Shijia Liua,
Jiaqi Liub,
Jialin Liub and
Jiuye Chen*c
aSchool of OptoElectronic Engineering, Changchun University of Science and Technology, Changchun 130022, China
bHalifax Grammar School, Halifax B3H 2Y2, Canada
cSchool of Science, Changchun University of Science and Technology, Changchun 130022, China. E-mail: chenjiuye@mails.cust.edu.cn
First published on 9th December 2024
Ultrafast laser processing technology exhibits exceptional precision and irreplaceable functionality in the fabrication of micron and nanometer-scale devices. However, its short action time presents challenges for observing and studying the interactions between ultrafast lasers and materials. This study employs molecular dynamics simulations to specifically investigate the application of ultrafast laser processing in treating amorphous structural defects on Ni–Fe alloy surfaces. The simulations reveal the impact of energy deposition on the material's crystallization behavior on a nanosecond timescale. It was found that the crystallization temperature increases with the rising rate of temperature change, although the final crystal structure remains unchanged. Enhanced energy deposition accelerates lattice formation, improves atomic ordering, and reduces the crystallization time from 4.5 ns to 3.2 ns. The lattice phase transition is completed within 0.5 ns, and an increased incubation temperature effectively minimizes the proportion of the amorphous phase. The simulation results clearly illustrate the fundamental nucleation and growth mechanisms, providing valuable insights into the effects of ultrafast laser processing on surface lattice structures and atomic dynamics. Moreover, these findings establish a theoretical foundation and offer data support for developing future material processing methods.
However, despite the extensive application of ultrafast laser processing technology in material processing, studies on subtle structural changes in materials, particularly at the nanoscale, remain insufficient. The interaction time of ultrafast lasers with materials is on the order of nanoseconds (10−9 seconds), which poses challenges for accurately observing the entire processing process. During ultrafast laser irradiation, material surfaces may undergo nanosecond-scale melting and recrystallization processes,13 which directly affect the chemical composition, crystal structure, and morphology of the material surface.14 There is no doubt that different structures significantly influence the properties of materials. The mechanical properties of amorphous structures differ greatly from those of crystalline structures,15 the magneto crystalline anisotropy and saturation magnetization of alloys can change with variations in their crystalline phases.16 Therefore, it is crucial to observe and control the changes in various lattices during metal processing. Although theoretical models exist to explain the processing mechanism,17–19 there is still a lack of research on finer structures, particularly at the atomic level. Furthermore, research on nucleation behavior at ultrafast timescales is relatively scarce, making it particularly challenging to explore the interactions between ultrafast lasers and materials in depth. Fortunately, advancements in computational technology have made it possible to study atomic behavior at ultrafast timescales through computer simulations.20,21 Molecular dynamics (MD) simulations are commonly used to investigate phase transitions, especially the microscopic structural changes in metallic systems. Using the Large-scale Atomic/Molecular Massively Parallel Simulator (LAMMPS), it is possible to directly track and simulate nucleation behavior, and analyze and visualize phase structural transformations and related dynamic processes in material systems, providing significant support for the study of atomic dynamics in ultrafast laser processing.
This paper employs molecular dynamics simulations to investigate the impact of ultrafast laser processing on the surface morphology of metals under different parameters, using a Ni3Fe1 alloy as the sample. Ni3Fe1 alloy, a widely used alloy also found in space meteorites,22 is characterized by its distinct lattice structure: a perfect face-centered cubic structure (γ phase) in the stable state, a body-centered cubic phase (δ phase) in the metastable state, and an amorphous state with disordered atomic arrangements when defects are present or the processing is inadequate.23 The stable Ni3Fe1 alloy exhibits excellent soft magnetic properties and an extremely low thermal expansion coefficient,24 making it an ideal material for magnetic recording media, information storage, as well as MEMS and radio frequency devices.25 It is evident that the amorphous structure significantly affects the performance of alloys, particularly in the study of structural transformations in Ni–Fe alloys. Additionally, the distinct and easily discernible structures of Ni–Fe alloys enable their effective application in simulation calculations. Thus, this study simulates the application of ultrafast laser processing technology in addressing amorphous structural defects on material surfaces, observes the transition from amorphous to crystalline states in the alloy structure, and calculates the factors influencing phase transitions. MD simulations utilize embedded atom potentials covering spatial scales from a few angstroms to several hundred angstroms on nanosecond timescales, making this approach suitable for simulating micro-nano scale structural morphologies. Through MD simulations, a comprehensive understanding of phase transition processes in ultrafast micro-nano processing can be achieved, offering deeper insights into the microscopic structure and properties of materials, and providing important guidance and support for optimizing ultrafast laser processing technology. Future work will extend this research to explore the micro-nano processing mechanisms of different material systems under ultrafast laser irradiation, aiming to provide more valuable information and insights for material processing development and applications.
(1) |
The OVITO (2.9.0) software was utilized for processing atomic images.28 As shown in Fig. 1, our simulation designated an amorphous region as the defect area to be processed. By injecting laser pulses into this region, we observed the evolution of the lattice. Throughout this process, we output atomic distribution images and temperature data at each step. Additionally, we calculated the radial distribution function (RDF) of the nuclear structure in the model, providing distribution data for body-centered cubic (BCC), face-centered cubic (FCC), and amorphous structures. These lattice structure data and RDF images served as criteria for evaluating surface morphology changes.
In addition, in this simulation, the Ni–Fe potential is constructed based on the Embedded Atom Method (EAM), which is highly efficient and effective in describing atomic interactions in transition metal systems, particularly suitable for body-centered cubic (BCC) and face-centered cubic (FCC) metals.29 The EAM treats each atom in the solid as embedded in a host solid composed of the remaining atoms. The energy functional is approximated through a potential function with two terms: (i) a sum of pairwise interactions ϕ(rij) between atoms i and j, and (ii) a sum of embedding energies F, which is a function of the atomic density rho. Alpha and Beta are the element types of atoms i and j:
(2) |
From eqn (2), it can be found that the atoms are brought to local minima of the potential energy surface, therefore avoiding potentials overcome limitations associated with simple pairwise interatomic potentials in simulations of metallic systems.
As we know, the Radial-Distribution-Function g(r) is commonly used to characterize whether a structure is crystalline or amorphous (hereafter referred to as RDF). It represents the distribution of other particles at a distance r from a given atom. To calculate the RDF, each atom is imagined to be at the center of a series of concentric spheres. RDF can then be obtained by the following formula:30
(3) |
(4) |
ln(β/T2) = −Ec/RT + constant | (5) |
Based on the simulation results presented in Fig. 1, it is evident that the energy input during laser processing facilitates a transition of atoms from a disordered state to an ordered state. Molecular dynamics simulations provide valuable insights into the specific mechanisms underlying this transformation. The simulation data reveal that the resulting lattice structure aligns with findings from existing studies,23,33–35 particularly regarding the formation of the face-centered cubic (FCC) crystal structure. As shown in Fig. 2d, the radial distribution function of the crystal structure closely approximates that of the ideal FCC structure, exhibiting similar peak positions, which suggests successful FCC structure formation and indicates that laser processing effectively promotes the reorganization of the lattice structure in amorphous materials. However, although the crystallization results match well with the ideal outcome at the positions of the first few nearest-neighbor peaks, there are slight differences in the peak values of the distribution function, revealing some imperfection in the crystal structure. This phenomenon may be due to the inclusion of defect regions in the statistical process of the distribution function, leading to a decrease and broadening of the radial distribution function peaks. Interestingly, the energy input during the simulation was insufficient to cause sample melting. Our simulation data also show that the sample temperature did not reach the melting point, which is crucial for the application of laser processing. In molten metals, there are many disordered atoms, and the properties exhibited by this disordered state are significantly different from those of the desired stable face-centered cubic structure. Therefore, it is preferable to avoid the formation of a molten state as much as possible without introducing additional complex structures. Since the melting point was not reached, the local structure of the sample can change without complete melting, demonstrating the unique advantage of laser processing in surface defect repair. This processing method is akin to an “annealing” process, allowing for localized improvement of the material's structure without disrupting its overall properties, thus enhancing material performance and longevity. Atomic projection maps at the interface at 300 K and 1000 K are also constructed in Fig. 2e and f. It is evident that the amorphous region (red area in Fig. 2e) gradually becomes more ordered with increasing temperature, with atoms starting to rearrange into an ordered lattice (red area in Fig. 2f). Under the same processing conditions, the crystallized regions are minimally affected, and the existing energy input remains insufficient to damage the already-formed lattice structure. Therefore, this processing method primarily targets the amorphous regions and promotes their crystallization through localized heating.
Notably, a body-centered cubic (BCC) structure appears in the orange region at the upper right of Fig. 2f. This phenomenon can be attributed to the lower number of atoms in the amorphous region within the same volume, resulting in a lower atomic density. Due to the lower coordination number of the BCC structure, this region is more likely to form such a structure rather than the FCC structure. Although this region has shown some ordered state structurally, it is still considered a “defect” compared to the expected FCC structure. This observation indicates that structural transformation issues exist during laser processing, which requires further investigation to address these “defects”. Therefore, we plan to investigate the treatment of the amorphous regions by studying both isothermal and non-isothermal crystallization methods.
The simulation results reveal that all energy curves exhibit a single exothermic crystallization peak associated with the crystallization process around 900 K as shown by the green band in Fig. 3. As the heating rate increases, the crystallization temperature rises from 855.2 K to 975.1 K, as detailed in Table 1. Consistent with experimental results on the crystallization process of amorphous alloys,36 the simulation results demonstrate a relationship between crystallization temperature and heating rate. We believe that the crystallization behavior has significant kinetic characteristics.37,38 As the temperature increases from 250 K to 1100 K, the atomic system acquires sufficient thermal energy to overcome activation energy barriers, facilitating nucleation. When the system temperature reaches the crystallization temperature, the potential energy of atoms drops sharply, and atoms in the amorphous state begin to progressively transform into an ordered crystal structure. The simulation results show that an increase in heating rate leads to a rise in crystallization temperature, which is directly related to the increased activation energy barrier caused by lattice thermal fluctuations. Therefore, the crystallization temperature of the Ni–Fe alloy model increases significantly with rising heating rates. To further confirm the effect of heating rate on the crystallization of the Ni–Fe alloy model, atomic correlations at different heating rates were analyzed using the RDF. Fig. 3 shows the RDF of the Ni3Fe1 alloy at heating rates of (a2) 90 K ns−1, (b2) 70 K ns−1, (c2) 50 K ns−1, (d2) 30 K ns−1, and (e2) 10 K ns−1. From Fig. 3a2–e2, all curves exhibit the typical face-centered cubic (FCC) structure, with the atomic structure showing long-range ordered characteristics. The first peak is sharp and has a higher intensity (red band), while the second and third neighbor peaks (gray and yellow bands) have lower intensities, as observed in the Ni–Fe alloy,39 and are consistent with the RDF of the ideal FCC lattice shown in Fig. 2d. Although the heating rate significantly affects the crystallization temperature of the Ni–Fe alloy, it has a negligible effect on lattice formation. This indicates that while the heating rate significantly alters the crystallization temperature, it does not noticeably affect the distribution of atomic distances. This can also be clearly observed from Table 1, indicating that the stable lattice structure of the atoms is less affected by the heating rate.
Heating rate (K ns−1) | FCC (%) | BCC (%) | Others (%) | Crystallization temperature (K) |
---|---|---|---|---|
10 | 70.3 | 1.1 | 22.3 | 855.2 |
30 | 70.7 | 1.0 | 22.5 | 899.8 |
50 | 70.1 | 1.1 | 22.7 | 939.9 |
70 | 69.3 | 0.9 | 23.4 | 950.2 |
90 | 70.0 | 0.9 | 23.2 | 975.1 |
Table 1 presents the relationship between heating rate and lattice formation. The data indicate that the proportion of face-centered cubic (FCC) lattice remains relatively unchanged as the temperature change rate increases. This suggests that optimizing amorphous defects through adjustments to the heating rate at a fixed final temperature has limited efficacy. During the crystallization process, the atomic distribution in the simulation system remains disordered below the crystallization temperature. Upon reaching the crystallization point, several clusters with initial lattice structures emerge as nuclei, leading to a process akin to heterogeneous nucleation. As the temperature increases, these nuclei gradually grow via diffusion. To enhance visualization, we expanded the simulation view to a 35 × 35 lattice region. Fig. 4 illustrates this crystallization process over time, with atoms color-coded by their lattice structures, where the FCC structure is indicated in green. It is evident that clusters near the FCC nucleus in the lower right corner gradually expand until all grains coalesce. This transition marks the disappearance of the amorphous structure and the emergence of a crystalline structure, confirming that this processing method aligns with the conventional grain boundary nucleation mechanism.40,41
It is evident that the rate of temperature change has a certain impact on the activation energy of crystallization, which can lead to fluctuations in the crystallization temperature within a certain range. Specifically, the rate of temperature change affects the rate of heat transfer within the material and the uniformity of the local temperature. This results in uneven heat distribution and thermal gradients within the material, which may cause fluctuations in the local crystallization temperature. Such fluctuations can alter the conditions for crystal formation to some extent, leading to variations in the crystallization temperature. This simulation has shown that a faster heating rate results in a higher crystallization temperature. However, the impact of temperature change rate on the nucleation rate and lattice fraction is relatively minor. As illustrated by eqn (4), the variation in nucleation rate is primarily influenced directly by temperature. With increasing temperature, more nucleation sites can form within a shorter period. Therefore, although the rate of temperature change does influence activation energy, the nucleation and growth processes remain more directly governed by temperature.
It can be inferred that temperature regulation has a more significant effect on the nucleation rate and lattice structure than adjustments in the rate of temperature change. To gain a deeper understanding of this process, we will investigate the changes in the critical crystallization temperature starting from the minimum crystallization temperature of 850 K obtained from Table 1. This temperature is generally considered the critical threshold for the transition of the material from an amorphous to a crystalline state. Building on this, we will further examine the crystallization of the amorphous structure under different temperature conditions, particularly analyzing the dynamic process of crystallization and the effect of temperature on the uniformity of the crystal structure. Through these studies, we aim to achieve a more comprehensive understanding of the temperature dependence in the crystallization process, which will contribute to optimizing the performance and processing techniques of crystalline materials.
Fig. 5a illustrates how the energy of the Ni–Fe alloy varies over time at temperatures of 820 K, 830 K, 840 K, 850 K, and 860 K. At 820 K, the energy remains nearly constant as incubation time increases, indicating that the alloy remains in an amorphous state due to insufficient atomic activation energy to transition from the disordered to the ordered phase.35 Consequently, the potential energy of the Ni–Fe alloy model is unaffected by incubation time at 820 K, suggesting there is a minimum threshold for laser energy input. Specifically, the laser energy must elevate the sample locally above 820 K to facilitate defect repair.
Fig. 5 (a) Potential energy vs. time for Ni–Fe alloy at different isothermal temperatures. (b) The radial distribution function (RDF) of the Ni–Fe alloy at different temperatures. |
At an isothermal temperature of 830 K, however, the model's overall energy shows a marked decrease with longer incubation times. This decline in potential energy is attributed to the transition of the lattice structure from disordered to ordered. Crystallization appears to be complete around 4.5 nanoseconds, with the phase transition occurring in approximately 0.5 nanoseconds, assuming the incubation period is disregarded. Notably, as the isothermal temperature rises, the incubation time decreases from 4.5 nanoseconds to 3.2 nanoseconds, as indicated in Table 2. This observation suggests that the nucleation and growth of the Ni–Fe alloy are highly sensitive to temperature changes. The expansion of the nucleation and growth regions leads to greater atomic ordering, aligning with the findings presented in Fig. 4.
Temperature (K) | FCC (%) | BCC (%) | Others (%) | Crystallization time (ns) |
---|---|---|---|---|
820 | 53.0 | 2.2 | 33.7 | — |
830 | 60.9 | 2.2 | 26.3 | 4.5 |
840 | 66.8 | 1.3 | 23.8 | 3.8 |
850 | 65.7 | 1.1 | 24.0 | 3.6 |
860 | 67.6 | 1.4 | 23.9 | 3.2 |
Additionally, as shown in Table 2, with increasing incubation temperature, the proportion of the FCC lattice in the overall structure rises, and the proportion of face-centered cubic structure also increases from 60.9% to 67.6% with the increase in temperature, indicating a strong temperature dependency. This suggests that precise control of lattice growth via temperature is a feasible approach. To further confirm the effect of temperature on atomic ordering, the radial distribution function (RDF) of the isothermal crystallization of the Ni–Fe alloy model was examined, as illustrated in Fig. 5b. At 820 K, the structure of the Ni–Fe alloy model is amorphous, exhibiting short-range order with a clearly defined first-neighbor peak, while the second and third neighbor peaks are obscured. As the isothermal temperature increases from 830 K to 860 K, changes in the RDF pattern are observed: the first peak narrows (red band in Fig. 5), and the intensity of the second and third neighbor peaks increases (yellow and gray bands in Fig. 5). This indicates a transition from short-range to long-range order in the alloy, with noticeable crystallization behavior kinetically. The appearance and growth of critical nuclei play a crucial role in the crystallization process of the alloy model. Critical nuclei exhibit strong spatial correlation with surrounding nuclei before crystallization begins, potentially leading to rapid subsequent crystallization. Therefore, with increasing isothermal temperature, the increased intensity and reduced width of the second and third neighbor peaks in the Ni–Fe alloy model confirm the feasibility of using ultrafast laser processing. Appropriate energy input can induce crystallization in amorphous regions of the alloy, ensuring lattice structure consistency and avoiding melting or phase explosion phenomena, thus providing a potential method for non-destructive alloy surface processing.
This journal is © The Royal Society of Chemistry 2024 |