Processing math: 5%
Tingting Li, Jian Yang, Yinhui Zhang, Han Sun, Yanli Chen, and Yuqi Zhang, Effect of Al content on nanoprecipitates, austenite grain growth and toughness in coarse-grained heat-affected zones of Al–Ti–Ca deoxidized shipbuilding steels, Int. J. Miner. Metall. Mater., 32(2025), No. 4, pp.879-891. https://doi.org/10.1007/s12613-024-2967-8
Cite this article as: Tingting Li, Jian Yang, Yinhui Zhang, Han Sun, Yanli Chen, and Yuqi Zhang, Effect of Al content on nanoprecipitates, austenite grain growth and toughness in coarse-grained heat-affected zones of Al–Ti–Ca deoxidized shipbuilding steels, Int. J. Miner. Metall. Mater., 32(2025), No. 4, pp.879-891. https://doi.org/10.1007/s12613-024-2967-8
Research Article

Effect of Al content on nanoprecipitates, austenite grain growth and toughness in coarse-grained heat-affected zones of Al–Ti–Ca deoxidized shipbuilding steels

Author Affilications
  • Corresponding author:

    Jian Yang      E-mail: yang_jian@t.shu.edu.cn

  • This work focuses on the influence of Al content on the precipitation of nanoprecipitates, growth of prior austenite grains (PAGs), and impact toughness in simulated coarse-grained heat-affected zones (CGHAZs) of two experimental shipbuilding steels after being subjected to high-heat input welding at 400 kJ·cm−1. The base metals (BMs) of both steels contained three types of precipitates: Type I: cubic (Ti,Nb)(C,N), Type II: precipitate with cubic (Ti,Nb)(C,N) core and Nb-rich cap, and Type III: ellipsoidal Nb-rich precipitate. In the BM of 60Al and 160Al steels, the number densities of the precipitates were 11.37 × 105 and 13.88 × 105 mm−2, respectively. The 60Al and 160Al steel contained 38.12% and 6.39% Type III precipitates, respectively. The difference in the content of Type III precipitates in the 60Al steel reduced the pinning effect at the elevated temperature of the CGHAZ, which facilitated the growth of PAGs. The average PAG sizes in the CGHAZ of the 60Al and 160Al steels were 189.73 and 174.7 µm, respectively. In the 60Al steel, the low lattice mismatch among Cu2S, TiN, and γ-Al2O3 facilitated the precipitation of Cu2S and TiN onto γ-Al2O3 during welding, which decreased the number density of independently precipitated (Ti,Nb)(C,N) particles but increased that of γ-Al2O3–TiN–Cu2S particles. Thus, abnormally large PAGs formed in the CGHAZ of the 60Al steel, and they reached a maximum size of 1 mm. These PAGs greatly reduced the microstructural homogeneity and consequently decreased the impact toughness from 134 (0.016wt% Al) to 54 J (0.006wt% Al) at −40°C.

  • Welding efficiency has been improved by high-heat input welding (HHIW) technologies, which can attain a heat input over 400 kJ·cm−1. The applications of HHIW technologies include shipbuilding, construction, machinery, and other fields [12]. However, with increased heat input, the peak temperature of coarse-grained heat-affected zones (CGHAZs) rises, and the exposure to high temperatures is prolonged, resulting in brittle microstructures, coarse grains, and deteriorated CGHAZ toughness [35]. Oxide metallurgy with Ca addition can be potentially used to improve the impact toughness of CGHAZ through promoting the formation of intragranular acicular ferrite (IAF) and refinement of prior austenite grains (PAGs) [610], which has been confirmed by JFE Steel [11] and Kobe Steel [12]. We previously showed that the number density (ND) of complex oxysulfides increases with the increase in the Ca content, which promotes IAF nucleation and the precipitation of nanosized TiN particles and considerably improves toughness in the CGHAZ [6,8,13].

    The size and distribution of austenite contribute to the toughness of CGHAZ, with large PAGs promoting the formation of brittle microstructures and inhomogeneous microstructure distribution, both of which negatively affect toughness [1415]. During HHIW, PAG growth suffers from the inhibition caused by the dragging force of the alloying elements segregated at grain boundaries (solute drag) and the pinning force of fine particles, with the latter exhibiting a notably stronger effect [1618]. These processes, such as particle dissolution and reprecipitation related to PAG growth, cause substantial alterations in the chemistry, size distribution, and volume fraction of precipitates during the welding process [1920]. Especially under HHIW, the CGHAZ experiences prolonged exposure to elevated temperatures, which exacerbates the dissolution or coarsening of pinning particles.

    Given the high thermal stability of fine TiN precipitates at high temperatures, dispersed TiN precipitates are used to pin grain boundaries and control the size of PAG in the CGHAZ of thick steel plates under HHIW [2122]. The addition of microalloying elements, such as Nb, V, B, and Mo, considerably affects PAG growth, microstructural evolution, and impact toughness of the CGHAZ of alloyed steels as a result of the influence of microalloying elements on the precipitation and thermal stability of TiN. Wu et al. [23] reported that the increase in the Mo content of steels not only increases the ND but also reduces the size of precipitates, which favors the improvement of impact toughness. Our previous work also reported that with the increase in Mo content, the concentration of Mo in Ti(C,N) particles and the thermal stability of TiN precipitates in high-Mo steel increase [24]. However, the addition of Nb results in coarse TiN precipitates in the CGHAZ of Mg/Ca-deoxidized shipbuilding steels and decreases the thermal stability of TiN due to the interaction between Ti and Nb atoms [1,25]. Pan et al. [26] demonstrated that the decrease in the Al content caused the refinement of precipitates in the CGHAZ of thick steel plates, which enabled the faster dissolution of the precipitates and growth of PAGs at a high rate during welding.

    The very strong chemical attraction of Al to oxygen allows its frequent use as a deoxidizer in the steel deoxidation process. In the CGHAZ of Ca-treated low-alloy steels with a low Al content, effective inclusions and IAFs exhibit a higher likelihood of formation than steels with a high Al content [2728]. Our previous work indicated that during HHIW, the microstructures of the CGHAZ of Ca-treated thick steel plates changed from IAF to bainitic ferrite with the increase in the Al content, and this condition resulted in a decrease in impact toughness [29].

    Studies rarely focused on the evolution of nanoprecipitates in the CGHAZ of Al–Ti–Ca-deoxidized shipbuilding steels with various Al contents. At a low Al content, Ti and Ca are likely to participate in the deoxidation process, which affects not only the development of inclusions and microstructures but also the nucleation of nanoprecipitates. In this work, experiments and theoretical calculations are performed to investigate the influence of Al content on the nanoprecipitates, PAG growth, and impact toughness in CGHAZ. Therefore, the mechanism of toughness improvement in Al–Ti–Ca-deoxidized shipbuilding steels with various Al contents is clarified.

    The studied steels contained various amounts of Al and were prepared via a 50-kg vacuum induction melting furnace. Argon gas was blown into the furnace during the melting process. The melt was added with appropriate amounts of deoxidizers and alloying elements to achieve the desired steel composition. The melt was then cast into ingots (120 mm × 180 mm × 240 mm). Table 1 lists the chemical compositions of the studied steels analyzed through spark emission spectrometry (ARL 4460, Thermo Fisher, USA) and inductively coupled plasma atomic emission spectrometry (ICP-AES). The experimental steels were designated 60Al and 160Al based on their Al contents. The thermomechanical control process (TMCP) was performed to hot roll ingots at a thickness of 50 mm. The roughing and finishing rolling temperatures were above 930 and 770°C, respectively, with reduction ratios over 30%. Eventual cooling down of the steel plates was achieved from 750 to 400°C at 10°C·s−1.

    Table  1.  Chemical compositions of shipbuilding steels investigated in the present study wt%
    Steel C Si Mn P S Nb Ni Cu Al Ti Ca O N Cr
    60Al 0.075 0.21 1.6 0.0070 0.0050 0.015 0.36 0.10 0.0060 0.0090 0.0017 0.0028 0.0033 0.013
    160Al 0.080 0.24 1.6 0.0070 0.0050 0.017 0.35 0.10 0.016 0.0085 0.0022 0.0024 0.0040 0.013
     | Show Table
    DownLoad: CSV

    The welding simulation involved obtaining 11 mm × 11 mm × 71 mm specimens obtained at 1/4 of the thickness and 1/4 of the width of the as-rolled steel plates (Fig. 1(a)). HHIW thermal cycle of the specimens was performed at 400 kJ·cm−1 using a Glebble-3800 simulator (Dynamic Systems Inc., Poestenkill, NY, USA). The thermal cycle, represented by the red line in Fig. 1(b), was calculated using the Rykalin 3D model embedded in the welding software.

    Fig. 1.  (a) Heating and cooling curves of CGHAZ simulation and in situ observation, (b) sample used in the simulated HHIW experiment, and (c) sample used in the in situ observation experiment.

    In situ observations of the growth behavior of PAGs were conducted via HTLSCM. After machining the samples from the TMCP plates, they were mechanically polished into metallographic specimens. The samples measured 7.5 mm in diameter and 3 mm in height (Fig. 1(c)). During simulation, the samples were heated to 1400°C, held at this peak temperature (Tp1) for 3 s, and then cooled sequentially to 800 and 500°C at the cooling rates of 3.41 and 0.78°C·s−1, respectively (blue line in Fig. 1(b)). A cooling time of 385 s was observed from 800 to 500°C (t8/5). Afterward, the standard V-notch specimens with dimensions of 10 mm × 10 mm × 55 mm were prepared for impact testing. The ASTM E23 standard was used to test the specimens at −40°C. Each impact toughness value was calculated from three test findings.

    Optical microscopy (DM2700M, Leica Microsystems) was performed for the observation of austenite grains in the simulated CGHAZ and HTLSCM samples. The metallographic specimens were polished and etched in a 4vol% nital solution for approximately 15 s. The average size of the PAGs was determined through measurement of more than 300 grains across more than 10 photographs at 50× magnification.

    Replica studies and twin-jet electropolishing experiments were performed to analyze the morphology and size of nanoscale precipitates [1,25]. The precipitates on carbon replicas were investigated through transmission electron microscopy (TEM, JEM-2010F, Japan) at an accelerating voltage of 120 kV. The chemistry and crystallographic characteristics of the precipitates were analyzed via energy-dispersive X-ray spectroscopy (EDS, Oxford, UK) and the selected area diffraction pattern (SADP). To study the size distribution and ND, we measured more than 1000 precipitates per unit area of each sample using Image Pro Plus. Each sample underwent a replica study and precipitate analysis three times.

    Nonaqueous electrolytic experiments were performed to determine the soluble Nb and Al contents in the matrix. Electrolytic experiments were conducted in a solution containing 2 mL triethanolamine + 1 g tetramethyl-ammonium chloride-methanol + 98 mL methanol with a current density of 10 mA·cm−2 at approximately 0°C. After its filtration using a polycarbonate filter membrane, the electrolyte was subjected to microwave digestion and analyzed via inductively coupled plasma (ICP) atomic emission spectroscopy (Agilent 5800 ICP-OES, USA). The nonaqueous electrolysis and ICP experiments were repeated twice for each sample, and the findings were averaged.

    Table 2 provides a summary of the mechanical properties of the base metals (BMs) and CGHAZs. The simulated CGHAZ of 160Al steel had a Charpy impact energy of 134 J, which is higher than that of 60Al steel (54 J); this finding indicates that the addition of an appropriate amount of Al favored improvement of the CGHAZ impact toughness. Supplementary Material contains further details on the fracture morphology analysis results.

    Table  2.  Mechanical properties of BM and CGHAZ of the two steels
    Steel BM CGHAZ
    Transverse yield strength / MPa Longitudinal yield strength / MPa Transverse tensile strength / MPa Longitudinal tensile strength / MPa Transverse toughness / J Longitudinal toughness / J Toughness / J
    60Al 478 464 595 575 160 226 54 ± 25
    160Al 540 544 643 645 151 239 134 ± 13
     | Show Table
    DownLoad: CSV

    Fig. 2 shows the PAG growth behavior at various temperatures during the thermal cycle. Fig. 2(a), (b), (d), and (e) reveals that a notable acceleration in PAG growth occurred during heating from 1300 to 1400°C and during cooling. Abnormal-growth PAGs were observed on the 60Al steel, whereas the PAG size was more uniform in 160Al steel (Fig. 2(c) and (f)). The abnormally grown PAG contained more bainite than normal grains. The relative difference (RD) described by Eq. (1) [30] can be used to determine abnormally large grains.

    Fig. 2.  SEM images of PAGs in (a) 60Al and (d) 160Al steels after heating to 1300°C, (b) 60Al and (e) 160Al steels after cooling to 900°C, and (c) 60Al and (f) 160Al steels after cooling to room temperature. (g) Average PAG size of both steels at various temperatures; (h) size distribution of PAGs after cooling to room temperature.
    RD=GSabnormal – GSnormalGSnormal (1)

    where GSabnormal and GSnormal denote the sizes of abnormally large and normal grains, respectively. If RD exceeded 0.9, the grains were considered to have abnormal growth [30]. In the 60Al steel, the largest grain was approximately 1 mm in diameter, which is almost twice the size of the largest normal grains. The abnormal grain growth in the 60Al steel substantially reduced the toughness of CGHAZ, and such a result can be attributed to the decreased pinning effect of the precipitates on grain boundaries.

    Measurement and statistical analysis of the equivalent diameters of PAGs at various temperatures of the 60Al and 160Al steels were performed (Fig. 2(g)). Both steels exhibited slow-growing PAGs until a temperature of 1300°C. When the temperature was increased to 1100°C, PAGs in the 60Al steel attained an average size of 17.13 µm, which is slightly larger than that in the 160Al steel (16.37 µm). As the temperature increased to 1300°C, the average sizes of PAGs in the 60Al and 160Al steels were 47.18 and 44.82 µm, respectively, and the average sizes of PAGs in both steels were coarsened by about 1.75 times. When the temperature dropped below 1100°C, no evident changes in the PAG size occurred. The 60Al and 160Al steels achieved average PAG sizes of 189.73 and 174.70 µm, respectively, after cooling to room temperature. PAGs in both steels presented similar size distributions at room temperature when PAGs were smaller than 450 μm (Fig. 2(h)).

    Based on the morphologies and compositions of precipitates in the BMs, the precipitates were categorized into ellipsoidal, mixed morphology, and cubic. Fig. 3 illustrates the morphologies and SADP and EDS analysis results on the three types of precipitates. All three were found in the BMs of both steels. Type I precipitate has a cubic structure, Type II precipitate comprises a cubic precipitate as a core with caps on its faces (Fig. 3(b)), and Type III precipitate possesses an ellipsoidal shape that is finer than those of Types I and II precipitates and contains a higher Nb atomic ratio (Fig. 3(c)). The EDS analysis revealed that all three particle types consisted of (Ti,Nb)(C,N). However, the proportions of Ti and Nb varied within (Ti,Nb)(C,N) precipitates of various morphologies. The Type I precipitates attained a lattice parameter of 4.24 Å, which is close to that of TiN (lattice parameter = 4.239 Å) [31]. In the case of Type II precipitate, lattice parameters of 4.26 and 4.38 Å were observed for the core and cap, respectively (Fig. 3(b)). The lattice parameter for the cap was closer to that of NbC (lattice parameter = 4.47 Å) and larger than that of the core; this result indicates a higher Nb content in the cap, in agreement with the results by Moon et al. [31]. The precipitates exhibited an ellipsoidal shape (Type III) when Nb(C,N) precipitated or distributed at the periphery of Ti(C,N), with a lattice parameter equal to 4.58 Å. With the increase in the Nb content of the precipitates, the lattice parameters of the precipitates increase correspondingly.

    Fig. 3.  Bright-field TEM micrographs patterns of three types of precipitates in the carbon extraction replicas of the BM before the thermal cycle, associated with SADPs and EDS analyses: (a) cubic (Ti,Nb)(C,N) precipitate, Type I precipitate; (b) cubic (Ti,Nb)(C,N) core with Nb-rich cap, Type II precipitate; (c) ellipsoidal Nb-rich precipitate, Type III precipitate.

    The CGHAZ of the two steels mostly contained Types I and II precipitates. However, the CGHAZ of the 60Al steel exhibited many irregular particles exceeding 80 nm. Several irregular particles were selected for the composition and structural analysis (Fig. 4). Fig. 4(a)–(c) displays the bright-field micrographs and high-angle annular dark field (HADDF) micrographs of irregular particles showing the growth of square particles from the substrates. Fig. 4(d)–(h) displays the SADP pattern, darkfield images, and element mappings for the irregular particles. The results presented in element mapping indicate that the particles comprised alumina inclusion (white line in Fig. 4(g)), Cu sulfide (Fig. 4(f)), and Ti–Nb carbonitride (red line in Fig. 4(g)). SADP analysis revealed the cubic structure of the different parts of composite particles. Cu sulfide had a lattice parameter of 5.62 Å, which suggests that Cu sulfide is Cu2S (lattice parameter a = 5.55 Å) [32]. Ti–Nb carbonitride had a lattice parameter of 4.32 Å, which is close to that of TiN. The alumina inclusions detected in the CGHAZ of the 60Al steel presented a lattice parameter of 7.87 Å. These inclusions were assumed to include γ-Al2O3 (lattice parameter = 7.94 Å) due to their cubic structure [33]. The diffraction patterns showed the cube–cube orientation relationship between Cu2S and (Ti,Nb)(C,N) and the cube-on-edge relationship observed between γ-Al2O3 and Cu2S/(Ti,Nb)(C,N).

    Fig. 4.  TEM analysis of irregular precipitates in the CGHAZ of the 60Al steel after the HHHIW thermal cycle: (a, b) bright-field micrographs and (c) HADDF micrograph of Al2O3-based precipitates; (d) bright-field micrographs with the high-resolution TEM micrograph of the precipitate. (e) Corresponding diffraction of incident beam direction parallel to [001] {}_{{\bf {Cu}}_{\boldsymbol{2}} {\bf {S}}} //[001](Ti,Nb)(C,N)//[114]γ {}_{-{\bf{Al}}_2{{\bf{O}}_{\bf{3}}}} , (f, g) dark-field images obtained from the precipitate, and (h) element mappings of the precipitate.

    Fig. 5 revealed the cluster and dispersed distributions in the BM and CGHAZ. The characteristic of cluster distribution is that the precipitates belongs to a group or is in contact with each other (Fig. 5(a)). Notably, most of the precipitates that formed clusters were Type III precipitates. This pattern was commonly observed in the BM of the 60Al steel but was uncommon in the 160Al steel. The precipitates were predominantly dispersed Types I and II precipitates in the CGHAZ of both steels.

    Fig. 5.  TEM images of the replica showing precipitates in the (a, b) BM and (c, d) CGHAZ after the HHIW thermal cycle: (a, c) 60Al and (b, d) 160Al steel.

    Fig. 6 displays the TEM micrographs of precipitates in the CGHAZ of the 60Al steel. The coarsened Type I and Nb-rich (Type III) precipitates aggregated within the grain and showed no pinning effect on PAG boundaries (Fig. 6(a)). Fig. 6(b) demonstrates the distribution and elemental mappings of Al2O3-based particles in the CGHAZ, with most of the coarsened particles contained within the grains. For the three particles distributed at the grain boundaries, P1 exhibited no pinning effect, and P2 demonstrated a relatively evident pinning effect on grain boundaries, followed by P3. This observation aligns with the findings, implying a strong pinning effect on small precipitates [34].

    Fig. 6.  TEM images of thin foils revealed precipitates in the CGHAZ of the 160Al steel: (a) coarsened Types I and III precipitates that aggregated within the grains; (b) dispersed composite precipitates along the boundaries and elemental mappings.

    Table 3 lists the changes in average size, ND, and proportion of precipitates observed from the BM to the CGHAZ of the two steels. The precipitates had an average size of 16.29 nm in the BM of the 60Al steel and 13.48 nm in 160Al steel. After being subjected to HHIW, both steels experience a substantial increase in the precipitate size, with an average of 58.91 nm for 60Al steel and 44.64 nm for 160Al steel. In the BM of both steels, the precipitates attained NDs of 11.37 × 105 and 13.88 × 105 mm−2, respectively. In the CGHAZ, the NDs decreased substantially to 6.12 × 105 mm−2 in the 60Al steel and to 9.28 × 105 mm−2 in the 160Al steel, respectively. In addition, Type III precipitates accounted for 38.12% of the total in the BM of the 60Al steel and 6.39% of that in the 160Al steel. Therefore, the BM of the 160Al steel were mostly Types I and II precipitates, with far fewer Type III precipitates than those in the 60Al steel.

    Table  3.  Average size, ND and proportion of various precipitates in the BMs and CGHAZs of both steels
    SteelsBMCGHAZ
    Average size / nmND / (105 mm−2)Proportion of Type III
    precipitates / %
    Average size / nmND / (105 mm−2)Proportion of Al2O3-
    based precipitates / %
    60Al16.2911.3738.1258.916.1229.78
    160Al13.4813.886.3944.649.280.00
     | Show Table
    DownLoad: CSV

    Fig. 7 illustrates the distribution of precipitate diameters for both steels. The BM of both steels showed unimodal and similar distributions of precipitates over the entire diameter range (Fig. 7(a)). After welding, a bimodal precipitate diameter distribution was observed in the CGHAZ of the 60Al steel. Peak 2 (Fig. 7(b)) in the CGHAZ of the 60Al steel was caused by the Al2O3-based precipitates, which accounted for 30.93% of the total in the CGHAZ (Table 3). Excluding the Al2O3-based precipitates, pure carbonitrides in the CGHAZ of the 60Al steel had an average size of 46.21 nm, similar to that of the 160Al steel. However, the ND of precipitates in the CGHAZ of 60Al steel (6.12 × 105 mm−2) decreased more substantially than that of the 160Al steel (9.28 × 105 mm−2) after welding.

    Fig. 7.  Distribution of precipitate diameters observed in the (a) BM and (b) CGHAZ of 60Al and 160Al steels.

    Table 4 lists the elemental concentrations measured by ICP in the matrix and precipitates. The Nb content in the precipitates exhibited a decrease from 0.010wt% in the 60Al steel to 0.0054wt% in the 160Al steel, which suggests that the increase in Al content inhibited the precipitation of Nb-rich precipitates in the steel, which is the same trend as the results displayed in Fig. 7. The inclusions of both steels contained very close amounts of Al.

    Table  4.  Element concent in the matrix and precipitates wt%
    Steel Nb Al
    Matrix Precipitates Matrix Oxides
    60Al 0.0046 0.010 0.0053 0.00068
    160Al 0.011 0.0054 0.015 0.00084
     | Show Table
    DownLoad: CSV

    Factsage 8.1 software was used to perform equilibrium calculations for the inclusions and carbonitride precipitates (Fig. 8). When the temperature decreased from 1500 to 800°C, both steels revealed the formation of various inclusions, such as perovskite (Ca2Ti2O(5,6)), Al2O3, CaO·6Al2O3, sulfides (MnS and CaS), and carbonitride precipitates (Fig. 8(a)). In the 60Al steel, the perovskite (green solid line) transformed to carbonitrides (blue solid line) and Al2O3 (cyan solid line) at 1388°C, and the CaO·6Al2O3 achieved the stable phase from 1080 to 1237°C. With the increase in the Al content, the predicted phase changed in the following manner: (1) The temperature at which perovskite (green dashed line) transformed into carbonitrides (blue dashed line) and calcium aluminates (red dashed line) increased to 1438°C. (2) The Al2O3 became the stable phase at temperatures lower than 1264°C, whereas CaO·6Al2O3 was the stable phase at temperatures above this threshold.

    Fig. 8.  Thermodynamic calculation results: (a) phase transformation in two steels during solidification; (b) changes in the specific compositions of carbonitrides (blue carbonitrides in (a)) with temperature.

    Fig. 8(b) shows that the TiN particles were the main phase of carbonitrides during the decrease in the temperature from the peak temperature to 1000°C. This finding suggests that the TiN precipitates exerted a pinning effect at high CGHAZ temperatures, and the Nb-rich particles exhibited poor high-temperature stability and thus cannot pin PAG boundaries. As the TiN content in the 60Al steel was lower than that in the 160Al steel at temperatures above 1100°C, the precipitates in 60Al exhibited a weaker pinning effect on grain boundaries compared with those in the 160Al steel. This finding indicates a faster growth rate of normal PAGs in the 60Al steel than in the 160Al steel. Based on the formation sequence shown in Fig. 8(b), the precipitate cores and caps observed in Fig. 3(b) corresponded to the TiN and NbC phases, respectively. This sequence is consistent with the larger size of Ti-rich precipitates and the smaller size of Nb-rich precipitates.

    According to the theory of Lifshitz, Slyozov, and Wagner, the coarsening of Tinitride precipitates can be calculated using Eq. (2) [3536]:

    {{r}}^{{3}}-{}{{r}}_{{0}}^{{3}}{}={}\frac{{8}\gamma {{V}}_{\rm{m}}{{C}}_{\rm{m}}{{D}}_{\rm{m}}}{{9}{RT}} (2)

    where {r} and {{r}}_{\text{0}} denote the average and initial radius of the precipitates, respectively, \gamma represents the interphase surface energy between the matrix and precipitate, {{V}}_{\text{m}} represents the molar volume of precipitates (11.53 cm3·mol−1 for TiN precipitates), {{C}}_{\text{m}} indicates the solute concentration in the matrix, and {{D}}_{\text{m}} refers to the diffusivity of solute atoms (cm2·s−1), t (s) and T (K) represent the time and absolute temperature, respectively. As calculated using Eq. (2), the degree of precipitate coarsening was primarily influenced by {{V}}_{\text{m}} and {{C}}_{\text{m}} under given t and T conditions. As the precipitates in both steels were (Ti,Nb)(C,N), both steels were assumed to have the same {{V}}_{\text{m}} value. As the Nb-rich particles failed to pin the PAG boundaries at 1100°C, only the coarsening of Ti-rich particles was considered during calculations. The red lines in Fig. 8(b) indicate the higher {{C}}_{\text{m}} of the 60Al steel than the 160Al steel. This result indicates the slightly elevated coarsening rate of carbonitride precipitates in 60Al steel. Therefore, the (Ti,Nb)(C,N) observed in the CGHAZ of 60Al steel were slightly larger than those of the 160Al steel.

    According to experimental results, in the CGHAZ of 60Al steel, γ-Al2O3 served as a nucleation site for (Ti,Nb)(C,N) and Cu2S (Fig. 4). From a thermodynamic perspective, if the γ-Al2O3 particles act as nucleation sites, γ-Al2O3 should form prior to the precipitation of (Ti,Nb)(C,N) and Cu2S. Fig. 8(a) reveals that during the cooling process of HHIW, Al2O3 formed in the 60Al and 160Al steels at temperatures around 1080 and 1264°C, respectively. However, a large number of (Ti,Nb)(C,N) had already precipitated (Fig. 8(b)). Therefore, the Al2O3 formed during the cooling process failed to satisfy the thermodynamic requirement. During the heating process, CaO·6Al2O3 in the 60Al steel transformed into Al2O3 at 1237°C (Fig. 8(a)), and this phenomenon was not observed in the 160Al steel. In the subsequent cooling process of the thermal cycle, TiN particles precipitated again with a decrease in the temperature. Therefore, in the 60Al steel, the Al2O3 particles that formed during the heating process met the thermodynamic requirements as nucleation sites for TiN.

    Based on the above analysis and the results in Fig. 8(a), during the heating process, the Al2O3 (cyan solid line) first transformed into CaO·6Al2O3 in the 60Al steel, followed by a retransformation back into Al2O3. Therefore, Ca reacted with Al2O3 in the oxides, which resulted in the increased CaO content in the oxide and Al diffusion into the matrix (Part I in Fig. 9). Subsequently, with the increase in the temperature, Al replaced Ca in the oxides again and formed Al2O3. Although transformations of oxide compositions in solid steel during heating or cooling processes have been previously studied [3738], little attention has been given to the formation of nanoscale Al2O3 particles. Al-deoxidation experiments revealed the nano Al2O3 particles formed by O in oxides and dissolved Al in the matrix [39]. In the present paper, the Al2O3 nanoparticles were formed by the Al isolated from the oxides and O in the matrix. The transformation of Al2O3 into CaO·6Al2O3 can be described using Eq. (3), and that of CaO·6Al2O3 to Al2O3 using Eqs. (4) and (5). In Eq. (5), the elemental Al in the alloy was mainly supplied by the Al2O3 in the oxide according to Eq. (3). The Supplementary Information further discusses the formation process of nanoscale alumina.

    Fig. 9.  Schematic of transformation of precipitates and PAGs in the CGHAZ of the 60Al and 160Al steels.
    \begin{aligned}[b] \;&{\text{3[Ca]}}_{\text{(Matrix)}}+{{\text{19Al}}_2{\text{O}}_3}_{\text{(Particle)}}=\\ & \quad 3{\text{CaO·}{\text{6Al}}_{\text{2}}{\text{O}}_{\text{3}}}_{\text{(Oxide)}}+ \text{2}{\text{[Al]}}_{\text{(Matrix)}} \end{aligned} (3)
    \begin{aligned}[b] & 3{\text{CaO·}{\text{6Al}}_2{\text{O}}_3}_{\text{(Oxide)}}+{2\text{[Al]}}_{\text{(Matrix)}}=\\ & \quad {3\text{[Ca]}}_{\text{(Matrix)}}+{19{\text{Al}}_2{\text{O}}_3}_{\text{(Oxide)}} \end{aligned} (4)
    \text{2}{\text{O}}_{\text{(Matrix)}}+{\text{3Al}}_{\text{(Matrix)}}={{\text{Al}}_{\text{2}}{\text{O}}_{\text{3}}}_{\text{(Particle)}} (5)

    In the present experiment, the detection of Cu–sulfide in the precipitates extracted from CGHAZ samples of 60Al steel by carbon replica deviated from the predictions obtained through thermodynamic equilibrium calculations. This phenomenon can be explained by the classical nucleation theory. Phase transition free energy per unit volume ( {\Delta }{{G}}_{\text{v}} ), which can be expressed using Eq. (6), is one of the primary factors controlling the nucleation energy ({\Delta }{{G}}^{\text{*}} ):

    {\Delta }{{G}}_{\text{v}}\;\text{≈}\; {Q}\frac{{\Delta }{T}}{{{T}}_{\text{e}}} (6)

    where {{T}}_{\text{e}} denotes the equilibrium temperature; {\Delta }{T} and {Q} are the undercooling below the {{T}}_{\text{e}} and free energy of the solution, respectively. For the solubility product of MX in Fe, {Q} can be evaluated using Eq. (7) and parameter A in Eq. (8) [40]:

    {Q}=\left(\frac{{R}}{\text{2}}\right){A}\cdot\text{ln}\text{10} (7)
    \text{lg}\text{(}\text{[\%M]}\text{[\%X]}\text{)}=-\frac{{A}}{{T}}+{B} (8)

    where R is the universal gas constant, 8.314 J·mol−1·K−1. The solubilities of MnS and Cu2S in γ-Fe were determined as follows [4041]:

    \text{lg}\text{(}\text{[\%Mn][\%S]}\text{)}=\text{–}\frac{\text{9020}}{{T}}+\text{2.929}-\left(-\frac{\text{215}}{{T}}+\text{0.097}\right)\text{[\%Mn]} (9)
    \text{lg}\left({\text{[}\text{%Cu]}}^{\text{2}}\text{[\%S]}\right)=-\frac{\text{44971}}{{T}}+\text{26.31} (10)

    Based on the present steel compositions, the calculated {{Q}}_{\text{MnS}} was 83235 J/mol for [%Mn] = 1.55, and {{Q}}_{{\text{Cu}}_{\text{2}}\text{S}} was 286900 J/mol [42]. The calculated equilibrium precipitation temperature ( {{T}}_{\text{e}} ) for MnS and Cu2S reached 1434 and 1257°C, respectively. Fig. 10 shows the variation in the calculated {\Delta }{{G}}_{\text{v}} with the temperature for MnS and Cu2S in γ-Fe. At temperatures below 1200°C, the {\Delta }{{G}}_{\text{v}} for Cu2S was larger than that for MnS, which means that Cu2S exerted a greater driving force for nucleation.

    Fig. 10.  Driving force for the nucleation of MnS and Cu2S at various temperatures.

    The mismatch theory can accurately reflect the capacity for heterogeneous nucleation. The interfacial energy ( \gamma ), which determines substrate capability for promoting nucleation, was influenced by the lattice mismatch between the substrate and the nucleating parameter (δ) proposed by Bramfitt [43] was used to calculate the mismatch using Eq. (11):

    {\delta }_{{\text{(}{hkl}\text{)}}_{{\rm n}}}^{{\text{(}{hkl}\text{)}}_{\text{s}}}=\sum _{i=1}^{3}\left[\frac{{{d}}_{{\text{[}{uvw}\text{]}}_{\text{s}}}^{i}\text{cos}\,\theta-{{d}}_{{\text{[}{uvw}\text{]}}_{\text{n}}}^i}{{{d}}_{{[{uvw}]}_{\text{n}}}^{{i}}}\right]\times \frac{1}{3}\times\text{100\%} (11)

    where {\text{(}{hkl}\text{)}}_{\text{s}} refers to a low-index lattice plane of the substrate (nucleating agent), {\text{(}{hkl}\text{)}}_{\text{n}} indicates a low-index lattice plane of the nucleating phase, {\text{[}{uvw}\text{]}}_{\text{s}} means a low-index direction in the {\text{(}{hkl}\text{)}}_{\text{s}} plane, {\text{[}{uvw}\text{]}}_{\text{n}} indicates the low-index direction in the {\text{(}{hkl}\text{)}}_{\text{n}} plane, {{d}}_{{\text{[}{uvw}\text{]}}_{\text{s}}}^{{i}} represents the atomic spacing along {{[uvw]}}_{\text{s}} , {{d}}_{{\text{[}{uvw}\text{]}}_{\text{n}}}^{{i}} means the atomic spacing along {\text{[}{uvw}\text{]}}_{\text{n}} , and θ is the angle between the {\text{[}{uvw}\text{]}}_{\text{s}} and {\text{[}{uvw}\text{]}}_{\text{n}} . The substrate can play a very effective role in heterogeneous nuclei when \delta is less than 6.0%; \delta between 6% and 12% indicates a moderate degree of effectiveness, and \delta above 12% is invalid [43].

    According to Eq. (11), the mismatch values between MnS and γ-Al2O3 for various interfaces were calculated (detailed results are provided in the Supplementary Information). The mismatch for the (110)MnS// {\text{(110)}}_{\text{γ-}{\text{Al}}_{\text{2}}{\text{O}}_{\text{3}}} interface reached 7.449%, which is smaller than those for other planes. This result suggests that the γ-Al2O3 can potentially induce the heterogeneous nucleation of MnS. The lattice parameters of TiN were used for the mismatch calculation between (Ti,Nb)(C,N) and γ-Al2O3. The SADP results indicate that {\text{(1}\overline{\text{1}}\text{0)}}_{\text{γ-}{\text{Al}}_{\text{2}}{\text{O}}_{\text{3}}} was a low-index plane that matched with the {\text{(1}\overline{\text{1}}\text{0)}}_{{\text{Cu}}_{\text{2}}\text{S}} and (1 \overline{\text{1}} 0)(Ti,Nb)(C,N) plane. Therefore, the {\text{(1}\overline{\text{1}}\text{0)}}_{\text{γ-}{\text{Al}}_{\text{2}}{\text{O}}_{\text{3}}} //(1 \overline{\text{1}} 0)(Ti,Nb)(C,N)// {\text{(1}\overline{\text{1}}\text{0)}}_{{\text{Cu}}_{\text{2}}\text{S}} interface was applied in \delta calculations. Fig. 11 displays the atomic configurations of the {\text{(1}\overline{\text{1}}\text{0)}}_{\text{γ-}{\text{Al}}_{\text{2}}{\text{O}}_{\text{3}}} , (1 \overline{\text{1}} 0)(Ti,Nb)(C,N) and {\text{(1}\overline{\text{1}}\text{0)}}_{{\text{Cu}}_{\text{2}}\text{S}} planes. The atoms used for \delta calculation between γ-Al2O3 and (Ti,Nb)(C,N) or Cu2S are indicated with red and blue dashed lines in these configurations, respectively. The calculated \delta between γ-Al2O3 and (Ti,Nb)(C,N) on the {\text{(1}\overline{\text{1}}\text{0)}}_{\text{γ-}{\text{Al}}_{\text{2}}{\text{O}}_{\text{3}}} //(1 \overline{\text{1}} 0)(Ti,Nb)(C,N) interface was 6.409% (<12%), which suggests that γ-Al2O3 can provide effective nucleation conditions for the precipitation of (Ti,Nb)(C,N). The calculated \delta between γ-Al2O3 and Cu2S on the {\text{(1}\overline{\text{1}}\text{0)}}_{\text{γ-}{\text{Al}}_{\text{2}}{\text{O}}_{\text{3}}} // {\text{(1}\overline{\text{1}}\text{0)}}_{{\text{Cu}}_{\text{2}}\text{S}} interface was 1.088% (<6%), which implies the higher likelihood of heterogeneous nucleation between γ-Al2O3 and Cu2S than that between γ-Al2O3 and MnS. The (Ti,Nb)(C,N) can also act as a core for the heterogeneous nucleation of Cu2S. Therefore, a lower interfacial energy was observed between γ-Al2O3 and Cu2S than between γ-Al2O3 and MnS.

    Fig. 11.  Atomic configurations of {{\boldsymbol{(1}}\overline{{\boldsymbol{1}}}}{\boldsymbol{0}})_{\text{γ-}{\bf Al}_{\boldsymbol{2}}{\bf {O}}_{\boldsymbol{3}}} , {({\boldsymbol{1}}\overline{\boldsymbol{1}}}{\boldsymbol{0}})_{{\bf {Cu}}_{\boldsymbol{2}}\bf {S}} , and {({\boldsymbol{1}}\overline{\boldsymbol{1}}}{\boldsymbol{0}}) (Ti,Nb)(C,N) planes.

    Based on the above {\Delta }{{G}}_{\text{v}} and \delta results, the heterogeneous nucleation activation energy for Cu2S on γ-Al2O3 was lower than that for MnS, which implies that Cu2S was more easily nucleated than MnS at temperatures below 1200°C. Moreover, the good lattice matching relationship promoted the precipitation of (Ti,Nb)(C,N) and Cu2S on γ-Al2O3. Thus, the larger-sized complex precipitates (γ-Al2O3–(Ti,Nb)(C,N)–Cu2S) formed at high temperatures. Considering that TiN is the main component of carbon nitride, complex precipitates were expressed as γ-Al2O3–TiN–Cu2S. Reduction in the amount of independently precipitated (Ti,Nb)(C,N) and the formation of large-sized γ-Al2O3–TiN–Cu2S drastically reduced the pinning force of the precipitates, which led to abnormal grain growth.

    Fig. 9 shows the transformation of precipitates and PAGs in CGHAZ after the HHIW for the 60Al and 160Al steels. In the CGHAZ, controlling the growth of PAG was crucial, as grain refinement played a crucial role in the enhancement of steel toughness. Experimental evidence indicates that the growth behavior of PAGs was influenced by the presence of Nb- and Ti-rich precipitates. Specifically, during welding, more Nb-rich precipitates in the 60Al steel dissolved at temperatures below 1100°C compared with that in the 160Al steel, which promoted normal PAG growth (Figs. 2 and 8(b)). The normal growth of PAGs was affected by the dissolution and reprecipitation of TiN precipitates during the heating process from 1100 to 1400°C and the subsequent cooling process. When the temperature dropped from the peak temperature of 1400°C, abnormally large grains appeared in the CGHAZ of the 60Al steel but not in the 160Al steel, and this finding was attributed to the formation of γ-Al2O3 nanoparticles in the former. These γ-Al2O3 particles served as nucleation sites for TiN and Cu2S, which facilitated the heterogeneous nucleation of TIN and Cu2S and reduced the amount of independent TiN precipitates. When the complex precipitates γ-Al2O3–TiN–Cu2S were located near the grain boundaries, the driving force for grain growth exceeded the pinning force exerted by precipitates on such areas, which led to abnormal grain growth. Therefore, in the 60Al steel, large PAG grains and abnormal-growth PAGs contributed to the heterogeneity of grain size distribution. The enlargement of grains increased the hardening capacity, which rendered the material nonuniform and potentially reduced its toughness due to the presence of an inhomogeneous microstructure [44]. Cracks in the 60Al steel were more prone to propagate in the grains, which resulted in a decreased impact toughness from 134 to 54 J at −40°C with a decrease in the Al content from 0.016wt% to 0.006wt% (Fig. 9).

    The present study examined the characteristics of precipitates and their influence on austenite grain growth and impact toughness in the CGHAZ of Al–Ti–Ca deoxidized shipbuilding steel plates (containing 0.006wt% and 0.016wt% Al) subjected to HHIW at 400 kJ·cm−1. The major conclusions are as follows:

    (1) Three types of precipitates formed in the BM of two steels: Type I: cubic (Ti,Nb)(C,N); Type II: precipitate with cubic (Ti,Nb)(C,N) core and Nb-rich cap; Type III: ellipsoidal Nb-rich precipitate. With the increase in the Al content, the ND of Type I precipitates in the matrix increased, which contributed to the suppression of austenite grain growth.

    (2) In the low-Al steel (0.006wt% Al), the good lattice mismatch between (Ti,Nb)(C,N) and γ-Al2O3 promoted the heterogeneous nucleation of TiN on the γ-Al2O3 surface during the welding thermal cycle. As a result, the ND of independently precipitated (Ti,Nb)(C,N) was reduced. This condition drastically lowered the pinning force of the precipitates on austinite grain boundaries.

    (3) The low nucleation energy of Cu2S and excellent lattice mismatch (1.088%) between Cu2S and γ-Al2O3 caused the precipitation of Cu2S on γ-Al2O3 and the formation of large-sized γ-Al2O3–TiN–Cu2S.

    (4) As the Al content was decreased, the reduction in the amount of independently precipitated (Ti,Nb)(C,N) and the formation of large-sized γ-Al2O3–TiN–Cu2S triggered the increase in austenite grain size and formation of nonuniform microstructures. Ultimately, the low-temperature impact toughness of the low Al steel was substantially reduced.

    The authors gratefully acknowledge financial support from the National Natural Science Foundation of China (No. U1960202) and the Opening Foundation from Shanghai Engineering Research Center of Hot Manufacturing, China (No. 18DZ2253400).

    Jian Yang is an editorial board member for this journal and was not involved in the editorial review or the decision to publish this article. The authors declare no conflict of interest.

    The online version contains supplementary material avail-able at https://doi.org/10.1007/s12613-024-2967-8.

  • [1]
    X.Q. Pan, J. Yang, and Y.H. Zhang, Microstructure evolution in heat-affected zone of shipbuilding steel plates with Mg deoxidation containing different Nb contents, Metall. Mater. Trans. A, 53(2022), No. 4, p. 1512. DOI: 10.1007/s11661-022-06617-1
    [2]
    X. Wang, C. Wang, J. Kang, G. Yuan, R.D.K. Misra, and G.D. Wang, Improved toughness of double-pass welding heat affected zone by fine Ti–Ca oxide inclusions for high-strength low-alloy steel, Mater. Sci. Eng. A, 780(2020), art. No. 139198. DOI: 10.1016/j.msea.2020.139198
    [3]
    D.K. Liu, J. Yang, Y.H. Zhang, and R.B. Li, Effect of C and Si contents on microstructure and impact toughness in CGHAZ of offshore engineering steel, Metall. Res. Technol., 119, No. 6, art. No. 615.
    [4]
    A.M. Guo, S.R. Li, J. Guo, et al., Effect of zirconium addition on the impact toughness of the heat affected zone in a high strength low alloy pipeline steel, Mater. Charact., 59(2008), No. 2, p. 134. DOI: 10.1016/j.matchar.2006.11.028
    [5]
    B.W. Zhou, G.Q. Li, X.L. Wan, Y. Li, and K.M. Wu, In-situ observation of grain refinement in the simulated heat-affected zone of high-strength low-alloy steel by Zr–Ti combined deoxidation, Met. Mater. Int., 22(2016), No. 2, p. 267. DOI: 10.1007/s12540-016-5301-9
    [6]
    Y.H. Zhang, J. Yang, D.K. Liu, X.Q. Pan, and L.Y. Xu, Improvement of impact toughness of the welding heat-affected zone in high-strength low-alloy steels through Ca deoxidation, Metall. Mater. Trans. A, 52(2021), No. 2, p. 668. DOI: 10.1007/s11661-020-06105-4
    [7]
    Z. Liu, B. Song, and J.H. Mao, Effect of Ca on the evolution of inclusions and the formation of acicular ferrite in Ti–Mg killed EH36 steel, Ironmaking Steelmaking, 48(2021), No. 9, p. 1115. DOI: 10.1080/03019233.2021.1932198
    [8]
    Y.H. Zhang, J. Yang, H.L. Du, Y. Zhang, and H. Ma, Effect of Ca deoxidation on toughening of heat-affected zone in high-strength low-alloy steels after large-heat-input welding, Metals, 12(2022), No. 11, art. No. 1830. DOI: 10.3390/met12111830
    [9]
    Y. Terazawa, K. Ichimiya, and K. Hase, Nucleation effect of Ca–oxysulfide inclusions of low carbon steel in heat affected zone by welding, Mater. Sci. Forum, 941(2018), p. 130. DOI: 10.4028/www.scientific.net/MSF.941.130
    [10]
    T.T. Li and J. Yang, Development in oxide metallurgy for improving the weldability of high-strength low-alloy steel: Combined deoxidizers and microalloying elements, Int. J. Miner. Metall. Mater., 31(2024), No. 6, p. 1263. DOI: 10.1007/s12613-023-2754-y
    [11]
    T. Kimura, H. Sumi, and Y. Kitani, High tensile strength steel plates and welding consumables for architectural construction with excellent toughness in welded joint: JFE EWEL technology for excellent quality in HAZ of high heat input welded joints, JFE Tech. Rep., (2005), No. 5, p. 45.
    [12]
    T. Kato, S. Sato, H. Ohta, and T. Shiwaku, Effects of Ca addition on formation behavior of TiN particles and HAZ toughness in large-heat-input welding, Kobelco Tech. Rev., (2011), No. 30, p. 76.
    [13]
    Y.H. Zhang, J. Yang, L.Y. Xu, et al., The effect of Ca content on the formation behavior of inclusions in the heat affected zone of thick high-strength low-alloy steel plates after large heat input weldings, Metals, 9(2019), No. 12, art. No. 1328. DOI: 10.3390/met9121328
    [14]
    A. Lambert-Perlade, A.F. Gourgues, J. Besson, T. Sturel, and A. Pineau, Mechanisms and modeling of cleavage fracture in simulated heat-affected zone microstructures of a high-strength low alloy steel, Metall. Mater. Trans. A, 35(2004), No. 13, p. 1039. DOI: 10.1007/s11661-004-1007-6
    [15]
    X.Q. Pan, J.J. Zhi, Z.J. Fan, Y.H. Zhang, and J. Yang, Relationship between the microstructure and impact toughness of the coarse-grained heat-affected zone for offshore engineering steels with different Mg contents, Steel Res. Int., 92(2021), No. 10, art. No. 2100099. DOI: 10.1002/srin.202100099
    [16]
    W.L. Wang, L.K. Wang, and P.S. Lyu, Kinetics of austenite growth and bainite transformation during reheating and cooling treatments of high strength microalloyed steel produced by sub-rapid solidification, Int. J. Miner. Metall. Mater., 30(2023), No. 2, p. 354. DOI: 10.1007/s12613-022-2548-7
    [17]
    Y.K. Yang, D.P. Zhan, H. Lei, et al., Coupling effect of prior austenite grain size and inclusion characteristics on acicular ferrite formation in Ti–Zr deoxidized low carbon steel, Metall. Mater. Trans. B, 51(2020), No. 2, p. 480. DOI: 10.1007/s11663-020-01785-0
    [18]
    N. Fujiyama, T. Nishibata, A. Seki, H. Hirata, K. Kojima, and K. Ogawa, Austenite grain growth simulation considering the solute-drag effect and pinning effect, Sci. Technol. Adv. Mater., 18(2017), No. 1, p. 88. DOI: 10.1080/14686996.2016.1244473
    [19]
    K. Poorhaydari and D.G. Ivey, Precipitate alterations in the heat-affected zone of a Grade 100 microalloyed steel, J. Mater. Sci., 46(2011), No. 14, p. 4953. DOI: 10.1007/s10853-011-5412-0
    [20]
    L.Y. Xu, J. Yang, J. Park, and H. Ono, Mechanism of improving heat-affected zone toughness of steel plate with Mg deoxidation after high-heat-input welding, Metals, 10(2020), No. 2, art. No. 162. DOI: 10.3390/met10020162
    [21]
    Y. Jian, Z. Kai, W.R. Zhi, and S.J. Guo, Excellent heat affected zone toughness technology improved by use of strong deoxidizers. J. Iron Steel Res. Int., 18(2011), No. Supplement 2, p. 141.
    [22]
    Z.X. Zhu, L. Kuzmikova, M. Marimuthu, H.J. Li, and F. Barbaro, Role of Ti and N in line pipe steel welds, Sci. Technol. Weld. Join., 18(2013), No. 1, p. 1. DOI: 10.1179/1362171812Y.0000000067
    [23]
    X.Y. Wu, P.C. Xiao, S.J. Wu, et al., Effect of molybdenum on the impact toughness of heat-affected zone in high-strength low-alloy steel, Materials, 14(2021), No. 6, art. No. 1430. DOI: 10.3390/ma14061430
    [24]
    D.K. Liu, J. Yang, Y.H. Zhang, and L.Y. Xu, Effect of Mo content on nano-scaled particles, prior austenite grains and impact toughness of CGHAZ in offshore engineering steels, J. Iron Steel Res. Int., 29(2022), No. 5, p. 846. DOI: 10.1007/s42243-021-00639-6
    [25]
    T.T. Li, J. Yang, Y.H. Zhang, Y.L. Chen, and Y.Q. Zhang, Particles, microstructures, and impact toughness of CGHAZ of Ca deoxidation shipbuilding steel plates with different Nb contents, Steel Res. Int., 94(2023), No. 8, art. No. 2300020. DOI: 10.1002/srin.202300020
    [26]
    X.Q. Pan, J. Yang, Y.H. Zhang, L.Y. Xu, and R.B. Li, Effects of Al addition on austenite grain growth, submicrometre and nanometre particles in heat-affected zone of steel plates with Mg deoxidation, Ironmaking Steelmaking, 48(2021), No. 4, p. 417. DOI: 10.1080/03019233.2020.1801302
    [27]
    X. Wang, C. Wang, J. Kang, G.D. Wang, D. Misra, and G. Yuan, Relationship between impact toughness and microstructure for the as-rolled and simulated HAZ of low-carbon steel containing Ti–Ca oxide particles, Metall. Mater. Trans. A, 51(2020), No. 6, p. 2927. DOI: 10.1007/s11661-020-05753-w
    [28]
    Z.H. Wu, W. Zheng, G.Q. Li, H. Matsuura, and F. Tsukihashi, Effect of inclusions’ behavior on the microstructure in Al–Ti deoxidized and magnesium-treated steel with different aluminum contents, Metall. Mater. Trans. B, 46(2015), No. 3, p. 1226. DOI: 10.1007/s11663-015-0311-4
    [29]
    T.T. Li, J. Yang, Y.H. Zhang, et al., Inclusions and microstructures in coarse-grained heat-affected zone of Al–Ti–Ca deoxidized shipbuilding steels with different Al contents after high-heat input welding, J. Mater. Res. Technol., 30(2024), p. 174. DOI: 10.1016/j.jmrt.2024.02.223
    [30]
    D.S. Martín, F.G. Caballero, C. Capdevila, and C. García de Andrés, Austenite grain coarsening under the influence of niobium carbonitrides, Mater. Trans., 45(2004), No. 9, p. 2797. DOI: 10.2320/matertrans.45.2797
    [31]
    J. Moon, S. Kim, H. Jeong, J. Lee, and C. Lee, Influence of Nb addition on the particle coarsening and microstructure evolution in a Ti-containing steel weld HAZ, Mater. Sci. Eng. A, 454-455(2007), p. 648. DOI: 10.1016/j.msea.2006.11.125
    [32]
    T. Barth, Die regulaere kristallart von kupferglanz, Zentralblatt fuer Mineralogie und Geologie , A, 26(1926), p. 284.
    [33]
    Ľ. Smrčok, V. Langer, and J. Křesťan, γ-alumina: A single-crystal X-ray diffraction study, Acta Crystallogr., Sect. C: Cryst. Struct. Commun., 62(2006), No. 9, p. i83. DOI: 10.1107/S0108270106026850
    [34]
    J. Moon, J. Lee, and C. Lee, Prediction for the austenite grain size in the presence of growing particles in the weld HAZ of Ti-microalloyed steel, Mater. Sci. Eng. A, 459(2007), No. 1-2, p. 40. DOI: 10.1016/j.msea.2006.12.073
    [35]
    I.M. Lifshitz and V.V. Slyozov, The kinetics of precipitation from supersaturated solid solutions, J. Phys. Chem. Solids, 19(1961), No. 1-2, p. 35. DOI: 10.1016/0022-3697(61)90054-3
    [36]
    C. Wagner, Theorie der alterung von niederschlägen durch umlösen (Ostwald–Reifung), Zeitschrift für Elektrochemie, Berichte der Bunsengesellschaft für physikalische Chemie, 65(1961), No. 7-8, p. 581.
    [37]
    Y.P. Chu, W.F. Li, Y. Ren, and L.F. Zhang, Transformation of inclusions in linepipe steels during heat treatment, Metall. Mater. Trans. B, 50(2019), No. 4, p. 2047. DOI: 10.1007/s11663-019-01593-1
    [38]
    Q. Ren, L.F. Zhang, Y.B. Liu, L.X. Cui, and W. Yang, Transformation of cerium-containing inclusions in ultra-low-carbon aluminum-killed steels during solidification and cooling, J. Mater. Res. Technol., 9(2020), No. 4, p. 8197. DOI: 10.1016/j.jmrt.2020.05.077
    [39]
    H. Lei, Y.Y. Xiao, G.C. Wang, H.W. Zhang, W. Jin, and L.F. Zhang, Thermodynamic insight into the growth of nanoscale inclusion of Al-deoxidation in Fe–O–Al melt, Sci. Rep., 10(2020), No. 1, art. No. 16909. DOI: 10.1038/s41598-020-73317-4
    [40]
    T. Nishizawa, Thermodynamics of microstructure control by particle dispersion, ISIJ Int., 40(2000), No. 12, p. 1269. DOI: 10.2355/isijinternational.40.1269
    [41]
    H.A. Wriedt and H. Hu, The solubility product of manganese sulfide in 3 pct silicon–iron at 1270 to 1670 K, Metall. Trans. A, 7(1976), No. 5, p. 711. DOI: 10.1007/BF03186803
    [42]
    Z.Z. Liu, Y. Kobayashi, and K. Nagai, Crystallography and precipitation kinetics of copper sulfide in strip casting low carbon steel, ISIJ Int., 44(2004), No. 9, p. 1560. DOI: 10.2355/isijinternational.44.1560
    [43]
    B.L. Bramfitt, The effect of carbide and nitride additions on the heterogeneous nucleation behavior of liquid iron, Metall. Trans., 1(1970), No. 7, p. 1987. DOI: 10.1007/BF02642799
    [44]
    A. Królicka, K. Radwański, A. Ambroziak, and Ż. Andrzej, Analysis of grain growth and morphology of bainite in medium-carbon spring steel, Mater. Sci. Eng. A, 768(2019), art. No. 138446. DOI: 10.1016/j.msea.2019.138446
  • Related Articles

    [1]Jing-ling Zhang, Qian-ying Guo, Yong-chang Liu, Chong Li, Li-ming Yu, Hui-jun Li. Effect of cold rolling and first precipitates on the coarsening behavior of γ″-phases in Inconel 718 alloy [J]. International Journal of Minerals, Metallurgy and Materials, 2016, 23(9): 1087-1096. DOI: 10.1007/s12613-016-1326-9
    [2]Wen-wen Mao, An-gang Ning, Han-jie Guo. Nanoscale precipitates and comprehensive strengthening mechanism in AISI H13 steel [J]. International Journal of Minerals, Metallurgy and Materials, 2016, 23(9): 1056-1064. DOI: 10.1007/s12613-016-1323-z
    [3]Xiang-liang Wan, Kai-ming Wu, Gang Huang, Ran Wei, Lin Cheng. In situ observation of austenite grain growth behavior in the simulated coarse-grained heat-affected zone of Ti-microalloyed steels [J]. International Journal of Minerals, Metallurgy and Materials, 2014, 21(9): 878-885. DOI: 10.1007/s12613-014-0984-8
    [4]Jing-jing Yang, Run Wu, Wen Liang, Meng-xia Tang. Effect of nano-sized precipitates on the crystallography of ferrite in high-strength strip steel [J]. International Journal of Minerals, Metallurgy and Materials, 2014, 21(5): 448-454. DOI: 10.1007/s12613-014-0928-3
    [5]Eslam Ranjbarnodeh, Stefanie Hanke, Sabine Weiss, Alfons Fischer. Effect of welding parameters on the heat-affected zone of AISI409 ferritic stainless steel [J]. International Journal of Minerals, Metallurgy and Materials, 2012, 19(10): 923-929. DOI: 10.1007/s12613-012-0648-5
    [6]Mai-wen Zhou, Hao Yu. Effects of precipitates and inclusions on the fracture toughness of hot rolling X70 pipeline steel plates [J]. International Journal of Minerals, Metallurgy and Materials, 2012, 19(9): 805-811. DOI: 10.1007/s12613-012-0632-0
    [7]Li-ying Li, Yong Wang, Tao Han, Chao-wen Li. Microstructure and embrittlement of the fine-grained heat-affected zone of ASTM4130 steel [J]. International Journal of Minerals, Metallurgy and Materials, 2011, 18(4): 419-423. DOI: 10.1007/s12613-011-0456-3
    [8]Ran Wei, Cheng-jia Shang, Kai-ming Wu. Grain refinement in the coarse-grained region of the heat-affected zone in low-carbon high-strength microalloyed steels [J]. International Journal of Minerals, Metallurgy and Materials, 2010, 17(6): 737-741. DOI: 10.1007/s12613-010-0382-9
    [9]Gaofeng Tian, Chengchang Jia, Ying Wen, Benfu Hu. Effect of solution cooling rate on the γ′ precipitation behaviors of a Ni-base P/M superalloy [J]. International Journal of Minerals, Metallurgy and Materials, 2008, 15(6): 729-734. DOI: 10.1016/S1005-8850(08)60278-9
    [10]Lin Zhao, Wuzhu Chen, Xudong Zhang, Jiguo Shan. Study on laser welded heat-affected zone in new ultralow carbon bainitic steel [J]. International Journal of Minerals, Metallurgy and Materials, 2007, 14(2): 136-140. DOI: 10.1016/S1005-8850(07)60027-9

Catalog

    Figures(11)  /  Tables(4)

    Share Article

    Article Metrics

    Article views (290) PDF downloads (33) Cited by()

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return