Next Article in Journal
Spin-Current Oscillations in Diluted Magnetic Semiconductor Multibarrier GaMnAs/GaAs: Role of Temperature and Bias Voltage
Next Article in Special Issue
Microstructure and Micro-Hardness of Dissimilar Metal Cladding from a Pipe–Nozzle Mockup for PWR
Previous Article in Journal
Comparison of Dust Deposition Reduction Performance by Super-Hydrophobic and Super-Hydrophilic Coatings for Solar PV Cells
Previous Article in Special Issue
Obtaining Uniform High-Strength and Ductility of 2A12 Aluminum Alloy Cabin Components via Predeformation and Annular Channel Angular Extrusion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Annealing and Oxidation on the Microstructure Evolution of Hot-Dipped Aluminide Q345 Steel with Silicon Addition

1
School of Materials Science and Engineering, Changzhou University, Changzhou 213164, China
2
Jiangsu Key Laboratory of Material Surface Science and Technology, Changzhou University, Changzhou 213164, China
*
Authors to whom correspondence should be addressed.
Coatings 2022, 12(4), 503; https://doi.org/10.3390/coatings12040503
Submission received: 14 February 2022 / Revised: 20 March 2022 / Accepted: 30 March 2022 / Published: 8 April 2022
(This article belongs to the Special Issue Corrosion and Degradation of Materials)

Abstract

:
Q345 steel was coated by hot dipping into molten pure aluminum and Al-Si baths. The coatings were annealed at 800 and 900 °C for 1–3 h and subsequently oxidized at 900 °C for 15 h in air. The results revealed that the thickness of the intermetallic layer increased with increasing hot-dipping time in the range of 700–750 °C, while it decreased when the hot-dipping aluminizing temperature was 800 °C. As the silicon content in the aluminum bath increased, the thickness of the intermetallic layer decreased, and the intermetallic layer/steel-substrate interface transformed from an irregular morphology into a flat morphology. The hot-dipped Al-2.5Si samples were subjected to annealing; the higher the annealing temperature and longer the annealing time, the faster the transformation of the intermediate phase in the coating. The Fe2Al5 phase was fully transformed into the ductile FeAl phase after the hot-dipped samples annealed at 900 °C for 3 h. When the outermost layer of Q345 steel was the FeAl phase, oxidation resistance of the oxide was the best.

1. Introduction

In order to reduce room-temperature gas emissions and oil consumption, the use of bioethanol in the automotive industry is gradually increasing; bioethanol and the concentration of ethanol used as a bioadditive in gasoline has increased to about 20% in EU countries [1,2]. It is well-known that after burning bioethanol in engines, the flue gas consists of large amounts of H2O and CO2, which can significantly increase the reaction rate of steel in motorcycle- and automotive-exhaust-system components, resulting in the components of the exhaust system being subjected to massive oxidation and severe corrosion of steel and its alloys [3,4,5,6]. Low-alloy high-strength steels (Q345) are increasingly used in the automotive industry because of their high strength, good welding properties, and better corrosion resistance, which can save metal materials and reduce the weight of equipment. However, after long-term use of low-alloy high-strength steel in a high-temperature environment with water vapor, the steel surface can very easily form iron oxide, which seriously reduces the surface quality of the steel and limits its use in high-temperature conditions.
Hot-dipping aluminizing is an effective and inexpensive method to form a protective coating on the steel surface [7,8]. By inhibiting the rapid diffusion of aluminum atoms in Fe2Al5 and reducing the thickness of the alloy layer, the addition of a certain amount of silicon to the aluminum bath is beneficial to improve the quality of the coating [9,10,11,12,13]. The mechanism of the reduction of the intermetallic layer by the addition of silicon to the hot-dip-coated aluminum solution has been studied widely [9,10,13,14]. Heumann and Dittrich [9] stated that the reduction in the Fe2Al5 layer is due to the fact that silicon occupies vacant sites in the c-axis of the Fe2Al5 crystal structure, thus slowing down the growth of the Fe2Al5 layer. In contrast, Lainer and Kurakin [13] proposed that the growth of the intermetallic compound layer can be slowed down due to the nucleation rate of the Fe-Al-Si intermetallic compound layer. Studies have shown that the high-temperature oxidation resistance of aluminized steel can be improved by using a suitable diffusion-annealing treatment to produce a phase-microstructure transformation in the coating [15,16,17,18,19,20,21]. The mechanism to improve the oxidation resistance of aluminized steel is the formation of a dense and stable Al2O3 protective layer on the surface of the aluminized layer during the oxidation process, accompanied by a phase transition from the Al-rich phase to the Fe-rich intermetallic compound in the aluminized layer. The improvement of high-temperature oxidation of low-carbon steel exposed to the ethanol combustion product by hot-dip-aluminized coating has been investigated widely [21,22,23]; however, there are few studies on hot-dip-aluminizing silicon of Q345 steel and its high-temperature oxidation properties. Therefore, it is important to study how to improve the high-temperature resistance of low-alloy high-strength steels in order to reduce the effect of alternative fuels on the high-temperature corrosion of low-alloy high-strength steels.
Previous studies indicated that the mechanical properties of the Fe-Al intermetallic compounds were different [24,25,26], in which the Fe3Al and FeAl2 phases show plastic characteristics, and FeAl, Fe2Al5, FeAl3, and Fe4Al13 show brittle characteristics. The coatings evolved from the plastic phase had better adhesion. Therefore, the purpose of this paper is to study the effects of hot-dipping aluminizing and diffusion annealing on the evolution of the Fe-Al phase, and to explore ways of improving the high-temperature oxidation resistance of Q345 steel.

2. Experimental Materials and Methods

2.1. Experimental Materials

The substrate used in this experiment was Q345 steel plate; rectangular specimens were cut to the dimensions of 10 mm × 10 mm × 3 mm by a water-cooled cutting machine. The composition of Q345 high-strength steel is shown in Table 1. The hot-dipping Al-Si solution was prepared by pure aluminum and Al-15Si master alloy.

2.2. Hot-Dipping Aluminizing

The samples were soaked in 15 wt.% sodium hydroxide aqueous solution at 70 °C for 3 min, and then washed with water and soaked in 15 wt.% hydrochloric acid solution for pickling for 3 min. After washing with water, the samples were treated in 200~300 g/L KCl-K2ZrF6-NaF solution at 90 °C for 3 min with a mass ration of KCl:K2ZrF6:NaF of 100:100:1. Then, the samples were removed and dried at 100 °C for 8 min. The hot-dipping aluminizing was carried out in a bath that was placed in a KGPS-30-8 medium-frequency induction furnace (Shenguang electric furnace, Ningbo, China) at 700, 750, and 800 °C, respectively. The working thermocouple was periodically tested against a calibrated standard thermocouple. The overall temperature accuracy was estimated to be ±5 °C. After the samples were immersed in the hot-dipping bath for 30 s and 3 min, respectively, they were quickly removed from the bath and cooled in room-temperature water.

2.3. Hot-Dipping Al-Si

The Al-Si molten bath was prepared from industrial pure Al and Al-15Si master alloy. The experiments were divided into six groups, with nominal composition of each group was Al-0.5Si, Al-1Si, Al-2.5Si, Al-5Si, Al-7.5Si, and Al-10Si, weighing 200 g in each. The hot-dip temperature was 750 °C, and the hot-dip time was 30 s.

2.4. Annealing and Oxidation

After removing the surface oxides, the hot-dipped samples were sealed in evacuated quartz tubes and placed in SK2-4-12 tubular resistance furnace (Shanghai Y-feng Electrical Furnace Co. Ltd., Shenyang, China) for diffusion annealing. The annealing temperatures were 800 °C and 900 °C, and the annealing times were 1 h, 2 h, and 3 h, respectively. The oxidation experiments were carried out in three groups: the first group of Q345 steel was hot-dipped in Al-2.5Si solution at 750 °C for 30 s and then directly oxidized in air at 900 °C for 15 h; the second group of Q345 steel was hot-dipped in Al-2.5Si solution at 750 °C for 30 s and then annealed in vacuum at 800 °C for 1 h; followed by oxidation in air at 900 °C for 15 h, and the third group of Q345 steel was hot-dipped in Al-2.5Si solution at 750 °C for 30 s and then annealed in vacuum at 900 °C for 3 h, followed by oxidation in air at 900 °C for 15 h. The three groups of samples were noted as D1, D2, and D3, respectively.

2.5. Characterization

Surface morphology, cross-sectional microstructure, and chemical composition of the coating were analyzed by a JSM-6510 Scanning Electron Microscope (SEM) (JEOL, Tokyo, Japan) equipped with Energy Dispersive X-ray Spectrometer (EDS) (Oxford Instrument, Cambridge, UK). In the observation of SEM, a focused electron beam with energy of 20 keV and working distance of 15 mm interacted with the sample in a teardrop-shape (around 1 µm deep into the sample).

3. Results and Discussion

3.1. Cross-Sectional Microstructure Analysis of Hot-Dipping Pure Aluminum

Figure 1 shows the cross-sectional microstructure of Q345 steel hot-dipped in aluminum solution at 700, 750, and 800 °C for 30 s and 3 min, respectively. It can be seen that the cross-sectional structure of the coating was divided into three different areas: external aluminum layer, intermetallic compound layer, and matrix steel. Composition of the intermetallic layer determined by EDS are listed in Table 2. The representative positions of the components of each phase determined in Figure 1 are marked with symbols a, b, c, respectively. In fact, the composition of each phase was the average of the components at five similar positions. According to the EDS results and combined with the Fe-Al phase diagram, the intermetallic compound layer consisted of two sublayers: the inner part at the metal-coating interface was the Fe2Al5 phase, and the composition determined by EDS point analyses was 70.9–71.9 at.% Al, 27.9–28.8 at.% Fe, and ≤0.6 at.% Mn. The outer part was the FeAl3 phase, and the composition found by EDS point analyses was 75.6–75.9 at.% Al, 23.8–24.1 at.% Fe, and 0.3 at.% Mn. The Fe2Al5 phase was formed by reactive diffusion of iron atoms and aluminum atoms at the interface during hot dipping [4,6]. There was an irregular interface between the alloy layer and the steel matrix, which was in a tongue shape [9,10]. However, the FeAl3-phase layer was not so obvious until the hot-dipping temperature increased to 800 °C. In addition, it could be found that there was also a small amount of discrete-phase precipitation in the external aluminum layer. According to EDS analysis, this discrete phase was the FeAl3 phase, which was precipitated through eutectic reaction in the solidification structure as the cooling speed was not quick enough while taking out the sample from the hot-dipping bath. Because the precipitated FeAl3 phase was relatively small, the average Al content measured by EDS was relative higher than that in the FeAl3 sublayer.
The intermetallic layer thickness of Q345 steel hot-dipped in pure aluminum solution at 700, 750, and 800 °C for 30 s and 3 min is shown in Figure 2. Comparing Figure 1a,b, it can be found that when the hot-dip temperature was 700 °C, the external pure aluminum layer was relatively thick, while the Fe2Al5 phase area was relatively thin. The reason for this obvious difference was that at this temperature, the thermal diffusion was relatively slow and the fluidity of the aluminum solution was poor, resulting in a considerable part of thick pure aluminum layer still adhering to the sample when the sample was taken out after hot dipping. At the same time, it can be found that when the hot-dip temperature was 700 °C, the type and thickness of the intermetallic layer did not change significantly with the extension of hot-dip time. When the dipping time was extended from 30 s to 3 min, the thickness of the intermetallic layer increased from 17 to 20 μm.
By observing Figure 1c,d, it can be found that when the hot-dip temperature was increased to 750 °C, the fluidity of aluminum solution was improved, so the external pure aluminum layer of the sample was significantly reduced compared with Figure 1a,b. At the same time, due to the increase of temperature, the diffusion of aluminum and iron was accelerated, and the intermetallic layer was obviously thickened. At 750 °C, the thickness of the intermediate alloy layer increased greatly with the extension of hot-dip time. When the dipping time was 30 s, the thickness was about 31 μm. When the dipping time was 3 min, the thickness was about 56 μm. The thickness increased by about 25 μm.
When the hot-dip temperature was raised to 800 °C, as shown in Figure 1e,f, the thickness and type of intermetallic layer changed significantly. When the dipping time was 30 s, the thickness of the intermetallic layer was about 70 μm. Compared with 750 °C, the thickness of the intermetallic layer had increased further. However, with the extension of the hot-dipping time to 3 min, the thickness of the intermetallic layer decreased. According to the results reported by Chen et al. [27], the average thickness of Fe2Al5 also decreased with increasing reaction time when studying the interfacial reaction between liquid aluminum and ST12 cold-roll steel sheets, though the growth of Fe2Al5 almost followed the parabolic law at 700, 800 and 900 °C. They had attributed the decrease in thickness into the austenitizing of the substrate (austenitizing temperature of ST12 was 935 °C). The crystal structure of steel transformed from the body-centered cubic phase (α-Fe) into the face-centered cubic phase (γ-Fe) when the temperature of steel exceeded the austenitizing temperature. The diffusion coefficient of Al atoms in the γ-Fe phase was significantly lower than that in the α-Fe phase [28], which means that the migration of the Fe2Al5/γ-Fe interface was suppressed relative to that of the Fe2Al5/α-Fe interface. Thus, the growth of Fe2Al5 was hindered. Yousaf et al. had also found that the intermetallic layer decreased radically or even ceased when the relative amount of α-Fe decreased substantially or when the single γ-Fe phase came in contact with the growing front of the intermetallic layer, in the temperature range 775–950 °C [29]. In the present study, as the critical temperatures for Q345 steel undergoing phase transition were Ac1 = 756 °C and Ac3 = 900 °C [30], the Q345 substrate was undergoing austenitizing while hot-dipping at 800 °C; thus, the migration of the Fe2Al5/γ-Fe interface was suppressed. With the extension of hot-dipping time, Al atoms reacted with the Fe2Al5 phase to form the FeAl3 phase, which consumed the Fe2Al5 phase, leading to further thinning of Fe2Al5 phase layer.

3.2. Effect of Si Addition in Molten Bath on Microstructure of Coating

Since Q345 steel had a moderate, flat, and uniform diffusion layer after hot dipping in pure aluminum at 750 °C for 30 s, so the temperature of hot dipping aluminum-silicon was also set to be 750 °C and the dipping time was 30 s. Figure 3 shows the cross-sectional microstructure of Q345 steel hot-dipped in different Al-Si molten pools. Phase analysis by EDS in Figure 3 is listed in Table 3. Similarly, the typical positions (a, b, c…) used to determine the composition of each phase were also added. It should be noted that the compositions of two ternary phases (Al2Fe3Si3 and Al4Fe1.7Si) in Figure 3 are not listed in Table 3. This is mainly because their size was too small to be analyzed by EDS. These two phases were inferred by combining the phase diagram and literature, which will be discussed in detail later. Figure 4 shows the relationship between the thickness of the intermetallic layer and the silicon content. When the Si content was 0.5 at.%, the thickness of the alloy layer was about 19 μm. When the Si content was increased to 10 at.%, the thickness of the alloy layer was reduced to about 7 μm. It is also found that when the Si content was 0.5 at.%, the interface between the alloy layer and the matrix was still in a classic tongue shape, but when the Si content was increased to 2.5 at.% or above, the interface between the alloy layer and the matrix gradually changed from tongue-like to flat. These observations indicated that the Si addition in pure Al reduced the growth rate of the intermetallic layer and changed its morphology.
The intermetallic layer of hot-dip pure aluminum at 750 °C consisted mainly of the Fe2Al5 phase, while the intermetallic layer of hot-dipped Al-0.5Si consisted of the outer FeAl3 phase layer and the inner Fe2Al5 phase layer. According to the Al-Si binary-phase diagram [31], the addition of Si elements can lower the melting point of the Al-Si alloy and makes the microstructure formation temperature of Al-0.5Si hot-dip coating lower than that of pure aluminum hot-dip coating. At the same time, the lower temperature during the hot dip of the Al-Si alloy slowed down the interdiffusion between the steel substrate and the bath, resulting in a thinner alloy-layer thickness. In addition, when the Si content was added to 2.5 at.%, white granular intermetallic compounds appeared in the inner layer of the coating. The size of the white granular phase was so small that it was difficult to analyze its exact composition by EDS. According to the Al-Fe-Si ternary-phase diagram [32], at 700–800 °C, Al2Fe3Si3 is in equilibrium with Fe2Al5 and FeAl3 phases; therefore, these white intermetallic-compound phases were probably the Al2Fe3Si3 phase. Meanwhile, a new phase layer appeared in the outer layer of coating, and this new phase was distributed between the (Al) phase and the FeAl3 phase. By a similar analysis of the Al-Fe-Si ternary-phase diagram, it can be estimated that this intermetallic-compound phase was the Al4Fe1.7Si phase. It can be seen that when the Si content was added above 2.5 at.%, the coating consisted of the Al4Fe1.7Si phase, the FeAl3 phase, the Fe2Al5 phase, and the Al2Fe3Si3 phase. When the Si content was increased to 10 at.%, the coating consisted mainly of the FeAl3 phase and the Al4Fe1.7Si phase. Based on the above analysis, a plausible reason for reducing the growth rate of the intermetallic layer can be attributed to the formation of new phases due to the Si addition in molten aluminum. Meanwhile, a small amount of Si was dissolved into the Fe2Al5 phase, as listed in Table 3, so the reduction of the Fe2Al5 layer might also be due to the fact that silicon occupies vacant sites in the c-axis of the Fe2Al5 crystal structure, thus hindering the diffusion of Al to the interface Fe2Al5/substrate.

3.3. Effect of Annealing Temperature and Time on the Phase Evolution of the Coating

Since the thickness of the diffusion layer of Q345 steel was moderate, flat, and uniform after hot dipping in Al-2.5Si solution at 750 °C for 30 s, this group of samples was selected for the subsequent diffusion-annealing treatment to study the effects of annealing temperature and annealing time on the evolution of microstructure in the coating. Figure 5 shows the cross-sectional microstructures of hot-dipped samples annealed at 800 °C and 900 °C for 1, 2, and 3 h, respectively. Phase analysis by EDS in Figure 5 is listed in Table 4. The typical positions used to determine the composition of each phase were also added in Figure 5. As the composition contrast between the α-Fe phase layer and the Fe-Al phases shown in BSE was not obvious in Figure 5, EDS line profiles were supplemented to distinguish these phase boundaries. In addition, compositions of each phase were determined by point analysis of EDS, and the typical positions marked as a, b, c, and d were also added in the figure.
Figure 5a shows the cross-sectional microstructure of Q345 steel after being hot-dipped in Al-2.5Si solution and annealed at 800 °C for 1 h. According to the EDS results and the analysis of the Fe-Al binary-phase diagram, it is found that the gray phase in the outermost layer of the coating was the FeAl2 phase. There was a small amount of light-gray circle areas, which were composited of the FeAl2 phase according to the EDS analysis. This was due to the inward diffusion of Al elements and the outward diffusion of Fe elements from the matrix during annealing at 800 °C, resulting in the transformation of the Fe2Al5 phase in the original coating into the FeAl2 phase. Moreover, the Fe2Al5 phase had a rhombohedral crystal structure with an atomic saturation concentration of only 70% [15], which allowed the inner Fe atoms to diffuse through the phase and further accelerates transformation of the external Fe2Al5 phase. Between the inner FeAl2 phase layer and the substrate in Figure 5a, there were another two layers, though the color contrast was not obvious. According to the EDS analysis, these two layers consisted of the FeAl phase and the α-Fe phase, respectively.
Figure 5b shows the cross-sectional microstructure of Q345 steel after being hot-dipped in Al-2.5Si solution and annealed at 800 °C for 2 h. The outermost light-gray region and the middle gray-circle region were composed of the FeAl phase. Compared with Figure 5a, it can be found that the thickness of the FeAl phase presented between the α-Fe phase and the FeAl2 phase increased and the outermost region gradually changed from the FeAl2 phase to the FeAl phase. This indicates that the rate of diffusion of Fe atoms into the outer FeAl2 phase layer increased with time. In Figure 5c, the thickness of the FeAl phase layer between the α-Fe phase and the FeAl2 phase further increased.
Figure 5d shows the cross-sectional microstructure of Q345 steel after being hot-dipped in Al-2.5Si solution and annealed at 900 °C for 1 h. According to the EDS analysis and its corresponding line profiles of elements Fe, Al, and Si distributions across the coating layer on the steel substrate, the outermost layer was the FeAl2 phase, inwardly followed by the FeAl phase, the α-Fe phase, and the matrix. In Figure 5e, the outermost layer was the FeAl phase, inwardly followed by the FeAl2 phase, the FeAl phase, the α-Fe phase, and the matrix. In Figure 5f, the FeAl2 phase disappeared and all of them were transformed into the FeAl phase. By analyzing the experimental results in Figure 5, it is known that in order to transform the brittle Fe2Al5 phase in the coating into the FeAl phase with better toughness, it is better to anneal the coating at 900 °C for 3 h.

3.4. Effect of Oxidation Process on the Formation of Al2O3 Film

Figure 6 shows the surface morphology and cross-sectional microstructure of these three groups of samples after oxidation and the corresponding EDS line profiles of elements Fe (red line), Al (green line), O (blue line), and Si (purple line) distributions across the coating layer on the steel substrate. Compositions of each phase are listed in Table 5. In Figure 5a–c were typical positions used to determine the composition of each phase.
Figure 6a,b show the surface morphology and cross-sectional microstructure of the D1 sample after oxidation at 900 °C for 15 h in air. The phase near the matrix in the cross section contained about 76 at.% Fe, 22 at.% Al, and 1 at.% Si, which was determined to be the α-Fe phase, and the phase near the outer part in the cross section contained about 64 at.% Fe, 35 at.% Al and 1 at.% Si, which was determined to be the FeAl phase according to the Fe-Al-Si phase diagram [32]. According to the EDS pattern shown in Figure 6g, the dark-gray area was the Al2O3 film that formed in the surface, which had poor continuity and was not dense, with a thickness of about 4–6 μm.
Figure 6c,d show the surface morphology and cross-sectional microstructure of the D2 sample after oxidation at 900 °C in air for 15 h. From the figures, it can be found that the Al2O3 film on the surface of this group of samples was denser after oxidation. In the surface morphology shown in Figure 6c, Al2O3 was covered with particles in the outermost layer, and in the cross-sectional microstructure shown in Figure 6d, the light-gray phase contained about 74 at.% Fe, 25 at.% Al and 1 at.% Si, so this phase was the α-Fe phase. The dark-gray area was the Al2O3 film formed in the surface, which had good continuity and denseness and was about 6–8 μm thick.
Figure 6e,f show the surface morphology and cross-sectional organization of the D3 sample after oxidation at 900 °C for 15 h in air. It can be found that Al2O3 was covered in the outermost layer in the form of strips, and the light-gray phase in the cross-section contained about 77 at.% Fe, 22 at.% Al, and 1 at.% Si, which was the α-Fe phase, and the dark gray area was the Al2O3 film formed in the surface layer, which had better continuity and denseness than the other two groups, with a thickness of about 5~7 μm.

3.5. High-Temperature Oxidation Resistance of Al2O3 Film

In order to test the high-temperature oxidation resistance of the oxidized samples, Q345 steel and oxidized D1, D2, and D3 samples were placed in corundum crucibles and oxidized at 900 °C in a chamber furnace. The weight increased was recorded at regular intervals, and each sample was oxidized for a total of 63 h. Figure 7 reveals the variation of the weight gain with time. It can be seen from the Figure 7 that the oxidation weight gain of the original Q345 steel was the most obvious as the high-temperature oxidation continued, while slopes of the oxidation weight-gain curves of D1, D2, and D3 samples were relatively small, indicating that the Al2O3 film on the surface of the samples could play a good role in inhibiting further high-temperature oxidation and protecting the Q345 steel substrate. Among the three oxidized samples, sample D3, i.e., the sample with Al2O3 film prepared after annealing at 900 °C for 3 h, showed the smallest weight gain in the subsequent high-temperature oxidation. This indicates that the Al2O3 film formed from the FeAl phase in the outermost layer of the coating has the best resistance to high-temperature oxidation.

4. Conclusions

The addition of silicon in aluminum baths changes the microstructure and phase constitution of the coating. The main findings of the study are as follows:
(1)
When Q345 steel was hot-dipped in pure aluminum, the intermediate phase layer was mainly composed of the Fe2Al5 phase. When the hot-dipping temperature was 750 °C and hot-dipping time was 30 s, the interface between the intermetallic phase layer and the substrate was relatively flat.
(2)
As the silicon content in the aluminum bath increased, the thickness of the intermetallic layer decreased, and the intermetallic layer/steel-substrate interface transformed from an irregular morphology into a flat morphology.
(3)
The annealing time affected the transformation of the intermediate phase in the coating. In addition, the longer the time, the more favorable the transformation of the brittle Fe2Al5 phase into the FeAl phase with good toughness. When the aluminized sample was annealed at 900 °C for 3 h, the intermediate phase in the diffusion layer was completely transformed into the FeAl phase.
In conclusion, the addition of Si into the aluminum bath made the thickness of the aluminizing layer decrease, and the outermost aluminized layer could be transformed into FeAl phase by annealing. This was found to be beneficial to improve the oxidation resistance of Q345 steel. Although annealing at 900 °C was beneficial to promote the transformation from the Fe2Al5 phase to the FeAl phase, the effect of high-temperature annealing on the mechanical properties of the steel matrix should be further studied.

Author Contributions

Y.M.: investigation, writing—original draft; B.Y.: investigation; Y.L.: conceptualization, methodology, investigation, writing—review and editing; J.W.: investigation; X.S.: supervision, validation. All authors have read and agreed to the published version of the manuscript.

Funding

Financial supports from the National Nature Science Foundation of China (Grant Nos. 51871030 and 52171003) and a project funded by the Priority Academic Program Development of Jiangsu higher education institutions are greatly acknowledged.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shafiee, S.; Topal, E. When will fossil fuel reserves be diminished? Energy Policy 2009, 37, 181–189. [Google Scholar] [CrossRef]
  2. The Council of the European Union. Directive 2003/30/EC of the European Parliament and of the Council of 8 May 2003 on the promotion of the use of biofuels or other renewable fuels for transport. Off. J. Eur. Union 2003, 123, 43–46. [Google Scholar]
  3. Grzesik, Z.; Smoła, G.; Adamaszek, K.; Jurasz, Z.; Mrowec, S. Thermal shock corrosion of valve steels utilized in automobile industry. Oxid. Met. 2013, 80, 147–159. [Google Scholar] [CrossRef] [Green Version]
  4. Sazzad, B.S.; Fazal, M.A.; Haseeb, A.; Masjuki, H.H. Retardation of oxidation and material degradation in biodiesel: A review. RSC Adv. 2016, 6, 60244–60263. [Google Scholar] [CrossRef] [Green Version]
  5. Drożdż, M.; Kyzioł, K.; Grzesik, Z. Chromium-based oxidation-resistant coatings for the protection of engine valves in automotive vehicles. Mater. Tehnol. 2017, 51, 603–607. [Google Scholar] [CrossRef]
  6. Badaruddin, M.; Kurniawan, R.D.; Pambudi, A. Low cycle fatigue properties of aluminizing coating on cold-drawn AISI 1018 steel. IOP Conf. Ser. Mater. Sci. Eng. 2020, 807, 012019. [Google Scholar] [CrossRef]
  7. Ryabov, V.R. Aluminizing of Steel (Translated from the Russian Work: Alitirovanie Stali); Amerind Publishing Co. Pvt. Ltd.: New Delhi, India, 1985. [Google Scholar]
  8. Bahadur, A.; Mohanti, O.N. Structural studies of hot dip aluminized coatings on mild steel. Mater. Trans. JIM 1991, 32, 1053–1061. [Google Scholar] [CrossRef] [Green Version]
  9. Heumann, T.; Dittrich, N.A. Structure character of the Fe2Al5 intermetallics compound in hot dip aluminizing process. Z Met. 1959, 50, 617–623. [Google Scholar]
  10. Eggeler, G.; Auer, W.; Kaesche, H. On the influence of silicon on the growth of the alloy layer during hot dip aluminizing. J. Mater. Sci. 1986, 21, 3348–3350. [Google Scholar] [CrossRef]
  11. Dey, P.P.; Modak, P.; Chakrabarti, D.; Banerjee, P.S.; Ghosh, M. A Study on the phase formation and physical characteristics of hot-dip aluminized coating at 750 °C. Metallogr. Microstruct. Anal. 2021, 10, 823–838. [Google Scholar] [CrossRef]
  12. Richards, R.W.; Jones, R.D.; Clements, P.D.; Clarke, H. Metallurgy of continuous hot dip aluminizing. Int. Mater. Rev. 1994, 39, 191–212. [Google Scholar] [CrossRef]
  13. Lainer, D.I.; Kurakin, A.K. Mechanism of the effect of silicon in aluminum on the process of reactive diffusion of iron into aluminum. Fiz. Metal. Metalloved. 1964, 18, 145. [Google Scholar]
  14. Zarei, F.; Nuranian, H.; Shirvani, K. Effect of Si addition on the microstructure and oxidation behaviour of formed aluminide coating on HH309 steel by cast-aluminizing. Surf. Coat. Technol. 2020, 394, 125901. [Google Scholar] [CrossRef]
  15. Wang, C.J.; Chen, S.M. The high-temperature oxidation behavior of hot-dipping Al–Si coating on low carbon steel. Surf. Coat. Technol. 2006, 200, 6601–6605. [Google Scholar] [CrossRef]
  16. Huang, X.; Xiao, K.; Fang, X.; Xiong, Z.; Wei, L.; Zhu, P.; Li, X. Oxidation behavior of 316L austenitic stainless steel in high temperature air with long-term exposure. Mater. Res. Express 2020, 7, 066517. [Google Scholar] [CrossRef]
  17. Abro, M.A.; Hahn, J.; Lee, D.B. High temperature oxidation of hot-dip aluminized T92 steels. Met. Mater. Int. 2018, 24, 507–515. [Google Scholar] [CrossRef]
  18. Zhang, G.; Yang, F.; Lu, G.; Xiang, X.; Tang, T.; Wang, X. Fabrication of Al 2 O 3/FeAl coating as tritium permeation barrier on tritium operating component on quasi-CFETR scale. J. Fusion Energy 2018, 37, 317–324. [Google Scholar] [CrossRef]
  19. Chen, J.; Li, X.; Hua, P.; Wang, C.; Chen, K.; Wu, Y.; Zhou, W. Growth of inter-metallic compound layers on CLAM steel by HDA and preparation of permeation barrier by oxidation. Fusion Eng. Des. 2017, 125, 57–63. [Google Scholar] [CrossRef]
  20. Cheng, W.J.; Wang, C.J. Study of microstructure and phase evolution of hot-dipped aluminide mild steel during high-temperature diffusion using electron backscatter diffraction. Appl. Surf. Sci. 2011, 257, 4663–4668. [Google Scholar] [CrossRef]
  21. Badaruddin, M. Improvement of high temperature oxidation of low carbon steel exposed to ethanol combustion product at 700 C by hot-dip aluminizing coating. Makara J. Technol. 2012, 15, 137–141. [Google Scholar] [CrossRef]
  22. Badaruddin, M.; Suharno, S.; Wijaya, H.A. Isothermal oxidation behavior of aluminized AISI 1020 steel at the temperature of 700 °C. J. Tek. Mesin 2014, 15, 15–19. [Google Scholar]
  23. Hakam, M.; Wang, C.J.; Wardono, H.; Asmi, D. High temperature oxidation of low carbon steel with and without an Al coating in an atmosphere containing burning ethanol. AIP Conf. Proc. 2018, 1983, 050002. [Google Scholar]
  24. Xue, Z.; Hao, X.; Huang, Y.; Gu, L.; Ren, Y.; Zheng, R. Wear resistance and wear mechanism of a hot dip aluminized steel in sliding wear test. Surf. Rev. Lett. 2016, 23, 1550098. [Google Scholar] [CrossRef]
  25. Li, Y.; Liu, Y.; Yang, J. First principle calculations and mechanical properties of the intermetallic compounds in a laser welded steel/aluminum joint. Opt. Laser Technol. 2020, 122, 105875. [Google Scholar] [CrossRef]
  26. Xu, G.; Wang, K.; Dong, X.; Yang, L.; Ebrahimi, M.; Jiang, H.; Wang, Q.; Ding, W. Review on corrosion resistance of mild steels in liquid aluminum. J. Mater. Sci. Technol. 2021, 71, 12–22. [Google Scholar] [CrossRef]
  27. Chen, S.; Yang, D.; Zhang, M.; Huang, J.H.; Zhao, X.K. Interaction between the growth and dissolution of intermetallic compounds in the interfacial reaction between solid iron and liquid aluminum. Metall. Mater. Trans. A 2016, 47, 5088–5100. [Google Scholar] [CrossRef]
  28. Zhang, Q.; Chen, J.H.; Li, Y.J. Microstructure and oxidationresistance property of aluminizing layer of two kinds of high-chromium steel. Heat Treat. Met. 2000, 10, 7–10. [Google Scholar]
  29. Yousaf, M.; Iqbal, J.; Ajmal, M. Variables affecting growth and morphology of the intermetallic layer (Fe2Al5). Mater. Charact. 2011, 62, 517–525. [Google Scholar] [CrossRef]
  30. Lin, W.; Zhang, X.; Zhao, T.K.; Li, H.Y. Continuous cooling transformation curve of under cooling austenite about Q345 steel. Mater. Sci. Technol. 2009, 17, 247–250. [Google Scholar]
  31. Massalski, T.B. Binary Alloy Phase Diagrams; ASM International: Phoenix, AZ, USA, 1986. [Google Scholar]
  32. Marker, M.; Skolyszewska-Kühberger, B.; Effenberger, H.S.; Schmetterer, C.; Richter, K.W. Phase equilibria and structural investigations in the system Al–Fe–Si. Intermetallics 2011, 19, 1919–1929. [Google Scholar] [CrossRef]
Figure 1. Cross-sectional microstructure of coating of Q345 steel hot-dipped in pure aluminum solution at 700, 750, and 800 °C for 30 s and 3 min, respectively. (a) at 700 °C for 30 s, (b) at 700 °C for 3 min, (c) at 750 °C for 30 s, (d) at 750 °C for 3 min, (e) at 800 °C for 30 s, (f) at 800 °C for 3 min.
Figure 1. Cross-sectional microstructure of coating of Q345 steel hot-dipped in pure aluminum solution at 700, 750, and 800 °C for 30 s and 3 min, respectively. (a) at 700 °C for 30 s, (b) at 700 °C for 3 min, (c) at 750 °C for 30 s, (d) at 750 °C for 3 min, (e) at 800 °C for 30 s, (f) at 800 °C for 3 min.
Coatings 12 00503 g001
Figure 2. Thickness of intermetallic layer in the coating of Q345 steel after hot dipping at different temperature for 30 s and 3 min.
Figure 2. Thickness of intermetallic layer in the coating of Q345 steel after hot dipping at different temperature for 30 s and 3 min.
Coatings 12 00503 g002
Figure 3. Cross-sectional micrographs of coatings of Q345 steel hot-dipped in baths with different Si addition at 750 °C for 30 s. (a) 0.5Si, (b) 1Si, (c) 2.5Si, (d) 5Si, (e) 7.5Si, and (f) 10Si.
Figure 3. Cross-sectional micrographs of coatings of Q345 steel hot-dipped in baths with different Si addition at 750 °C for 30 s. (a) 0.5Si, (b) 1Si, (c) 2.5Si, (d) 5Si, (e) 7.5Si, and (f) 10Si.
Coatings 12 00503 g003aCoatings 12 00503 g003b
Figure 4. Relationship between intermetallic thickness of Q345 steel after hot-dip aluminum with different silicon content.
Figure 4. Relationship between intermetallic thickness of Q345 steel after hot-dip aluminum with different silicon content.
Coatings 12 00503 g004
Figure 5. Cross-sectional microstructure of Q345 after hot dipping in Al-2.5Si solution and annealed at 800 °C or 900 °C for different times, and its corresponding EDS line profiles of elements Fe (red line), Al (green line), and Si (blue line) distributions across the coating layer on the steel substrate. (a) at 800 °C for 1 h, (b) at 800 °C for 22 h, (c) at 800 °C for 3 h, (d) at 900 °C for 1 h, (e) at 900 °C for 2 h, (f) at 900 °C for 3 h.
Figure 5. Cross-sectional microstructure of Q345 after hot dipping in Al-2.5Si solution and annealed at 800 °C or 900 °C for different times, and its corresponding EDS line profiles of elements Fe (red line), Al (green line), and Si (blue line) distributions across the coating layer on the steel substrate. (a) at 800 °C for 1 h, (b) at 800 °C for 22 h, (c) at 800 °C for 3 h, (d) at 900 °C for 1 h, (e) at 900 °C for 2 h, (f) at 900 °C for 3 h.
Coatings 12 00503 g005
Figure 6. Surface morphology and cross-sectional microstructure of Q345 steel and its corresponding EDS line profiles of elements Fe (red line), Al (green line), O (blue line), and Si (purple line) distributions across the coating layer on the steel substrate after oxidation at 900 °C for 15 h with different preoxidation heat treatment. (a,b) Oxidation directly after hot dip in Al-2.5Si; (c,d) oxidation after hot dip in Al-2.5Si and annealing at 800 °C for 1 h; (e,f) oxidation after hot dip in Al-2.5Si and annealing at 900 °C for 3 h; and (g) the typical EDS pattern of the Al2O3 phase.
Figure 6. Surface morphology and cross-sectional microstructure of Q345 steel and its corresponding EDS line profiles of elements Fe (red line), Al (green line), O (blue line), and Si (purple line) distributions across the coating layer on the steel substrate after oxidation at 900 °C for 15 h with different preoxidation heat treatment. (a,b) Oxidation directly after hot dip in Al-2.5Si; (c,d) oxidation after hot dip in Al-2.5Si and annealing at 800 °C for 1 h; (e,f) oxidation after hot dip in Al-2.5Si and annealing at 900 °C for 3 h; and (g) the typical EDS pattern of the Al2O3 phase.
Coatings 12 00503 g006
Figure 7. Relationship of weight gain of Q345 steel with oxidation time: (a) pure Q345 steel; (b) D1, D2, and D3 samples prepared by hot dipping and annealing.
Figure 7. Relationship of weight gain of Q345 steel with oxidation time: (a) pure Q345 steel; (b) D1, D2, and D3 samples prepared by hot dipping and annealing.
Coatings 12 00503 g007
Table 1. Comparison of chemical composition (wt.%) of Q345 steel in this study with standard values.
Table 1. Comparison of chemical composition (wt.%) of Q345 steel in this study with standard values.
SourcesFeMnSSiCP
This studyBal.1.30.0150.30.160.015
StandardBal.≤1.70≤0.035≤0.50≤0.20≤0.035
Table 2. Phase Analysis by EDS in Figure 1 (at.%).
Table 2. Phase Analysis by EDS in Figure 1 (at.%).
No.Typical PositionsFeAlMnPhase
(a)a28.271.20.6Fe2Al5
b24.275.30.5FeAl3
c23.376.40.3FeAl3
(b)a28.870.90.3Fe2Al5
b24.175.30.6FeAl3
c23.476.10.5FeAl3
(c)a28.571.40.1Fe2Al5
b24.375.30.4FeAl3
c23.376.20.5FeAl3
(d)a28.171.60.3Fe2Al5
b24.575.10.4FeAl3
c23.276.20.6FeAl3
(e)a27.971.90.2Fe2Al5
b23.875.90.3FeAl3
c23.176.30.6FeAl3
(f)a28.171.50.4Fe2Al5
b24.175.60.3FeAl3
c23.176.50.4FeAl3
Table 3. Phase analysis by EDS in Figure 3 (at.%).
Table 3. Phase analysis by EDS in Figure 3 (at.%).
No.Typical PositionsFeAlSiPhase
(a)a28.170.51.4Fe2Al5
b24.574.31.2FeAl3
c23.674.91.5FeAl3
(b)a27.370.62.1Fe2Al5
b23.374.91.8FeAl3
c22.975.21.9FeAl3
(c)a26.570.43.1Fe2Al5
b23.374.32.4FeAl3
c22.375.22.5FeAl3
(d)a27.769.13.2Fe2Al5
b23.473.53.1FeAl3
c23.474.13.5FeAl3
(e)a27.668.24.2Fe2Al5
b24.269.76.1FeAl3
c23.070.46.6FeAl3
(f)a24.168.67.3FeAl3
Table 4. Phase Analysis by EDS in Figure 5 (at.%).
Table 4. Phase Analysis by EDS in Figure 5 (at.%).
No.Typical PositionsFeAlSiPhase
(a)a74.622.33.1α-Fe
b45.241.912.9FeAl
c52.136.711.2FeAl
d31.767.21.1FeAl2
(b)a78.219.12.7α-Fe
b46.645.57.9FeAl
c46.447.56.1FeAl
d30.267.52.3FeAl2
(c)a75.222.12.7α-Fe
b48.544.17.4FeAl
c44.87.148.1FeAl
d31.167.21.7FeAl2
(d)a78.818.72.5α-Fe
b56.940.42.7FeAl
c62.335.62.1FeAl
d33.865.50.7FeAl2
(e)a78.818.82.4α-Fe
b50.446.53.1FeAl
c49.444.95.7FeAl
d34.164.81.1FeAl2
(f)a78.321.30.4α-Fe
b52.546.41.1FeAl
Table 5. Phase Analysis by EDS in Figure 6 (at.%).
Table 5. Phase Analysis by EDS in Figure 6 (at.%).
No.Typical PositionsFeAlSiOPhase
(b)a76.622.21.2-α-Fe
b63.935.30.8-FeAl
c2.932.72.661.8Al2O3
(d)a78.219.12.7-α-Fe
b2.629.22.865.4Al2O3
(f)a72.924.92.2-α-Fe
b2.429.82.665.2Al2O3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ma, Y.; Yuan, B.; Liu, Y.; Wang, J.; Su, X. Effect of Annealing and Oxidation on the Microstructure Evolution of Hot-Dipped Aluminide Q345 Steel with Silicon Addition. Coatings 2022, 12, 503. https://doi.org/10.3390/coatings12040503

AMA Style

Ma Y, Yuan B, Liu Y, Wang J, Su X. Effect of Annealing and Oxidation on the Microstructure Evolution of Hot-Dipped Aluminide Q345 Steel with Silicon Addition. Coatings. 2022; 12(4):503. https://doi.org/10.3390/coatings12040503

Chicago/Turabian Style

Ma, Yujian, Binbin Yuan, Ya Liu, Jianhua Wang, and Xuping Su. 2022. "Effect of Annealing and Oxidation on the Microstructure Evolution of Hot-Dipped Aluminide Q345 Steel with Silicon Addition" Coatings 12, no. 4: 503. https://doi.org/10.3390/coatings12040503

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