Next Article in Journal
Experimental Study on the Anti-Erosion of the Exterior Walls of Ancient Rammed-Earth Houses in Yangjiatang Village, Lishui
Next Article in Special Issue
Electric Field-Assisted Chemical Bath Deposition of ZnO Thin Films: Effects of Field Intensity, Polarity Inversion, and Air Agitation on Film Properties
Previous Article in Journal
The Charge Transport Properties of Polycrystalline CVD Diamond Films Deposited on Monocrystalline Si Substrate
Previous Article in Special Issue
Formation of Protective Coatings on TZM Molybdenum Alloy by Complex Aluminosiliconizing and Application of a Preceramic Layer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

First-Principles Study on the Microheterostructures of N-GQDs@Si3N4 Composite Ceramics

1
College of Mechanical and Electrical Engineering, Shaanxi University of Science & Technology, Xi’an 710021, China
2
Shaanxi Agricultural Products Processing Technology Research Institute, Xi’an 710021, China
3
Key Laboratory of Materials and Surface Technology (Ministry of Education), School of Materials Science and Engineering, Xihua University, Chengdu 610039, China
*
Author to whom correspondence should be addressed.
Coatings 2025, 15(10), 1172; https://doi.org/10.3390/coatings15101172
Submission received: 27 August 2025 / Revised: 29 September 2025 / Accepted: 29 September 2025 / Published: 7 October 2025

Abstract

In the previous research that aimed to enhance the toughness and tribological properties of silicon nitride ceramics, a lignin precursor was added to the ceramic matrix, which achieved conversion through pyrolysis and sintering, resulting in a silicon nitride-based composite ceramic containing nitrogen-doped graphene quantum dots (N-GQDs). This composite material demonstrated excellent comprehensive mechanical properties and friction-wear performance. Based on the existing experimental results, the first-principles plane wave mode conservation pseudopotential method of density functional theory was adopted in this study to build a microscopic heterostructure model of Si3N4-based composite ceramics containing N-GQDs. Meanwhile, the surface energy of Si3N4 and the system energy of the N-GQDs@Si3N4 heterostructure were calculated. The calculation results showed that when the distance between N-GQDs and Si3N4 in the heterostructure was 2.3 Å, the structural energy was the smallest and the structure was the steadiest. This is consistent with the previous experimental results and further validates the coating mechanism of N-GQDs covering the Si3N4 column-shaped crystals. Simultaneously, based on the results of the previous experiments, the stress of the heterostructure composed of Si3N4 particles coated with different numbers of layers of nitrogen quantum dots was calculated to predict the optimal lignin doping amount. It was found that when the doping amount was between 1% and 2%, the best microstructure and mechanical properties were obtained. This paper provides a new method for studying the graphene quantum dot coating structure.

1. Introduction

As an excellent engineering ceramic material, Si3N4 ceramics are widely used in tribological components and the aerospace and automotive industries (e.g., in aviation bearings, tools, etc.) due to their unique thermodynamic, tribological and chemical properties [1]. In practical applications, various sizes of defects and cracks inevitably form in Si3N4 ceramic materials, showing the inherent brittleness of Si3N4 ceramic materials. On the other hand, the dry tribological properties of single-phase Si3N4 against metal pairs are not ideal [2,3,4]. Many researchers have proposed that graphene materials should be added to the silicon nitride matrix in order to enhance the toughness of the ceramic material and reduce friction [5,6]. However, the preparation process of pure graphene is usually very complicated, and graphene is prone to agglomeration in ceramic matrices. Therefore, directly adding graphene to enhance the mechanical properties and tribological properties of ceramic materials still poses some potential problems.
Scholars have found that lignin, as a well-dispersed precursor, can be converted into graphene materials such as graphene quantum dots (GQDs) in situ during the process of pyrolysis conversion [7,8,9]. GQDs are a subset of carbon quantum dots (CDs) derived from graphene or graphene oxide, with a diameter of 2~20 nm. As a new type of carbon nanomaterial, it has excellent properties such as low cytotoxicity, stable photoluminescence and excellent solubility. Some researchers have been focusing on the preparation of GQDs and their application as functional materials in recent years [10,11,12]. In addition, GQDs can be monodispersed well when used as a coating material for Ag. Ag@GQD nanoparticles have higher surface-enhanced Raman scattering (Sers) activity, and its peak intensity is 6–8 times that of Ag nanoparticles [13]. However, the effect of GQDs on the tribological properties and mechanical properties of matrix materials has not been reported by other scholars.
Recently, our research group [14,15,16,17] was surprised to find that, when nano-scale lignin was added to Si3N4 ceramics as a precursor, a core–shell structure of N-GQD-coated β-Si3N4 columnar crystals appeared after pyrolysis and hot-pressing sintering. The interlayer spacing of the N-GQDs was 2.2~2.3 Å, and the cross-sectional size was 8.99 nm2. Meanwhile, the ceramic composite presented excellent mechanical and tribological properties. However, the research on the special core–shell structure of N-GQD-coated β-Si3N4 columnar crystals is still in the experiment stage using SEM, TEM and other characterization methods (core–shell phenomenon and core–shell effect) and has failed to explain the formation mechanism.
The formation of the coated structure is closely related to the surface properties of the Si3N4 grains. Therefore, in order to study the coating mechanism of N-GQDs to cover Si3N4, it is necessary to start with studying its surface properties. Due to the anisotropy of the material, different planes in the bulk material exhibit inconsistent surface characteristics, and the surface energy can be considered as a physical variable to explain these surface characteristics [18]. However, the experimental determination of surface energy is difficult [19], and the surface properties of Si3N4 have not been studied. Therefore, many calculation and simulation methods have been proposed to calculate the surface energy of anisotropic materials, including first-principles methods [20,21,22], tight-binding parameterizations [23,24] and semi-empirical methods [25,26,27]. Meanwhile, the first-principles method is considered to be an effective tool for studying the surface properties of materials by accurately calculating the surface energy [28,29,30,31].
Some scholars have studied the stability of the various surfaces of anisotropic materials using first-principles calculations of surface energy [19,32,33]. The lower the surface energy, the more stable the surface is and vice versa. For example, Bao et al. [19] systematically studied the structure and surface energy of Mg2Pb (001), (110) and (111) surfaces using first-principles calculations. The calculated results showed that the surface energies of the Mg2Pb (111)-Mg and Mg2Pb (110)-MgPb surfaces were the lowest, indicating that the Mg2Pb (111)-Mg and Mg2Pb (110)-MgPb surfaces were the most stable. Sun et al. [32] studied the surface stability and equilibrium morphology of MoO3 using first-principles calculations. The results showed that the surface energy of MoO3 was (010) < (101) < (001) < (100). This is consistent with Tokarz-Sobieraj‘s result [33]: the (010) surface had the lowest surface energy, indicating that the (010) surface should have the lowest reactivity. The higher the surface energy, the smaller the equilibrium morphology area and the higher the chemical activity.
Based on the discussion above, the goal of this study was to select the appropriate crystal plane by calculating the surface energy of β-Si3N4, and then build the N-GQDs@β-Si3N4 heterogeneous interface structure to calculate the energy to verify the coating mechanism and calculate the stress of the heterostructure of Si3N4 grains coated with different numbers of layers of N-GQDs to optimize the lignin doping amount. The surface energies of β-Si3N4 along three low Miller index surfaces (100), (110) and (111) and three high-peak crystal surfaces (200), (210) and (101) were studied using first-principles calculations. The two active surfaces (111) and (200) of Si3N4 were obtained and used to build heterostructures with N-GQDs. By calculating the energy of N-GQDs@β-Si3N4 heterostructures with different spacings, it was verified that the cladding structure produced by the experiment had the most stable structure. Different numbers of layers of N-GQDs were added to the (200) crystal plane of β-Si3N4 to obtain heterostructures. The mechanical properties were predicted by calculating the stress of the system, and then the lignin doping amount of was optimized. This study provides theoretical guidance for further study of the surface coating structure of GQDs.

2. Calculation Model and Calculation Method for Core–Shell Structure

2.1. Core–Shell Structure Theoretical Model Construction

It is well known that there are three common crystal structures of Si3N4: the α, β and γ phases. The β phase belongs to the hexagonal system, and the space group of β-Si3N4 is P63/m [34]. The lattice constants are a = 7.66 Å, b = 7.66 Å and c = 2.93 Å. The spatial position of each atom in the unit cell is shown in Table 1. According to the experimental results, the Si3N4 matrix of the composite material is basically β phase, so based on the above data, this study built the crystal model of β-Si3N4 as shown in Figure 1a (blue atoms represent nitrogen and yellow atoms represent silicon).
The lattice constants of graphite are a = 2.46 Å, b = 2.46 Å, c = 6.8 Å, α = β = 90 ° and γ = 120 °. In order to match the actual results, the N-GQD model adopts the approximate structure of graphene doped with nitrogen atoms. The graphite structure shown in Figure 1b is cut by 0.5 layers along the (001) direction, and the obtained structure is graphene.

2.2. Structure Optimization and Surface Energy Calculation Method

The calculation first builds a β-Si3N4 crystal model and optimizes the structure, and then calculates the surface energy of β-Si3N4 along three low Miller index surfaces ((100), (110) and (111)) and three high-peak crystal surfaces ((200), (210) and (101)) to find the most active surface. Finally, the heterostructure model was built with N-GQDs, and the formation mechanism of the core–shell structure was revealed by calculating the energy of the N-GQDs@β-Si3N4 heterostructures with different spacings.

2.2.1. Structural Optimization Method

The reason why the crystal can maintain a stable shape and volume is that there are two types of attraction and repulsion between atoms. In order to construct the surface model of different crystal planes, the structure of Si3N4 was optimized first so that the atomic spacing reaches the equilibrium distance, reduces the interaction between atoms, and maintains a stable state. We chose the structure with the smallest volume change rate. At this time, the overall spatial structure and the relative position between atoms did not change much during the optimization process, which ensured the physical stability of the structure. In addition, the structure with a small volume change rate was often in a relatively low energy and stable state. The volume change rate of structural optimization was defined as follows:
Deviation   ( % )   = V Final   V Initial V Initial
In the formula, Deviation represents the volume deviation, V I n i t i a l represents the volume before structural optimization, and V F i n a l represents the volume after structural optimization.
In this paper, the calculation was carried out using CASTEP [35] based on the density functional theory (DFT). The exchange–correlation functional was represented by the Perdew–Burke–Ernzerhof (PBE) functional [36] of the generalized gradient approximation (GGA). The valence electrons involved in the calculation for β-Si3N4 were 3s23p2 of Si and 2s22p3 of N. The LBFGS algorithm [37] was used to optimize the structure of the unit cell model, and the pseudopotential was based on the model-conserving pseudopotential. In order to achieve sufficient convergence accuracy, the self-consistent cyclic convergence accuracy was 5.0 × 10−7 eV/atom, the stress error between atoms was less than 0.02 GPa, the maximum displacement convergence error between atoms was less than 5.0 × 10−4 Å, and the convergence criterion of total energy was less than 5 × 10−6 eV/atom.

2.2.2. Surface Energy Calculation Method

Surface energy is an important parameter for studying surface properties. It refers to the energy required per unit area when the crystal is dissociated into two semi-infinite crystals under the action of an external force, which can reflect the surface stability [38]. Due to the anisotropy of the Si3N4 material lattice, different surface characteristics would be generated on different surfaces. The anisotropy makes the surface energy of different crystal planes different. The crystal surface with a high surface energy is in a high energy state. In order to reduce the energy, it can react with the external material to form a more stable surface structure. The crystal surface with a low surface energy is relatively stable and the reactivity is low. Since the coating process would preferentially occur on the active surface of the Si3N4 crystal and based on the actual test results, three low Miller index surfaces ((100), (110) and (111)) and three high-peak crystal surfaces ((200), (210) and (101)) were selected in this study. The active surface of the Si3N4 crystal was found by using surface energy calculations as the basis for the subsequent heterostructure construction. In order to describe the stability of the surface based on the structural optimization and according to the characteristics of different crystal planes of β-Si3N4, six crystal cells with different crystal planes were constructed to study the surface energy of the (100), (110), (111), (200), (210) and (101) surfaces.
When the crystal faces in these cells were in contact with other atoms, considering the short-range interaction between the atoms and the surface, we only considered five adjacent atomic layers near the (100), (110), (111), (200), (210) and (101) surfaces. When the atomic layer was equal to or greater than the fifth layer, the interaction energy between the atom and the surface was considered to be approximately zero due to the long-range effect. A 15 Å vacuum layer was applied to separate the model in the Z direction to prevent unnecessary interactions along the model surface.
In order to calculate the surface energy of each crystal face, the variable Esurf was defined as the surface energy of Si3N4, which was used to describe the minimum energy in the process of forming two free surfaces. It can be expressed by the following formula:
Esurface = [Eslab − nEbulk]/2A
where Esurface represents the surface energy of the unit cell, Esalb is the total energy of the lamellar crystal model, Ebulk is the total energy of the bulk crystal, n is the number of bulk crystals contained in the surface model, and A is the surface area of the surface model.

3. Results and Analysis

3.1. Energy Calculation of β-Si3N4 Low-Index Crystal Face Heterostructure

3.1.1. Structural Optimization

The convergence test of the plane wave cutoff energy and k-point density of the β-Si3N4 crystal was carried out to complete the structural optimization of β-Si3N4. Firstly, the convergence test of plane wave truncation energy was carried out with a Brillouin zone sampling of 2 × 2 × 6. It was found that when the cutoff energy of the plane wave substrate was 780 eV, the structure had the smallest volume change rate of 0.63663%. The specific data are shown in Table 2. Subsequently, the cut-off energy of the plane wave substrate was set to 780 eV, and the convergence test of the k-point density was performed. It was found that when the Brillouin zone sampling was 2 × 2 × 5, the structure had the smallest volume change rate of 0.63357%. The specific data are shown in Table 3.
In summary, when the cutoff energy of the plane wave substrate was 780 eV and the Brillouin zone sampling was 2 × 2 × 5, the structure had the smallest volume change rate, which was closest to the experimental value. Thus, the structural optimization of the β-Si3N4 crystal model was completed and the most stable structure was obtained. It has a plane wave energy close to the cutoff of 770 eV and the k-point density of 2 × 2 × 6 adopted by Jiang et al. [34].
In the theoretical study, the surface energy is the energy consumed in order to destroy the chemical bonds between molecules on the surface of the material. Therefore, the surface energy is an index to measure the surface stability. The smaller the surface energy, the better the surface stability. Because the surface formed by the core–shell structure must be the most active surface of the β-Si3N4 crystal, the surface with the largest surface energy was obtained by calculating and comparing the surface energy of different crystal planes of the β-Si3N4 crystal to determine the most active surface of the β-Si3N4 crystal. Combined with the results of the actual test (shown in Figure 2), three low Miller index surfaces (100), (110) and (111) were selected in this study. The active surface of the β-Si3N4 crystal was found by calculating the surface energy of the above surfaces.
Due to the long-range effect, when the number of atomic layers was equal to or greater than five layers, the interaction energy between the internal atoms and the surface was considered to be approximately zero so the maximum thickness of this study was only calculated to five adjacent atomic layers near the surface. The unit cell model of the different crystal planes of the four-layer β-Si3N4 crystal is shown in Figure 3.

3.1.2. Surface Energy Calculation of Low Miller Index Surfaces

In order to select the crystal plane with the highest surface energy, the surface energy convergence test with one to five layers was carried out on the (100), (110) and (111) surfaces of the β-Si3N4 crystal. The surface energy data of the (100), (110) and (111) surfaces are shown in Table 4.
The results show that the surface energy of β-Si3N4 was (100) < (110) < (111). The surface energy of the (111) surface was the highest, which indicates that the (111) surface is more active, the surface interaction is stronger, and it is easier to bind to graphene quantum dots. It can be seen from Table 4 that the surface energy of the (111) surface reached convergence in four layers, so the four-layer structure of the (111) surface was selected for the construction of the heterogeneous interface structure. In particular, when the number of layers of the β-Si3N4 (100) crystal plane was five, the surface energy was negative. Our analysis suggests that the core reason for the negative surface energy of the (100) crystal plane is that the model thickness does not meet the convergence requirements, resulting in a deviation between the volumetric energy simulation and the actual situation. The strong relaxation effect and low intrinsic surface energy characteristics that may exist in the β-Si3N4 (100) crystal plane itself eventually lead to the calculated value deviating from the physical reality. Negative values are not “negative surface energy” in the physical sense, but rather a manifestation of the limitations of the model.

3.1.3. Surface Relaxation of Low Miller Index Surfaces

During the surface cutting process, the symmetry of the surface layer is destroyed relative to the bulk interior, and the CN (coordination number) of the surface atoms is reduced, so the surface atoms will move to a new position [19]. Surface relaxation occurs because the force on the surface atom is different from the bulk structure so the surface atom layer moves relative to the inner atom to reduce the energy of the system. The relaxation degree of the crystal surface plays an important role in the properties of the surface [38]. Therefore, it is necessary to perform surface relaxation on the surface structure.
The surface relaxation was performed on the four-layer cell of the (111) crystal plane of β-Si3N4, the surface relaxation was performed on the surface of the five-layer atoms of the cell, and the remaining part was fixed as a bulk structure.
The convergence test of the plane wave cutoff energy and k-point density was carried out on the unit cell with four layers on the (111) surface of the β-Si3N4 crystal to complete the surface relaxation of the unit cell model with four layers on the (111) surface of the β-Si3N4 crystal. Firstly, using a Brillouin zone sampling of 2 × 2 × 5, the convergence test of plane wave truncation energy was carried out. It was found that when the cutoff energy of the plane wave substrate was 930 eV, the structure had the smallest volume change rate of 0.01520%. The specific data are shown in Table 5.
Subsequently, the cutoff energy of the plane wave substrate was set to 930 eV, and the convergence test of the k-point density was performed. It was found that when the Brillouin zone sampling was 2 × 2 × 5, the structure had the smallest volume change rate of 0.01520%. The specific data are shown in Table 6.
Combined with the above calculation results, when the cutoff energy was 930 eV and the k-point density was 2 × 2 × 5, the unit cell structure of the (111) surface of the β-Si3N4 crystal with four layers had the smallest volume change rate and the system energy converged well.

3.1.4. Energy Calculation of Heterostructures

In terms of crystal structure, N-GQDs are graphene fragments with a size of less than 100 nm and have obvious graphene lattice properties. For the modeling of N-GQDs@β-Si3N4 heterostructures, the four-layer structure model of the (111) plane of the β-Si3N4 crystal and the N-GQD structure model were combined in this study.
In this study, the optimized lattice constant of graphene was a = b = 2.46 Å, and the optimized lattice constant of the β-Si3N4 crystal (111) crystal plane was a =b = 8.22 Å. The modeling method of the N-GQDs@β-Si3N4 heterostructure is as follows. The 3 × 3 supercells on the (111) surface of the β-Si3N4 crystal are used as the bottom layer, and the GQDs form 10 × 10 supercells. Following the method of a related study [39], three C atoms are substituted with N atoms on each 5 × 5 supercell model to construct the N-GQD model (as shown in Figure 4a). Finally, the double-layer N-GQDs are matched with the (111) surface of the β-Si3N4 crystal to form the N-GQDs@β-Si3N4 heterostructure (as shown in Figure 4b). The interlayer spacing between N-GQDs was 2.3 Å.
During the formation of the heterostructures, the N-GQDs were subjected to 0.077% and 0.008% strain on the a and b lattice parameters when adapting to the lattice of the heterostructures, and 0.150% and 0.016% strain was exerted on the (111) surface of the β-Si3N4 crystal. The model is periodic in two dimensions (a and b), and the third dimension is associated with the thickness of the system. When the double-layer N-GQDs were matched with the (111) plane of the β-Si3N4 crystal, the resulting lattice mismatch was tolerable for the system. The optimized lattice parameters of the N-GQDs (001) and β-Si3N4 (111) surfaces are reported in Table 7 along with the lattice parameters of the heterojunctions.
In this study, the energy of different distances between the N-GQDs and β-Si3N4 heterostructure interface was calculated as the starting point. The distances between the N-GQDs and β-Si3N4 were 1.0, 1.5, 2.3, 3.0, 4.0 and 10.0 Å. The energy calculation results for the heterostructures with different spacings between the N-GQDs and β-Si3N4 are shown in Table 8.
The results show that, as expected, the calculated data from the heterostructure are consistent with the experimental data. Compared with other spacings, when the spacing between the N-GQDs and β-Si3N4 was 2.3 Å, the minimum energy of the system was −165,087.23 eV, indicating that this heterostructure system was the most stable at this time. Further research shows that the energy of the N-GQDs@β-Si3N4 heterostructure basically showed an increasing trend as the distance between the N-GQDs and β-Si3N4 increased, which means that the microstructure of the N-GQDs and β-Si3N4 dispersed with each other was unstable. Therefore, in the pyrolysis–hot pressing sintering coupling process, from an energy point of view, in order to obtain a stable crystal structure, the composite material is bound to spontaneously form a core–shell structure.

3.2. Energy Calculation of β-Si3N4 High Peak Crystal Plane Heterostructure

In order to further study the actual crystal plane of β-Si3N4 combined with N-GQDs, the high peak surface energy of the crystal plane was calculated. The peak values of the β-Si3N4 (200), (210) and (101) crystal planes were the highest (as shown in Figure 2), indicating that they were the most unstable crystal planes. Among them, the surface energy of the crystal plane with a high crystal plane index was high, and its particle density was small, the spacing was large, and it was unstable, making it easier to combine with N-GQDs. Therefore, the surface energy of the three β-Si3N4 crystal planes (200), (210) and (101) was calculated, and the heterostructure was further built to calculate the energy of the system

3.2.1. Surface Energy Calculation of High-peak Crystal Surfaces

The unit cell models of the β-Si3N4 (101), (200) and (210) crystal planes are shown in Figure 5. The surface energies of the β-Si3N4 (101), (200) and (210) crystal planes with different thicknesses are shown in Table 9.
The results show that the surface energy of β-Si3N4 was (210) < (101) < (200). The surface energy of the (200) surface was the highest, which indicates that the (200) surface is more active, the surface interaction is stronger, and it is easier for it to bind to N-GQDs.
A sufficiently thick Slab supercell has very similar internal properties and bulk material properties so the surface energy will tend to be stable with an increased Slab thickness. If the accuracy of Eslab and Ebulk is exactly the same, Equation (2) can be used to accurately calculate the surface energy. However, there are some inevitable differences in the calculation (e.g., the k-point grid cannot be exactly the same), and the accuracy of the bulk and surface calculations cannot be completely consistent, which will cause the surface energy to deviate linearly with increasing n. It can be seen from Table 9 that the surface energy of the (200) surface tended to diverge as the number of layers increased.
In order to calculate the surface energy more accurately, the surface energy of the (200) surface was calculated using the following method [40]: The bulk energy Ebulk is not calculated using a separate bulk phase and the energy Ebulk (n) corresponding to the gradually increasing number of layers Slab is calculated first, and the relationship between Ebulk (n) and the number of layers n is fitted. When n is large enough, the bulk energy Ebulk is obtained using the fitted linear slope, which avoids the error caused by inconsistent parameters. For the (200) surface, by calculating the surface energy corresponding to the Slab supercell with 1–5 atomic layers, the change in the surface energy with the thickness of Slab is calculated. The fluctuation in the surface energy was also very small and it converged easily (as shown in Table 10). Therefore, the β-Si3N4 (200) crystal plane with four atomic layers was selected for subsequent heterostructure construction and calculation.

3.2.2. Surface Relaxation of High-peak Crystal Surfaces

The plane wave truncation energy and the convergence of the k-point density of the unit cell with four layers of thickness on the (200) surface of the β-Si3N4 crystal were tested, and the surface relaxation of the unit cell model with four layers on the (200) surface of the β-Si3N4 crystal was completed. Firstly, the Brillouin zone sampling was set to 2 × 2 × 5 and the convergence test of the plane wave truncation energy was carried out. It was found that when the cutoff energy of the plane wave substrate was 830 eV, the structure had the smallest volume change rate of 0.61589%. The specific data are shown in Table 11.
Subsequently, the cutoff energy of the plane wave substrate was set to 830 eV, and the convergence test of the k-point density was performed. It was found that when the Brillouin zone sampling was 2 × 2 × 5, the structure had the smallest volume change rate of 0.61589%. The specific data are shown in Table 12.
Combined with the above calculation results, when the cutoff energy was 830 eV and the k-point density was 2 × 2 × 5, the unit cell structure of the (200) surface of the β-Si3N4 crystal with four layers had the smallest volume change rate, and the system energy converged well.

3.2.3. Energy Calculation of Heterostructures

According to the modeling method for the N-GQDs and β-Si3N4 heterostructures, the β-Si3N4 (111) surface carries a 1 × 3 supercell as the bottom layer, and the GQD (001) surface carries a 3 × 2 supercell. One C atom on each 3 × 2 supercell model is substituted with a N atom, and the doping ratio is 4.17%. As shown in Figure 6a, 3 × 2 double-layer N-GQDs were selected to match the surface of β-Si3N4, resulting in a lattice mismatch that is tolerable for the system. The optimized lattice parameters of the N-GQDs (001) and β-Si3N4 (200) surfaces are reported in Table 13 and compared with the lattice parameters of the heterojunctions. In the formation of the heterostructures, the N-GQDs (001) were subjected to 1.9% and 0.52% strain on the a and b lattice parameters when adapting to the lattice of the heterostructures, and the β-Si3N4 (200) surface was subjected to 3.6% and 1% strain in turn.
Figure 6 b,c show heterostructure models composed of two layers of N-GQDs and β-Si3N4 (200) crystal planes. The interlayer spacing between the N-GQDs is 2.3 Å, and the spacings between the N-GQDs and β-Si3N4 are 2.3 and 10 Å, respectively. When the distance between the N-GQDs and β-Si3N4 is 2.3 Å, it represents the experimental coating structure, and when the distance is 10 Å, it is not coated. The energy calculation results for the N-GQDs and β-Si3N4 heterostructures with different spacings are shown in Table 14.
The results show that, as expected, the calculated data from the heterostructure are consistent with the experimental phenomenon data. Compared with the distance of 10 Å, when the distance between the N-GQDs and β-Si3N4 was 2.3 Å, the system energy was smaller than −24,433.42 eV, indicating that the heterostructure system was more stable at this time.
In short, Si3N4 has a high surface energy, and the surface is in an unstable state with high energy. When N-GQDs are in contact with Si3N4, N-GQDs will cover the surface of Si3N4, reducing the contact area between the surface of Si3N4 and the external environment, thereby reducing the energy of the whole system and making the system reach a more stable state.

3.2.4. Stress Calculation of Graphene and β-Si3N4 Heterostructure with Different Numbers of Layers

Stress describes the internal forces per unit area generated within a crystalline material due to external forces (e.g., lattice distortion) or internal states (such as thermal vibration or defects). Through the calculation of stress, the changes in the macroscopic mechanical properties of the material can be reflected to a certain extent [41]. The changes in the mechanical properties of pure Si3N4 with a (200) crystal plane and graphene with different numbers of layers are shown in Table 15. The crystal structure model of the pure Si3N4 (200) crystal plane and Si3N4 with different numbers of layers of graphene is shown in Figure 7.
It can be seen from Table 15 that the maximum stress was observed with two layers and the minimum was observed with four layers. The smaller the stress, the lower the hardness, the greater the toughness, the greater the bending strength, and the greater the density while the bulk density is constant. The change in the apparent porosity depends on the properties of the material and the specific application conditions. Therefore, when the number of layers of N-GQDs was two, the hardness of the composite was the highest. When the number of layers was four, the toughness, bending strength and density of the composite were the highest.
The experimental results show that the lignin doping amount was proportional to the number of layers of N-GQDs. When the nano-lignin doping amount was 2 wt.%, the best comprehensive performance was obtained. At this time, the number of layers of N-GQDs was 6–8 layers. Figure 7 shows the crystal structure model of Si3N4-based composites with different numbers of layers of graphene. It is predicted that in order to improve the mechanical properties of the composites, the doping amount should be between 1 wt.% and 2 wt.%.

4. Analysis of Theoretical Calculation Results and Experimental Results

In previous studies [14,15,16,17] using transmission electron microscopy, the composite material was shown to be composed of N-GQDs coated on the surface of β-Si3N4 columnar crystals. The size of the N-GQDs was 8.99 nm2, the number of layers was 6~8, and the interlayer spacing was 2.2~2.3 Å (as shown in Figure 8b).
Previous research [42] showed that the lower the energy of the system, the more stable the system is. Therefore, in order to make the structure of the system more stable, the material will spontaneously form a microstructure with the lowest energy. Therefore, the stability can be verified by calculating the structural energy, and the formation mechanism of the core–shell structure can be further revealed. In this chapter, the energy of N-GQDs@β-Si3N4 heterostructure is calculated. It was found that the system had the smallest energy and the most stable microstructure when the distance between the two was 2.3 Å (as shown in Figure 4b and Figure 6b).
The theoretical calculation results show that the N-GQDs@β-Si3N4 heterostructure had the smallest overall energy and the most stable microstructure when the spacing was 2.3 Å. The actual test results show that the microstructure of the composite material formed a core–shell structure with a layer spacing of 2.2~2.3 Å after the pyrolysis–hot pressing sintering coupling process. The theoretical calculation results are consistent with the actual test results. Therefore, the formation mechanism of the core–shell structure can be obtained. In order to obtain a more stable microstructure, during hot-pressing sintering, N-GQDs spontaneously coat the surface of β-Si3N4 columnar crystals, thus forming a lower energy and more stable core–shell structure.
Previous experimental results [17] showed that the core–shell structure formed is also different when different amounts of nano-lignin are doped. When the doping amount is too much, the thickness of the N-GQD aggregation also increases. From the calculation results, it can be seen that when the number of N-GQDs layers is too high, the adsorption capacity of the active surface of the β-Si3N4 columnar crystal for external N-GQDs is reduced so it cannot be subjected to uniform stress, resulting in an uneven coating.

5. Conclusions

In this study, a β-Si3N4 crystal model was built and the structure was optimized. Then, the surface energy of the low-plane index crystal plane and the high-peak crystal plane of the β-Si3N4 crystal was calculated and studied. The (111) crystal plane in the low-plane index crystal plane and the (200) crystal plane in the high-peak crystal plane were determined to be the most active surfaces. Finally, the heterostructure model was built by combining the crystal model with N-GQDs. The formation mechanism of the core–shell structure was revealed by calculating the energy of the heterostructures with different spacings of N-GQDs@β-Si3N4. The results showed the following:
(1)
Si3N4 had a high surface energy, and surface was in an unstable state with high energy. When N-GQDs were in contact with Si3N4, the N-GQDs would cover the surface of Si3N4, reducing the contact area between the surface of Si3N4 and the external environment, thereby reducing the energy of the whole system and making the system reach a more stable state.
(2)
Comparing the energy of N-GQDs@β-Si3N4 heterostructures with different spacings, the lowest system energy was obtained and the heterostructure system was the most stable when the spacing between the N-GQDs and β-Si3N4 was 2.3 Å. This spacing was consistent with the previous experimental data. As the distance between the N-GQDs and β-Si3N4 increased, the energy of the N-GQDs@β-Si3N4 heterostructure basically increased, indicating that the microstructure of the N-GQDs and β-Si3N4 dispersed with each other was unstable. Because the material always spontaneously changed into a stable structure, the N-GQDs would spontaneously coat on the surface of the β-Si3N4 crystal during the pyrolysis–hot pressing sintering coupling process so that the system had the minimum energy, and the composite material also formed the most stable microstructure-core–shell structure.
(3)
N-GQDs formed a core–shell structure by adsorbing onto the surface of the β-Si3N4 crystal, and the content of nano-lignin had a certain influence on the performance of the core–shell structure. It was predicted that the best mechanical properties of the composites would be obtained with a lignin doping amount between 1 wt.% and 2 wt.%.

Author Contributions

Conceptualization, W.C. and E.X.; methodology, Y.M.; software, Y.L. and Y.M.; validation, Y.M., Z.S. and Y.T.; formal analysis, Y.L.; investigation, E.X.; resources, W.C.; data curation, Y.T.; writing—original draft preparation, Y.M.; writing—review and editing, W.C. and Y.L.; visualization, R.L.; supervision, J.Z.; project administration, J.Z.; funding acquisition, W.C. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to acknowledge the support from the National Natural Science Foundation of China (No. 52175178), National Natural Science Foundation of China (No. 22378252), Natural Science Foundation of Shaanxi Province (No. 2018JM5056), Key Laboratory of Materials and Surface Technology, Ministry of Education (No. 2024-yb007) and Shaanxi Agricultural Products Processing Technology Research Institute (No. L2024-ZDYF-ZDYF-GY-0044).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

This already includes all the data of the research work.

Conflicts of Interest

The authors declare no conflicts of interest. The manuscript was approved by all authors for publication. The work described in this article is original research that has not been published previously and is not under consideration for publication elsewhere.

References

  1. Mir, A.H.; Ahmad, S.N. A study on fabrication of silicon nitride-based advanced ceramic composite materials via spark plasma sintering. J. Mater. Des. Appl. 2021, 235, 1739–1756. [Google Scholar] [CrossRef]
  2. Zhang, J.; Liu, J.; Wang, Z.; Chen, W.; Hu, B.; Zhang, Y.; Liao, H.; Ma, S. Tribological behavior and lubricating mechanism of Si3N4 in artificial seawater. Ceram. Int. 2020, 46, 14361–14368. [Google Scholar] [CrossRef]
  3. Wang, W.; Wen, H.; He, N.; Chen, W. Effect of load on tribological properties of silicon nitride/steel under rolling-sliding contact condition. Tribol. Int. 2018, 125, 27–38. [Google Scholar] [CrossRef]
  4. Wu, C.; Liu, Z.; Zhao, H.; Yang, H.; Li, X.; Ni, J. Effect of the grease thickener on tribological properties of Si3N4/GCr15 contact interface and the performance in hybrid ceramic ball bearing. Ceram. Int. 2023, 49, 16857–16867. [Google Scholar] [CrossRef]
  5. Almansoori, A.; Balázsi, K.; Balázsi, C. Advances, Challenges, and Applications of Graphene and Carbon Nanotube-Reinforced Engineering Ceramics. Nanomaterials 2024, 14, 1881. [Google Scholar] [CrossRef]
  6. Chen, F.; Yan, K.; Hong, J.; Song, J. Synergistic effect of graphene and β-Si3N4 whisker enables Si3N4 ceramic composites to obtain ultra-low friction coefficient. Tribol. Int. 2023, 178, 108045. [Google Scholar] [CrossRef]
  7. Zhao, Y.; Wen, M.; He, C.; Liu, C.; Li, Z.; Liu, Y. Preparation of Graphene by Catalytic Pyrolysis of Lignin and Its Electrochemical Properties. Mater. Lett. 2020, 274, 128047. [Google Scholar] [CrossRef]
  8. Liu, W.; Ning, C.; Sang, R.; Hou, Q.; Ni, Y. Lignin-derived graphene quantum dots from phosphous acid-assisted hydrothermal pretreatment and their application in photocatalysis. Ind. Crop. Prod. 2021, 171, 113963. [Google Scholar] [CrossRef]
  9. Ding, Z.Y.; Wang, X.L.; Sun, R.C. Gram-scale synthesis of single-crystalline graphene quantum dots derived from lignin biomass. Green. Chem. 2018, 20, 1383–1390. [Google Scholar] [CrossRef]
  10. Cho, H.; Bae, G.; Hong, B.H. Engineering functionalization and properties of graphene quantum dots (GQDs) with controllable synthesis for energy and display applications. Nanoscale. 2024, 16, 3347–3378. [Google Scholar] [CrossRef]
  11. El Haddad, Y.; Ouarrad, H.; Drissi, L.B. Underlying mechanism of hetero-ring doping GQDs for OLEDs, photovoltaic and nanomedical applications. Opti. Quant. Electron. 2024, 56, 966. [Google Scholar] [CrossRef]
  12. Bhuvaneswari, C.; Elangovan, A.; Sudhan, N.; Vinodhkumar, G.; Saravanan, S.; Balasubramanian, V.; Sharmila, C.; Karuppaiah, S. A low-cost hybrid GQDs/Fe3O4/polypyrrole nanocomposite based chemo-sensor for electrochemical non-enzymatic selective determination of creatinine in biological samples. Microchem. J. 2023, 194, 109259. [Google Scholar] [CrossRef]
  13. Wang, J.; Gao, X.; Sun, H.; Su, B.; Gao, C. Monodispersed graphene quantum dots encapsulated Ag nanoparticles for surface-enhanced Raman scattering. Mater. Lett. 2016, 162, 142–145. [Google Scholar] [CrossRef]
  14. Chen, W.; Xu, E.; Liu, X.; Zhao, Z.; Li, Z.; Zhu, P.; Lou, R.; Zhang, Z.; Li, H. Study on a novel Si3N4-based composite with the incorporation of N-GQDs produced from nano-lignin. Ceram. Int. 2024, 50, 19534–19542. [Google Scholar] [CrossRef]
  15. Chen, W.; Ma, Y.; Liu, X.; Xu, E.; Yang, W.; Jia, J.; Lou, R.; Zhu, C.; Wu, C.; Zhao, Z. Effect of load on the tribological properties of Si3N4-based composite with N-GQDs. Ind. Lubr. Tribol. 2024. [CrossRef]
  16. Chen, W.; Zhao, Z.; Lou, R.; Li, H.; He, N. Study on self-derived products of nanometer lignin in silicon nitride ceramics during sintering process. Results Mater. 2021, 12, 100228. [Google Scholar] [CrossRef]
  17. Chen, W.; Xu, E.; Zhao, Z.; Wu, C.; Zhai, Y.; Liu, X.; Jia, J.; Lou, R.; Li, X.; Yang, W.; et al. Study on mechanical and tribological behaviors of N-GQDs@Si3N4 composite ceramics. Tribol. Int. 2023, 179, 108095. [Google Scholar] [CrossRef]
  18. Tyson, W.R.; Miller, W.A. Surface free energies of solid metals: Estimation from liquid surface tension measurements. Surf. Sci. 1977, 62, 267–276. [Google Scholar] [CrossRef]
  19. Bao, L.; Kong, Z.; Qu, D.; Duan, Y. First-principles investigation on the structures, energies and electronic properties of low-index surfaces of Mg2Pb. Mater. Chem. Phys. 2021, 260, 124028. [Google Scholar] [CrossRef]
  20. Skriver, H.L.; Rosengaard, N.M. Surface energy and work function of elemental metals. Phys. Rev. B 1992, 46, 7157–7168. [Google Scholar] [CrossRef]
  21. Methfessel, M.; Hennig, D.; Scheffler, M. Trends of the surface relaxations, surface energies, and work functions of the 4d transition metals. Phys. Rev. B 1992, 46, 4816. [Google Scholar] [CrossRef] [PubMed]
  22. Kollár, J.; Vitos, L.; Skriver, H.L. Surface energy and work function of the light actinides. Phys. Rev. B 1994, 49, 11288. [Google Scholar] [CrossRef] [PubMed]
  23. Mehl, M.J.; Papaconstantopoulos, D.A. Applications of a tight-binding total energy method for transition and noble metals: Elastic Constants, Vacancies, and Surfaces of Monatomic Metals. Phys. Rev. B 1996, 54, 4519–4530. [Google Scholar] [CrossRef] [PubMed]
  24. Barreteau, C.; Spanjaard, D.; Desjonquères, M.C. Electronic structure and energetics of transition metal surfaces and clusters from a new spd tight-binding method. Surf. Sci. 1999, 433–435, 751–755. [Google Scholar] [CrossRef]
  25. Foiles, S.M.; Baskes, M.I.; Daw, M.S. Embedded-atom-method functions for the fcc metals Cu, Ag, Au, Ni, Pd, Pt, and their alloys. Phys. Rev. B 1986, 33, 7983. [Google Scholar] [CrossRef]
  26. Wolf, D. Correlation between energy, surface tension and structure of free surfaces in fcc metals. Surf. Sci. 1990, 226, 389–406. [Google Scholar] [CrossRef]
  27. Baskes, M.I. Modified embedded-atom potentials for cubic materials and impurities. Phys. Rev. B. 1992, 46, 2727–2742. [Google Scholar] [CrossRef]
  28. Xiong, H.; Yang, S.; Zhu, K.; Yang, J.; Feng, J.; Zhou, X. Interface energy, adhesion work and electronic structure of Al/PdCu interface calculated from first principles. Mater. Today. Commun. 2025, 45, 112385. [Google Scholar] [CrossRef]
  29. Wang, Z.; Chen, D.; Fang, Q.; Chen, H.; Fan, T.; Liu, B.; Liu, F.; Tang, P. Effects of finite temperature on the surface energy in Al alloys from first-principles calculations. Appl. Surf. Sci. 2019, 479, 499–505. [Google Scholar] [CrossRef]
  30. Wang, L.; Lai, K.C.; Huang, L.; Evans, J.W.; Han, Y. Low-index surface energies, cleavage energies, and surface relaxations for crystalline NiAl from first-principles calculations. Surf. Sci. 2020, 695, 121532. [Google Scholar] [CrossRef]
  31. Zhong, M.; Zeng, W.; Liu, F.-S.; Tang, B.; Liu, Q.-J. First-principles study of the atomic structures, electronic properties, and surface stability of BaTiO3 (001) and (011) surfaces. Surf. Interface. Anal. 2019, 51, 1021–1032. [Google Scholar] [CrossRef]
  32. Sun, S.P.; Zhu, J.L.; Gu, S.; Li, X.P.; Lei, W.N.; Jiang, Y.; Yi, D.Q.; Chen, G.H. First principles investigation of the surface stability and equilibrium morphology of MoO3. Appl. Surf. Sci. 2019, 467–468, 753–759. [Google Scholar] [CrossRef]
  33. Tokarz-Sobieraj, R.; Hermann, K.; Witko, M.; Blume, A.; Mestl, G.; Schlögl, R. Pro perties of oxygen sites at the MoO3(0 1 0) surface: Density functional theory cluster studies and photoemission experiments. Surf. Sci. 2001, 489, 107–125. [Google Scholar] [CrossRef]
  34. Jiang, W.L.; Dai, T.; Zhang, F.; Zhang, Y.L.; Liu, Z.T.; Du, Y.H.; Liu, Q.J. First-principles calculations of electronic, optical and thermodynamic properties of α and β-Si3N4. J. At. Mol. Phys. 2023, 40, 8. [Google Scholar] [CrossRef]
  35. Segall, M.D.; Lindan, P.J.D.; Probert, M.J.; Pickard, C.J.; Hasnip, P.J.; Clark, S.J.; Payne, M.C. First-principles simulation: Ideas, illustrations and the CASTEP code. J. Phys. Condens. Mat. 2002, 14, 2717–2744. [Google Scholar] [CrossRef]
  36. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1998, 77, 3865–3868. [Google Scholar] [CrossRef]
  37. Fischer, T.H.; Almlof, J. General methods for geometry and wave function optimization. J. Phys. Chem. 1992, 96, 9768–9774. [Google Scholar] [CrossRef]
  38. Wei, L.T.; Tong, Z.F.; Wu, D.H.; Zhou, S.C.; Wang, H.C.; Luo, T.P.; Tang, B.Y. First-principles study of surface properties of calcium oxide (100), (110) and Ca-terminated (111). Guangxi Sci. 2015, 22, 670–674. [Google Scholar] [CrossRef]
  39. Hou, B.P.; Gan, Z.L.; Lei, X.L.; Zhong, S.Y.; Xu, B.; Ouyang, C.Y. First-principles study on the mechanism of nitrogen-doped graphene as a cathode material for lithium-air batteries to reduce oxygen molecules. Acta Phys. Sin. 2019, 68, 303–312. [Google Scholar] [CrossRef]
  40. Boettger, J.C.; Birkenheuer, U. Extracting convergent surface formation energies from slab.calculations. J. Phys. Condens. Mat. 1996, 8, 6525–6529. [Google Scholar] [CrossRef]
  41. Zhang, S.; Rao, L.; Shao, W.; He, Q.; Xing, X.; Zhou, Y.; Yang, Q. Electronic, Tensile, and Sliding Characteristics of the C/Ti Interface: A First-Principles Study. Langmuir 2022, 38, 15113–15120. [Google Scholar] [CrossRef]
  42. Patra, S.; Das, P.; Rajbhar, M.K. Formation of core-shell nanostructure through wrapping of cuprous oxide nanowires by hydrogen titanate nanotubes. Radiat. Phys. Chem. 2022, 196, 110103. [Google Scholar] [CrossRef]
Figure 1. (a) β-Si3N4 and (b) graphene crystal models.
Figure 1. (a) β-Si3N4 and (b) graphene crystal models.
Coatings 15 01172 g001
Figure 2. XRD patterns of different crystal planes of N-GQDs@Si3N4 composite ceramics.
Figure 2. XRD patterns of different crystal planes of N-GQDs@Si3N4 composite ceramics.
Coatings 15 01172 g002
Figure 3. Cell model of β-Si3N4 (a) (100), (b) (110) and (c) (111) crystal planes.
Figure 3. Cell model of β-Si3N4 (a) (100), (b) (110) and (c) (111) crystal planes.
Coatings 15 01172 g003
Figure 4. (a) N-GQD model and (b) heterostructure model of N-GQDs@β-Si3N4.
Figure 4. (a) N-GQD model and (b) heterostructure model of N-GQDs@β-Si3N4.
Coatings 15 01172 g004
Figure 5. Cell model of β-Si3N4 (a) (101), (b) (200) and (c) (210) crystal planes.
Figure 5. Cell model of β-Si3N4 (a) (101), (b) (200) and (c) (210) crystal planes.
Coatings 15 01172 g005
Figure 6. (a) N-GQD model; (b,c) heterostructure model of β-Si3N4 and N-GQDs.
Figure 6. (a) N-GQD model; (b,c) heterostructure model of β-Si3N4 and N-GQDs.
Coatings 15 01172 g006
Figure 7. (a) Crystal structure of pure Si3N4 (200) plane and Si3N4 with (b) 2 layers, (c) 4 layers and (d) 7 layers of N-GQDs.
Figure 7. (a) Crystal structure of pure Si3N4 (200) plane and Si3N4 with (b) 2 layers, (c) 4 layers and (d) 7 layers of N-GQDs.
Coatings 15 01172 g007
Figure 8. Comparison between test results and calculation results [14,16,17].
Figure 8. Comparison between test results and calculation results [14,16,17].
Coatings 15 01172 g008
Table 1. Atomic coordinates of β-Si3N4 crystal.
Table 1. Atomic coordinates of β-Si3N4 crystal.
WyckoffElementxyz
2dN2/31/31/4
6hSi0.825140.593891/4
6hN0.670060.700741/4
Table 2. Volume deviation of different plane wave cutoff energies of β-Si3N4 crystal.
Table 2. Volume deviation of different plane wave cutoff energies of β-Si3N4 crystal.
Cutoff Energy (eV)760770780790
Deviation (%)0.643420.639780.636630.63691
Table 3. Volume deviation of different k-point densities of β-Si3N4 crystal.
Table 3. Volume deviation of different k-point densities of β-Si3N4 crystal.
K-Point Density2 × 2 × 42 × 2 × 52 × 2 × 63 × 3 × 6
Deviation (%)0.634060.633570.636630.63927
Table 4. Surface energy of (100), (110) and (111) β-Si3N4 surfaces with different thicknesses.
Table 4. Surface energy of (100), (110) and (111) β-Si3N4 surfaces with different thicknesses.
N (Layer Number)S1S2S3S4S5
ESurf
(J/M2)
(100)2.2181.6460.9230.274−0.343
(110)2.3572.5702.0891.7181.347
(111)2.1791.5433.2413.2293.162
Table 5. Volume deviation of different plane wave cutoff energies of 4-layer β-Si3N4 crystal ((111) crystal plane).
Table 5. Volume deviation of different plane wave cutoff energies of 4-layer β-Si3N4 crystal ((111) crystal plane).
Cutoff Energy (eV)830880930980
Deviation (%)0.024450.026070.015200.01521
Table 6. Volume deviation of different k-point densities of 4-layer β-Si3N4 crystal ((111) crystal plane).
Table 6. Volume deviation of different k-point densities of 4-layer β-Si3N4 crystal ((111) crystal plane).
K-Points3 × 3 × 12 × 2 × 56 × 6 × 2
Deviation (%)0.025660.015200.01551
Table 7. The calculated structural properties of the N-GQDs (001) and β-Si3N4 (111) parts compared with the calculated structural properties of the N-GQDs@β-Si3N4 model. We report the cell vector (a, b), angle (γ) and surface area (S).
Table 7. The calculated structural properties of the N-GQDs (001) and β-Si3N4 (111) parts compared with the calculated structural properties of the N-GQDs@β-Si3N4 model. We report the cell vector (a, b), angle (γ) and surface area (S).
Systema (Å)b (Å)γ (°)S (Å2)
N-GQDs (001)24.60024.600120.0492.710
β-Si3N4 (111)24.65624.594107.2561.807
N-GQDs@β-Si3N424.61924.598115.7519.743
Table 8. Energy of N-GQDs@β-Si3N4 (111) heterostructures with different spacings.
Table 8. Energy of N-GQDs@β-Si3N4 (111) heterostructures with different spacings.
Spacing (Å)1.01.52.33.04.010.0
E (eV)−165,086.85−165,086.92−165,087.23−165,086.43−165,086.93−165,085.86
Table 9. Surface energy of (101), (200) and (210) crystal planes of β-Si3N4 crystal with different thicknesses.
Table 9. Surface energy of (101), (200) and (210) crystal planes of β-Si3N4 crystal with different thicknesses.
N (Layer Number)S1S2S3S4S5
ESurf
(J/M2)
(101)0.19850.3260.3290.3270.326
(200)31.34262.58893.798125.013156.224
(210)0.0580.1370.1590.1260.110
Table 10. The change in surface energy of the β-Si3N4 (200) crystal plane with the number of layers.
Table 10. The change in surface energy of the β-Si3N4 (200) crystal plane with the number of layers.
Tier Number (n)nEbulk (eV)Eslab (eV)A (Å2)Esurf (J/M2)
1−1404.948−1402.17222.5270.062
2−2809.896−2808.61722.5270.028
3−4214.844−4216.71222.527−0.041
4−5619.792−5624.62022.527−0.107
5−7024.739−7032.66722.527−0.176
Table 11. Volume deviations of different plane wave cutoff energies of 4-layer β-Si3N4 crystal ((200) crystal plane).
Table 11. Volume deviations of different plane wave cutoff energies of 4-layer β-Si3N4 crystal ((200) crystal plane).
Cutoff Energy (eV)730780830880
Deviation (%)0.648010.627420.615890.61641
Table 12. Volume deviations of different k-point densities of 4-layer β-Si3N4 crystal ((200) crystal plane).
Table 12. Volume deviations of different k-point densities of 4-layer β-Si3N4 crystal ((200) crystal plane).
K-Points4 × 10 × 12 × 2 × 57 × 19 × 2
Deviation (%)0.710970.615890.70273
Table 13. The calculated structural properties of the N-GQDs (001) and β-Si3N4 (200) surface compared with the calculated structural properties of the N-GQDs@β-Si3N4 model. We report the cell vector (a), (b), angle (γ) and surface area (S).
Table 13. The calculated structural properties of the N-GQDs (001) and β-Si3N4 (200) surface compared with the calculated structural properties of the N-GQDs@β-Si3N4 model. We report the cell vector (a), (b), angle (γ) and surface area (S).
Systema (Å)b (Å)γ (°)S (Å2)
N-GQDs (001)7.3798.5209062.865
β-Si3N4 (200)7.7998.65289.999767.478
N-GQDs@β-Si3N47.5198.56489.999964.390
Table 14. Energy of N-GQDs@β-Si3N4 (200) crystal plane heterostructures with different spacings.
Table 14. Energy of N-GQDs@β-Si3N4 (200) crystal plane heterostructures with different spacings.
Spacing (Å)2.310
E (eV)−24,433.42−24,433.29
Table 15. The stress of heterostructures of N-GQDs with different numbers of layers and β-Si3N4 (200) crystal plane.
Table 15. The stress of heterostructures of N-GQDs with different numbers of layers and β-Si3N4 (200) crystal plane.
GQD Content (Number of Layers)0123457
Stress (GPa)−0.012 2−0.2270.367 10.0590.006−0.0340.027
1 Positive values indicate compressive stress. 2 Negative values indicate compressive stress.
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

Chen, W.; Li, Y.; Ma, Y.; Xu, E.; Lou, R.; Sun, Z.; Tian, Y.; Zhang, J. First-Principles Study on the Microheterostructures of N-GQDs@Si3N4 Composite Ceramics. Coatings 2025, 15, 1172. https://doi.org/10.3390/coatings15101172

AMA Style

Chen W, Li Y, Ma Y, Xu E, Lou R, Sun Z, Tian Y, Zhang J. First-Principles Study on the Microheterostructures of N-GQDs@Si3N4 Composite Ceramics. Coatings. 2025; 15(10):1172. https://doi.org/10.3390/coatings15101172

Chicago/Turabian Style

Chen, Wei, Yetong Li, Yucheng Ma, Enguang Xu, Rui Lou, Zhuohao Sun, Yu Tian, and Jianjun Zhang. 2025. "First-Principles Study on the Microheterostructures of N-GQDs@Si3N4 Composite Ceramics" Coatings 15, no. 10: 1172. https://doi.org/10.3390/coatings15101172

APA Style

Chen, W., Li, Y., Ma, Y., Xu, E., Lou, R., Sun, Z., Tian, Y., & Zhang, J. (2025). First-Principles Study on the Microheterostructures of N-GQDs@Si3N4 Composite Ceramics. Coatings, 15(10), 1172. https://doi.org/10.3390/coatings15101172

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