3.2. EBSD Analysis
The EBSD technique can also be used to count the grain size of steel. It has been shown that large angular grain boundaries with an orientation difference of more than 15° can split the crystal structure and form so-called ‘equivalent grains’, whose diameter is known as the effective grain size.
Figure 6 shows the inverse polefigure (IPF) of each specimen obtained using EBSD. When statistically analyzing the effective grain size of steel, use a 15° misorientation as the criterion for delineation, exclude boundary grains and disregard the influence of twin boundaries in special grain boundaries. According to statistical analysis, the average effective grain size for each specimen, as detailed in
Table 5, is determined by considering the distribution and characteristics of grain sizes. The average effective grain size of each specimen is within 0.2 μm of the size measured by the Nano Measurer, which verifies the reliability of the Nano Measurer’s measurement results.
After this, the geometric dislocation density of each specimen was quantitatively calculated using KAM plots [
5], and the results are shown in
Figure 7 and
Table 6. It can be seen that the solid solution treatment can effectively reduce the dislocation density of the specimens in the hot rolled state, and the degree of reduction in dislocation density is greater with the increase in solid solution temperature and the prolongation of solid solution time.
3.3. Study on Precipitation Behavior Under Different Solid Solution Processes
In 200-series austenitic stainless steels, precipitated phases can seriously affect the mechanical properties and corrosion resistance of the materials. In order to study the morphology and distribution of precipitated phases under different solid solution treatment conditions, this study used transmission electron microscopy (TEM) to observe the distribution and type of precipitated phases in the original specimen and the solid solution-treated specimen and combined with energy spectroscopy (EDS) analysis to determine the chemical composition of the precipitated phases.
Figure 8 shows the bright field TEM results of the precipitated phases in the hot rolled specimen and the carbon film extracted replica samples of different solid solution-treated test steels. From the TEM results, it can be found that there are more precipitated phases in the original specimen and the size is larger, most of the precipitated phases are more than 200 nm in size, and, at the same time, there are many precipitation aggregation areas, which affect the mechanical properties and corrosion resistance of stainless steel hot rolled sheet. After solid solution treatment, the precipitated phase distribution is uniform, and the size is obviously smaller. Only sample #1 has several precipitated phases with a size larger than 60 nm, and the other four samples of the precipitated phase size are below 60 nm. The average size of the precipitated phase particles, the density of the precipitated phase and the volume fraction of the precipitated phase were statistically analyzed. The formula for calculating the volume fraction of the precipitated phase is as follows [
32].
where d
i is the diameter of individual precipitated phase particles in the field of view, nm; A is the actual area of the field of view, nm2; t
v + d is the thickness of the sample, and t
v is taken as 100 nm;
is the average size of the precipitated phase in the field of view, nm. Substituting the average diameter of the precipitated phase in the different solid solution processes in the statistics into Equation (3), the volume fractions of precipitated phases of individual samples were calculated as shown in
Table 7. From the table, it can be seen that the solid solution temperature has a great influence on the volume fraction of the precipitated phase; the volume fraction of the precipitated phase of the specimen held at 1040 °C for 5 min is more than 5 times that of the specimen held at 1080 °C for 5 min and is more than 30 times that of the specimen held at 1120 °C for 5 min.
A certain magnification revealed that the precipitated phases in the test steels used in this study were mainly of spherical and ellipsoidal morphology. The precipitated phases are mainly precipitated along the grain boundary, as shown in
Figure 9. To identify the types of precipitated phases, a transmission electron microscope (TEM) coupled with an energy-dispersive X-ray (EDX) spectrometer was utilized to perform surface scanning on various precipitated phases across random areas.
Figure 10 shows the results of the EDS surface scan, the upper left is the dark field TEM image, and the rest is the distribution of the elements. From the results of the surface scan it can be seen that the precipitated phases are mainly chromium carbides and nitrides.
Figure 11 shows the transmission electron micrographs of the precipitated phase in the test steel, the Selected Area Diffraction Pattern (SADP) of the precipitated phase and the calibration of the diffraction pattern. Among them, due to the small particle size of the precipitated phase in
Figure 11c, the arrangement of atoms in the precipitated phase was obtained using high-resolution transmission electron microscopy (HRTEM) images, and then the HRTEM images were subjected to the DigitalMicrograph (Version 3.22.1461.0) software. Fast Fourier transform (FFT) was applied to the HRTEM images using DigitalMicrograph software to obtain the same images as the diffraction spots, and more information could be obtained by calibrating them to determine the type of the precipitated phase. In the lower left corner of each image, the diffraction spots of the precipitated phases are displayed along with their calibration results, confirming the presence of primarily Cr
7C
3, Cr
23C
6 and Cr
2N. Cr
7C
3, a wear-resistant chromium carbide alloy developed by Beijing Naimo Company, boasts a surface hardness of HV1250. Cr
23C
6, with a density of 6.68 g/cm
3 and a melting point of 1890 °C, is known for its excellent wear, corrosion and oxidation resistance at high temperatures. Cr
2N is a nitride that precipitates at grain boundaries during heat treatment above 1210 °C and can transform into an ordered structure with increasing temperature. The reliability of the results was confirmed by comparison with the literature [
33,
34,
35]. Combined with the surface scanning results of EDS, it can be determined that the precipitated phases of this steel grade are mainly of the Cr
xC
y and Cr
xN
y types after hot rolling and solid solution treatment. No Mo-rich precipitates have been detected in the EDS analysis and diffraction spot calibration, which should be related to the low content of Mo.
3.4. Strengthening Mechanism Analysis
It is generally accepted that the strengthening mechanisms of metallic materials can be grouped into four categories: grain boundary strengthening, precipitation strengthening, dislocation strengthening and solid solution strengthening. A large number of previous statistics and studies have pointed out that the overall contribution of different strengthening mechanisms to yield strength can be calculated by the Hall−Petch correction formula [
32,
36,
37]. The calculation equation is as follows:
where
is the theoretically calculated yield strength, MPa;
is the lattice friction stress, taken as 54 MPa [
32];
is the grain boundary strengthening increment, MPa;
is the precipitation strengthening increment, MPa;
is the dislocation strengthening increment, MPa; and
is the solid solution strengthening increment, MPa.
Grain boundary strengthening is one of the important mechanisms to increase the yield strength of austenitic stainless steel. According to the Hall−Petch relationship [
3], the yield strength increment caused by grain boundary strengthening is inversely proportional to the grain size as shown in Equation (5).
where K is the Hall−Petch coefficient and d is the average grain size.
For austenitic stainless steels, the Hall−Petch coefficient is typically 316.15 MPa μm
1/2 [
38]. From the above statistics for grain size, it can be seen that the average grain sizes of specimens #1, #2, #3, #4 and #5 are 11.42 μm, 12.70 μm, 13.37 μm, 13.90 μm and 15.12 μm, respectively, and therefore the amount of change in yield strength due to grain boundary strengthening is shown in
Table 8.
Precipitation phases make an important contribution to the enhancement of material strength, which is mainly achieved by pinning dislocations and hindering dislocation movement. The precipitated phases, such as Cr
23C
6 and Cr
2N in this study, can effectively inhibit dislocation slip through pinning, thus increasing the yield strength and tensile strength of the material. According to the Orowan strengthening mechanism, the increment
of precipitation relative to the yield strength can be calculated using Equation (6) [
32].
where V
f is the volume fraction of the precipitated phase and d is the average size of the precipitated phase particles in μm. For individual particles, larger particles (>60 nm) provide more strength than smaller particles. However, for a given volume fraction, the strengthening of the small particle number effect may outweigh the strengthening of particle size on the whole due to the greater number of small particles, so their overall contribution to strengthening may outweigh the effect of larger particles, which is why #1 is lower than #2 precipitation strengthening increments. Substituting the average diameter of the precipitated phase and the calculated volume fraction into Equation (6), the precipitation strengthening increment of each specimen was calculated, as shown in
Table 9.
Dislocations, a common type of line defect in metal materials, significantly impact the material’s organizational structure and mechanical properties. As they move, dislocations interact with neighboring ones, creating a strengthening effect. The higher the dislocation density, the more intense the interaction and the more pronounced the subsequent strengthening effect. The dislocation strengthening increment
of the test steel can be calculated by the Bailey−Hirsch formula [
39], as shown in Equation (7).
where M is the Taylor factor, taken as 3.06; α is the scale factor related to the crystal structure, taken as 0.38; G is the shear modulus, taken as 79 GPa for the austenitic matrix; b is the Burgers vector, taken as 0.248 nm for Fe; and ρ is the dislocation density, m
−2. Substituting the dislocation density of each specimen in
Table 6, the dislocation enhancement increment for each specimen was calculated as shown in
Table 10.
The solid solution strengthening formula σs for experimental steels can be calculated using the following equation [
40].
where [M] (M = C, Si, Mn…) is the content of elements solidly dissolved in the austenitic matrix, wt.%. According to
Table 7, the volume fraction of precipitated phase under different solid solution processes is 0.077%, 0.051%, 0.023% and 0.013%, respectively. After calculation, if all elements in the precipitated phase are solidated, the strengthening increment on the matrix will not exceed 3.0 MPa at most. Therefore, it can be approximated that all elements are almost completely solidated in the matrix at normal temperature. Utilizing the mechanical performance calculation formula for non-quenched and tempered steels, the solid solution strengthening increment for each specimen was determined to be 185.99 MPa, as per Equation (8).
The strengthening increments of the above strengthening mechanisms are summarized and substituted into Equation (4) to calculate the theoretical yield strength of each experimental steel, and the results are shown in
Figure 12 It can be seen that the calculated yield strengths of the experimental steels are very close to their measured yield strengths, with a difference of 15 MPa or less, indicating that the contribution of each strengthening mechanism to the yield strength obtained through the calculation has a certain degree of confidence, and the linear superposition of each strengthening increment is applicable to the current alloy system. Furthermore, upon comparing the contributions of various strengthening mechanisms, we identify solid solution strengthening and grain boundary strengthening as the primary mechanisms in the experimental steels. As the solid solution temperature rises from 1040 °C to 1080 °C, both grain boundary strengthening and dislocation strengthening in the experimental steels decrease, but the precipitates become smaller and more homogeneous, resulting in an increase in the precipitation strengthening, which is the reason why the experimental steels #1 and #2 have the same mechanical properties despite different conditions of solid solution treatment.