Next Article in Journal
Cooling Behavior and Microstructure of Semisolid A201 Aluminum Alloy Prepared by the SEED Process
Previous Article in Journal
Developing Improved Mechanical Property and Corrosion Resistance of Mg-9Li Alloy via Solid-Solution Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Composition Optimum Design and Strengthening and Toughening Mechanisms of New Alumina-Forming Austenitic Heat-Resistant Steels

1
College of Materials Science and Engineering, Taiyuan University of Technology, Taiyuan 030024, China
2
Technology Center, Taiyuan Iron and Steel (Group) Co., Ltd., Taiyuan 030003, China
*
Authors to whom correspondence should be addressed.
Metals 2019, 9(9), 921; https://doi.org/10.3390/met9090921
Submission received: 29 June 2019 / Revised: 9 August 2019 / Accepted: 19 August 2019 / Published: 23 August 2019

Abstract

:
In order to promote the development of ultra-supercritical technology, the optimum composition design of three new alumina-forming austenitic heat-resistant steels, based on Fe–22Cr–25Ni (wt. %), with low cost and excellent performance, and used for 700 °C ultra-supercritical unit was carried out using Thermo-Calc software. A comparison of the mechanical properties presented that with increasing Al content, the plasticity of the system was further improved. Based on the composition system, a systematic investigation regarding the structure stability, thermodynamic properties, and mechanical properties of these new steels was carried out to reveal possible strengthening and toughening mechanisms by employing the first-principles method. Calculation results showed that when Al existed in the Fe–Cr–Ni alloy system as a solid solution, the new structures were stable, especially under high temperature. The solution of Al and Al + Si could increase the value of B/G, namely improving the plasticity of the system, particularly in case of alloying with Al + Si. The inclusion of Si in the Fe–Cr–Ni–Al system was conducive to further improving the plasticity without affecting the strength, which provided references for the subsequent optimum composition design and performance regulation of alumina-forming austenitic heat-resistant steels.

1. Introduction

Environmental protection and energy conservation have been regarded as international issues in recent years. As one of the most important methods used to generate electricity, coal-fired thermal power generation often causes two key issues: environmental pollution and low thermal efficiency. Due to the increasing demands for electric power around the world, the development of clean and efficient coal-fired thermal power generation with lower fuel costs and lower emissions is being pursued in the medium- to long-term future [1]. The efficiency of electricity generation is positively related to the steam conditions (temperatures and pressures) [2]. However, increases in operating temperature and pressure present many challenges for materials with both excellent oxidation and mechanical properties. Conventional austenitic heat-resistant steels, such as HR3C, TP310HCbN, SUPER304H, and NF709, are commonly used in 600–650 °C ultra-supercritical (USC) conditions as superheater tubes and reactor pressure vessels with excellent high-temperature creep property and oxidation stability [3,4,5,6,7]. However, these steels face challenges in water–steam environments when the service temperature is higher than 650 °C, including the sharply deteriorated oxidation resistance caused by the formation of volatile CrO2(OH)2 species and destruction of the Cr2O3-based protective scale [8,9,10], and poor microstructure stability [11,12], thus limiting their use at temperatures higher than 650 °C. Facing the demand for development of higher grade supercritical units, it is urgent to develop heat-resistant materials with better high-temperature oxidation resistance and creep resistance [13].
The new alumina-forming austenitic (AFA) [14,15,16,17,18,19,20,21,22] steels with excellent high-temperature oxidation resistance through the formation of protective Al2O3 and Cr2O3 scales and creep resistance via the dispersive precipitation of nanoscale NbC-type carbide strengthening will be the high-temperature candidate materials for the next generation of USC units [14]. Furthermore, the new AFA steels have shown promising yield and fracture stress and elongation at high temperatures, making them promising for industrial applications [16,23,24,25,26]. The key challenge for AFA steels is to maintain a single austenitic matrix phase and to form continuous protective Al2O3 and Cr2O3 scales [15]. As we know, Al and Cr can promote the formation of oxide scales; however, high contents may weaken the stability of the austenitic phases. Ni as an austenitic stabilizer can change the amount of the NiAl precipitated phases and alter the Al content in the matrix. Thus, the equilibrium concentrations of Al, Cr, and Ni are important for obtaining a single austenitic matrix and forming continuous protective Al2O3 and Cr2O3 scales [17,18,27,28,29].
At present, the research on AFA steels has mainly focused on the alloy system high-temperature ultrafine precipitation-strengthened (HTUPS) steels Fe–(12–15)Cr–20Ni and NF709 steels Fe–20Cr–25Ni, while less focus has been turned to AFA steels used in USC conditions higher than 650 °C. For HR3C steel (Fe–25Cr–20Ni–Nb–N), commonly used in 600–650 °C USC conditions, structural stability is poor and the precipitated phase amounts are increased under higher service temperatures [7,30]. Thus, it is urgent to develop new AFA heat-resistant steels with low cost and excellent performance, to be used for higher grade USC conditions. In this paper, three new high Cr and high Ni AFA heat-resistant steels based on the Fe–25Cr–20Ni (wt. %) were optimally designed using Thermo-Calc software. According to the compositions of the new steels, experimental steels with different amounts of Al were prepared through melting, casting, and hot rolling. After solution-treating, tensile tests were carried out to analyze their mechanical properties. In order to reveal possible strengthening and toughening mechanisms at the atomic scale of these new steels, the effects of alloy elements on their structural stabilities, thermodynamic properties, mechanical properties, and electronic properties were thoroughly investigated using the first-principles method. The results could provide suggestions for the rational design of high-performance stainless steel to be used in advanced ultra-supercritical boilers.

2. Experiment and Calculation Details

According to the compositions of the three new AFA heat-resistant steels designed using Thermo-Calc software (Thermo-Calc Software 2015b, Solna, Sweden), experimental steels with different amounts of Al were prepared in a 50 kg vacuum induction furnace (DDVIF-50-120-2.5, Dongda Advanced Material Preparation Technology Engineering Research Center, Shenyang, China) using commercially pure elements, and named 1.5Al, 2.5Al, and 3.5Al steels. After melting and casting, the ingots were hot rolled to slap with a thickness of 20 mm. The alloys were eventually solution-treated at 1260 °C for 2 h, and subsequently water-quenched to eliminate precipitation of the second phase. The chemical compositions of the three new AFA steels were analyzed using an inductively coupled plasma (ICP) spectrometer (Agilent 7500, Agilent Technology, Palo Alto, CA, USA). The mechanical tensile tests were according to the GB/T228.1-2010 standard. The size of the test samples is shown in Figure 1. Dog-bone-shaped specimens with a gauge cross section of 7 mm × 3.5 mm and a gauge length of 40 mm were cut from the solution-treated AFA steels using a wire-cut electrical discharge machine (DK7720, Huadong Automation Co., Ltd., Taizhou, China), and then the surface was polished to a mirror-like finish. The tensile tests were carried out on an Instron 5582 testing machine (Instron, Boston, MA, USA) at a strain rate of 2.5 × 10−4 s−1; the corresponding loading rate was 0.6 mm/min. Moreover, three identical specimens were measured to ensure that accurate experimental data were obtained.
First-principles calculations were performed using the Cambridge Serial Total Energy Package (CASTEP)CASTEP code based on density functional theory (DFT) [31]. The Perdew–Burke–Ernzerhof (PBE) functional within the generalized gradient approximation (GGA) was applied to describe the exchange-correlation potential [32]. The states of Fe 3d64s2, Cr 3d54s1, Ni 3d84s2, Al 3s23p1, and Si 3s23p2 were treated as valence states. Interactions between the valence electrons and the ionic cores were described by the Ultrasoft pseudopotentials. A kinetic cutoff energy of 500 eV was used for plane wave expansions. Finally, a uniform k-point grid of 4 × 4 × 7 for the Brillouin zone was used for all calculations [33]. In calculations, all atoms were fully relaxed until the convergence parameters satisfied the following conditions: total energy tolerance, 1.0 × 10−5 eV/atom; maximum force tolerance, 0.3 eV/nm; maximal stress component, 0.03 GPa; and maximal displacement, 1 × 10−4 nm. Using the current theoretical scheme, the calculated lattice constant (a) of the face-centered cubic (fcc) Fe solid was 3.44 Å, which was in good agreement with the previously reported values of 3.45 Å [34] and 3.43 Å [35]. The bulk modulus (B) was 348.51 GPa, the shear modulus (G) was 243.34 GPa, and the Young’s modulus (E) was 592.74 GPa, which were close to the values reported in another paper [34], which reported the bulk modulus (B) 317 GPa, the shear modulus (G) 240 GPa, and the Young’s modulus (E) 575 GPa. The good agreement in the crystal lattice constants and elastic modulus thereby validated the application of the methodology to study the structural stability and strengthening and toughening mechanisms.

3. Results and Discussion

3.1. Composition Optimum Design

Based on the Fe–22Cr–25Ni (wt. %) system, the optimum composition design of three new alumina-forming austenitic heat resistant steels by adjusting the proportion of alloy elements such as Nb, Al, Cu, and Si, which was carried out using Thermo-Calc software. The calculated chemical compositions of these new steels were Fe–22Cr–25Ni-0.45Nb–0.80Mn–2.75Cu–0.30Si–0.07C–(1.5, 2.5, 3.5)Al (wt. %). The equilibrium volume fractions of the phases, calculated as a function of the solution temperatures of the three new steels, are shown in Figure 2. Al and Si mainly played the role of solid solution strengthening, while Nb and Cu mainly formed NbX (X = C, N) and copper nanoscale precipitation phases [24,36,37]. At wide high-temperature ranges, only single austenite phases were observed in the three new steels, without any other precipitation phases. This was beneficial for the hot rolling process. At lower temperatures, body-centered cubic (bcc) phases, σ phases, and M23C6 could precipitate in the new steels, which reduced the stability of the austenite phase. Specifically, the content of Al had a significant influence on the precipitation temperature and amount; the larger the content of Al, the higher the precipitated temperature region and larger the precipitated phase amounts. Fortunately, σ phases, the hardest and most brittle, precipitated mainly at 450–700 °C. The precipitation temperatures were all outside the range of the designed operating temperature.
According to the compositions designed above, the corresponding experimental steels were prepared in a vacuum induction furnace. In order to weaken the elements’ segregation and reduce the harmful influence of precipitated phases, the steels were eventually solution-treated at 1260 °C for 2 h. The microstructures of the new AFA steels after solid solution treatment are presented in Figure 3. It can be seen that all the matrices of the steels were composed of a single austenite phase. The analyzed chemical compositions of the new AFA steels, which were confirmed by an ICP spectrometer, are shown in Table 1.

3.2. Mechanical Properties

The stress–strain curves under room temperature are presented in Figure 4. The tensile curves of the three steels were basically similar. For 1.5Al, 2.5Al, and 3.5Al steels, the average ultimate tensile strengths were 559.6 MPa, 561.8 MPa, and 569.8 MPa, respectively. The uniform elongations were 56.6%, 57.3%, and 60.2%, respectively. On the whole, with increased Al content, the strengths and plasticities of the three steels increased slightly. The strengths of our designed steels were close to those of the heat-resistant steels used in ultra-supercritical units at 600 °C [38]. The enhanced phase was mainly composed of niobium carbide, copper nanoscale precipitation phases, and another nanoscale precipitated phases, while other alloying elements, such as Cr, Ni, Al, and Si, mainly stabilized the austenite phases and strengthen by solid solution. There have been many studies on the precipitated phase of niobium carbide and copper nanoscale precipitation phases [19,24,36,37,39]. These precipitated phases were also relatively stable under service conditions. While the roles of alloying elements in the austenite matrix were lacking for these new AFA heat-resistant steels, the mechanism at the atomic level was particularly unclear. Therefore, it is necessary to make a pertinent study of the effects of various elements on the structure and mechanical properties of the austenite matrix, especially the action mechanism of Al.

3.3. Structure Stability and Strengthening and Toughening Mechanisms

The above experimental results indicated that the matrix austenite phases of the three new AFA heat-resistant steels were mainly composed of Cr, Ni, Al, and Si. In this section, the effects of these alloy elements on the structure stabilities, thermodynamic properties, and mechanical properties of the three new steels were thoroughly investigated to reveal possible strengthening and toughening mechanisms at the atomic scale.

3.3.1. Calculation Models

As an austenitic stainless steel, Fe crystallizes in the fcc structure. In order to keep up the fcc crystal lattice periodicity, an expanded lattice cell with 16 metallic atoms (supercell: 2 × 2 × 1), Fe8Cr4Ni4, was chosen for the calculations corresponding to the composition of Fe–22Cr–25Ni (wt. %). In the case of the new AFA heat-resistant steels, Al and Si atoms were introduced into the systems by replacing Fe atoms, corresponding to Fe7Cr4Ni4Al, Fe6Cr4Ni4Al2, Fe5Cr4Ni4Al3, and Fe6Cr4Ni4AlSi; all crystals belonging to the same space group (tetragonal system). The models are shown in Figure 5. The grey spheres represent Fe atoms, blue spheres are Cr atoms, pink spheres stand for Ni atoms, green spheres stand for Al atoms, and yellow spheres correspond to Si atoms. Using the equilibrium structure, the elastic constants and modulus of different systems were calculated through the stress–strain method based on a generalized Hooke’s Law [40]: σij = Cijklεkl, where σij is the tensile stress and εkl is the longitudinal strain. The thermodynamic properties of the considered system under high temperatures were calculated based on the frequency analysis.

3.3.2. Structural Stability

The energetic relative stabilities of Fe8Cr4Ni4, Fe7Cr4Ni4Al, Fe6Cr4Ni4Al2, Fe5Cr4Ni4Al3, and Fe6Cr4Ni4AlSi can be discussed considering our calculations. The formation energies of different systems can be defined by Equation (1):
E f   =   1 Σ N i   [ E total     Σ ( N i E atom ) ]
where Etotal is the calculated total energy of the system, Ni denotes the number of atomic species i (i = Fe, Cr, Ni, Al, or Si) in the unit cell, and Eatom is the energy per atom i at their most stable state (pure constituents). The pure constituents of Fe (fcc), Cr (bcc), Ni (fcc), Al (fcc), and Si (diamond structure) were used to calculate the forming energy. When Ef is negative, it indicates that the system is easy to form, i.e., a spontaneous process, and the greater the absolute value, the easier it is to form. On the contrary, when Ef is positive, it means that the system is not easy to form, and the formation process is a process that absorbs energy. The calculated formation energies are presented in Figure 6a. It is worth noting that the formation energies were all negative, which showed that these systems were easily crystallized. As different contents of Al were introduced to the Fe8Cr4Ni4 system, the values of Fe7Cr4Ni4Al, Fe6Cr4Ni4Al2, and Fe5Cr4Ni4Al3 were lower than that of Fe8Cr4Ni4, which indicated that the substitution element Al was easy to dissolve in the Fe8Cr4Ni4 system. With an increasing number of introduced Al atoms, the formation energy became much smaller, showing that these alloys became much more easily crystallized. As Al and Si were introduced into the system simultaneously, the value of Ef was lower than for of Fe8Cr4Ni4 and Fe7Cr4Ni4Al, which indicated that the coordination of Si and Al was easier to dissolve in the austenite phases.
To further understand the structural stability of the different systems, we calculated their bonding energies, defined as Equation (2):
E b   =   1 Σ N i   [ E total     Σ ( N i E iso i ) ]
where E iso i is the energy of isolated atom i (i = Fe, Cr, Ni, Al, or Si). The total energies of isolated atoms were calculated by putting a Fe, Cr, Ni, Al, or Si atom in the middle of a lattice constant of a 10 Å cubic unit cell and taking values from the CASTEP output files directly. The term “bonding energy” refers to the energy that is released upon the creation of a bound state and which is used to compare the relative stabilities of structural properties. A negative value of bonding energy indicates that the structure is stable. The greater the absolute value, the more stable the structure is, and the stronger the ability to resist external damage will be. To a certain extent, the bonding energy has a positive correlation with the bulk modulus. The greater the absolute value of the bonding energy, the greater the bulk modulus [41]. The calculated bonding energies are shown in Figure 6b. It was clearly seen that all bonding energies were negative, which indicated that these solid solutions were structurally stable. As different contents of Al were introduced to the Fe8Cr4Ni4 system, the bonding energy values increased slightly, which indicated that the bonding between different atoms became weaker, finally decreasing the stability of the structure. With an increasing number of introduced Al atoms, the bonding energy values of the Fe–Cr–Ni–Al systems became larger, showing that these alloys became less stable and the interactions between atoms in the structures became much weaker. As Al and Si were introduced into the system simultaneously, the value of Eb was close to that of Fe7Cr4Ni4Al. However, the values of Eb were all scaled between −10.0 and −8.5 eV/atom, showing that the structures were also very stable.

3.3.3. Thermodynamic Stability

Thermodynamic properties are used to describe the structural stabilities of different systems with elevated temperatures, especially the Gibbs free energy. The calculated Gibbs free energies of different structures as a function of temperature from 0 to 1000 K are shown in Figure 7. It was found that the value of the Gibbs free energy gradually decreased, and the difference between different systems was very small. Comparing the Gibbs free energies of different structures at the same temperature, the Gibbs free energy of Fe8Cr4Ni4 was the smallest, then Fe7Cr4Ni4Al, Fe6Cr4Ni4AlSi, Fe6Cr4Ni4Al2, and finally, Fe5Cr4Ni4Al3. As we know, the smaller the value of the Gibbs free energy, the better the thermal stability of the system [42,43]. A negative Gibbs free energy indicates that the structural stability exists; while a positive Gibbs free energy indicates that structural instability exists. Hence, the calculated results showed that the thermal stability of these structures gradually increased and the stable existing temperature gradually increased with elevated temperature in the order of Fe7Cr4Ni4Al, Fe6Cr4Ni4AlSi, Fe6Cr4Ni4Al2, and Fe5Cr4Ni4Al3. The results showed that the addition of Al and Si weakened the thermal stability of the Fe8Cr4Ni4 system slightly. The results were in good agreement with the bonding energy analyses in Figure 6b.

3.3.4. Mechanical Properties

Except for in the energy favorable condition, mechanical stability is also a necessary condition to examine the feasibility of crystal existence. The criteria of mechanical stability can be evaluated by single-crystal elastic constants according to the Born–Huang criterion, which is dependent on the crystal system. For the different crystal structures, the criteria of the mechanical stability are given as follows [44]: (i) C11 > 0, C44 > 0, C11 > |C12|, and C11 + 2C12 > 0 for cubic crystals; (ii) C11 > 0, C44 > 0, C11 > C12, and (C11 + 2C12)C33–C122 > 0 for hexagonal crystals; (iii) C11 > 0, C22 > 0, C33 > 0, C44 > 0, C55 > 0, C66 > 0, C11 + C22 + C33 + 2(C12 + C13 + C23) > 0, C11 + C22–2C12 > 0, C11 + C33–2C13 > 0, and C22 + C33–2C23 > 0 for orthorhombic crystals; (iv) C11 > 0, C33 > 0, C44 > 0, C66 > 0, C11–C12 > 0, C11 + C33–2C12 > 0, and 2(C11 + C12) +C33 + 4C13 > 0 for tetragonal crystals. In our calculations, all crystals were tetragonal crystals. The calculated elastic constants Cij for the different structures re listed in Table 2. They showed that all calculated systems were stable according to the mechanical stability criteria. In general, all of the mechanically stable phases showed relatively higher values of C11, C22, and C33 than that of C44, indicating relatively higher incompressibility along the a-, b-, and c-axes.
Compared with single-crystal elastic properties, polycrystalline elastic properties (such as bulk modulus, shears modulus, Young’s modulus and Poisson’s ratio) usually have a higher practical application value. In order to better demonstrate the mechanical properties of polycrystalline Fe–Cr–Ni–Al(Si) systems, the bulk modulus (B) and shear modulus (G) were first calculated from the single-crystal elastic constants using the Voight–Reuss–Hill (VRH) average scheme [45].
B   =   1 9   ( C 11   +   C 22   +   C 33 )   +   2 9   ( C 12   +   C 13   +   C 23 )
G   =   1 15   ( C 11   +   C 22   +   C 33     C 12     C 13     C 23 )   +   1 5   ( C 44   +   C 55   +   C 66 )
The Young’s modulus (E) and Poisson’s ratio (ν) of these structures could then be obtained by the following equations.
E   =   9 GB G   +   3 B
ν   =   3 B     2 G 2 ( G   +   3 B )
With the bulk modulus and shear modulus, the macroscopic mechanical properties of different systems, such as hardness, brittleness, and toughness, could be evaluated. The Vickers hardness of different systems could be calculated by the following equation [46]:
H V   =   2 ( G K 2 ) 0.585     3
where K = B/G. According to the empirical Pugh formula, the B/G ratio is a good indicator used to distinguish a material is brittle or ductile [47], and the critical value which separates ductile and brittle materials is approximately 1.75. Ductile materials should meet B/G > 1.75, while values below that indicate brittleness.
Due to the anisotropy of active materials, micro-cracks could be induced in austenitic stainless steels during the practical engineering application process. To study the influence of single-crystal elastic anisotropy on their mechanical properties, we calculated the elastic anisotropy of different systems using the Zener’s anisotropy parameter AZ [46]:
A Z   =   2 C 44 C 11   C 12
All calculated mechanical properties are summarized in Figure 8a, which shows the elastic modulus of different systems. The bulk modulus B and shear modulus (G) of Fe8Cr4Ni4 were found to be in the order of 250.40 GPa and 397.73 GPa; these values agreed to a large extent with those reported in Reference [34,48]. It was found that as different contents of Al atoms were introduced into the Fe8Cr4Ni4 system, the bulk modulus (B), the shear modulus (G), and Young’s modulus (E) all decreased slightly, demonstrating the weaker tendency of the resistant ability to deformation and the increased elasticity and compressibility in the structures. When increasing the number of introduced Al atoms, the bulk modulus (B), shear modulus (G), and Young’s modulus (E) of Fe–Cr–Ni–Al systems seemed to be more sensitive to the Al content; the elastic constants became smaller, showing that the elasticity and compressibility of these alloys became much greater. As Al and Si were introduced into the system simultaneously, the bulk modulus (B), shear modulus (G), and Young’s modulus (E) of the Fe6Cr4Ni4AlSi system became smaller, indicating that the synergistic effect of Si and Al was more pronounced.
The calculated Vickers hardness of different systems is shown in Figure 8b. Fe8Cr4Ni4 had the highest hardness, 20.31 GPa, among these structures [49]. It could be seen that our calculated results were in good agreement with previously experimental results. As Al and Al + Si atoms were introduced into the Fe8Cr4Ni4 systems, the hardness values became lower, showing that the strengths of these materials became weakened, and with increasing Al contents, the hardness values were much lower. The hardness results were in good agreement with the elastic constant analyses in Figure 8a.
Poisson’s ratio is a measure of the expansion of a material when it is compressed in the transverse direction, and is sometimes used as an approximate measure of plastic ductility. Figure 8c shows the Poisson’s ratios of different systems. As Al and Al + Si atoms were introduced into the Fe8Cr4Ni4 systems, the Poisson’s ratio generally increased, indicating an increase of plasticity and metallicity in these new structures. As the Al content increased, the Poisson’s ratios of the Fe–Cr–Ni–Al systems also increased, demonstrating the greater plastic ductility.
The B/G ratio is a good indicator as to whether a material is brittle or ductile. Values above 1.75, the threshold value, indicate ductility, while values below indicate brittleness. It can be seen in Figure 8d that the values of B/G were less than 1.75 (the purple line) for Fe8Cr4Ni4 system, thus, it was a brittle material. As Al and Al + Si atoms were introduced into the Fe8Cr4Ni4 system, the values of B/G increased, and the calculated ratio B/G was 1.94 for Fe6Cr4Ni4AlSi, corresponding to ductile behavior. Similarly, as the Al content increased, the values of B/G (1.57, 1.61, and 1.93) increased and the calculated ratio B/G was 1.93 for Fe5Cr4Ni4Al3, corresponding to ductile behavior. The results indicated that Al and Al + Si atoms could improve the plasticity and ductility of the Fe–Cr–Ni system, and the steel with Al + Si showed the most prominent effect; with increasing Al contents, the effect become more obvious. The results of B/G were in good agreement with the experimental data of elongations of the three AFA steels (56.6%, 57.3%, and 60.2%).
The degree of deviation of Zener’s anisotropy parameter from unity 1 indicates the degree of elastic anisotropy. As shown in Figure 8e, a high degree of elastic anisotropy was found in Fe8Cr4Ni4 with A = 2.34, while, as Al and Al + Si atoms were introduced into the Fe8Cr4Ni4 system, the values of AZ became larger, indicating more anisotropy. With increasing Al contents, the values of AZ of the Fe–Cr–Ni–Al systems also increased, and the Fe5Cr4Ni4Al3 had the largest value (10.51), showing the most anisotropy.

3.3.5. Electronic Properties

To further understand the nature of the structural stabilities and mechanical properties of the five different structures, the electronic properties were also analyzed. Charge density distribution can reflect the bonding characteristics between atoms. Figure 9 displays the charge density distribution maps of Fe8Cr4Ni4, Fe7Cr4Ni4Al, Fe6Cr4Ni4Al2, Fe5Cr4Ni4Al3, and Fe6Cr4Ni4AlSi. For Fe8Cr4Ni4, the valence charge densities were distributed uniformly around the Fe, Cr, and Ni atoms, and the near-spherical charge density represented the metallic bonds. As Al atoms were added into the Fe8Cr4Ni4 system, it was seen that the area with lower electron density (the orange area) around the substitutional Al atoms became larger. With an increase of Al content, the area with lower electron density around the Al atoms became much larger, and the bonding strengths were gradually weakened. As Al and Si were introduced into the system simultaneously, the area with lower electron density around the substitutional Al and Si atoms became larger than Fe7Cr4Ni4Al. Fe6Cr4Ni4AlSi showed the most obvious change. These phenomena suggested that there were significant depletions of charges around substitutional Al and Si atoms in these new systems, indicating that the bonding strengths, structural stabilities, and mechanical strengths became weakened. In summary, the charge density distribution results were accordant with those of the above structural stability and mechanical property analyses.
We investigated the electrochemical stabilities of the different structures mentioned above by analyzing the total density of states (TDOS), as shown in Figure 10a. These structures showed very similar patterns. The finite and non-zero values of TDOS at the Fermi level indicated the metallic character of these structures. As we know, electrochemical stability is closely related to the number of electrons at the Fermi level; fewer electrons implies that higher electrochemical stability and a more stable structure. TDOS analysis suggested that the numbers of electrons at the Fermi level had significant changes with the addition of Al and Al + Si atoms. As different contents of Al were introduced to the Fe8Cr4Ni4 system, the numerical values decreased for Fe7Cr4Ni4Al (20.8 electrons/eV), Fe6Cr4Ni4Al2 (19.4 electrons/eV), and Fe5Cr4Ni4Al3 (18.1 electrons/eV), which indicated that as Al atoms were added into the Fe8Cr4Ni4 system, the electron densities were decreased and the electrochemical stabilities were enhanced. With increasing Al contents, the numbers of electrons at the Fermi level became much smaller, and the electron densities were decreased and the electrochemical stabilities were enhanced. As Al and Si were introduced into the system simultaneously, the numerical value decreased for Fe6Cr4Ni4AlSi (19.4 electrons/eV), lower than that of Fe7Cr4Ni4Al (20.8 electrons/eV). These results indicated that as Al and Al + Si atoms were added into the Fe8Cr4Ni4 system, the electron densities decreased and the electrochemical stabilities were enhanced.
The partial densities of states (PDOS) of the different systems are shown in Figure 10b–f. For Fe8Cr4Ni4, it was clearly seen that the part of TDOS below the Fermi level was mainly contributed to by Fe 3d, Cr 3d, and Ni 3d electrons, while the part of TDOS above the Fermi level was mainly occupied by Fe 3d and Cr 3d electrons. With the introduction of Al and Si atoms into the system, the Al–Fe (or Al–Cr, Al–Ni, Si–Fe, Si–Cr, Si–Ni) interactions were relatively weak because of the lack of overlap between the PDOS of Al and Fe (Cr, Ni) atoms (or Si and Fe atoms). The pseudo-gap is defined as the trough of the Fermi level, where there are two peaks of the density of state (DOS) at each side of the non-zero Fermi level. There are two mechanisms for the formation of the pseudo-gap in alloys. One is ionic origin and the other is due to hybridization effects [50,51,52,53]. The wider the splitting shape of the pseudo-gap, the stronger the hybridization effects [51,52]. As can be seen from Figure 8b, the pseudo-gap for Fe8Cr4Ni4 was exactly at the Fermi level and was the widest among the five structures, so the hybridization effect between Fe 3d, Cr 3d, and Ni 3d electrons was the strongest, which indicated the most stable structure and the largest elastic modulus [51]. As Al and Si atoms were added into the Fe8Cr4Ni4 system, the pseudo-gaps became smaller and the hybridization effect between Fe 3d, Cr 3d, and Ni 3d electrons became weaker. With increasing Al contents, the pseudo-gaps became much smaller and the hybridization effect between Fe 3d, Cr 3d, and Ni 3d electrons became weaker. For Fe5Cr4Ni4Al3 and Fe6Cr4Ni4AlSi, the pseudo-gap had disappeared, showing that the hybridization effects between Fe 3d, Cr 3d, and Ni 3d electrons had reached the minimum and the bond interactions were the weakest, which further resulted in their lowest strength but greatest plasticity and the ductility. The results were in accordance with those of the above analyses.

4. Conclusions

On the basis of a Fe–22Cr–25Ni (wt. %) system, the optimum composition design of three new AFA heat resistant steels was carried out by adjusting the proportion of Nb, Al, Cu, and Si using Thermo-Calc software. Based on the composition system, a systematic investigation regarding the structural stability, thermodynamic properties, and mechanical properties of these new steels was carried out by employing the first-principles method to reveal possible strengthening and toughening mechanisms at the atomic scale. A comparison of the mechanical properties revealed that with increasing Al content, the plasticity of the system was further improved. The calculated formation energies, bonding energies, and Gibbs free energies indicated that Al and Al + Si were easy to dissolve in Fe8Cr4Ni4 systems and that the structures were all stable, especially under high temperatures. A comparison of the mechanical properties showed that the solutions of Al and Al + Si could increase the value of B/G, i.e., improve the plasticity of the system, particularly in the case of alloying with Al + Si. The calculated results were in good agreement with the experimental tensile test results. The electronic structure and DOS analysis showed excellent consistency with the computation results of the energy and mechanical properties analyses. Above all, the inclusion of Si in the Fe–Cr–Ni–Al system was conducive to further improving the plasticity without affecting the strength, which can provide reference for the subsequent optimum composition design and performance regulation of AFA heat-resistant steels.

Author Contributions

Conceptualization, N.D.; methodology, N.D. and X.F.; software, N.D. and R.J.; writing—original draft preparation, N.D.; writing—review and editing, G.F. and P.H.; supervision, J.W. and P.H.; funding acquisition, N.D. and P.H.

Funding

This research was funded by the National Natural Science Foundation of China (51871159), the Natural Science Foundation of Shanxi province (201801D221125), and the Graduate Education Innovation Project of Shanxi Province (2016JD20).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hussain, T.; Syed, A.U.; Simms, N.J. Fireside Corrosion of Superheater Materials in Coal/Biomass Co-fired Advanced Power Plants. Oxid. Met. 2013, 80, 529–540. [Google Scholar] [CrossRef]
  2. Maziasz, P.J. Development of creep-resistant and oxidation-resistant austenitic stainless steels for high temperature applications. JOM 2018, 70, 66–75. [Google Scholar] [CrossRef]
  3. Takahashi, T.; Sakakibara, M.; Kikuchi, M.; Ogawa, T.; Sakurai, H.; Nagao, K.; Yasuda, H. Development of High-Strength 20Cr-25Ni (NF709) Steel for USC Boiler Tubes; Nippon Steel Technical Report No. 38; Nippon Steel Corp: Tokyo, Japan, 1998. [Google Scholar]
  4. Viswanathan, R.; Sarver, J.; Tanzosh, J. Boiler Materials for Ultra-Supercritical Coal Power Plants—Steamside Oxidation. J. Mater. Eng. Perform. 2006, 15, 255–274. [Google Scholar] [CrossRef]
  5. Zhao, Y.; Zhao, J.; Li, X.N. Microstructural evolution and change in hardness during creep of NF709 austenitic stainless steel. Acta Metall. Sin. 2011, 24, 220–224. [Google Scholar]
  6. Dudziak, T.; Łukaszewicz, M.; Simms, N.; Nicholls, J.R. Steam oxidation of TP347HFG, super 304H and HR3C-analysis of significance of steam flowrate and specimen surface finish. Corros. Eng. Sci. Technol. 2015, 50, 272–282. [Google Scholar] [CrossRef]
  7. Vujic, S.; Sandström, R.; Sommitsch, C. Precipitation evolution and creep strength modelling of 25Cr20NiNbN austenitic steel. Mater. High Temp. 2015, 32, 607–618. [Google Scholar] [CrossRef]
  8. Wright, I.G.; Dooley, R.B. A review of the oxidation behavior of structural alloys in steam. Int. Mater. Rev. 2010, 55, 129–167. [Google Scholar] [CrossRef]
  9. Saunders, S.; Monteiro, M.; Rizzo, F. The oxidation behaviour of metals and alloys at high temperatures in atmospheres containing water vapour: A review. Prog. Mater. Sci. 2008, 53, 775–837. [Google Scholar] [CrossRef]
  10. Jung, K.; Kim, C.S.; Pettit, F.S.; Meier, G.H. Interfacial failure via encapsulation of external particulates in an outward-growing thermal oxide. J. Power Sources 2011, 196, 4686–4694. [Google Scholar] [CrossRef]
  11. Chai, G.; Boström, M.; Olaison, M.; Forsberg, U. Creep and LCF Behaviors of Newly Developed Advanced Heat Resistant Austenitic Stainless Steel for A-USC. Procedia Eng. 2013, 55, 232–239. [Google Scholar] [CrossRef] [Green Version]
  12. Iseda, A.; Okada, H.; Semba, H.; Igarashi, M. Long term creep properties and microstructure of SUPER304H, TP347HFG and HR3C for A-USC boilers. Energy Mater. 2007, 2, 199–206. [Google Scholar] [CrossRef]
  13. Liu, P.; Chu, Z.K.; Yuan, Y.; Wang, D.H.; Cui, C.Y.; Hou, G.C.; Zhou, Y.Z.; Sun, X.F. Microstructures and mechanical properties of a newly developed austenitic heat resistant steel. Acta Metall. Sin. 2019, 32, 517–525. [Google Scholar] [CrossRef]
  14. Yamamoto, Y.; Brady, M.P.; Lu, Z.P.; Maziasz, P.J.; Liu, C.T.; Pint, B.A.; More, K.L.; Meyer, H.M.; Payzant, E.A. Creep-resistant, Al2O3-forming austenitic stainless steels. Science 2007, 316, 433–436. [Google Scholar] [CrossRef] [PubMed]
  15. Brady, M.; Yamamoto, Y.; Santella, M.; Pint, B.; Brady, M.; Pint, B. Effects of minor alloy additions and oxidation temperature on protective alumina scale formation in creep-resistant austenitic stainless steels. Scr. Mater. 2007, 57, 1117–1120. [Google Scholar] [CrossRef]
  16. Xu, X.Q.; Zhang, X.F.; Chen, G.L.; Lu, Z.P. Improvement of high-temperature oxidation resistance and strength in alumina-forming austenitic stainless steels. Mater. Lett. 2011, 65, 3285–3288. [Google Scholar] [CrossRef]
  17. Brady, M.; Magee, J.; Yamamoto, Y.; Helmick, D.; Wang, L. Co-optimization of wrought alumina-forming austenitic stainless steel composition ranges for high-temperature creep and oxidation/corrosion resistance. Mater. Sci. Eng. A 2014, 590, 101–115. [Google Scholar] [CrossRef]
  18. Yanar, N.M.; Lutz, B.S.; Garcia-Fresnillo, L.; Brady, M.P.; Meier, G.H.; Brady, M. The Effects of Water Vapor on the Oxidation Behavior of Alumina Forming Austenitic Stainless Steels. Oxid. Met. 2015, 84, 541–565. [Google Scholar] [CrossRef]
  19. Zhou, D.; Zhao, W.; Mao, H.; Hu, Y.; Xu, X.; Sun, X.; Lu, Z.; Mao, H. Precipitate characteristics and their effects on the high-temperature creep resistance of alumina-forming austenitic stainless steels. Mater. Sci. Eng. A 2015, 622, 91–100. [Google Scholar] [CrossRef]
  20. Gao, Q.; Qu, F.; Zhang, H.; Huo, Q. Austenite grain growth in alumina-forming austenitic steel. J. Mater. Res. 2016, 31, 1732–1740. [Google Scholar] [CrossRef]
  21. Yu, Z.D.; Chen, M.H.; Shen, C.B.; Zhu, S.L.; Wang, F.H. Oxidation of an austenitic stainless steel with or without alloyed aluminum in O2+10% H2O environment at 800 °C. Corro. Sci. 2017, 121, 105–115. [Google Scholar] [CrossRef]
  22. Baker, I.; Afonina, N.; Wang, Z.; Wu, M. Preliminary creep testing of the alumina-forming austenitic stainless steel Fe-20Cr-30Ni-2Nb-5Al. Mater. Sci. Eng. A 2018, 718, 492–498. [Google Scholar] [CrossRef]
  23. Brady, M.P.; Yamamoto, Y.; Santella, M.L.; Maziasz, P.J.; Pint, B.A.; Liu, C.T.; Lu, Z.P.; Bei, H.; Brady, M.; Maziasz, P. The development of alumina-forming austenitic stainless steels for high-temperature structural use. JOM 2008, 60, 12–18. [Google Scholar] [CrossRef]
  24. Bei, H.; Yamamoto, Y.; Brady, M.; Santella, M.; Brady, M. Aging effects on the mechanical properties of alumina-forming austenitic stainless steels. Mater. Sci. Eng. A 2010, 527, 2079–2086. [Google Scholar] [CrossRef]
  25. Wang, M.; Sun, H.Y.; Phaniraj, M.P.; Han, H.N.; Jang, J.S.; Zhou, Z.J. Evolution of microstructure and tensile properties of Fe-18Ni-12Cr based AFA steel during aging at 700 °C. Mater. Sci. Eng. A 2016, 672, 23–31. [Google Scholar] [CrossRef]
  26. Hu, B.; Baker, I. The effect of thermo-mechanical treatment on the high temperature tensile behavior of an alumina-forming austenitic steel. Mater. Sci. Eng. A 2016, 651, 795–804. [Google Scholar] [CrossRef] [Green Version]
  27. Peraldi, R.; Pint, B.A. Effect of Cr and Ni Contents on the Oxidation Behavior of Ferritic and Austenitic Model Alloys in Air with Water Vapor. Oxid. Met. 2004, 61, 463–483. [Google Scholar] [CrossRef]
  28. Sun, Y.F.; Lv, Y.Z.; Zhang, Y.; Zhao, J.Y.; Wu, Y. Microstructural and properties evolution of austenitic heat resistant steel after addition of aluminium. Mater. Sci. Technol. 2013, 29, 511–516. [Google Scholar] [CrossRef]
  29. Guo, X.; Chen, K.; Gao, W.; Shen, Z.; Zhang, L. Corrosion behavior of alumina-forming and oxide dispersion strengthened austenitic 316 stainless steel in supercritical water. Corros. Sci. 2018, 138, 297–306. [Google Scholar] [CrossRef]
  30. Wang, B.; Liu, Z.D.; Cheng, S.C.; Liu, C.M.; Wang, J.Z.; Liu, Z.C. Microstructure Evolution and Mechanical Properties of HR3C Steel during Long-term Aging at High Temperature. J. Iron Steel Res. Int. 2014, 21, 765–773. [Google Scholar] [CrossRef]
  31. Segall, M.D.; Lindan, P.J.D.; Probert, M.J.; Pickard, C.J.; Hasnip, P.; Clark, S.J.; Payne, M.C. First-principles simulation: Ideas, illustrations and the CASTEP code. J. Phys. Condens. Matter 2002, 14, 2717–2744. [Google Scholar] [CrossRef]
  32. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  34. Guo, G.Y.; Wang, H.H. Gradient-corrected density functional calculation of elastic constants of Fe, Co and Ni in bcc, fcc and hcp structures. Chin. J. Phys. 2000, 38, 949–961. [Google Scholar]
  35. Haglund, J.; Guillermet, A.F.; Grimvall, G.; Körling, M. Theory of bonding in transition-metal carbides and nitrides. Phys. Rev. B 1993, 48, 11685–11691. [Google Scholar] [CrossRef] [PubMed]
  36. Zhao, B.; Chang, K.; Fan, J.; Chen, Z.; Dong, X.; Zhang, L. Annealing effects on precipitation and high-temperature properties of a Cu-containing alumina-forming austenitic steel. Mater. Lett. 2016, 176, 83–86. [Google Scholar] [CrossRef]
  37. Zhao, B.; Fan, J.; Chen, Z.; Dong, X.; Sun, F.; Zhang, L. Evolution of precipitates in a Cu-containing alumina-forming austenitic steel after short-term mechanical tests. Mater. Charact. 2017, 125, 37–45. [Google Scholar] [CrossRef]
  38. Chi, C.Y.; Yu, H.Y.; Xie, X.S. Research and development of austenitic heat-resistant steels for 600 °C superheat/reheater tubes of USC power plant boilers. World Iron Steel 2012, 4, 50–65. [Google Scholar]
  39. Moon, J.; Lee, T.H.; Heo, Y.U.; Han, Y.S.; Kang, J.Y.; Ha, H.Y.; Suh, D.W. Precipitation sequence and its effect on age hardening of alumina-forming austenitic stainless steel. Mater. Sci. Eng. A 2015, 645, 72–81. [Google Scholar] [CrossRef]
  40. Fast, L.; Wills, J.M.; Johansson, B.; Eriksson, O. Elastic constants of hexagonal transition metals: Theory. Phys. Rev. B 1995, 51, 17431–17438. [Google Scholar] [CrossRef]
  41. Meradji, H.; Drablia, S.; Ghemid, S.; Belkhir, H.; Bouhafs, B.; Tadjer, A. First-principles elastic constants and electronic structure of BP, BAs, and BSb. Phys. Status Solidi (b) 2004, 241, 2881–2885. [Google Scholar] [CrossRef]
  42. Zhou, D.; Liu, J.; Xu, S.; Peng, P. Thermal stability and elastic properties of Mg3Sb2 and Mg3Bi2 phases from first-principles calculations. Phys. B Condens. Matter 2010, 405, 2863–2868. [Google Scholar] [CrossRef]
  43. Huang, Z.W.; Zhao, Y.H.; Hou, H.; Han, P.D. Electronic structural, elastic properties and thermodynamics of Mg17Al12, Mg2Si and Al2Y phases from first-principles calculations. Phys. B 2002, 407, 1075–1081. [Google Scholar] [CrossRef]
  44. Du, J.; Wen, B.; Melnik, R.; Kawazoe, Y. Phase stability, elastic and electronic properties of Cu–Zr binary system intermetallic compounds: A first-principles study. J. Alloy. Compd. 2014, 588, 96–102. [Google Scholar] [CrossRef]
  45. Wu, Z.J.; Zhao, E.J.; Xiang, H.P.; Hao, X.F.; Liu, X.J.; Meng, J. Crystal structures and elastic properties of superhardIrN2andIrN3from first principles. Phys. Rev. B 2007, 76, 1–15. [Google Scholar] [CrossRef]
  46. Zhu, Y.; Yan, M.; Zhang, Y.; Zhang, C. First-principles investigation of structural, mechanical and electronic properties for Cu–Ti intermetallics. Comput. Mater. Sci. 2016, 123, 70–78. [Google Scholar] [CrossRef]
  47. Pugh, S. XCII. Relations between the elastic moduli and the plastic properties of polycrystalline pure metals. Lond. Edinb. Dublin Philos. Mag. J. Sci. 1954, 45, 823–843. [Google Scholar] [CrossRef]
  48. Benyelloul, K.; Aourag, H. Elastic constants of austenitic stainless steel: Investigation by the first-principles calculations and the artificial neural network approach. Comput. Mater. Sci. 2013, 67, 353–358. [Google Scholar] [CrossRef]
  49. Fang, Y.; Zhao, J.; Li, X. Precipitates in hr3c steel aged at high temperature. Acta Met. Sin. 2010, 46, 844–849. [Google Scholar] [CrossRef]
  50. Li, K.; Xie, H.; Liu, J.; Ma, Z.; Zhou, Y.; Xue, D. From chemistry to mechanics: Bulk modulus evolution of Li–Si and Li–Sn alloys via the metallic electronegativity scale. Phys. Chem. Chem. Phys. 2013, 15, 17658–17663. [Google Scholar] [CrossRef] [PubMed]
  51. Zhang, P.; Ma, Z.; Wang, Y.; Zou, Y.; Lei, W.; Pan, Y.; Lu, C. A first principles study of the mechanical properties of Li–Sn alloys. RSC Adv. 2015, 5, 36022–36029. [Google Scholar] [CrossRef]
  52. Yu, R.; Chong, X.; Jiang, Y.; Zhou, R.; Yuan, W.; Feng, J. The stability, electronic structure, elastic and metallic properties of manganese nitrides. RSC Adv. 2015, 5, 1620–1627. [Google Scholar] [CrossRef]
  53. Zhang, B.; Wu, L.; Wan, B.; Zhang, J.; Li, Z.; Gou, H. Structural evolution, mechanical properties, and electronic structure of Al–Mg–Si compounds from first principles. J. Mater. Sci. 2015, 50, 6498–6509. [Google Scholar] [CrossRef]
Figure 1. The size of test sample (units: mm).
Figure 1. The size of test sample (units: mm).
Metals 09 00921 g001
Figure 2. Equilibrium phase diagrams of new alumina-forming austenitic heat-resistant steels: (a) 1.5Al; (b) 2.5Al; (c) 3.5Al.
Figure 2. Equilibrium phase diagrams of new alumina-forming austenitic heat-resistant steels: (a) 1.5Al; (b) 2.5Al; (c) 3.5Al.
Metals 09 00921 g002
Figure 3. Equilibrium phase diagrams of new alumina-forming austenitic heat-resistant steels: (a) 1.5Al; (b) 2.5Al; (c) 3.5Al.
Figure 3. Equilibrium phase diagrams of new alumina-forming austenitic heat-resistant steels: (a) 1.5Al; (b) 2.5Al; (c) 3.5Al.
Metals 09 00921 g003
Figure 4. Stress–strain curves of new alumina-forming austenitic heat-resistant steels at room temperature.
Figure 4. Stress–strain curves of new alumina-forming austenitic heat-resistant steels at room temperature.
Metals 09 00921 g004
Figure 5. Schematic diagram of models for calculations: (a) Fe8Cr4Ni4; (b) Fe7Cr4Ni4Al; (c) Fe6Cr4Ni4Al2; (d) Fe5Cr4Ni4Al3; (e) Fe6Cr4Ni4AlSi.
Figure 5. Schematic diagram of models for calculations: (a) Fe8Cr4Ni4; (b) Fe7Cr4Ni4Al; (c) Fe6Cr4Ni4Al2; (d) Fe5Cr4Ni4Al3; (e) Fe6Cr4Ni4AlSi.
Metals 09 00921 g005
Figure 6. (a) Formation energy Ef and (b) binding energy Eb of Fe8Cr4Ni4, Fe7Cr4Ni4Al, Fe6Cr4Ni4Al2, Fe5Cr4Ni4Al3, and Fe6Cr4Ni4AlSi.
Figure 6. (a) Formation energy Ef and (b) binding energy Eb of Fe8Cr4Ni4, Fe7Cr4Ni4Al, Fe6Cr4Ni4Al2, Fe5Cr4Ni4Al3, and Fe6Cr4Ni4AlSi.
Metals 09 00921 g006
Figure 7. Relationships of Gibbs free energies with temperature of Fe8Cr4Ni4, Fe7Cr4Ni4Al, Fe6Cr4Ni4Al2, Fe5Cr4Ni4Al3, and Fe6Cr4Ni4AlSi.
Figure 7. Relationships of Gibbs free energies with temperature of Fe8Cr4Ni4, Fe7Cr4Ni4Al, Fe6Cr4Ni4Al2, Fe5Cr4Ni4Al3, and Fe6Cr4Ni4AlSi.
Metals 09 00921 g007
Figure 8. Calculated mechanical properties of different systems: (a) The bulk modulus (B), shear modulus (G), and Young’s modulus (E); (b) the Vicker’s hardness; (c) the Poisson’s ratio values; (d) the B/G ratio values; (e) the Zener’s anisotropy parameter AZ.
Figure 8. Calculated mechanical properties of different systems: (a) The bulk modulus (B), shear modulus (G), and Young’s modulus (E); (b) the Vicker’s hardness; (c) the Poisson’s ratio values; (d) the B/G ratio values; (e) the Zener’s anisotropy parameter AZ.
Metals 09 00921 g008
Figure 9. Charge density distribution maps of different systems in the (002) plane: (a) Fe8Cr4Ni4; (b) Fe7Cr4Ni4Al; (c) Fe6Cr4Ni4Al2; (d) Fe5Cr4Ni4Al3; (e) Fe6Cr4Ni4AlSi.
Figure 9. Charge density distribution maps of different systems in the (002) plane: (a) Fe8Cr4Ni4; (b) Fe7Cr4Ni4Al; (c) Fe6Cr4Ni4Al2; (d) Fe5Cr4Ni4Al3; (e) Fe6Cr4Ni4AlSi.
Metals 09 00921 g009
Figure 10. (a) Total density of states (TDOS) of different systems; partial density of states (PDOS) of different systems: (b) Fe8Cr4Ni4; (c) Fe7Cr4Ni4Al; (d) Fe6Cr4Ni4Al2; (e) Fe5Cr4Ni4Al3; and (f) Fe6Cr4Ni4AlSi.
Figure 10. (a) Total density of states (TDOS) of different systems; partial density of states (PDOS) of different systems: (b) Fe8Cr4Ni4; (c) Fe7Cr4Ni4Al; (d) Fe6Cr4Ni4Al2; (e) Fe5Cr4Ni4Al3; and (f) Fe6Cr4Ni4AlSi.
Metals 09 00921 g010
Table 1. Chemical compositions of the new alumina-forming austenitic heat-resistant steels (wt. %).
Table 1. Chemical compositions of the new alumina-forming austenitic heat-resistant steels (wt. %).
AlloyCrNiAlNbMnCuSiCFe
1.5Al21.1925.121.500.440.802.600.300.07Bal
2.5Al21.3624.592.480.420.772.650.280.07Bal
3.5Al22.1425.293.510.400.782.630.290.07Bal
Table 2. Calculated single-crystal elastic constants for different systems.
Table 2. Calculated single-crystal elastic constants for different systems.
Elastic ConstantsC11C12C13C33C44C66
Fe8Cr4Ni4371.484204.667170.828417.936195.492209.082
Fe7Cr4Ni4Al323.440186.422157.412354.115184.691174.932
Fe6Cr4Ni4Al2271.592169.982149.656298.950170.620147.576
Fe5Cr4Ni4Al3225.141199.483126.016297.716134.792103.835
Fe6Cr4Ni4AlSi298.478211.721178.404364.793168.795133.751

Share and Cite

MDPI and ACS Style

Dong, N.; Jia, R.; Wang, J.; Fan, G.; Fang, X.; Han, P. Composition Optimum Design and Strengthening and Toughening Mechanisms of New Alumina-Forming Austenitic Heat-Resistant Steels. Metals 2019, 9, 921. https://doi.org/10.3390/met9090921

AMA Style

Dong N, Jia R, Wang J, Fan G, Fang X, Han P. Composition Optimum Design and Strengthening and Toughening Mechanisms of New Alumina-Forming Austenitic Heat-Resistant Steels. Metals. 2019; 9(9):921. https://doi.org/10.3390/met9090921

Chicago/Turabian Style

Dong, Nan, Ruirui Jia, Jian Wang, Guangwei Fan, Xudong Fang, and Peide Han. 2019. "Composition Optimum Design and Strengthening and Toughening Mechanisms of New Alumina-Forming Austenitic Heat-Resistant Steels" Metals 9, no. 9: 921. https://doi.org/10.3390/met9090921

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