon may be attributed to the reaction between H2O2, tyrosine, and the material surface, forming a soft layer. Fig. 3(g) provides strong evidence of the significant impact of the synergistic effect of H2O2 and tyrosine on Sa. While this finding is important, the underlying mechanism of this synergistic effect still requires further investigation and analysis. With the inclusion of CeO2 abrasives, Sa reached its lowest value, as displayed in Fig. 3(h). This result was closely related to the favorable chemical reaction between CeO2 and Si. The optical images clearly revealed numerous scratches on the surface before polishing, while the surface became extremely smooth after polishing, as shown in Fig. 3(i) and (j). The AFM analysis in Fig. 3(k) and (l) demonstrated the exceptional surface quality achieved after polishing, with a measured Sa of 0.201 nm and peak-to-valley height variation of approximately 0.12 nm. Extensive experimental results confirmed that the optimized CMP slurry successfully balanced the competing demands of minimal Sa and enhanced MRR, which is a critical requirement for precision surface finishing.
To evaluate the surface and subsurface features of AlSi10Mg before and after polishing, TEM samples were prepared using a focused ion beam (FIB) equipment. The cross-sectional view of the unpolished surface obtained via FIB preparation is shown in Fig. 4(a). Similarly, Fig. 4(b) revealed a damaged surface, with the thickness of the damage layer measuring approximately 315 nm. A comparison between the damaged and undamaged regions pinpointed a structural transformation from the original single-crystalline phase to a polycrystalline structure, as illustrated in Fig. 4(c) and (d). Meanwhile, a precipitate phase of Si was also identified in the undamaged area, which was confirmed through single-crystal electron diffraction patterns, as evident from Fig. 4(e) and (f). Further observations of the damaged region in Fig. 4(g) showed a high density of crystal defects in the outermost layer of the damage zone. High-resolution imaging of this damaged region, illustrated in Fig. 4(h), exposed a high density of edge dislocations within individual crystalline grains. These observations demonstrate that coarse grinding induces significant grain fragmentation and stress concentration at grain boundaries.46 Notably, the residual stress exhibited a gradual attenuation with increasing depth from the surface. Post-polishing characterization showed remarkable improvement, with the damage layer thickness being reduced to 3.6 nm. The high-resolution images also displayed well-defined lattice fringes with a spacing of 0.268 nm, as depicted in Fig. 4(i). This confirms that there is minimal subsurface impact after polishing and that the crystal integrity is maintained, emphasizing the superior performance of slurry S8 in the CMP process.
A high-resolution scanning transmission electron microscopy (STEM) image of the longitudinal section is depicted in Fig. 5(b), highlighting the directional cellular growth along the build path. The corresponding STEM-EDS elemental mappings in Fig. 5(c) and (d) further confirmed the segregation behavior during solidification: Al dominated the cell interiors, while Si was enriched along the boundaries. The localized enrichment of Si at cell boundaries introduced chemical and mechanical heterogeneity across the surface, which was a critical factor during the CMP process.
From a CMP perspective, the heterogeneous microstructure of the AlSi10Mg alloy implied that its surface comprised regions with varying hardness and chemical reactivity.48 The relatively soft α-Al matrix was more prone to mechanical abrasion and chemical oxidation, whereas the Si-rich boundaries, being solid and chemically inert, resisted removal and potentially acted as polishing endpoints or defect initiation sites. Therefore, understanding the distribution and morphology of these phases is essential for optimizing the slurry chemistry and process parameters. These observations highlight the necessity of tailoring CMP strategies for Al–Si alloys, particularly those fabricated by additive manufacturing, as microstructural anisotropy can further complicate surface planarization.
To elucidate the formation mechanism of subsurface damage layers during the CMP process of the AlSi10Mg alloy, abrasive particle cutting simulations were performed on the Al–Si alloy model. The SEM and STEM observations guided the establishment of this model. The Al–Si alloy model illustrated in Fig. 6 was constructed based on the characteristic microstructure of the AlSi10Mg alloy. The atomic-scale simulation results were post-processed and rendered using the Open Visualization Tool (OVITO) for defect analysis and visualization.49 A virtual abrasive particle with a radius of 35 Å was employed in the simulation. In the cutting simulation, the virtual abrasive particle initially penetrated the material along the negative z-axis at a velocity of 25 m s−1 until it reached the designated depth. Subsequently, it was transported laterally along the negative x-axis at a velocity of 100 m s−1. To examine the effect of the indentation depth on the cutting behavior and material response, simulations were conducted at six distinct depths: 0.5 nm, 1.0 nm, 1.5 nm, 2.0 nm, 2.5 nm, and 3.0 nm.
Fig. 7 displays the tangential and normal forces exerted on the abrasive particle as it traverses across the AlSi10Mg surface at varying indentation depths. The x-axis denotes the lateral displacement of the abrasive, corresponding to its trajectory across different material regions, while the y-axis represents the measured forces during the cutting process. The force profiles demonstrated distinct mechanical responses across these regions. In Al domains (0–8 nm and 13–21 nm), both tangential and normal forces exhibited relatively smooth and moderate increases with greater penetration depth, indicating ductile cutting behavior. In contrast, a sharp spike in both forces was observed in the Si region (8–13 nm), particularly at deeper indentation depths. This abrupt rise reflected the high resistance of a solid and more brittle Si phase, suggesting localized stress concentration and potential initiation of brittle fracture.50 Such segmented force evolution underscores the challenges of planarizing dual-phase systems like AlSi10Mg, where the embedded Si phase limits the ability to achieve low Sa and efficient material removal. The increased cutting resistance in the Si zone confirmed that Si was significantly more difficult to remove through purely mechanical means, which can lead to subsurface damage or inefficient planarization if not appropriately addressed. These findings emphasize the need for a selective chemical approach to assist in the removal of the Si phase during CMP.
![]() | ||
| Fig. 7 Stress curve of the abrasive grain in the cutting process. (a) Tangential force curve. (b) Normal force curve. | ||
Fig. 8 illustrates the surface topography (a–f) and corresponding von Mises stress distributions (g–l) of the AlSi10Mg alloy under varying abrasive penetration and cutting depths. As the penetration and cutting depth increased, a clear progression in material deformation and damage morphology was observed.51 At shallow depths (0.5–1.0 nm), the surface displayed only minor ploughing marks with limited material pile-up. The contact zone remained relatively localized, suggesting primarily elastic to incipient plastic deformation. The associated von Mises stress distribution was concentrated directly beneath the abrasive tip, exhibiting a hemispherical stress field confined within the Al matrix.
![]() | ||
| Fig. 8 Surface morphologies at depth of cut of (a) 0.5, (b) 1, (c) 1.5, (d) 2, (e) 2.5, (f) 3 nm and their corresponding stress distributions (g)–(l). | ||
As the cutting depth increased from 1.5 to 2.0 nm, the surface features became more pronounced, exhibiting broader material uplift and asymmetric groove profiles. This indicated enhanced plastic flow and directional shearing. In such cases, the von Mises stress fields expanded both laterally and vertically, forming stress “tails” that extended toward the Al–Si interface. Notably, the stress intensity was higher near the Si region, implying that interface mismatch began to influence the local mechanical response. At the maximum cutting depth between 2.5 and 3.0 nm, severe material distortion was observed. The grooves became deeper, with sharp material extrusion and increased surface roughness. The corresponding von Mises maps showed extensive stress accumulation along both the cutting direction and beneath the abrasive tip. The accumulation of significant stress directly affected the nature and extent of subsurface damage in the material, thereby influencing its performance during service.
The effect of stress accumulation on SSD was further analyzed using the lattice structure diagram of the model. The lattice structure analysis presented in Fig. 9 visually revealed SSD, including dislocations, dislocation loops, vacancies, atomic clusters, and stacking fault defects. As the cutting depth increased, the nature and extent of subsurface damage became increasingly complex. At lower cutting depths (0.5–1.0 nm), the primary damage modes observed in the Al phase were stacking faults and incipient dislocation loops. These features emerged due to the relatively soft and ductile nature of Al, which allowed for plastic accommodation through dislocation nucleation and glide.52 The presence of stacking faults—particularly noticeable near the cutting path—indicated that partial dislocations were activated in response to shear stresses generated by the abrasive particle.
![]() | ||
| Fig. 9 Distribution and types of subsurface damage under different cutting depths. (a) 0.5 nm. (b) 1 nm. (c) 1.5 nm. (d) 2 nm. (e) 2.5 nm. (f) 3 nm. | ||
As the cutting depth increased to 1.5–2.0 nm, dislocation loops and stacking faults became more prominent and began to penetrate deeper into the material. Notably, closed dislocation rings were observed around the cutting trajectory in both the Al and Si regions, suggesting the presence of localized stress concentration and zones of plastic relaxation. This transition marked the onset of cross-interface interaction, where the stress field began to significantly affect the more brittle Si phase.
At the highest cutting depths of 2.5–3.0 nm, extensive lattice distortion was observed in both phases. In the Al region, dislocation tangles and dense stacking fault networks developed, indicating significant plastic deformation. Meanwhile, in the Si region, the damage mode transitioned from isolated dislocations to partial amorphization near the cutting zone. The loss of long-range order, as evident by irregular atomic arrangements and the absence of clear lattice planes, suggested that a mechanically induced amorphous layer was formed.53 This phenomenon is characteristic of severe shear deformation in brittle materials subjected to high contact stress.
In summary, the formation of stacking faults, dislocation rings, and amorphous regions is strongly dependent on the cutting depth. These defects initiate sequentially with increasing depth, reflecting a transition from elastic/plastic deformation to fracture-dominated removal, particularly in the Si phase. The evolution of these subsurface features offers critical insight into the material removal mechanism in Al–Si alloys and reinforces the need for chemical-assisted cutting strategies when targeting the embedded Si phase.
![]() | ||
| Fig. 10 Electrochemical tests of AlSi10Mg specimens in different slurries. (a) Tafel plots. (b) Corrosion current densities and corrosion potentials. (c) Nyquist plots. (d) Bode plots. | ||
The EIS results presented in Fig. 10(c) and (d) further corroborate the protective capabilities of different CMP slurries. In the Nyquist plot, the smallest semicircle observed for the S4 slurry indicated the lowest charge transfer resistance (Rct), suggesting severe corrosion under alkaline conditions.55 The addition of H2O2 and tyrosine in the S5 and S6 slurries increased Rct, reflecting the formation of passivation or complexation layers. Remarkably, the S7 slurry exhibited the largest Rct and impedance modulus in the Bode plot, along with the highest phase angle, confirming that the synergistic effect of H2O2 and tyrosine led to the formation of a compact and protective interfacial layer that effectively inhibited the corrosion of AlSi10Mg.
The XPS full spectra of AlSi10Mg analyzed before and after immersion in the CMP slurry are shown in Fig. 11(a). After immersion in slurry S5, a significant increase in oxygen (O) intensity was observed, confirming the strong oxidative effect of H2O2. Following immersion in slurry S6, a noticeable reduction in the Al signal was observed, indicating that the slurry constituents chemically interacted with the Al surface. Furthermore, the appearance of a nitrogen (N) peak was attributed to the presence of tyrosine.56 A significant change in the content of Si and SiO2 before and after polishing suggests that the CMP process effectively removed the surface oxide layer on Si, revealing a smoother underlying Si surface, as presented in Fig. 11(b) and (c).
The Al 2p spectra of the samples before and after immersion in the CMP slurry are shown in Fig. 11(d)–(f). The peaks located at 74.38, 74.58, and 74.58 eV indicated the presence of Al3+, whereas those at 71.88, 72.28, and 72.18 eV were associated with metallic Al.57 The decreased intensity of Al indicates that the Al surface was either oxidized or complexed. Both H2O2 and tyrosine exerted beneficial regulatory effects on Al. Similarly, as shown in Fig. 11(g)–(i), the peaks located at 531.28, 531.78, and 531.68 eV were associated with Al2O3, and those at 531.98, 532.08, and 532.18 eV were attributed to Al(OH)3. Before immersion, a dense oxide layer composed primarily of Al2O3 dominated the surface of AlSi10Mg. However, the introduction of H2O2 reversed this trend, resulting in a predominance of Al(OH)3.41 The peak at 535.08 eV corresponded with the complexation between Al and tyrosine. Notably, the enhanced intensity of the Al–tyrosine complex and the reduced intensity of Al2O3 suggested that tyrosine dissolved the dense oxide layer before forming a chelate with Al ions. This reaction occurred continuously, enabling the rapid removal of surface asperities, which contributed to improved Sa and MRR values.
The surface chemical states of the AlSi10Mg alloy were analyzed both before and after immersion in the CMP slurry using XPS, FTIR, and Raman spectroscopy, as illustrated in Fig. 12. The survey spectra in Fig. 12(a) showed enhanced signals related to Ce following immersion, indicating the adsorption or deposition of Ce-containing species on the alloy's surface. This supports the role of CeO2 abrasive in the surface modification process. The Ce 3d spectra in Fig. 12(b) exhibited characteristic peaks for both Ce3+ and Ce4+ after immersion, suggesting the presence of redox-active Ce species.58 These redox interactions facilitated the chemical softening and removal of native silicon oxide layers, thereby accelerating the polishing of Si-rich regions and improving planarization. The chemical reaction equation between the CMP slurry and AlSi10Mg can be expressed as follows:59
| 2Al + 3H2O2 → Al2O3 + 3H2O | (4) |
| Al2O3 + 3H2O → 2Al(OH)3 | (5) |
| Al(OH)3 + OH− → [Al(OH)4]− | (6) |
| Si + H2O2 → SiO2 + H2↑ | (7) |
| Ce4+ + Si → Ce3+ + SiO2 | (8) |
| CeO2 + SiO2 + H2O2 → Ce3+ + SiO2·H2O | (9) |
Fig. 12(c) displays the FTIR spectra of AlSi10Mg surfaces treated with S8 slurries. The broad absorption features around 3200 cm−1 and 1450 cm−1 were respectively associated with the O–H stretching and bending vibrational modes. Moreover, the absorption bands around 2930 cm−1, 1580 cm−1, and 1320 cm−1 corresponded to the stretching vibrations of N–H, C
O, and C
C bonds, respectively.60,61 These bands implied the existence of functional groups originating from tyrosine on the polished surface, indicating successful complexation and adsorption. Additionally, novel peaks observed at ∼560 cm−1 were linked to Ce–O vibrations, further confirming CeO2 deposition.62 These complementary spectroscopic results collectively demonstrated that the CeO2 abrasive in the S8 slurry actively participated in surface reactions, particularly with Si components, while tyrosine and H2O2 contributed to aluminum surface modification. Likewise, Fig. 12(d) presents the Raman spectra of the surface before and after immersion. Subsequent immersion in the S8 slurry led to a significant change in the peak intensity and position, including the emergence of new peaks associated with Al–O and Ce–O bonds.63 Overall, these spectroscopic results provide strong evidence for the synergistic roles of H2O2, tyrosine, and CeO2 in forming a chemically active, protective, and removable surface layer that enhances CMP performance.
The complexation process between Al ions and varying numbers of tyrosine molecules is illustrated in Fig. 13. This interaction is pivotal for understanding the mechanism of corrosion inhibition, as the coordination of Al3+ with tyrosine molecules contributed to the formation of a protective layer on the AlSi10Mg surface. With the increasing number of tyrosine molecules involved, the stability of the resulting complexes was enhanced, effectively reducing the risk of further oxidation or corrosion of the Al substrate. This coordination was primarily expedited through the carboxyl (–COOH) and phenolic hydroxyl groups (–OH) of tyrosine, which exhibited strong affinity for Al ions. In some instances, the amino group (–NH2) of tyrosine also participated in coordination, leading to a tridentate binding mode. Consequently, stable chelate rings were formed, effectively isolating the Al3+ ions from the corrosive environment.
The electron-donating capacity of the tyrosine functional groups not only contributed to complex stability but also altered the local electronic structure of the aluminum surface, further enhancing passivation. Such a multilayered complexation and adsorption model provided a plausible explanation for the synergistic inhibition effect observed in the presence of both tyrosine and H2O2, as validated by the electrochemical and XPS analyses. This proposed mechanism underscores the significance of molecular design in CMP slurry formulations, where tailored ligands like tyrosine can serve dual functions: corrosion inhibition and surface modification. The multidentate complexation of tyrosine not only enhances aluminum ion immobilization but also favors the formation of a dense organic–inorganic hybrid film, which explains the improved surface smoothness (low Sa) and MRR after CMP using the S7 slurry.
The CMP mechanism of the AlSi10Mg alloy is illustrated in Fig. 14, based on findings from electrochemical testing, XPS, FTIR, and Raman spectroscopy. The CMP process for this Al–Si alloy, utilizing a composite slurry, operated through a synergistic removal mechanism governed by both chemical and mechanical interactions. Initially, H2O2 oxidized the surface of the material, creating a softened layer. Both Na2CO3 and Y2O3 significantly enhanced the dispersion of SiO2 particles, which helped minimize abrasive agglomeration. This reduction in agglomeration decreased localized defects and enabled consistent material removal across the Al and Si regions.
The chemical contributions of tyrosine and cerium oxide further modulated the CMP behavior. Tyrosine formed stable coordination complexes with surface Al atoms, weakening the native oxide layer and promoting efficient removal of the Al matrix under mild mechanical loading. This chelation-assisted polishing mechanism was manifested in the smoother Al regions shown in the magnified left inset of Fig. 12. In contrast, CeO2 enhanced the oxidative dissolution of the more brittle Si phase through redox interactions with Si–O bonds. The right inset revealed local amorphization and nanoscale material removal in Si-rich areas, indicating the active chemical participation of CeO2. Together, these processes created a balanced CMP mechanism wherein chelation facilitated Al removal, oxidation accelerated surface degradation of Si, and uniformly dispersed abrasives provided the mechanical force necessary for material removal. This integrated approach ensures high-quality surface finishing while maintaining selectivity between the softer Al matrix and the harder Si inclusions.
(1) The optimized slurry yielded a minimum Sa of 0.187 nm within a 50 × 50 μm2 area and a high MRR of 17.23 μm h−1, meeting the demanding surface requirements for aerospace optical components.
(2) Under alkaline conditions, Y2O3 enhances the stability and dispersion of SiO2 abrasives, as evidenced by a significant increase in zeta potential and a reduction in polydispersity index. This enhancement mitigated abrasive agglomeration and reduced surface defects during polishing.
(3) CeO2 exhibited redox-assisted reactivity with Si phases, promoting selective material removal at phase interfaces and contributing to overall surface uniformity and polishing efficiency.
(4) The combination of hydrogen peroxide and tyrosine ensured a balanced oxidation–dissolution mechanism. Electrochemical tests confirmed their synergistic corrosion inhibition, while XPS, FTIR, and Raman analyses revealed the formation of surface oxide and complexation layers.
(5) TEM characterization and MD simulations revealed the evolution of subsurface damage under different cutting depths. Stacking faults, dislocation loops, and amorphization zones were observed, particularly near phase interfaces, offering valuable insight into phase-specific deformation responses during polishing.
Overall, this work presents a sustainable and high-efficiency CMP strategy tailored for Al–Si alloys with complex microstructures. The findings provide a robust framework for future slurry design and polishing optimization in high-precision optical and aerospace applications.
| This journal is © The Royal Society of Chemistry 2025 |