Next Article in Journal
An Experimental Investigation about the Dimensional Accuracy and the Porosity of Copper-Filled PLA Fused Filament Fabrication Parts
Next Article in Special Issue
Effects of Finish Rolling Temperature on the Critical Crack Tip Opening Displacement (CTOD) of Typical 500 MPa Grade Weathering Steel
Previous Article in Journal
Fatigue Behavior of Cu-Zr Metallic Glasses under Cyclic Loading
Previous Article in Special Issue
Evolutions of Microstructure and Impact Toughness of Submerged Arc Weld Metal via Introducing Varied Si for Weathering Bridge Steel
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Heat Input on Microstructure and Tensile Properties in Simulated CGHAZ of a V-Ti-N Microalloyed Weathering Steel

1
State Key Laboratory of Metastable Materials Science and Technology, Yanshan University, Qinhuangdao 066004, China
2
Hebei Key Lab for Optimizing Metal Product Technology and Performance, Yanshan University, Qinhuangdao 066004, China
*
Author to whom correspondence should be addressed.
Metals 2023, 13(9), 1607; https://doi.org/10.3390/met13091607
Submission received: 20 August 2023 / Revised: 12 September 2023 / Accepted: 14 September 2023 / Published: 17 September 2023
(This article belongs to the Special Issue Advances in Weathering Bridge Steels)

Abstract

:
The mechanical properties of a coarse-grained heat-affected zone (CGHAZ) are affected by welding thermal cycling with varied heat input (Ej), but its effect on tensile properties is rarely studied. In the present work, Ej = 15, 35, 55, 75 kJ/cm CGHAZ samples were prepared via GleebleTM (St. Paul, MN, USA) for a novel V-Ti-N microalloyed weathering steel. The tensile properties of CGHAZ with varied Ej were evaluated. The results indicated that mixed microstructures dominated by lath bainitic ferrite (LBF) and granular bainitic ferrite (GBF) were obtained at Ej = 15 and 35 kJ/cm, respectively, while a mixed microstructure composed of GBF, intragranular acicular ferrite (IGAF), and polygon ferrite (PF) formed at Ej = 55 and 75 kJ/cm, apart from martensite/austenite (M/A) constituents in each Ej condition. The above variation tendency in the microstructure with the increase in Ej led to coarsening of low-angle grain boundaries (LAGBs) and a decrease in dislocation density, which in turn resulted in a yield strength (YS) decrease from 480 MPa to 416 MPa. The mean equivalent diameter (MED), defined by the misorientation tolerance angles (MTAs) ranging from 2–6°, had the strongest contribution to YS due to their higher fitting coefficient of the Hall–Petch relationship. In addition, the increase in the average size (dM/A) of M/A constituents from 0.98 μm to 1.81 μm and in their area fraction (fM/A) from 3.11% to 4.42% enhanced the strain-hardening stress. The yield strength ratio (YR) reduced as the Ej increased, and the lower density and more uniform dislocations inside the ferrite led to a uniform elongation (uE) increase from 9.5% to 18.6%.

1. Introduction

Weathering steel (WS) is, essentially, an atmospheric-corrosion-resistant steel, usually containing no more than 0.20 wt % C, 0.25–0.55 wt % Si, and 0.8–1.5 wt % Mn, as well as one or more other alloy elements, such as Cu, Cr, Ni, with a total alloy content of 1.00–5.00 wt % [1,2,3]. WS normally possesses a stable rust layer to prevent erosion, having many advantages over laborious, costly, and high-emission coatings, and it has been increasingly used for important applications, such as uncoated bridges, buildings, and transmission towers, through the traditional arc welding process [4,5,6]. However, during the welding of WS, due to the influence of thermal cycling, the strength and toughness reduction in CGHAZ near the fusion line cannot be ignored [7,8]. Since weathering steel contains more Si [9], Cr [4,10], and other alloy elements, after subjection to the welding thermal cycle with a peak temperature higher than 1300 °C, hard brittle phase GBF and M/A constituents are enriched in CGHAZ [11,12]. When the Ej increases, the above-mentioned hard and brittle phases will coarsen and increase in quantity [13,14], further increase the reduction in toughness and strength.
Extensive investigations [15,16] have reported that the nucleation of heteronucleated ferrites on inclusions or precipitated particles can reduce the content of GBF, ferrite side plate, etc., refining the CGHAZ microstructure and improving its mechanical properties. V-Ti-N microalloyed steels with high ferritic heterogeneous nucleation capacity have received extensive attention. These utilize Ti (C, N) particles with high thermal stability to pin prior austenite grain boundaries (PAGBs) and refine them. VN and/or V (C, N) particles from the solid γ-phase reduce the mismatch with ferrite and promote the heterogeneous nucleation of ferrites [17,18]. These studies provide good microalloyed methods to improve the CGHAZ mechanical properties of weathering steel. Zhang and Hu et al. [19,20] studied the effect of Ej on the CGHAZ microstructure and impact properties of V-Ti-N microalloyed C-Mn steel; the results showed that with the increase in Ej, the ability of V-rich particles to promote heterogeneous nucleation of ferrite was enhanced, and the corresponding impact energy increased. Shi et al. [21] also obtained similar research conclusions: as the time of T8/5 is increased from 45 s to 285 s, the CGHAZ toughness of V-Ti-N microalloyed steel significantly improved; however, the YS of CGHAZ decreased from 385 MPa to 335 MPa due to the formation of a large amount of intragranular polygon ferrite (IGPF) with the extension of T8/5. Moreover, Wang et al. [22] investigated the CGHAZ tensile properties of V-Ti-N microalloyed steel containing 0.28 wt % Mo under varied Ej, during the process of increasing Ej from 35 kJ/cm to 65 kJ/cm and 120 kJ/cm; the YS of CGHAZ showed a trend of first increasing from 545 MPa to 610 MPa, and then decreasing to 394 MPa, respectively, the main reason being that the CGHAZ main microstructure transformed from GBF to IGAF and IGPF, respectively.
Considering that the influences of heat inputs on CGHAZ microstructure and the mechanical properties of V-Ti-N microalloyed steels with varied compositions are different, for the V-Ti-N microalloyed weathering steel, there may still be complex connections between the heat inputs and CGHAZ microstructure and mechanical properties. The effect of heat inputs on their CGHAZ impact property was reported by the author [23], because the change in CGHAZ strength can reflect its tendency towards softening and hardening, while also affecting the fatigue performance of the joint [24]. It is necessary to further evaluate the strength of CGHAZ with different Ej for V-Ti-N microalloyed weathering steel, and to the authors’ knowledge, research on heat inputs and CGHAZ strength has been scarcely reported previously.
In addition, during the continuous cooling process of the welding thermal cycle, bainitic ferrite, IGAF, IGPF, and other mixed microstructures are produced in the CGHAZ of V-Ti-N microalloyed weathering steel. It is difficult to establish relationship between complex microstructure and tensile properties [25,26]. In recent years, electron backscatter diffraction (EBSD) technology has been developed to analyze the influence of complex microstructure on tensile properties based on the variation of MTAs of different grain boundaries [25,27]. Fan et al. [28] showed that grain boundaries with MTAs ranging from 2–6° are the main boundaries controlling the tensile properties for low-alloy steel, while Zhu et al. [29] pointed out that the MED defined by MTAs smaller than 10° effectively governed the tensile properties. Olasolo et al. [30] indicated that MED defined by LAGBs less than 4° strongly contributes to tensile properties. Further, Li et al. [31] found the closest Hall–Petch relationship between the YS and the ferrite grains defined by LAGBs ranging from 2–12° in C-Mn steel. Hence, complex CGHAZ microstructure and corresponding tensile properties need further study. Moreover, the evolution of Ej will lead to differences in M/A constituents, dislocation structure, and density in CGHAZ microstructure, and its influences on strain-hardening behavior and elongation need to be elaborated.
In the present work, a novel V-Ti-N microalloyed weathering steel was designed and prepared. The steel was subjected to a CGHAZ welding thermal cycle with Ej = 15, 35, 55, 75 kJ/cm on a Gleeble-3500, and then the CGHAZ microstructure under each heat input condition was characterized and its tensile properties were tested. Moreover, the CGHAZ microstructure was further analyzed and quantitatively characterized by BESD and TEM, the relationship between the complex CGHAZ microstructure and tensile properties was evaluated, and the most effective structural unit for controlling YS was determined. In addition, the influence of CGHAZ microstructure on strain-hardening ability and elongation under different heat input conditions had also been studied. The research in this article will further clarify the influence of heat input on the microstructure and tensile properties of the welding heat-affected zone and understand the tensile fracture behavior with multiphase microstructure.

2. Materials and Experimental Method

The steel used in this article is a novel V-Ti-N microalloyed weathering steel. The specific chemical composition is shown in Table 1, and the mechanical properties of the test steel are shown in Table 2; this indicates that this novel weathering steel has the advantages of high strength, high toughness, and low yield ratio.
Round bars with sizes φ15 × 80 (rolling direction) mm2 were machined from the test steel. The simulated CGHAZ with Ej = 15, 35, 55, and 75 kJ/cm were conducted in a Gleeble-3500 system (Figure 1a). The samples were heated to 1320 °C at 100 °C/s and cooled based on the thermal cycle curve developed via the Rykalin-2D mode; the model formula is shown in Equation (1), the T8/5 time corresponding to Ej = 15, 35, 55, and 75 kJ/cm is 18.9 s, 50.6 s, 112.3 s, and 232.5 s, respectively, and the thermal cycle curves are shown in Figure 1b. Taking samples from a 2 mm area on both sides of the thermocouple of the round bars, the CGHAZ microstructure and tensile properties of each Ej sample was detected, as shown in Figure 1c. Tensile properties are tested on an MTS servo hydraulic testing machine at room temperature with a strain of 0.25 mm/min, and the specimens obtained for each Ej were tested three times. The YS was determined by the 0.2% offset flow stress. In addition, a Vickers hardness tester (HV-10Z) was used to test the hardness of different heat input samples, and the average value was measured 5 times under each condition.
The microstructure was characterized via optical microscopy (OM; Axiover-200MA T, ZEISS, Jena, Germany), scanning electron microscopy (SEM; Hitachi-SU5000, Hitachi, Tokyo, Japan), and transmission electron microscopy (TEM; Talos-F200X, FEI, Hillsboro, USA) for each Ej sample. OM and SEM specimens were polished and etched with 4% nitrate alcohol. TEM samples for observing precipitated particles at different Ej were prepared by carbon extraction replication. The film samples of TEM were obtained by sanding to 30 μm and subsequently electrolytic polishing, and the polishing occurred in a solution of 5% perchloric acid in ethyl alcohol with v = 25 V, i = 55–65 mA, and T = 25 °C. The OM samples were re-polished and then etched with LePera’s reagent. The M/A constituents were corroded into bright white and subsequently observed by OM, then Image-Pro Plus (IPP) software (Media Cybernetics, Austin, USA) was used to evaluate the size and area fraction of M/A constituents for each Ej through more than 1000 M/A constituents in 15 fields of view. Samples from each Ej were ground and then electrolytically polished in 90% methanol solution of perchlorate with v = 18 V, I = 1 A, and t = 30 s, and evaluated by SEM equipped with a TSL electron backscatter diffraction (EBSD); the scanning step was 0.20 μm, and the MED of grains at different MTAs was determined by EBSD. In addition, an XRD (Rigaku D/max-2500 PC, Rigaku, Tokyo, Japan) with Cu K-α radiation was employed to determine the dislocation density for each Ej sample, and the specimens were scanned at a step width of 0.02° and counting time of 2 s over the range of 40°–105°, with three tests per sample.
E j = 4 π k c ρ Δ t 1 / T 2   - T 0 2   -   1 / T 1   -   T 0 2 × δ
where Ej is heat input; c is the specific heat capacity, 0.586 J/(g·°C); ρ is the density of iron, 7.800 g/cm3; k is the heat conductivity coefficient, 0.360 J/(cm·s·°C); δ is the thickness of plating, 0.24 cm; T1 and T2 are the defined cooling time temperatures T1 = 800 °C, T2 = 500 °C, respectively; T0 is the preheating temperature, 25 °C; Δt is temperature cooling time from T1 to T2.

3. Experimental Result

3.1. Microstructural Characteristics

OM observations of simulated CGHAZ samples under different Ej are displayed in Figure 2; there are obvious differences in the morphology of the microstructures. When the Ej = 15 and 35 kJ/cm, the prior austenite grain boundaries (PAGBs) can be observed, and the PAGBs may be further divided by lath bainitic ferrite (LBF) and granular bainitic ferrite (GBF). However, the proportion of LBF is higher at Ej = 15 kJ/cm, while a microstructure of GBF is dominant at Ej = 35 kJ/cm. As the Ej further increased to 55 and 75 kJ/cm with two large heat inputs, PAGBs were almost covered by growth of ferrite and the CGHAZ microstructures were mainly intragranular acicular ferrite (IGAF) and polygon ferrite (PF). These types of microstructures occupy a higher proportion at Ej = 75 kJ/cm. In addition, developed GBF can also be found in CGHAZ microstructure of two large heat inputs.
Moreover, Figure 3 characterizes the variation in M/A constituents of simulated CGHAZ under different Ej; the M/A constituents appear bright white after being corroded by LePera’s reagent. With the increase in Ej, the macroscopic morphology of M/A constituents gradually transforms from elongated to blocky. Meanwhile, according to the statistics of the IPP software, the average size (dM/A) and area fraction (fM/A) of M/A constituents increase with the increase in Ej. Specifically, dM/A increases from 0.98 μm to 1.13 μm, 1.68 μm, and 1.81 μm, and fM/A increases from 3.11% to 3.56%, 4.21%, and 4.42%, respectively.
The effect of Ej on CGHAZ microstructure is further characterized by TEM. With the increase in Ej from 15 kJ/cm to 75 kJ/cm, the ferrite morphology changes from parallel lath bundles to blocky, and the size of ferrite coarsens significantly, with the results shown in Figure 4a–d. The black island structures can be found in each Ej sample, and their size increases with the increase in Ej. The island structure is identified as M/A constituent by light and dark fields and selected area electron diffraction (SAED) pattern (Figure 4e–e2). Moreover, the dislocation structures inside the ferrite are also characterized; when the Ej = 15 and 35 kJ/cm, there are high-density dislocation structures in the ferrite. In the partial enlarged view of Figure 4a,b, dislocation tangles (DTs) occur around the elongated M/A constituents (Figure 4a1,b1). With a further increase in Ej to 55 kJ/cm, dislocation density decreases significantly compared to Ej = 15 and 35 kJ/cm, but compared to the relatively uniformly distributed dislocation lines (UDDLs) at Ej = 75 kJ/cm (Figure 4d1), the internal dislocation lines (DLs) are densely distributed (Figure 4c1).
In addition, XRD patterns of samples that are subjected to different Ej are displayed in Figure 5a. The dislocation density can be calculated by Equation (2) according to Ref. [32]; the relevant calculation results are shown in Figure 5b. With the increase in Ej from 15 kJ/cm to 75 kJ/cm, the dislocation density decreases from 2.20 × 1014 m−2 to 0.73 × 1014 m−2.
ρ = k ε 2 / F b 2
where ρ is the dislocation density, k is the Burgers vector of dislocations, ɛ is microstrain, (b) along <111> for body-centered cubic metals; the value of F is assumed to be 1.
Moreover, TEM characterizations of CGHAZ carbon extraction replica samples are shown in Figure 6. When the Ej = 15 and 35 kJ/cm (Figure 6a,b), the precipitated particles are mainly square particles. When the Ej increases to 55 and 75 kJ/cm (Figure 6c,d), in addition to square particles, there are also round particles in the microstructure, and the density and size of precipitated particles is enhanced with the increase in Ej. Further, the typical precipitated particles under each Ej condition are analyzed by energy dispersive spectroscopy (EDS), as shown in Figure 6I–IV. The EDS results indicate that the square particles are Ti (C, N) particles at low Ej = 15 and 35 kJ/cm (Figure 6I,II). However, when the Ej increases to 55 and 75 kJ/cm, based on the results of EDS analysis, the circular precipitated particles appearing in the sample are V (C, N) particles (Figure 6III), and the chemical composition of the square particles transforms into V-rich (Ti, V) (C, N) particles (Figure 6IV); this is also consistent with the research results of Zhang and Hu et al. [19,33]. In addition, 20 square particles and round particles under each heat input condition are selected respectively, and their EDS data were further processed and the atomic fraction of elements were statistically analyzed; the results are shown in Table 3. The results indicate that with the increase in Ej, the atomic fraction of V in square and circular precipitates is increased.
The EBSD inverse pole figure maps and the corresponding kernel average misorientation (KAM) maps of the welding thermal cycle at Ej = 15–75 kJ/cm are shown in Figure 7a–d and Figure 7e–h, respectively. The high-angle grain boundaries (HAGBs) are defined by misorientation tolerance angles (MTAs) of boundaries higher than 15° and marked by black lines, while the low-angle grain boundaries (LAGBs) are marked by white lines, which are defined by MTAs of boundaries ranging from 2°–15° [34]. The mean equivalent diameter (MED) of grains with varied MTA is evaluated, with the results counted in Table 4. With the increase in Ej, coarsening occurs in the MED defined by MTAs of 2°, 4°, 6°, 8°, 10° and 12°. Moreover, the KAM maps can be employed to estimate the degree of plastic deformation and the density of geometrically necessary dislocations (GNDs) [35,36]. The average KAM value (KAMave) decreases from 0.83° to 0.55° with the increase in Ej; this trend is also consistent with the results of TEM and XRD observations, where an increase in Ej leads to a decrease in dislocation density.

3.2. Mechanical Properties of CGHAZ

The tensile properties and Vickers hardness of CGHAZ samples with different welding heat inputs were tested, and the relevant results are shown in Table 5, and the typical CGHAZ stress-strain curves and detailed tensile properties of each Ej sample are shown in Figure 8a and Figure 8b, respectively. With the increase in Ej, the hardness of the CGHAZ reduced from 238.4 HV10 to 201HV10, and the YS and TS decreased from 480 MPa to 416 MPa and from 615 MPa to 574 MPa, respectively, while the YR decreased monotonically from 0.78 to 0.72. In addition, the uE increased significantly from 9.5% to 20.1%.

4. Discussion

4.1. Effect of Ej on CGHAZ Microstructure

With the increase in Ej from 15 kJ/cm to 35 kJ/cm and 55/75 kJ/cm, the main CGHAZ microstructure changes from LBF to GBF, and further changes to a mixture of PF and IGAF. The evolutions in microstructure are related to the difference in the cooling process caused by varied Ej. When the Ej = 15 kJ/cm to 35 kJ/cm, the high-temperature residence time is shorter and the cooling rate is faster, which inhibits the diffusion of C atoms, and the supercooling degree is larger and hence promotes the transformation of bainitic ferrite [37]. The PAGBs are known to be the nucleation sites of bainitic ferrite due to the irregular atomic arrangement and lattice distortion [11]. The γ (austenite)→LBF transition tends to require a greater degree of supercooling than the γ→GBF transition [8]; therefore, with the increase in Ej, the nucleation of LBF is inhibited, resulting in a large amount of GBF at Ej = 35 kJ/cm, while the microstructure is mainly LBF at Ej = 15 kJ/cm.
In addition, it is noteworthy that uniform nucleation of ferrite and heterogeneous nucleation of ferrite are very difficult to detect under small heat inputs, which is related to limited carbon diffusion and high undercooling in favor of bainite ferrite transformation, as well as to heteronuclear particles. Although the nucleation temperatures of GBF and IGAF are both between 400–600 °C according to Ref. [38], at small heat inputs, the precipitated particles that reach the critical nucleation size (particles with an equivalent circular diameter higher than 100 nm have nucleation ability according to Refs. [38,39]) are Ti (C, N) particles (Figure 6a,b). Previous studies [40,41] have reported that the ability of Ti (C, N) particles as heterogeneous nucleation sites is weak; as shown in Figure 9a, Ti (C, N) particles fail to promote heterogeneous nucleation of ferrite. The competitive nucleation and growth of GBF and IGAF under small heat inputs are mainly dominated by GBF, while IGAF is inhibited. Furthermore, increasing the Ej to 55/75 kJ/cm, the high-temperature residence time is extended, the corresponding undercooling decreases, and the diffusion of carbon atoms is promoted, all of which are beneficial for γ→α (ferrite) phase transition to occur at a higher temperature. Meanwhile, the precipitated particles transform from Ti (C, N) to V-rich (Ti, V) (C, N) particles (Figure 6c,d). Compared with Ti (C, N) particles, the mismatch between V-rich (Ti, V) (C,N) particles and ferrite is lower, which can significantly promote the heterogeneous nucleation of ferrite [42], as shown in Figure 9b; the heterogeneous nucleation of the ferrite on the V-rich (Ti, V) (C, N) particles is evident. Accordingly, the nucleation at PAGBs is mainly GBPF, while the heterogeneous nucleation of IGAF and IGPF occurs in the grain; thereby high proportions of PF and IGAF are formed in the CGHAZ when Ej = 55 and 75 kJ/cm.
In fact, the γ→α phase transformation is accompanied by the diffusion of C atoms from α to γ due to the higher solubility of C atoms at γ, and the C-rich retained austenite (γ’) forms during this period. γ’ can be stable at relatively high temperature because of the high C content, and then transform into M/A constituents in the air cooling process [43]. With the increase in Ej from 15 kJ/cm to 75 kJ/cm, the high-temperature residence time gradually prolongs, and the diffusion of C becomes more sufficient. Therefore, with the increase in Ej, the mean size and area fraction of M/A constituents increase.

4.2. Effect of Ej on CGHAZ Tensile Properties

With the increase in Ej, the grain size, the dislocation density, and the size and volume fraction of precipitated particles in CGHAZ will all change, and these will have an effect on the strength. According to previous research [28,44], the YS of low-alloy steel is generally determined by the cumulative contributions of various strengthening mechanisms, and the main strengthening modes are shown in Equation (3):
σ y = σ 0 + σ s s + σ d + σ ρ + σ p p + σ M / A
Here, σy is the YS, σ0 represents the Peierls–Nabarro stress or lattice friction stress, σss represents the solid solution strengthening, σd represents the boundary strengthening, σρ represents the dislocation strengthening, σpp represents the precipitation strengthening, and σM/A represents the M/A constituent strengthening. Among these strengthening mechanisms, the boundary strengthening exerted by LAGBs and the dislocation strengthening exerted by dislocation structures in the microstructure contribute the greatest [28,45]. With the increase in Ej, the MED defined by LAGBs are coarsened (Figure 7 and Table 4), while the dislocation density decreases (Figure 4 and Figure 5), resulting in a monotonic decrease in YS of CGHAZ.
Typically, the LAGBs defined by MTAs ranging from 2° to 15° have the most effective contribution to boundary strengthening [22], the HAGBs (θ ≥ 15°) have the most significant contribution to blocking crack propagation [21,46], and the remaining boundaries (θ < 2°) contribute to dislocation strengthening [47]. YS and grain size can be described by the Hall–Petch equation, as shown in Equation (4).
σ y = σ 0 * +   k hp d - 1 / 2
Here, σy is the YS, σ0* is the is the sum of other strengthening methods except the boundary strengthening, khp is the structural constant, and d is the grain size corresponding to MED defined by varied MTAs. It is worth noting that the microstructure is complex; under different Ej conditions, there are obvious differences in the size, morphology and distribution of CGHAZ microstructures. In this paper, with the change in Ej, the YS have been plotted as a function of the MED−1/2 with MTAs ranging from 2° to 12°, with the results shown in Figure 10. The results indicate that the MEDMTA≥2–6°−1/2 defined by MTAs ranging from 2° to 6° showed a fairly strong linear fit with YS, with the correlation factor of 0.99–0.96 (Figure 10a–c). However, Figure 10 d–f show the correlation factor was reduced to 0.82–0.76 for the YS and MEDMTA≥8–12° defined by MTAs ranging from 8° to 12°. Accordingly, the MED−1/2 defined by MTAs ranging from 2° to 6° had a higher correlation factor compared with MED−1/2 defined by MTAs ranging from 8° to 12°. This result made it clear that that MED defined by MTA ranging from 2° to 6°of LAGBs is the most efficient microstructural unit to controlling YS; this result is also consistent with that reported in related literature [28].
Furthermore, with the increase in Ej from 15 kJ/cm to 75 kJ/cm, the YR monotonically decreases, which means that an increase in Ej leads to an increase in strain-hardening ability. This phenomenon is mainly related to the area fraction and size of the M/A constituents. M/A constituents exist in ferrite matrixes as hard phases, and their hardnesses are higher than those of ferrite matrixes, and M/A constituents can be plastically deformed by two-stage or three-stage work hardening. This deformation mechanism is mainly controlled by high-strength local plastic deformation; the local internal stress applied from the M-A constituent hard phase acts on the ductile ferrite phase, thereby improving the strain-hardening ability [48,49]. Accordingly, with the increase in Ej from 15 to 75 kJ/cm, the dM/A increases from 0.98 μm to 1.81 μm, and fM/A increases from 3.11% to 4.42%, leading to an increase in the strain-hardening ability, and the YR monotonically decreases.
Furthermore, the uE increases significantly with the increase in Ej from 15 to 75 kJ/cm. The uE is closely related to slip of dislocation: once dislocations slip over long distances without encountering barriers, high uE is generated [50,51]. When the Ej = 15 and 35 kJ/cm, there are many DTs and ferrite lath boundaries in the microstructure, and the dislocation density is high (Figure 4a,b and Figure 7e,f). The dislocations will be significantly hindered after short-distance movement during the stretching process. While Ej = 55 and 75 kJ/cm, the proportion of PF in microstructure increases, relatively uniform dislocations are distributed in the microstructure, and the dislocation density is low (Figure 4c,d and Figure 7g,h), these are beneficial for long-distance slip of dislocations and obtaining higher uE. In addition, the corresponding characterization results of tensile fracture are shown in Figure 11. The tensile fractures under low Ej and large Ej are composed of two parts: fiber zone and shear-lip zone: the fiber zone is mainly composed of dimples, but deeper and smaller dimples appear at Ej = 75 kJ/cm. Further, although microvoids can be found in all fiber regions, the number and the size of microvoids are larger at Ej = 15 kJ/cm. Compared to deeper dimples, the appearance of large-sized microvoids indicates that microcracks are more likely to occur during the stretching process [52], which is not conducive to obtaining high uE. Accordingly, the characterizations of tensile fracture further demonstrate that increasing Ej can improve uE.

5. Conclusions

The effects of heat input on CGHAZ microstructure and tensile properties were studied for a novel V-N-Ti microalloyed weathering steel, and the following main conclusions can be drawn:
  • With the increase in Ej from 15 kJ/cm to 35 kJ/cm, and 55/75 kJ/cm, the dominant CGHAZ microstructure transforms from LBF to GBF, and then to a mixture of IGAF+PF. The size and area fraction of M/A constituents increase monotonically with the increase in Ej, the dM/A increases from 0.98 μm to 1.81 μm, and the fM/A increases from 3.11% to 4.42%.
  • With the increase in Ej, the increase in MED defined by the LAGBs and a reduction in dislocation density led to the decrease in YS. And the MED with MTAs ranging from 2°–6° strongly controlled the YS due to their higher fitting coefficient of the Hall–Petch relationship.
  • With the increase in Ej, the decrease in YR was mainly due to the increase in the size and area fraction of M/A constituents, which improved the strain-hardening ability. Meanwhile, the lower density and more uniform dislocation structures improved the uE.

Author Contributions

Q.W. (Qingfeng Wang) and B.H. conceived and designed the experiments; Q.W. (Qiuming Wang) and B.H. performed the experiments and analyzed the data; Q.W. (Qingfeng Wang) contributed the experimental material and analysis tools; and Q.W. (Qingfeng Wang) and B.H. wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (52127808), Innovation Ability Promotion Program of Hebei (22567609H), and National Key Research and Development Program of China (Grant No. 2017YFB0304800 and Grant No. 2017YFB0304802 for the second sub-project).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AcronymsFull Names
WSWeathering steel
CGHAZCoarse-grained heat-affected zone
EjHeat input
YSYield strength
TSTensile strength
YRYield strength ratio
uEUniform elongation
T8/5The time of cooling from 800 °C to 500 °C during the welding thermal cycle cooling stage
LAGBsLow-angle grain boundaries
HAGBsHigh-angle grain boundaries
MEDMean equivalent diameter
MTAsMisorientation tolerance angles
IPP softwareImage-Pro Plus software
EBSDElectron backscatter diffraction
dM/AThe average size of martensite/austenite constituents
fM/AArea fraction of martensite/austenite constituents
SAEDSelected area electron diffraction
LBFLath bainitic ferrite
GBFGranular bainitic ferrite
PFPolygon ferrite
IGAFIntragranular acicular ferrite
IGPFIntragranular polygon ferrite
M/A constituentsMartensite/austenite constituents
PAGBsPrior austenite grain boundaries
DTsDislocation tangles
UDDLsUniformly distributed dislocation lines
DLsDislocation lines
GNDsGeometrically necessary dislocations
KAMKernel average misorientation

References

  1. Wang, Z.; Zhang, X.; Cheng, L. Role of inclusion and microstructure on corrosion initiation and propagation of weathering steels in marine environment. J. Mater. Res. Technol. 2021, 10, 306–321. [Google Scholar] [CrossRef]
  2. Travassos, S.; Almeida, M.B.D.; Tomachuk, C. Non-destructive thickness measurement as a tool to evaluate the evolution of patina layer formed on weathering steel exposed to the atmosphere. J. Mater. Res. Technol. 2020, 9, 687–699. [Google Scholar] [CrossRef]
  3. Yamashita, M.; Nagano, H.; Misawa, T. Structure of protective rust layers formed on weathering steels by long-term exposure in the industrial atmospheres of Japan and North America. ISIJ Int. 1998, 38, 285–290. [Google Scholar] [CrossRef]
  4. Wang, D.; Zhong, Q.; Yang, J. Effects of Cr and Ni on the microstructure and corrosion resistance of high-strength low alloy steel. J. Mater. Res. Technol. 2023, 23, 36–52. [Google Scholar] [CrossRef]
  5. Morcillo, M.; Diaz, I.; Chico, B.; Cano, H. Weathering steels: From empirical development to scientific design. A review. Corros. Sci. 2014, 83, 6–31. [Google Scholar] [CrossRef]
  6. Prasad, S.N.; Mediratta, S.R.; Sarma, D.S. Influence of austenitisation temperature on the structure and properties of weather resistant steels. Mater. Sci. Eng. A 2003, 358, 288–297. [Google Scholar] [CrossRef]
  7. Amer, A.E.; Koo, M.Y.; Lee, K.H. Effect of welding heat input on microstructure and mechanical properties of simulated HAZ in Cu containing microalloyed steel. J. Mater. Sci. 2010, 45, 1248–1254. [Google Scholar] [CrossRef]
  8. Yang, X.; Di, X.; Liu, X.; Wang, D.; Li, C. Effects of heat input on microstructure and fracture toughness of simulated coarse-grained heat affected zone for HSLA steels. Mater. Charact. 2019, 155, 109818. [Google Scholar] [CrossRef]
  9. Cao, R.; Chan, Z.S.; Yuan, J.J. The effects of Silicon and Copper on microstructures, tensile and Charpy properties of weld metals by refined X120 wire. Mater. Sci. Eng. A 2018, 718, 350–362. [Google Scholar] [CrossRef]
  10. Jorge, J.; Souza, L.; Rebello, J. The effect of chromium on the microstructure/toughness relationship of C–Mn weld metal deposits. Mater. Charact. 2001, 47, 195–205. [Google Scholar] [CrossRef]
  11. Zhou, P.; Wang, B.; Liang, W. Effect of welding heat input on grain boundary evolution and toughness properties in CGHAZ of X90 pipeline steel. Mater. Sci. Eng. A 2018, 722, 112–121. [Google Scholar] [CrossRef]
  12. Bao, L.; Huang, Y.; Liu, F. Investigation on microstructure and impact toughness of simulated heat affected zone of high strength low alloy steels by laser-arc hybrid welding. Mater. Res. Express 2022, 9, 096519. [Google Scholar] [CrossRef]
  13. Wang, J.; Shen, Y.F.; Xue, W.Y. The significant impact of introducing nanosize precipitates and decreased effective grain size on retention of high toughness of simulated heat affected zone (HAZ). Mater. Sci. Eng. A 2021, 803, 140484. [Google Scholar] [CrossRef]
  14. Zhu, Z.; Liu, Y.; Gou, G. Effect of heat input on interfacial characterization of the butter joint of hot-rolling CP-Ti/Q235 bimetallic sheets by Laser+ CMT. Sci. Rep. 2021, 11, 10020. [Google Scholar] [CrossRef]
  15. Jia, X.; Chen, Y.; Li, H. Study on the mechanism of AF nucleation induced by complex oxide inclusions after LF refining in oxide metallurgical steel. Mater. Charact. 2023, 204, 113239. [Google Scholar] [CrossRef]
  16. Qi, H.; Pang, Q.; Li, W. In Situ observation of CGHAZ microstructure evolution of hot-rolled ship plate steel containing 0.012 yttrium under high-heat-input welding. Steel Res. Int. 2023, 98, 2200771. [Google Scholar] [CrossRef]
  17. Fadel, A.; Glisic, D.; Radovic, N.; Drobnjak, D. Intragranular ferrite morphologies in medium carbon vanadium-micro alloyed steel. J. Min. Metall. B 2013, 49, 237–244. [Google Scholar] [CrossRef]
  18. Medina, S.F.; Gómez, M.; Rancel, L. Grain refinement by intragranular nucleation of ferrite in a high nitrogen content vanadium microalloyed steel. Scr. Mater. 2008, 58, 1110–1113. [Google Scholar] [CrossRef]
  19. Zhang, J.; Xin, W.; Luo, G. Effect of welding heat input on microstructural evolution, precipitation behavior and resultant properties of the simulated CGHAZ in high-N V-alloyed steel. Mater. Charact. 2020, 162, 110201. [Google Scholar] [CrossRef]
  20. Hu, J.; Du, L.-X.; Wang, J.-J.; Xie, H.; Gao, C.-R.; Misra, R.D.K. High toughness in the intercritically reheated coarse-grained (ICRCG) heat-affected zone (HAZ) of low carbon microalloyed steel. Mater. Sci. Eng. A 2014, 590, 323–328. [Google Scholar] [CrossRef]
  21. Shi, Z.; Yang, C.; Wang, R.; Su, H.; Chai, F. Effect of nitrogen on the microstructures and mechanical properties in simulated CGHAZ of vanadium microalloyed steel varied with different heat inputs. Mater. Sci. Eng. A 2016, 651, 184. [Google Scholar] [CrossRef]
  22. Wang, L.; Fan, H.; Shi, G. Effect of ferritic morphology on yield strength of CGHAZ in a low carbon Mo-V-N-Ti-B steel. Metals 2021, 11, 1863. [Google Scholar] [CrossRef]
  23. Hu, B.; Shi, G.; Wang, Q. Elucidating the heat input on CGHAZ microstructure and its irregular effect on impact toughness for a novel V–N microalloying weathering steel. J. Mater. Res. Technol. 2023, 25, 5888–5906. [Google Scholar] [CrossRef]
  24. Wang, B.; Chen, N.; Cai, Y. Effect of crystallographic features on low-temperature fatigue ductile-to-brittle transition for simulated coarse-grained heat-affected zone of bainite steel weld. Int. J. Fatigue 2023, 170, 107523. [Google Scholar] [CrossRef]
  25. Zhang, X.; Gao, H.; Zhang, X. Effect of volume fraction of bainite on microstructure and mechanical properties of X80 pipeline steel with excellent deformability. Mater. Sci. Eng. A 2012, 531, 84–90. [Google Scholar] [CrossRef]
  26. Fu, Z.; Yang, B.; Shan, M. Hydrogen embrittlement behavior of SUS301L-MT stainless steel laser-arc hybrid welded joint localized zones. Corros. Sci. 2020, 164, 108337. [Google Scholar] [CrossRef]
  27. Guo, A.; Misra, R.D.K.; Liu, J. An analysis of the microstructure of the heat-affected zone of an ultra-low carbon and niobium-bearing acicular ferrite steel using EBSD and its relationship to mechanical properties. Mater. Sci. Eng. A 2010, 527, 6440–6448. [Google Scholar] [CrossRef]
  28. Fan, L.; Zhou, D.; Wang, T. Tensile properties of an acicular ferrite and martensite/austenite constituent steel with varying cooling rates. Mater. Sci. Eng. A 2014, 590, 224–231. [Google Scholar] [CrossRef]
  29. Zhu, K.; Bouaziz, O.; Oberbillig, C. An approach to define the effective lath size controlling yield strength of bainite. Mater. Sci. Eng. A 2010, 527, 6614–6619. [Google Scholar] [CrossRef]
  30. Olasolo, M.; Uranga, P.; Rodriguez-Ibabe, J.M. Effect of austenite microstructure and cooling rate on transformation characteristics in a low carbon Nb-V microalloyed steel. Mater. Sci. Eng. A 2011, 528, 2559–2569. [Google Scholar] [CrossRef]
  31. Li, C.; Duan, R.; Fu, W. Improvement of mechanical properties for low carbon ultra-high strength steel strengthened by Cu-rich multistructured precipitation via modification to bainite. Mater. Sci. Eng. A 2021, 817, 141337. [Google Scholar] [CrossRef]
  32. Williamson, G.K.; Smallman, R.E.J.P.M., III. Dislocation densities in some annealed and cold-worked metals from measurements on the X-ray debye-scherrer spectrum. Philos. Mag. A 1956, 1, 34–46. [Google Scholar] [CrossRef]
  33. Hu, J.; Du, X.; Wang, J.; Gao, R. Effect of welding heat input on microstructures and toughness in simulated CGHAZ of V-N high strength steel. Mater. Sci. Eng. A 2013, 577, 161–168. [Google Scholar] [CrossRef]
  34. Wang, X.L.; Wang, Z.Q.; Ma, X.P. Analysis of impact toughness scatter in simulated coarse-grained HAZ of E550 grade offshore engineering steel from the aspect of crystallographic structure. Mater. Charact. 2018, 140, 312–319. [Google Scholar] [CrossRef]
  35. Zhang, S.; Yang, M.; Yuan, F. Extraordinary fracture toughness in nickel induced by heterogeneous grain structure. Mater. Sci. Eng. A 2022, 830, 142313. [Google Scholar] [CrossRef]
  36. Wang, J.; Pan, Z.; Wang, Y. Evolution of crystallographic orientation, precipitation, phase transformation and mechanical properties realized by enhancing deposition current for dual-wire arc additive manufactured Ni-rich NiTi alloy. Addit. Manuf. 2020, 34, 101240. [Google Scholar] [CrossRef]
  37. Moeinifar, S.; Kokabi, A.H.; Hosseini, H. Influence of peak temperature during simulation and real thermal cycles on microstructure and fracture properties of the reheated zones. Mater. Des. 2010, 31, 2948–2955. [Google Scholar] [CrossRef]
  38. Baker, T.N. Microalloyed steels. Ironmak. Steelmak. 2016, 43, 264–307. [Google Scholar] [CrossRef]
  39. Tomita, Y.; Saito, N.; Tsuzuki, T. Improvement in HAZ toughness of steel by TiN-MnS addition. ISIJ Int. 1994, 34, 829–835. [Google Scholar] [CrossRef]
  40. Yang, L.; Peng, D.; Zhou, J. Effects of TiC on the microstructure and formation of acicular ferrite in ferritic stainless steel. Metall. Mater. 2019, 26, 1385–1395. [Google Scholar]
  41. Gregg, J.M.; Bhadeshia, H. Solid-state nucleation of acicular ferrite on minerals added to molten steel. Acta Mater. 1997, 45, 739–748. [Google Scholar] [CrossRef]
  42. Ishikawa, F.; Takahashi, T.; Ochi, T.J.M. Intragranular ferrite nucleation in medium-carbon vanadium steels. Metall. Mater. Trans. A 1994, 25, 929–936. [Google Scholar] [CrossRef]
  43. Tan, X.; Lu, W.; Guo, N. Effect of tempering and partitioning (T&P) treatment on microstructure and mechanical properties of a low-carbon low-alloy quenched and dynamically partitioned (Q-DP) steel. Mater. Sci. Eng. A 2023, 872, 144968. [Google Scholar]
  44. Sun, X.; Wang, Y.; Sun, D. Microstructure and mechanical properties of Si-rich H13 steel processed via austempering and tempering. Mater. Sci. Eng. A 2022, 834, 142616. [Google Scholar] [CrossRef]
  45. Iza-Mendia, A.; Gutierrez, I. Generalization of the existing relations between microstructure and yield stress from ferrite-pearlite to high strength steels. Mater. Sci. Eng. A 2013, 561, 40–51. [Google Scholar] [CrossRef]
  46. Chen, J.; Shuai, T.; Liu, Z.Y. Microstructural characteristics with various cooling paths and the mechanism of embrittlement and toughening in low-carbon high performance bridge steel. Mater. Sci. Eng. A 2013, 559, 241–249. [Google Scholar] [CrossRef]
  47. Gutierrez, I. Effect of microstructure on the impact toughness of Nb-microalloyed steel: Generalisation of existing relations from ferrite-pearlite to high strength microstructures. Mater. Sci. Eng. A 2013, 571, 57–67. [Google Scholar] [CrossRef]
  48. Hüper, T.; Endo, S.; Ishikawa, N. Effect of volume fraction of constituent phases on the stress-strain relationship of dual phase steels. ISIJ Int. 2007, 39, 288–294. [Google Scholar] [CrossRef]
  49. Li, Z.; Chen, Y.; Yuan, H. Microstructure and properties of low alloy Cr-Mo steel processed with varied isothermal temperatures. Mater. Today Commun. 2023, 35, 106158. [Google Scholar] [CrossRef]
  50. Sung, H.K.; Lee, D.H.; Lee, S. Correlation between microstructures and tensile properties of strain-based API X60 pipeline steels. Metall. Mater. Trans. A 2016, 47, 2726–2738. [Google Scholar] [CrossRef]
  51. Bandhu, D.; Vora, J.J.; Das, S. Experimental study on application of gas metal arc welding based regulated metal deposition technique for low alloy steel. Mater. Manuf. Process. 2022, 37, 1727–1745. [Google Scholar] [CrossRef]
  52. Noell, P.; Carroll, J.; Hattar, K. Do voids nucleate at grain boundaries during ductile rupture? Acta Mater. 2017, 137, 103–114. [Google Scholar] [CrossRef]
Figure 1. Welding thermal simulation diagram of the round bars: (a), the welding thermal cycle curves with different heat input (b) [23], microstructure observation and micro-tensile specimens taken from the thermocouple region (c).
Figure 1. Welding thermal simulation diagram of the round bars: (a), the welding thermal cycle curves with different heat input (b) [23], microstructure observation and micro-tensile specimens taken from the thermocouple region (c).
Metals 13 01607 g001
Figure 2. Optical micrographs of the simulated CGHAZ at Ej = 15 kJ/cm (a), 35 kJ/cm (b), 55 kJ/cm (c), and 75 kJ/cm (d). PF—intragranular polygon ferrite (IGPF) and/or grain boundary polygonal ferrite (GBPF).
Figure 2. Optical micrographs of the simulated CGHAZ at Ej = 15 kJ/cm (a), 35 kJ/cm (b), 55 kJ/cm (c), and 75 kJ/cm (d). PF—intragranular polygon ferrite (IGPF) and/or grain boundary polygonal ferrite (GBPF).
Metals 13 01607 g002
Figure 3. The observation of M/A constituents in CGHAZ after LePera’s agent corrosion, Ej = 15 kJ/cm (a), 35 kJ/cm (b), 55 kJ/cm (c), and 75 kJ/cm (d).
Figure 3. The observation of M/A constituents in CGHAZ after LePera’s agent corrosion, Ej = 15 kJ/cm (a), 35 kJ/cm (b), 55 kJ/cm (c), and 75 kJ/cm (d).
Metals 13 01607 g003
Figure 4. Observation of CGHAZ ferrite and M/A constituents via TEM at Ej = 15 kJ/cm (a,a1), 35 kJ/cm (b,b1), 55 kJ/cm (c,c1), and 75 kJ/cm (d,d1). The bright field (e) and dark field (e1) of M/A constituent at Ej = 15 kJ/cm and SAED pattern of M/A constituent (e2).
Figure 4. Observation of CGHAZ ferrite and M/A constituents via TEM at Ej = 15 kJ/cm (a,a1), 35 kJ/cm (b,b1), 55 kJ/cm (c,c1), and 75 kJ/cm (d,d1). The bright field (e) and dark field (e1) of M/A constituent at Ej = 15 kJ/cm and SAED pattern of M/A constituent (e2).
Metals 13 01607 g004
Figure 5. XRD patterns (a) and dislocation density of specimens after heat cycle with different Ej (b).
Figure 5. XRD patterns (a) and dislocation density of specimens after heat cycle with different Ej (b).
Metals 13 01607 g005
Figure 6. TEM observation results of precipitated particles at Ej = 15 kJ/cm (a), 35 kJ/cm (b), 55 kJ/cm (c), 75 kJ/cm (d), the inset images (IIV) in (ad) show the TEM-EDS of particles.
Figure 6. TEM observation results of precipitated particles at Ej = 15 kJ/cm (a), 35 kJ/cm (b), 55 kJ/cm (c), 75 kJ/cm (d), the inset images (IIV) in (ad) show the TEM-EDS of particles.
Metals 13 01607 g006
Figure 7. The inverse pole figure maps (ad) and the corresponding KAM maps (eh) of CGHAZ, Ej = 15 kJ/cm (a,e), 35 kJ/cm (b,f), 55 kJ/cm (c,g), and 75 kJ/cm (d,h), and the inset images in (eh) show the KAMave value of each Ej sample.
Figure 7. The inverse pole figure maps (ad) and the corresponding KAM maps (eh) of CGHAZ, Ej = 15 kJ/cm (a,e), 35 kJ/cm (b,f), 55 kJ/cm (c,g), and 75 kJ/cm (d,h), and the inset images in (eh) show the KAMave value of each Ej sample.
Metals 13 01607 g007
Figure 8. The typical stress-strain curves for each Ej sample (a) and statistical chart of CGHAZ tensile performance testing results (b).
Figure 8. The typical stress-strain curves for each Ej sample (a) and statistical chart of CGHAZ tensile performance testing results (b).
Metals 13 01607 g008
Figure 9. Observation of precipitated particles in CGHAZ microstructures at Ej = 35 kJ/cm (a) and the corresponding EDS analysis of particles (I); observation of precipitated particles in CGHAZ microstructures at Ej = 75 kJ/cm (b), and the corresponding EDS analysis of particles (II).
Figure 9. Observation of precipitated particles in CGHAZ microstructures at Ej = 35 kJ/cm (a) and the corresponding EDS analysis of particles (I); observation of precipitated particles in CGHAZ microstructures at Ej = 75 kJ/cm (b), and the corresponding EDS analysis of particles (II).
Metals 13 01607 g009
Figure 10. The YS as a function of the MED−1/2 that defined by MTA of 2° (a); the YS as a function of the MED−1/2 that defined by MTA of 4° (b); the YS as a function of the MED−1/2 that defined by MTA of 6° (c), the YS as a function of the MED−1/2 that defined by MTA of 8° (d), the YS as a function of the MED−1/2 that defined by MTA of 10° (e), and the YS as a function of the MED−1/2 that defined by MTA of 12° (f).
Figure 10. The YS as a function of the MED−1/2 that defined by MTA of 2° (a); the YS as a function of the MED−1/2 that defined by MTA of 4° (b); the YS as a function of the MED−1/2 that defined by MTA of 6° (c), the YS as a function of the MED−1/2 that defined by MTA of 8° (d), the YS as a function of the MED−1/2 that defined by MTA of 10° (e), and the YS as a function of the MED−1/2 that defined by MTA of 12° (f).
Metals 13 01607 g010
Figure 11. Typical observations of tensile fracture morphology at Ej = 15 kJ/cm (aa2) and Ej = 75 kJ/cm (bb2).
Figure 11. Typical observations of tensile fracture morphology at Ej = 15 kJ/cm (aa2) and Ej = 75 kJ/cm (bb2).
Metals 13 01607 g011
Table 1. Chemical composition of experimental steel (wt. %) [23].
Table 1. Chemical composition of experimental steel (wt. %) [23].
CSiMnPSCrCuNiMoTiVNFe
0.0460.191.300.0090.0050.450.320.370.050.0150.0240.0073Bal.
Table 2. Mechanical properties of experimental steel [23].
Table 2. Mechanical properties of experimental steel [23].
YS
ReL/MPa
TS
Rm/MPa
YR
ReL/Rm
Elongation
A/%
Energy Absorbed at
−40 °C AkV/J
Micro-Hardness
/HV10
4895940.8226.5238.4226.0
Table 3. The atomic fraction of each chemical element in particles (wt %).
Table 3. The atomic fraction of each chemical element in particles (wt %).
Ej kJ/cmType of Element in Square ParticlesType of Element in Circular Particles
TiVCNFeVCNFe
1523.15%0%41.81%24.15%10.89%////
3523.19%0.05%42.43%24.89%10.44%0.02%60.15%5.89%33.94%
5524.55%0.95%43.05%26.15%5.30%0.88%61.89%10.25%26.98%
7525.11%1.60%43.10%27.33%2.86%1.45%62.82%12.63%23.10%
Table 4. MED of grains with boundaries at MTAs from 2° to 15°.
Table 4. MED of grains with boundaries at MTAs from 2° to 15°.
Ej
kJ/cm
MEDMTA≥2°
/μm
MEDMTA≥4°
/μm
MEDMTA≥6°
/μm
MEDMTA≥8°
/μm
MEDMTA≥10°
/μm
MEDMTA≥12°
/μm
MEDMTA≥15°
/μm
152.753.153.454.155.215.476.36
353.163.64.134.965.625.766.73
554.034.545.125.675.795.846.15
754.695.565.685.765.875.915.96
MED—Mean equivalent diameter; MTAs—Misorientation tolerance angles; MEDMTA≥θ—The MED of grains with MTAs higher than θ.
Table 5. Summary of the tensile results at different Ej.
Table 5. Summary of the tensile results at different Ej.
Heat Input/kJ·cm−1YS/MPaTS/MPaYRuE/%Micro-Hardness
/HV10
15480 ± 5615 ± 40.789.5 ± 0.5238.4 ± 3.5
35461 ± 6605 ± 30.7613.2 ± 0.7222 ± 5.2
55437 ± 6585 ± 30.7516.4 ± 0.7210 ± 4.8
75416 ± 5574 ± 40.7218.6 ± 0.6201 ± 3.5
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hu, B.; Wang, Q.; Wang, Q. Effect of Heat Input on Microstructure and Tensile Properties in Simulated CGHAZ of a V-Ti-N Microalloyed Weathering Steel. Metals 2023, 13, 1607. https://doi.org/10.3390/met13091607

AMA Style

Hu B, Wang Q, Wang Q. Effect of Heat Input on Microstructure and Tensile Properties in Simulated CGHAZ of a V-Ti-N Microalloyed Weathering Steel. Metals. 2023; 13(9):1607. https://doi.org/10.3390/met13091607

Chicago/Turabian Style

Hu, Bing, Qiuming Wang, and Qingfeng Wang. 2023. "Effect of Heat Input on Microstructure and Tensile Properties in Simulated CGHAZ of a V-Ti-N Microalloyed Weathering Steel" Metals 13, no. 9: 1607. https://doi.org/10.3390/met13091607

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