Next Article in Journal
The Effect of Ce on the Microstructure, Superplasticity, and Mechanical Properties of Al-Mg-Si-Cu Alloy
Next Article in Special Issue
Predictive Modeling of Blast Furnace Gas Utilization Rate Using Different Data Pre-Processing Methods
Previous Article in Journal
Microstructure and Fracture Performance of Wire Arc Additively Manufactured Inconel 625 Alloy by Hot-Wire GTAW
Previous Article in Special Issue
Development of Bottom-Blowing Copper Smelting Technology: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Formation Mechanisms and Evolution of Multi-Phase Inclusions in Ti-Ca Deoxidized Offshore Structural Steel

1
Sustainable Minerals Institute, The University of Queensland, Brisbane 4072, Australia
2
HBIS Group Co., Ltd., Shijiazhuang 050000, China
3
Faculty of Materials Metallurgy and Chemistry, Jiangxi University of Science and Technology, Ganzhou 341000, China
*
Authors to whom correspondence should be addressed.
Metals 2022, 12(3), 511; https://doi.org/10.3390/met12030511
Submission received: 17 February 2022 / Revised: 11 March 2022 / Accepted: 14 March 2022 / Published: 17 March 2022
(This article belongs to the Special Issue Fundamentals of Advanced Pyrometallurgy)

Abstract

:
To understand and clarify the formation mechanisms and evolution of complex inclusions in Ti-Ca deoxidized offshore structural steel, inclusions in industrial steel were systematically investigated. The number density of total inclusions generally decreased from Ladle Furnace (LF), Vacuum Degassing (VD), Tundish to the final product except for Ti and Ca addition. The major inclusions during the refining process were CaO-Al2O3-SiO2-(MgO)-TiOx and CaO-Al2O3-SiO2. CaO-Al2O3-SiO2-(MgO)-TiOx inclusion initially originated from the combination of CaO-SiO2-(MgO) in refining slag or refractory and deoxidization product Al2O3 and TiO2. With the refining process proceeding and Ca addition, the Al2O3 concentration in the CaO-Al2O3-SiO2-(MgO)-TiOx inclusions gradually dropped while the CaO and TiO2 concentrations gradually increased. The CaO-Al2O3-SiO2 inclusions originally came from refining slag, existing as 2CaO∙ Al2O3∙ SiO2, and maintained a liquid state during the early stage of LF. After Ca treatment, it was gradually transferred to 2CaO∙ SiO2 due to Al2O3 continuously being reduced by Ca. The liquidus of 2CaO∙ SiO2 inclusion was higher than that of molten steel, so they presented as a solid-state during the refining process. After welding thermal simulation, CaO-Al2O3-SiO2-(MgO)-TiOx inclusions were proven effective for inducing intragranular acicular ferrite (IAF) while CaO-Al2O3-SiO2 was inert for IAF promotion. Additionally, Al2O3-MgO spinel in multiphase CaO-Al2O3-SiO2-(MgO)-TiOx inclusion has different formation mechanisms: (1) initial formation as individual Al2O3-MgO spinel as a solid-state in molten steel; (2) and it presented as a part of liquid inclusion CaO-Al2O3-SiO2-(MgO)-TiOx and firstly precipitated due to its low solubility.

1. Introduction

With the development of large-scale engineering structures, high heat input welding has been widely used in shipbuilding, marine engineering, and oil container fields due to its advantages, such as high efficiency and high stability. However, the increment in the heat input will result in the coarsening of grains in the heat affect zone (HAZ), the origin of a tiny cleavage crack, and a heterogeneous microstructure, thereby significantly deteriorating the toughness [1].
To address these problems, researchers have started to focus on inclusions that are inevitable in metal products. Some non-metallic inclusions can have positive impacts on the steel microstructure and mechanical properties [2,3,4]. Initially, in the 1970s, nanosized TiN particles were found to be effective at inhibiting the growth of austenite grains and beneficial for the mechanical property of steel [5]. They were formed during the cooling process after solidification due to their relatively low solubility. In the 1990s, titanium oxide-containing inclusions, formed in the molten steel and solidification process, were reported to act effectively as intragranular nucleation sites for acicular ferrite, contributing to a smaller grain size and consequently improving the weldability of steel. These practical inclusions were formed during the second refining process and had a relatively high liquidus, so they had excellent thermostability. The technology of titanium oxides and their positive effect on the steel microstructure and phase transformation behaviour was proposed and termed “oxide metallurgy” [6,7,8,9]. Since then, the understanding of inclusions in the steelmaking industry has been rejuvenated. Harmful inclusions, such as large-sized and Al2O3 inclusions that may cause clogging, should be removed as much as possible while those with a small size and particular chemical composition can be maintained and utilized to achieve outstanding mechanical properties [10,11]. Oxide metallurgy has been one of the most effective methods used to address the toughness problem in HAZ.
In recent decades, the use of strong deoxidizers, such as Ca, Mg, Zr, and rare Earth metals (REMs) [10,12,13,14,15], has attracted considerable research interest due to their stronger affinity with O and S and higher thermal stability of inclusions. Traditionally, Ca treatment is mostly used for modifying Al2O3 inclusions with a high melting point into CaO-Al2O3 with a lower melting point, thereby solving the nozzle clogging problem [16,17,18]. Kobe Steel also reported that modified inclusion of CaO-Al2O3 could help prohibit the coarse TiN precipitating on oxides and facilitate the formation and dispersive distribution of fine TiN particles [19]. So far, research has mainly focused on Ca treatment in Ti-bearing Al-killed steel. Wang et al. [20] compared Al-Ca deoxidized steel and Ti-Ca deoxidized steel in the lab, pointing out that CaO-Al2O3-TiO2-MnS-related inclusions were dominant and active in facilitating fine IAF formation [21,22]. CaO-Al2O3-TiO2 will act as a nucleation core while MnS will precipitate around the core during the cooling process and form a manganese depletion zone (MDZ) to induce IAF formation [23,24,25].
Although several investigations regarding Ti-Ca deoxidized steel have been reported, they mainly focused on the effect of different deoxidizers on inclusions in final products or the effect of inclusions on the microstructure in HAZ. However, few studies harnessing the integrated understanding of inclusion evolution in each step of actual industry practice are available. In this paper, the inclusion characteristics regarding categories, number density, and evolution were systematically investigated to understand and clarify the formation mechanisms and evolution of complex inclusions in Ti-Ca-treated offshore structural steel, which will benefit the translation Ti-Ca oxide metallurgy of offshore steels.

2. Materials and Methods

Steel samples were obtained from an industrial process for offshore structural steel, which followed the sequential steps through a basic oxygen furnace (BOF)→ladle furnace (LF)→vacuum degassing (VD)→continuous casting (CC)→thermomechanical control process (TMCP) in Wusteel, HBIS. Ferrosilicon (Si ≥ 75.0%, Al ≤ 1.5%, Fe ≥ 21.0%) and ferromanganese (65.0–72.0%Mn, Si ≤ 2.5%, C ≤ 7.0%) alloys together with lime were added to the molten steel during BOF tapping. Then, the molten steel was transferred into LF for refining and bottom argon blowing was adopted for homogenizing the steel composition and accelerating the removal of large-sized inclusions. LF refractories were composed of MgO-C bricks. After refining in LF at 1650 °C, the molten steel was transferred to VD at 1600 °C and was eventually transported into tundish for continuous casting. In particular, samples were collected at the beginning of LF 1 min after Ti addition (Ti-Fe: 25.0–35.0%Ti, Al ≤ 8.0%, Si ≤ 4.5%), 1 min after Ca addition, VD soft blowing, tundish, and hot-rolled plate (final products). The chemical compositions of the steel samples from different stages are shown in Table 1. The product sample underwent heat treatment as welding thermal simulation at a peak temperature of 1400 °C. The measured temperature curve of the welding thermal simulation is shown in Figure 1. The welding samples were held at 1400 °C for 180 s, and the cooling rates of the 1400–800 °C and 800–500 °C range were approximately 3 and 0.75 °C/s (t8/5 = 400 s), respectively.
Steel samples were mounted in resin and ground, polished, and etched for inclusion composition analysis and microstructure observation. Composition analysis of inclusions was conducted by using a JXA 8200 electron probe microanalyzer (EPMA, JEOL, Akishima, Japan) with wavelength-dispersive spectrometers (WDSs).
The EPMA operating details are as follows: the standard samples are CaSiO3 for Ca and Si, pure MgO for Mg, pure Al2O3 for Al, pure TiO2 for Ti, Spessartine for Mn, Fe2O3 for Fe, and FeS for S. An acceleration voltage of 15 kV, probe current of 15 nA, probe diameter of “zero” (the smallest operating probe diameter achieved by the focused electron beams), peak measuring time of 30 s, and background measuring time of 10 s were used. Backscattered electron imaging and EDS analysis were utilized to recognize the non-metallic inclusions, especially those complex inclusions with multiple phases. Then, WDS was conducted on the inclusions to analyze the inclusion composition accurately. EPMA can accurately determine the composition of inclusions exceeding 1.5 µm. Field emission scanning electron microscopy (FE-SEM) was used to analyze inclusions sized 1.5 µm or less by element spot analysis, line scanning, and map scanning to demonstrate the elementary distribution of multiphase inclusions.
To describe the number densities of inclusions, 20 photos were taken randomly under 500 magnifications for each sample to ensure reliable statistics. The area of each photo was 0.246 mm × 0.196 mm = 0.0482 mm2, so the total observation area of each sample was 0.964 mm2.

3. Results and Discussions

3.1. Characteristics of the Inclusions at Different Stages

Due to complex alloy systems including Si, Mn, Al, Ti, and Ca related to the deoxidization process, the inclusions can be classified into nine groups: Al2O3-MnO, Al2O3-SiO2-MnO, CaO-Al2O3-SiO2-MnO, CaO-SiO2, CaO-Al2O3-SiO2, CaO-Al2O3-SiO2-TiOx, Al2O3-MnO-TiOx, SiO2-MnO, and CaO-TiOx, as shown in Table 2. Generally, several valences of Ti, such as TiO, TiO2, Ti2O3, and Ti3O5, exist in steel, and it is assumed that Ti-oxide was mainly TiO2 when analyzing the inclusion composition, so a TiO2 standard sample was used during the EPMA analysis. Among these inclusions, Al2O3-MnO, Al2O3-SiO2-MnO, and CaO-Al2O3-SiO2-MnO are either primary deoxidization products or those that contacted with lime, which was added during the BOF tapping process. CaO-SiO2 and CaO-Al2O3-SiO2 are grouped into slag entrapment products (large-sized inclusions) and reduction products (small-sized inclusions). CaO-Al2O3-SiO2-TiOx and CaO-TiOx belong to the refining products. Al2O3-MnO-TiOx and SiO2-MnO inclusions are considered to be secondary oxidization products during the vacuum break of VD and Tundish processes.
The initial inclusions in the LF-entry sample include Al2O3-MnO, Al2O3-SiO2-MnO, CaO-Al2O3-SiO2-MnO, and CaO-SiO2. Among these inclusions, Al2O3-MnO and Al2O3-SiO2-MnO are primary deoxidization products with small sizes generally less than 3 µm. Although Fe-Si and Fe-Mn ferroalloys are mainly used as pre-deoxidizer in industrial practice, inevitably a small amount of Al impurities exist in alloys. As Al has a stronger ability to combine with oxygen than Si and Mn, almost all the primary deoxidization products contained Al2O3. CaO-Al2O3-SiO2-MnO can be considered as the coalescence between primary deoxidization products and CaO from lime and CaO-SiO2 from refining slag that was added to molten steel during BOF tapping. CaO-SiO2 inclusions in the LF entry stage were generally larger than 10 µm, but the proportion of CaO-SiO2 inclusions in this stage was very low. They originated from slag entrapment. After Ti was added to LF, the main inclusion types were CaO-Al2O3-SiO2 and CaO-Al2O3-SiO2-TiOx 1–10 µm in size. The dissolved aluminium coming from continuous reaction of ferroalloys with CaO-SiO2 from refining slag then formed CaO-Al2O3-SiO2 inclusions. The additive Ti entered molten steel, reacting with CaO-Al2O3-SiO2 inclusions and forming CaO-Al2O3-SiO2-TiO2 inclusions. After Ca was added to the molten steel, a considerable proportion of CaO-SiO2 inclusions was observed apart from CaO-Al2O3-SiO2 and CaO-Al2O3-SiO2-TiOx inclusions due to the severe slag entrapment caused by the splashing of the Ca addition, and the reduction of Al2O3 by additive calcium. In the following VD and Tundish, until the final products, CaO-SiO2, CaO-Al2O3-SiO2, and CaO-Al2O3-SiO2-TiOx inclusions remained the main types. The most significant difference was that Al2O3-MnO-TiOx and SiO2-MnO inclusions were observed after the vacuum break of VD and Tundish. In general, these were typical re-oxidization products due to the excessive surface turbulence during molten steel teeming into the Tundish ladle. However, these kinds of re-oxidization products were seldom found in the final products, which indicated that most of them floated upwards into the mold flux and were removed effectively. In the final products, CaO-TiOx inclusions were found apart from CaO-SiO2, CaO-Al2O3-SiO2, and CaO-Al2O3-SiO2-TiOx inclusions. They may be generated from either the enrichment of the CaO component and the reduction of Al2O3 and SiO2, or the precipitation of CaO-TiOx from CaO-Al2O3-SiO2-TiOx inclusions during the cooling process.

3.2. Number Density and Size Distribution of Inclusions

The number density of all inclusions in industrial steel is shown in Figure 2. The primary inclusions in the LF entry sample were generally primary deoxidization products, such as Al2O3-MnO and Al2O3-SiO2-MnO, featuring a high proportion of fine inclusions that were less than 1 µm and had a number density of 1560/mm2. With ladle furnace refining processing, when Fe-Ti alloy was added to the molten steel, the total amount of inclusions decreased dramatically to about 300/mm2, especially those with a small size (<1 µm). This is due to the coalescence of small inclusions, floating upward into the refining slag. When Si-Ca wire was added to the molten steel and the end of LF refining, the total amount of inclusions increased slightly to higher than 400/mm2 owing to the severe splashing during Si-Ca wire feeding. It also featured a higher proportion of 1–2 µm and 2–5 µm inclusions, and this can be explained by inclusion coalescence and growth to be larger ones. In the VD section, the total amount of inclusions dropped gradually to around 200 mm2, small-sizd inclusions of less than 1 µm. This is because small-sized inclusions kept coalescing and gathering during the vacuum break and soft blowing process of VD. From VD to Tundish and products, the total amount of inclusions decreased steadily, and the main factor was the decrease in the inclusions with small sizes less than 1 µm.
The oxide components in inclusions at different stages are shown in Figure 3. As can be seen, Al2O3 and MnO, as primary deoxidization products, accounted for the majority initially when LF began. With the LF refining processing, when Fe-Ti and Si-Ca alloys were added to the molten steel, CaO and SiO2 were substituted for Al2O3 and MnO, accounting for more than 50% and 30%, respectively. Therefore, the content of Al2O3 and MnO dramatically decreased to less than 10%. From the Ca addition to vacuum degassing, the CaO content dropped a little to 40% due to the evaporation effect of Ca, and accordingly, the Al2O3 content increased a little to about 20%. However, the CaO content increased gradually to about 50% after the vacuum break of VD, which was due to extra Si-Ca alloy being added to the molten steel. The TiO2 concentration fluctuated from 10% to 15% after Ti-Fe addition.
CaO-SiO2, CaO-Al2O3-SiO2, and CaO-Al2O3-SiO2-(MgO)-TiOx inclusions remained the main types among all the processes after Ti addition. Particularly, CaO-SiO2 and CaO-Al2O3-SiO2 can be described in a CaO-Al2O3-SiO2 system, thus CaO-Al2O3-SiO2-(MgO)-TiOx and CaO-Al2O3-SiO2 inclusions were focused on and their formation mechanisms and evolutions were investigated and are discussed here.

3.3. Formation Mechanism and Evolution of CaO-Al2O3-SiO2-(MgO)-TiOx Inclusion

The initial CaO-Al2O3-SiO2-(MgO)-TiOx (MgO was marked here for discussing the formation mechanism) inclusion was found after Ti addition. Three typical inclusions (size ranged from 1 to above 10 µm) with different sizes are shown in Figure 4. The size of the three inclusions were 10.8, 5.1, and 2.5 µm, respectively. All of them belonged to CaO-Al2O3-SiO2-(MgO)-TiOx inclusions, among which the MgO content was relatively low at about 5%. The detailed chemical composition is shown in Figure 5 to demonstrate the varying trend of each oxide component with different sizes.
It is obvious from Figure 5 that the mass proportion of CaO and SiO2 steadily dropped with the decrease in the inclusion size. The mass proportion of MgO remained stable at a relatively low level. Adversely, the mass proportion of Al2O3 and TiO2 showed a gradually increasing trend. CaO and SiO2 were generally the main contents of the refining slag, and MgO caneither come from refining slag or the corrosion of MgO-C refractory and Al2O3, and TiOx was a typical deoxidization product of Ti-Ca killed steel. Generally, deoxidization products are small in size while the size of particles from slag and refractory can range widely. When small-sized deoxidization products are combined with large-sized CaO-SiO2(-MgO) particles, the CaO-SiO2(-MgO) content in the average composition will be high, and Al2O3-TiO2 will be low. On the other hand, when a small-sized deoxidization product is combined with small- or medium-sized CaO-SiO2(-MgO) particles, the CaO-SiO2(-MgO) content in the average composition will be relatively lower, and Al2O3-TiO2 will be moderately higher. As a result, it can be concluded from Figure 5 that the CaO-Al2O3-SiO2-(MgO)-TiOx inclusion in LF originated from the reaction between CaO-SiO2 from slag, MgO from refractory, and the deoxidization products Al2O3 and TiOx.
The CaO-Al2O3-SiO2-MgO-TiOx inclusions’ composition at different stages was projected into pseudo-ternary phase diagrams of CaO-Al2O3-TiOx at fixed 20%SiO2 (14.8–23.2% SiO2) and 5%MgO (2.7–7.8%MgO), as shown in Figure 6. In Figure 6a, all the inclusions’ composition is projected, and relatively dispersive. Accordingly, the average composition of each stage was calculated and is shown in Figure 6b. After the addition of Ti in the LF refining, the main inclusions are generally located in the regions of CaO-Al2O3-MgO and melilite, and with Ca addition in LF and following VD, the main inclusions are basically located in melilite and Ca2SiO4 regions, and the inclusion composition in final product is eventually located in the CaTiO3 region.
The morphologies of CaO-Al2O3-SiO2-(MgO)-TiOx at different stages are shown in Figure 7. Generally, they were spherical. From the liquidus temperature in the pseudo-ternary phase diagram, the compositions of inclusions at different stages are located in the 1400–1500 °C liquidus region, which means they were in a liquid state in molten steel and thereby presented a sphere shape. The morphologies shown in Figure 7 indicate relatively fast cooling and the precipitation of different components; thus, the morphologies reflect the formation mechanism of inclusion. The component with a light grey colour is TiO2, and the area proportion of TiOx continuously increased with the refining process proceeding, which was in accordance with the TiOx component trend shown in the phase diagram. During the early stage of LF refining after Ti addition, large amounts of Al-Ti-oxide particles started to gather around and stick to the CaO-SiO2-(MgO) inclusion on its surface. Particularly, Al2O3 was from the oxidization of residual Al in the ferroalloys, and TiOx was from the oxidization of Ti in the. Ti-Fe alloy. These small deoxidization products did not have enough time to dissolve into the inclusion from the refining slag, so they stayed on the surface. With the refining process proceeding, Ti-oxide continuously increased and had sufficient time to gradually dissolve into the inclusion from the refining slag and formed new inclusions as a whole. As can be seen in the phase diagrams shown in Figure 6, the liquidus of the average composition of the inclusions was in the range from 1400 to 1450 °C. So, these new inclusions were mainly liquid inclusions in molten steel. During the cooling of the sample-taking process, different phases precipitated and formed multiphase inclusions, as shown in Figure 7. As a result, the morphology after cooling presented, such as the Ti-oxide component, “entered” the inclusion and formed the core of the multiphase inclusions. Meanwhile, the Al2O3 and SiO2 component generally decreased due to the reduction of Ca, so the average content of CaO showed an increasing trend when compared with that of Al2O3 and SiO2, as shown in Figure 6. The formation mechanism of CaO-Al2O3-SiO2-(MgO)-TiOx is in accordance with that of the CaO-Al2O3-SiO2-(MgO) inclusions in Si-Mn killed steel with a limited aluminum content [21].
In summary, the formation and evolution mechanism of CaO-Al2O3-SiO2-(MgO)-TiOx inclusion during the whole refining process can be described in a schematic diagram, as shown in Figure 8.

3.4. Formation Mechanism and Evolution of CaO-Al2O3-SiO2 Inclusion

There are generally two morphology types of CaO-SiO2-Al2O3 inclusions as shown in Figure 9. One is a single-phase inclusion with a smooth surface, and the other has a rough surface. These two types of CaO-SiO2-Al2O3 inclusions have different chemical compositions as shown in Table 3. The basicity CaO/SiO2 of them is similar at about 1.9–2.0. The difference is mainly the component of Al2O3: Al2O3 of the inclusion in Figure 9a is 24.5% while that of the inclusion in Figure 9b is only 2.8%, so the inclusion in Figure 9b can be defined as CaO-SiO2. The chemical composition distribution of CaO-SiO2-Al2O3 inclusions in steel samples taken from different stages of industrial manufacturing was projected into CaO-Al2O3-SiO2 ternary phase diagrams, as shown in Figure 10.
The CaO-Al2O3-SiO2 and CaO-SiO2 inclusions’ compositions at different stages were projected into ternary phase diagrams of CaO-Al2O3-SiO2, as shown in Figure 10. Since the inclusion composition of the industrial sample is distributed dispersedly (Figure 10a), the average composition of inclusions during different stages is shown in Figure 10b. The initial composition of the LF entry is located in the 2CaO∙ SiO2 region, and with the refining processing from LF→VD→TD→final products, the average composition moved from the 2CaO∙ SiO2 region to the 2CaO∙ Al2O3∙ SiO2 region (Gehlenite) and eventually returned to the 2CaO∙ SiO2 region (Ca2SiO4). First, additive lime during the BOF tapping process brought CaO into molten steel, and it met the primary deoxidization product SiO2 and formed 2CaO∙ SiO2. When Ti-Fe alloy was added, residual Al enterted the molten steel and formed the deoxidization product Al2O3. The combination of Al2O3 and Ca2SiO4 formed 2CaO∙ Al2O3∙ SiO2, and significantly decreased the liquidus of inclusion. Then, Ca treatment led to the reduction of Al2O3 and SiO2 in 2CaO∙ Al2O3∙ SiO2, so the Al2O3 and SiO2 content kept dropping while the CaO content gradually increased. When LF began, the average inclusion composition located in the 2CaO∙ SiO2 region and the CaO/SiO2 ratio was about 1.7–1.8, and its liquidus was around 2000 °C. After the Ti addition, the average inclusion composition was located in the Gehlenite region, and the liquidus significantly decreased to about 1500–1600 °C due to the increment of Al2O3. The inclusions after the Ti addition had a spherical and smooth shape because they remained liquid in molten steel. After Ca treatment, the CaO/SiO2 ratio tended to be slightly higher than the initial CaO/SiO2 ratio, and the average composition entered the Ca2SiO4 region, the liquidus of which was about 1700–1800 °C. As a result, the CaO-SiO2 inclusions from VD, tundish, and the final product presented rough surfaces, which indicated that they remained in a solid state in molten steel as shown in Figure 11. The reactions during these processes can be described as Formula (1)–(3):
2 CaO + SiO 2 2 CaO · SiO 2
2 CaO · SiO 2 + [ Al ] + [ O ] 2 CaO · Al 2 O 3 · SiO 2
2 CaO · Al 2 O 3 · SiO 2 + [ Ca ] CaO Al 2 O 3 SiO 2 + [ Al ] + [ Si ]

3.5. Effect of Inclusions on IAF Formation

As many researchers have reported, titanium oxide is the key part of “oxide metallurgy”. Generally, they perform as an effective and stable nucleation site for the induction of intragranular acicular ferrite (IAF). As discussed above, the main categories of inclusion in Ti-Ca industrial steel are CaO-Al2O3-SiO2-(MgO)-TiOx and CaO-Al2O3-SiO2 (CaO-SiO2 can be defined as one of CaO-Al2O3-SiO2). However, CaO-Al2O3-SiO2 may not be an effective inclusion for IAF promotion while CaO-Al2O3-SiO2-(MgO)-TiOx can be a potential nucleation site for IAF. As a result, the chemical composition of some typical inclusions was investigated using EDS mapping scanning of FE-SEM to clarify the element distribution in multiphase inclusions. Figure 12 shows the morphologies of typical Ti-oxide-containing inclusions in final products, and their chemical composition is shown in Table 4. They were generally CaO-Al2O3-(SiO2-MgO)-TiOx inclusions, with MnS precipitating around them.
Additionally, element mapping scanning was used to identify the chemical composition of different phases in multiphase inclusions. The element distribution of 2 typical multiphase inclusions is shown in Figure 13 and Figure 14. The size of the inclusion in Figure 13 is about 2 µm. The shape of the inclusion shows different parts combined, and it mainly consists of two parts: MgO-Al2O3 spinel formed the core with two small MgO + MgS particles attached to it. The core was half surrounded by the CaO-TiO2 layer, and the interface between the core and surrounding layer was relatively clear. The size of the inclusion in Figure 14 is about 1.5 µm. The shape looks like a whole, and it contains two phases: the left part is MgO-Al2O3 spinel, and the right part is CaO-TiOx and (Ca, Mn) S. Generally, MgO-Al2O3 spinel is formed by the combination of [Mg], [Al], and [O] in molten steel and has a very high melting point. However, different mechanisms of the formation of MgO-Al2O3 spinel in Ti-Ca-treated industrial steel may exist. As for the multiphase inclusion in Figure 13, it is indicated that MgO-Al2O3 spinel was first formed when dissolved [Al] from ferroalloy combined with [Mg] from the refractory. Then, spinel acted as a heterogeneous nucleation site, and MgS and CaO-TiOx started to precipitate around the core with the decreasing temperature of molten steel and the solidification process. As a result, clear interfaces between each phase existed. As for the multiphase inclusion in Figure 14, the explanation of its formation may be illustrated, as shown in Figure 8: liquid inclusion CaO-Al2O3-SiO2-(MgO)-TiOx was formed in the steel refining process, but with the decreasing temperature of molten steel and the solidification process, MgO-Al2O3 spinel firstly crystallized and precipitated due to its highest melting point, and then other phases, such as CaO-TiOx and (Ca, Mn) S, precipitated in succession.
A large-sized multiphase inclusion is shown in Figure 15 and the results of EPMA analysis for each phase are listed in Table 5. The main body of this large-sized inclusion consisted of CaO-Al2O3-SiO2 (red number 1) and CaO-TiOx (red number 2). It seemed that this large-sized inclusion was crushed and separated into several parts composed of (Ca, Mn)S (red number 3), CaO-Al2O3-MgO (red number 4) and CaO-SiO2 (red number 5) during the rolling process of TMCP, so it spread along the rolling direction.
After welding simulation, the inclusions and microstructure were investigated, and some typical inclusions that can effectively induce IAFs were observed and detected using EPMA. The morphologies and chemical composition of typical inclusions are shown in Figure 16 and Table 6, respectively. The induced IAFs are marked in red (AF1–AF6, AF1–AF3) in Figure 16. It can be confirmed that only TiO2 containing inclusions were effective for IAF promotion while CaO-Al2O3-SiO2 inclusions were not found to be the nucleation site of IAF formation. These effective IAF inclusions were generally multiphase, and their size ranged from 2 to 5 µm and consisted of CaO-Al2O3-(SiO2-MgO)-TiOx. This indicates that CaO-Al2O3-(SiO2-MgO)-TiOx was essential for inducing IAFs in Ti-Ca-treated offshore structural steel. Particularly, it can be found that the concentration of Ti-oxide in these effective IAF inclusions (20–30%wt) was significantly higher than the average concentration of Ti-oxide in all inclusions, as shown in Figure 13, which was about 15%wt. The higher concentration of Ti-oxide is generally considered to increase the possibility of IAF formation, thereby enhancing the mechanical property of steel. As a result, how to increase the proportion of CaO-Al2O3-(SiO2-MgO)-TiOx and how to increase the concentration of Ti-oxide can be a potential and meaningful research direction, and more work is yet to be done in the future.
To clearly identify different phases in these multi-phased inclusions and thoroughly clarify the formation mechanisms, mapping scanning of FE-SEM was used for detecting the elementary distribution of effective inclusions that induced IAFs (inclusion A, B, and C). The results are shown in Figure 17, Figure 18 and Figure 19.
As can be seen in Figure 17, the equivalent diameter of inclusion A is about 5.1 µm and one piece of acicular ferrite lath is induced by this inclusion. The morphology of the multi-phased inclusion has a spherical shape. This is a typical CaO-Al2O3-SiO2-MgO-TiOx inclusion, and its main body consists of CaO-Al2O3-SiO2-MgO, Al2O3-MgO, and CaO-TiOx, respectively. Particularly, CaO-Al2O3-SiO2-MgO is the main content of secondary refining slag and has a low liquidus, thereby presenting a liquid state in molten steel. An Al2O3-MgO spinel particle 1.2 µm in size is embedded into the CaO-Al2O3-SiO2-MgO liquid inclusion, and it has a very high liquidus temperature. CaO-TiO2 also has a higher liquidus temperature than that of molten steel except for the condition when TiO2 = 80%. As a result, it can be concluded that the CaO-Al2O3-SiO2-MgO-TiOx liquid inclusion was formed in molten steel as described in Figure 8. Then, Al2O3-MgO spinel and CaO-TiOx perovskite precipitated in order during the cooling process.
As shown in Figure 18, inclusion B also shows the spherical shape, and two pieces of acicular ferrite laths are induced by this inclusion. It consists of CaO-Al2O3-SiO2-MgO, (Ca, Mn) S, and CaO-TiOx. Particularly, the left part of the main body is CaO-Al2O3-SiO2-MgO, and the right part of the main body is (Ca, Mn) S while CaO-TiOx perovskite precipitates along the edge of the above two phases.
In Figure 19, five pieces of acicular ferrite laths are induced by inclusion C. Different from the above two inclusions, the morphology reveals that it is formed due to the coalescence of several particles: CaO-Al2O3-SiO2-MgO, (Ca, Mn) S, and CaO-TiOx. During the secondary refining process, phases with a higher liquidus temperature, such as (Ca, Mn) S and CaO-TiOx, are “captured” by liquid-phase CaO-Al2O3-SiO2-MgO. These particles do not melt and form a new liquid inclusion but mechanically coalesce together.
Almost all the detected inclusions effective at inducing IAFs are found to be CaO-Al2O3-SiO2-MgO-TiOx-based inclusions. It can be confirmed that CaO-Al2O3-SiO2-MgO-TiOx-based inclusions are effective nucleation sites for IAF promotion while CaO-Al2O3-SiO2-based inclusions are ineffective at inducing IAFs. The above three elementary distribution analyses also prove the formation mechanism of CaO-Al2O3-SiO2-MgO-TiOx-based inclusions due to the combination of particles from refining slag and deoxidization products. Then, different phases precipitate in order during the cooling process after secondary refining.

4. Conclusions

To clarify the formation mechanism and evolution of oxide inclusions in Ti-Ca-treated offshore structural steel, industrial sampling was conducted from LF, VD, and TD to final products. Continuous changes and correlations of various inclusions were explained by analysing the number density, morphology, and chemical composition using EPMA and FE-SEM. The primary findings are concluded as follows:
  • The evolution of inclusions during different stages in Ti-Ca-treated offshore structural steel is from primary deoxidization products (Al2O3-MnO, Al2O3-SiO2-MnO) and their combination with lime (CaO-Al2O3-SiO2-MnO and CaO-SiO2)→CaO-Al2O3-SiO2-(MgO)-TiOx and CaO-Al2O3-SiO2→CaO-Al2O3-SiO2-(MgO)-TiOx, CaO-Al2O3-SiO2, and secondary deoxidization products (Al2O3-MnO-TiOx and SiO2-MnO)→CaO-Al2O3-SiO2-(MgO)-TiOx and CaO-Al2O3-SiO2. The number density of the inclusions in Ti-Ca-treated industrial steel generally dropped from LF, VD, and Tundish to the final product, except for the Ti-Fe and Si-Ca addition in the LF, the number density slightly increased. The total decrease in the inclusion number density was mainly due to the significantly decreasing number density of small inclusions (<1 µm) during the refining process.
  • The formation mechanism of CaO-Al2O3-SiO2-(MgO)-TiOx inclusion was due to CaO-SiO2-(MgO) from refining slag and refractory combining with the deoxidization product Al2O3 and TiOx. With the refining process proceeding, Ti-oxide continuously increased and gradually “entered” the inclusion and formed the core of the multiphase inclusions while the Al2O3 component generally decreased due to the reduction of Ca, so the average content of CaO showed an adverse trend when compared with that of Al2O3.
  • The formation mechanism of CaO-Al2O3-SiO2 inclusions is the initial 2CaO∙Al2O3∙SiO2 inclusion came from the combination of CaO-SiO2 particles in refining slag and the deoxidization product Al2O3, and its liquidus was lower than that of molten steel, so it presented a liquid state in steel and had a smooth surface. After Ca addition, the initial 2CaO∙Al2O3∙SiO2 was gradually transferred to 2CaO∙ SiO2 with Al2O3 continuously reduced by Ca. 2CaO∙ SiO2 had a higher liquidus than that of molten steel, so it presented as a solid state in steel and had a rough surface.
  • In Ti-Ca-treated offshore structural steel, after welding simulation, CaO-Al2O3-SiO2 inclusions were not effective at inducing IAFs while CaO-Al2O3-SiO2-(MgO)-TiOx inclusions were proven to be effective nucleation sites for promoting IAFs. The Al2O3-MgO spinel component in welding samples may have different formation mechanisms: one is that it formed directly in molten steel as a solid state, and other phases and inclusions, such as CaO-TiOx and MnS, precipitated on Al2O3-MgO spinel, so the interface between each phase was clear. Another is that CaO-Al2O3-SiO2-(MgO)-TiOx as a whole formed in molten steel as a liquid state, and Al2O3-MgO spinel firstly precipitated due to its highest melting point and was followed by other phases, so the interface between each phase was not clear.

Author Contributions

Conceptualization, Z.R. and X.M.; methodology, Z.R., B.Z. and X.M.; software, Z.R.; validation, B.Z., X.M. and G.W.; formal analysis, Z.R.; investigation, Z.R. and X.M.; resources, H.L., P.Z. and F.W.; data curation, H.L., P.Z. and F.W.; writing—original draft preparation, Z.R.; writing—review and editing, Z.R, F.T.; visualization, Z.R.; supervision, X.M.; project administration, B.Z., X.M. and G.W.; funding acquisition, B.Z. and X.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Hebei Iron and Steel Group (HBIS), grant number ICSS2017-02.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Financial support for this research provided by Hebei Iron and Steel Group (HBIS) is greatly appreciated. The authors also acknowledge the facilities and the scientific and technical assistance of the Australian Microscopy& Microanalysis Research Facility at the Centre of Microscopy and Microanalysis at the University of Queensland.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wan, X.L.; Wu, K.M.; Wang, H.H.; Li, G.Q. Applications of Oxide Metallurgy on High Heat Input Welding Steel. China Metall. 2015, 25, 6–12. [Google Scholar]
  2. Villegas, R.; Redjaimia, A.; Confente, M.; Perrot-Simonetta, M.T. Fractal Nature of Acicular Ferrite, and Fine Precipitation in Medium Carbon Micro-Alloyed Forging Steels. Proc. New Developments on Metallurgy and Applications of High Strength Steels. Ed. Asoc. Argentina de Materiális. Buenos Aires. 2008. Available online: https://www.researchgate.net/publication/301840861_Proc_of_New_Developments_on_Metallurgy_and_Applications_of_High_Strength_Steels_Conference_Minerals_Wiley_John_Sons_Incorporated (accessed on 15 February 2022).
  3. Sarma, D.S.; Karasev, A.V.; Jönsson, P.G. On the Role of Non-metallic Inclusions in the Nucleation of Acicular Ferrite in Steels. ISIJ Int. 2009, 49, 1063–1074. [Google Scholar] [CrossRef] [Green Version]
  4. wan Kim, J.; koo Kim, S.; sik Kim, D.; deuk Lee, Y.; keun Yang, P. Formation Mechanism of Ca-Si-Al-Mg-Ti-O Inclusions in Type 304 Stainless Steel. ISIJ Int. 1996, 36, 140–143. [Google Scholar] [CrossRef]
  5. Kanazawa, S.; Nakashima, A.; Okamoto, K.; Kanaya, K. Improved Toughness of Weld Fussion Zone by Fine TiN Particles and Development of a Steel for Large Heat Input Welding. Tetsu Hagané 1975, 61, 2589–2603. [Google Scholar] [CrossRef] [Green Version]
  6. Takamura, J.; Mizoguchi, S. Roles of Oxides in Steels Performance. In Proceedings of the 6th International Iron and Steel Congress, ISIJ, Nagoya, Japan, 21–26 October 1990; Volume I, pp. 591–597. [Google Scholar]
  7. Mizoguchi, S.; Takamura, J. Control of Oxides as Inoculants-Metallurgy of Oxides in Steels. In Proceedings of the 6th International Iron and Steel Congress, ISIJ, Nagoya, Japan, 21–26 October 1990; Volume I, pp. 598–604. [Google Scholar]
  8. Sawai, T.; Wakoh, M.; Ueshima, Y.; Mizoguchi, S. Effects of Zirconium on the Precipitation of MnS in Low Carbon Steels. In Proceedings of the 6th International Iron and Steel Congress, ISIJ, Nagoya, Japan, 21–26 October 1990; Volume I, pp. 605–611. [Google Scholar]
  9. Ogibayashi, S.; Yamaguchi, K.; Hirai, M.; Goto, H. The Features of Oxides in Ti-deoxidized Steel. In Proceedings of the 6th International Iron and Steel Congress, ISIJ, Nagoya, Japan, 21–26 October 1990; Volume I, pp. 612–617. [Google Scholar]
  10. Kojima, A.; Yoshii, K.I.; Hada, T.; Saeki, O.; Ichikawa, K.; Yoshida, Y.; Shimura, Y.; Azuma, K. Development of High Toughness Steel Plates for Box Columns with High Heat Input Welding. Shinnittetsu Giho. 2004, 90, 39–44. [Google Scholar]
  11. Kojima, A.; Kiyose, A.; Uemori, R.; Minagawa, M.; Hoshino, M.; Nakashima, T.; Ishida, K.; Yasui, H. Super High HAZ Toughness Technology with Fine Microstructure Imparted by Fine Particles. Shinnittetsu Giho. 2004, 380, 2–6. [Google Scholar]
  12. Uemori, R. Analysis of the Crystallographic Direction of the Intra-granular Ferrite in Low Alloy Steel. CAMP-ISIJ 2001, 14, 1174. [Google Scholar]
  13. Kojima, A.; Tomoya, K.; Noboru, H. Effect of Microstructure and Chemical Composition on Laser Weld Metal Toughness. Welded Structure Symposium 2002. Collect. Lect. Pap. 2002, 11, 327. [Google Scholar]
  14. Kojima, A. High HAZ Toughness Steels with Fine Microstructure Imparted by Fine Particles. CAMP-ISIJ 2003, 16, 360. [Google Scholar]
  15. Shigesato, G. Progress of High-Performance Steel Plates with Excellent HAZ Toughness. Nippon. Steel Sumitomo Met. Tech. Rep. 2018, 119, 22–25. [Google Scholar]
  16. Ye, G.; Jönsson, P.; Lund, T. Thermodynamics and Kinetics of the Modification of Al2O3 Inclusions. ISIJ Int. 1996, 36, 105–108. [Google Scholar] [CrossRef]
  17. Ito, Y.I.; Suda, M.; Kato, Y.; Nakato, H.; Sorimachi, K.I. Kinetics of Shape Control of Alumina Inclusions with Calcium Treatment in Line Pipe Steel for Sour Service. ISIJ Int. 1996, 36, S148–S150. [Google Scholar] [CrossRef] [Green Version]
  18. Higuchi, Y.; Numata, M.; Fukagawa, S.; Shinme, K. Inclusion Modification by Calcium Treatment. ISIJ Int. 1996, 36, S151–S154. [Google Scholar] [CrossRef] [Green Version]
  19. Kato, T.; Sato, S.; Ohta, H.; Shiwaku, T. Effects of Ca Addition on Formation Behavior of TiN Particles and HAZ Toughness in Large-Heat-Input-Welding. Kobelco Technol. Rev. 2011, 30, 76–79. [Google Scholar]
  20. Wang, K.; Jiang, M.; Wang, X.; Wang, Y.; Zhao, H.; Cao, Z. Formation Mechanism of CaO-SiO2-Al2O3-MgO Inclusions in Si-Mn-Killed Steel with Limited Aluminum Content during Low Basicity Slag Refining. Metall. Mater. Trans. B 2016, 47, 282–290. [Google Scholar] [CrossRef]
  21. Wang, C.; Wang, X.; Kang, J.; Yuan, G.; Wang, G. Effect of Thermomechanical Treatment on Acicular Ferrite Formation in Ti-Ca Deoxidized Low Carbon Steel. Metals 2019, 9, 296. [Google Scholar] [CrossRef] [Green Version]
  22. Wang, X.; Wang, C.; Kang, J.; Yuan, G.; Misra, R.D.K.; Wang, G. 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 2020, 780, 139198. [Google Scholar] [CrossRef]
  23. Yamamoto, K.; Hasegawa, T.; Takamura, J.I. Effect of Boron on Intra-granular Ferrite Formation in Ti-Oxide Bearing Steels. ISIJ Int. 1996, 36, 80–86. [Google Scholar] [CrossRef]
  24. Mabuchi, H.; Uemori, R.; Fujioka, M. The Role of Mn Depletion in Intra-Granular Ferrite Transformation in the Heat Affected Zone of Welded Joints with Large Heat Input in Strucutral Steels. ISIJ Int. 1996, 36, 1406–1412. [Google Scholar] [CrossRef]
  25. Fukunaga, K.; Chijiiwa, R.; Watanabe, Y.; Kojima, A.; Nagai, Y.; Mamada, N.; Adachi, T.; Date, A.; Taniguchi, S.; Uemori, R.; et al. Advanced Titanium Oxide Steel with Excellent HAZ Toughness for Offshore Structures. In Proceedings of the ASME 2010 29th International Conference on Ocean, Offshore and Arctic Engineering OMAE2010, Shanghai, China, 6–11 June 2010; pp. 1–9. [Google Scholar]
Figure 1. The measured temperature curve of welding thermal simulation.
Figure 1. The measured temperature curve of welding thermal simulation.
Metals 12 00511 g001
Figure 2. The number density of inclusions at different stages.
Figure 2. The number density of inclusions at different stages.
Metals 12 00511 g002
Figure 3. Oxide component in the inclusions at different stages.
Figure 3. Oxide component in the inclusions at different stages.
Metals 12 00511 g003
Figure 4. Morphologies of typical inclusions with different sizes after Ti addition.
Figure 4. Morphologies of typical inclusions with different sizes after Ti addition.
Metals 12 00511 g004
Figure 5. Oxide component proportion in different sized CaO-Al2O3-SiO2-(MgO)-TiOx inclusions.
Figure 5. Oxide component proportion in different sized CaO-Al2O3-SiO2-(MgO)-TiOx inclusions.
Metals 12 00511 g005
Figure 6. Pseudo-ternary phase diagrams CaO-Al2O3-TiO2 at fixed 20%SiO2 (14.8–23.2% SiO2) and 5%MgO (2.7–7.8%MgO); (a) experimental data; (b) average data.
Figure 6. Pseudo-ternary phase diagrams CaO-Al2O3-TiO2 at fixed 20%SiO2 (14.8–23.2% SiO2) and 5%MgO (2.7–7.8%MgO); (a) experimental data; (b) average data.
Metals 12 00511 g006
Figure 7. The morphologies of CaO-Al2O3-SiO2-(MgO)-TiOx at different stages: (a) Ti addition; (b) Ca additon; (c) LF end; (d) VD; (e) Tundish; (f) product.
Figure 7. The morphologies of CaO-Al2O3-SiO2-(MgO)-TiOx at different stages: (a) Ti addition; (b) Ca additon; (c) LF end; (d) VD; (e) Tundish; (f) product.
Metals 12 00511 g007
Figure 8. The formation and evolution mechanism of CaO-Al2O3-SiO2-(MgO)-TiO2 inclusion during the whole refining process.
Figure 8. The formation and evolution mechanism of CaO-Al2O3-SiO2-(MgO)-TiO2 inclusion during the whole refining process.
Metals 12 00511 g008
Figure 9. Two morphology types of CaO-Al2O3-SiO2 inclusions. (a) CaO-Al2O3-SiO2, (b) CaO-SiO2.
Figure 9. Two morphology types of CaO-Al2O3-SiO2 inclusions. (a) CaO-Al2O3-SiO2, (b) CaO-SiO2.
Metals 12 00511 g009
Figure 10. CaO-Al2O3-SiO2 ternary phase diagrams.(a) experimental data, (b) average data.
Figure 10. CaO-Al2O3-SiO2 ternary phase diagrams.(a) experimental data, (b) average data.
Metals 12 00511 g010
Figure 11. The morphologies of CaO-Al2O3-SiO2 inclusions at different stages.
Figure 11. The morphologies of CaO-Al2O3-SiO2 inclusions at different stages.
Metals 12 00511 g011
Figure 12. The morphologies of typical Ti-oxide-containing inclusions in the final products.
Figure 12. The morphologies of typical Ti-oxide-containing inclusions in the final products.
Metals 12 00511 g012
Figure 13. The element distribution of typical multiphase inclusion.
Figure 13. The element distribution of typical multiphase inclusion.
Metals 12 00511 g013
Figure 14. The element distribution of typical multiphase inclusion.
Figure 14. The element distribution of typical multiphase inclusion.
Metals 12 00511 g014
Figure 15. The morphology of a large-sized multiphase inclusion.
Figure 15. The morphology of a large-sized multiphase inclusion.
Metals 12 00511 g015
Figure 16. The morphologies of three typical inclusions after welding simulation.
Figure 16. The morphologies of three typical inclusions after welding simulation.
Metals 12 00511 g016
Figure 17. EDS mapping analysis of one typical inclusion A effective for IAF nucleation in Ti-Ca deoxidized steel.
Figure 17. EDS mapping analysis of one typical inclusion A effective for IAF nucleation in Ti-Ca deoxidized steel.
Metals 12 00511 g017
Figure 18. EDS mapping analysis of one typical inclusion B effective for IAF nucleation in Ti-Ca deoxidized steel.
Figure 18. EDS mapping analysis of one typical inclusion B effective for IAF nucleation in Ti-Ca deoxidized steel.
Metals 12 00511 g018
Figure 19. EDS mapping analysis of one typical inclusion C effective for IAF nucleation in Ti-Ca deoxidized steel.
Figure 19. EDS mapping analysis of one typical inclusion C effective for IAF nucleation in Ti-Ca deoxidized steel.
Metals 12 00511 g019
Table 1. Chemical composition of samples taken from different stages (wt%).
Table 1. Chemical composition of samples taken from different stages (wt%).
No.CSiMnPSNbVAlOTiCa
LF begin0.0250.0891.290.0130.025--0.00620.0110--
Ti added0.0320.2011.310.0130.00700.020.0390.00650.00680.0042-
Ca added0.0480.2001.430.0130.00340.020.0390.00580.00310.0150.0015
VD0.0660.2141.470.0130.00220.0220.0430.00520.00350.0100.0026
Tundish0.0650.2161.470.0130.00210.0230.0430.00470.00320.0100.0016
TMCP0.0690.2221.520.0130.00210.0230.0440.00400.00310.0140.0014
Table 2. Classification of inclusions based on the chemical compositions (wt%).
Table 2. Classification of inclusions based on the chemical compositions (wt%).
InclusionsChemical Composition
CaOAl2O3SiO2MgOTiO2MnO
Al2O3-MnO<10%>60%<10%<10%<10%>10%
Al2O3-SiO2-MnO<10%>10%>20%<10%<10%>10%
CaO-Al2O3-SiO2-MnO>10%>10%>10%<10%<10%>10%
Al2O3-MnO-TiOx<10%>20%<10%<10%>30%>10%
SiO2-MnO<10%>30%<10%<10%<10%>30%
CaO-SiO2>40%<10%>10%<10%<10%<10%
CaO-Al2O3-SiO2>20%>20%>20%<10%<10%<10%
CaO-Al2O3-SiO2-TiOx>20%>10%>10%<10%>10%<10%
CaO-TiOx>30%<10%<10%<10%>30%<10%
Table 3. Chemical compositions of CaO-Al2O3-SiO2 inclusions.
Table 3. Chemical compositions of CaO-Al2O3-SiO2 inclusions.
No.CaOAl2O3SiO2MgOTiO2MnOFeOS
a40.625.520.52.32.90.27.00.8
b59.22.831.30.30.90.25.10.1
Table 4. The chemical composition of typical Ti-oxide-containing inclusions in the final products.
Table 4. The chemical composition of typical Ti-oxide-containing inclusions in the final products.
No.CaOAl2O3SiO2MgOTiO2MnOFeOS
122.715.70.76.615.316.112.111.0
227.719.26.99.222.82.86.64.7
Table 5. The chemical composition of each part of the large-sized multiphase inclusion.
Table 5. The chemical composition of each part of the large-sized multiphase inclusion.
No.CaOAl2O3SiO2MgOTiO2MnOFeOSComposition
140.529.224.93.00.30.12.40.0CaO-Al2O3-SiO2
240.31.40.50.354.70.02.90.0CaO-TiOx
353.40.10.00.30.06.79.130.3(Ca, Mn) S
415.146.23.219.60.92.88.13.6CaO-Al2O3-MgO
553.40.227.61.90.10.516.40.1CaO-SiO2
Table 6. The chemical composition of typical inclusions after welding simulation.
Table 6. The chemical composition of typical inclusions after welding simulation.
No.CaOAl2O3SiO2MgOTiO2MnOFeOS
113.429.38.024.819.11.610.60.1
211.716.516.320.830.11.73.30.2
33.718.20.33.836.13.237.31.0
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rong, Z.; Liu, H.; Zhang, P.; Wang, F.; Wang, G.; Zhao, B.; Tang, F.; Ma, X. The Formation Mechanisms and Evolution of Multi-Phase Inclusions in Ti-Ca Deoxidized Offshore Structural Steel. Metals 2022, 12, 511. https://doi.org/10.3390/met12030511

AMA Style

Rong Z, Liu H, Zhang P, Wang F, Wang G, Zhao B, Tang F, Ma X. The Formation Mechanisms and Evolution of Multi-Phase Inclusions in Ti-Ca Deoxidized Offshore Structural Steel. Metals. 2022; 12(3):511. https://doi.org/10.3390/met12030511

Chicago/Turabian Style

Rong, Zhe, Hongbo Liu, Peng Zhang, Feng Wang, Geoff Wang, Baojun Zhao, Fengqiu Tang, and Xiaodong Ma. 2022. "The Formation Mechanisms and Evolution of Multi-Phase Inclusions in Ti-Ca Deoxidized Offshore Structural Steel" Metals 12, no. 3: 511. https://doi.org/10.3390/met12030511

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