首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Abstract

The high solidification cracking susceptibility of low C steel weld metals was investigated using pure Fe model alloys containing 0–0·23%C, 0–5%Ni and 0–0·0144%B. In addition, a few Fe–C–Ni ternary alloys were also tested. Solidification cracking susceptibility was tested using longitudinal varestraint and transvarestraint tests. Cracking was evaluated using crack length and brittleness temperature range criteria. The Fe–C alloys showed high cracking tendency in two regimes, the first in the ultralow carbon range of 0·03–0·05%C and the second in a narrow band close to 0·1%C. The cracking was much more than that attributable to solute segregation. In Fe–Ni and Fe–B alloys, cracking was a function of alloy content. Solidification cracking due to C and Ni was higher in the ferritic mode of solidification compared to the austenitic, unlike in stainless steels, where the ferritic mode provides high resistance to cracking. In Fe-C-Ni ternary alloys, cracking could be better related to composition in terms of a variable coefficient for C in the Ni equivalent. In the vicinity of 0·1%C, cracking was attributable to shrinkage due to solid state transformation from δ to γ in the brittle temperature range, and is similar to that occurring during continuous casting of steel. However, this factor did not appear to play a role in cracking in the ultralow C range of 0·03–0·05%C.  相似文献   

2.
Abstract

In order to investigate the solidification cracking susceptibilities of SUS347H weld metal containing high niobium and carbon contents, the crystallisation behaviours of niobium carbide as well as those of and γ phases during welding solidification were numerically analysed to compare with the results from in situ observations of the behaviour of the phase selection for γ, and niobium carbide phases for Fe–18Cr–0·2C–1Nb–(5–12)Ni weld metals. The in situ observation was carried out during tungsten inert gas welding using synchrotron radiation. On the other hand, to evaluate the segregation in liquid phase, which has significant influences on the solidification cracking susceptibility, a numerical model to calculate the segregation in liquid phase considering the crystallisation of γ, and niobium carbide phases was developed. The developed numerical model was verified by the comparison with experimental results from the in situ observation. It was suggested that the solidification cracking susceptibilities of Fe–18Cr–0·2C–1Nb–(5–12)Ni weld metals estimated from the results of the developed numerical model coincided with the experimental results evaluated by transverse Varestraint test.  相似文献   

3.
Abstract

When austenitic high-alloy steel weld metal sustains single-phase solidification generally described as A-mode solidification, this is well-known to result in heightened solidification cracking susceptibility.1–3 To reduce the solidification cracking susceptibility of austenitic stainless steel, it is known to be effective to undertake component modification such as to obtain a solidification mode called the AF mode or FA mode involving the ä phase being crystallized or retained.1, 2 To obtain a complete γ solidification mode in the case of high Nibase alloys, such as Fe–36%Ni alloy, however, it is necessary to arrange for high Cr addition in order to achieve component modification facilitating AF or FA mode solidification such as affects austenitic stainless steel. The result of such addition, however, is that impairment of base metal properties also self-evidently cause heavy loss of hot workability in a way that makes this approach difficult to describe as effective.  相似文献   

4.
Peraldi  R.  Pint  B. A. 《Oxidation of Metals》2004,61(5-6):463-483
Ferritic and austenitic model alloys with various contents of Cr and Ni ranging between 10–20% and 0–30%, respectively, were oxidized in air + 10% water vapor during 1 hr cyclic oxidation at 650°C and 800°C. Depending on the alloy composition and temperature, either a thin protective oxide scale was observed or accelerated attack occurred which sometimes included spallation. For austenitic model alloys, increasing either the Cr or Ni contents delayed the accelerated attack. For lower Cr and Ni contents at 800°C, accelerated attack, including spallation, occurred at short exposure times. No spallation was observed for the ferritic model alloys. However, accelerated attack can occur quickly with low Cr contents. Increasing the temperature delayed the breakaway observed on ferritic alloys whereas it reduced the protective-oxide-growth stage for austenitic alloys.  相似文献   

5.
Abstract

Studies on the weldability of 17-4PH stainless steel, in the 621°C overaged condition, showed that Creq/Nieq ratio higher than 1·5 resulted in primary ferritic mode of solidification in the weld metal. Post-weld aging treatment at 482°C enhanced the strength of the weld joint with corresponding reduction in impact toughness of the weld metal while post-weld aging at 621°C caused marginal reduction in strength of the weld joint with significant increase in impact toughness of the weld metal.  相似文献   

6.
P91 ferritic‐martensitic steel, 17Cr–13Ni and alloy 800 austenitic stainless steels and Inconel 617 alloy have been aluminised to form Fe2Al5, (Fe,Ni)Al and Ni2Al3 aluminide coatings. These alloys and their corresponding coatings were subjected to corrosion in air by 50:50 mol/mol K2SO4/KCl deposits at 650 °C for 300 h. With the exception of the Inconel 617 alloy, significant metal losses (>180 µm) were recorded. These losses were planar for P91 alloy but involved internal corrosion for the two austenitic steels. The (Fe,Ni)Al and NiAl coatings on the austenitic steels and the Inconel 617 alloy were significantly corroded via intergranular and internal chloridation–sulphidation–oxidation. In contrast, the Fe2Al5 coating on the P91 alloy coating was virtually unattacked. For the alloys, the relative extents of corrosion damage can be explained in terms of the stability and volatility of metal chlorides formed. For the coatings, STEM/EDS analyses enable clear linkages to be made between the presence and number of Cr‐rich particles on coating grain boundaries and the corrosion damage observed for the coatings.  相似文献   

7.
Abstract

The pitting corrosion resistance of Fe18Cr10Mn(0·33–0·69)N, Fe18Cr10Mn1Ni(0·33–0·84)N, and Fe18Cr10Mn0·35N(0–3)Ni alloys were investigated. The pitting potential increased as the N content increased in both Fe18Cr10Mn(0·33–0·69)N and Fe18Cr10Mn1Ni(0·33–0·84)N alloys. The rise in the pitting potential was more pronounced in Fe18Cr10Mn1Ni(0·33–0·84)N alloys than in Fe18Cr10Mn(0·33–0·69)N alloys. However, it was found that Ni alone had no effect on the pitting corrosion resistance of Fe18Cr10Mn0·35N based alloys. Thus, it was concluded that the alloyed N worked synergistically with Ni to promote the pitting corrosion resistance in Fe18Cr10Mn based alloys. Analyses of passive films of Fe18Cr10Mn(0·33–0·69)N and Fe18Cr10Mn1Ni(0·33–0·84)N alloys revealed that N was incorporated into the passive film, with N enriched at the film/metal interface. However, the alloyed N increased the Cr cation fraction in passive films of Fe18Cr10Mn1Ni(0·33–0·84)N alloys, whereas N decreased in that of Fe18Cr10Mn(0·33–0·69)N alloys. This difference was considered as the reason for the synergistic effect between N and Ni in Fe18Cr10Mn based alloys.  相似文献   

8.
Abstract

The effect of addition of La to a filler metal on microcracking (ductility dip cracking) in the multipass weld metal of alloy 690 was investigated with the aim of improving its microcracking susceptibility. The susceptibility to ductility dip cracking in the reheated weld metal could be greatly improved by adding 0·01–0·02 wt-%La to the weld metal. Conversely, excessive La addition to the weld metal led to liquation and solidification cracking in the weld metal. Hot ductility of the weld metal at the cracking temperature was greatly improved by adding 0·01–0·02 wt-%La to the weld metal, implying that the ductility dip cracking susceptibility was decreased as a result of the desegregation of impurity elements of P and S to grain boundaries due to the scavenging effect of La. The liquation and solidification cracking resulting from excessive addition of La to the weld metal is attributed to the formation of liquefiable Ni–La intermetallic compound. A multipass welding test confirmed that microcracks in the multipass weldment were completely prevented by using a filler metal containing an addition of 0·01 wt-%La.  相似文献   

9.
The high temperature oxidation behaviour of FeCrNi austenitic alloys containing 1% Ti which, in some cases, had been converted into an oxide dispersion has been examined. The oxide dispersions were produced by an internal oxidation treatment using a 50/50 Cr/Cr2O3 powder mixture in a sealed quartz capsule at 1100°C: the samples were not in direct contact with the powders. Generally, the effect of the dispersed oxide was much less pronounced than in corresponding nickel-free, ferritic alloys. Nevertheless, the time-to-breakaway of the protective Cr3O3 scale which developed on Fe18CrNi alloys was substantially increased, although the differences between the untreated and the internally oxidized alloys reduced with increasing nickel content. An Fe14Cr20Ni alloy did not show any improvement after internal oxidation. Unlike the ferritic alloys, no coarsening of the dispersoid phase was observed during exposure.  相似文献   

10.
Abstract

The brittle to ductile transition (BDT) in nickel free high nitrogen austenitic stainless steel was investigated. Falling weight impact tests at 176, 273 and 336 K revealed that Fe–25Cr–1·1N (wt-%) austenitic steel exhibits a sharp BDT in spite of being a face centred cubic alloy. The plastic deformation observed following the impact tests indicated that the BDT is induced by poor ductility at low temperatures, as is the case with ferritic steels. To measure the activation energy for the BDT, the strain rate dependence of the BDT temperature was examined using four-point bending tests. The BDT temperature was found to be weakly dependent on strain rate. Arrhenius plots of the BDT temperature against strain rate showed that the activation energy for the BDT of Fe–25Cr–1·1N steel is much higher than that of low carbon ferritic steels. The origins of this distinctive BDT and the large value for its activation energy in this high nitrogen steel are discussed in terms of the reduction in dislocation mobility at low temperatures because of the interactions between the glide dislocations and the solute nitrogen atoms.  相似文献   

11.
The first layer of the deposit on the edges of low-alloy steels in producing welded joints with 08Cr18Ni10Ti steel is made using materials based on 02Cr24Ni13 composition (TsL-25L electrodes, welding wire SV-02Cr24Ni13) producing the deposited metal with a reduced carbon content and the required content of the ferritic phase (2–5%). Welding with these materials results in the required parameters of technological properties in welding and efficiency of the austenitic–ferritic deposited metal:
  • resistance to solidification cracking;

  • preventing the formation of a structure containing very hard brittle compounds;

  • preventing the formation of the sigma phase and, correspondingly, embrittlement in tempering;

  • the mechanical properties and fatigue strength satisfy the requirements of PNAE G-7-002–86.

Technical documents for the production of these welding materials have been compiled.  相似文献   

12.
Abstract

Solidification structure of a weld in a manual gas tungsten arc welding of 3 mm thick Ti–5Ta–1·8Nb plates has been examined using optical/scanning electron microscopy, EDX spectroscopy and electron backscattered diffraction techniques. The weld exhibited columnar grains of prior β, with a substructure of martensitic α′. The backscattered electron image revealed the solidification mode to be cellular–dendritic, and the EDX analysis confirmed the partitioning of Ta atoms between the dendrites. The crystal orientation of the product α′ obtained by electron backscattered diffraction and the Burgers orientation relationship (for βα′ transformation) were used to predict the orientation of parent β grains in the weld zone, and the long dendritic arms were found to be aligned parallel to β<100>. Analytical solution based on Rosenthal equation was used to calculate the temperature cycle of the welding process. The solidification mode predicted from calculated G/R ratio (where G is solidification gradient and R is rate of solidification) and Kurz–Fischer map matched with the observed cellular–dendritic solidification structure.  相似文献   

13.
Zhang  Y.  Pint  B. A.  Haynes  J. A.  Tortorelli  P. F. 《Oxidation of Metals》2004,62(1-2):103-120
The oxidation behavior of iron-aluminide coatings, Fe3Al or (Fe,Ni)3Al, produced by chemical-vapor deposition (CVD) was studied in the temperature range of 700–800°C in air + 10 vol.% H2O. A typical ferritic steel, Fe–9Cr–1Mo, and an austenitic stainless steel, 304L, were coated. For both substrates, the as-deposited coating consisted of a thin (<5μm), Al-rich outer layer above a thicker (30–50 μm), lower-Al-content inner layer. In addition to coated and uncoated Fe–9Cr–1Mo and 304L, cast Fe–Al model alloys with similar Al contents (13–20 at.%) to the CVD coatings were included in the oxidation exposures for comparison. The specimens were cycled to 1000 1 hr cycles at 700°C and 500 1 hr cycles at 800°C, respectively. The CVD coating specimens showed excellent performance in the water-vapor environment at both temperatures, while the uncoated alloys were severely attacked. These results suggest that an aluminide coating can substantially improve resistance to water-vapor attack under these conditions.  相似文献   

14.
This paper describes in detail the oxidation of two cast stainless steels in synthetic diesel and gasoline exhaust gases. One alloy was ferritic (Fe18Cr1.4Nb2.1Mn0.32C) and one austenitic (Fe20Cr9Ni1.9Nb2.7W0.47C). Polished sections were exposed, mostly for 50 h, at temperatures between 650 and 1,050 °C. The oxidation product was characterized by means of SEM/EDX, AES, and XRD. Inter-dendritic non-Cr carbides initiated thick oxides. The ferritic steel formed a rather thin and adherent oxide scale at all temperatures. It consisted of (Mn, Cr) oxide on top of Cr2O3 and, starting at 850 °C, a thin silica film at the metal–oxide interface. Chromium depletion triggered dissolution of carbides providing Cr to the oxide. Water vapor did not accelerate the attack since the outer (Mn, Cr) spinel oxide reduced the Cr evaporation. The austenitic grade was very sensitive to water vapor. Chromium segregation directed pitting to the dendrites up to 950 °C whereas uniform catastrophic oxidation occurred at 1,050 °C.  相似文献   

15.
In order to clarify the effect of tip velocity on the weld solidification process of hot-work tool steel (SKD61) during welding, information about microstructure evolution was obtained by the combination of a liquid tin quenching and time resolved X-ray diffraction technique using intense synchrotron radiation. From the experimental results, it was found that the solidification mode was changed from FA mode (L → L+δ → L+δ+γ → L+γ → γ) to A mode (L → L+γ → γ) at high tip velocity. Moreover, the effect of tip velocity on the microstructure selection during solidification between the primary δ, ferrite and the primary γ, austenite was theoretically proven by the Kurz, Giovanola and Trivedi model. Therefore, it was understood that the solidification cracking susceptibility of hot-work tool steel (SKD61) weld metal was increased due to the δ to γ transition of the primary phase.  相似文献   

16.
Metal dusting, i.e. disintegration into fine metal particles and carbon, was induced on a selection of chromia forming high temperature alloys in a flowing CO-H2-H2O atmosphere in exposures at 650°C, 600°C, 500°, and 450°C. The materials were pretreated by annealing in H2 at 1000°C and electropolishing, this leads to large grain size and low surface deformation, both is disadvantageous for formation of a Cr2O3 scale. The resistance to metal dusting is only dependent on the ability to form a protective Cr2O3 scale, thus the high Cr ferritic steels proved to be very resistant, the ferritic steels with 12–13% Cr were less resistant. Due to the lower Cr diffusivity in the austenitic steels, these were very susceptible, especially two alloys with about 30% Ni (Alloy 800, AC 66). The appearance of metal dusting was somewhat different for Ni-base materials but they were also attacked under pitting. The metal dusting is preceded in all cases by internal carburization whereby the chromium is tied up, afterwards the remaining Fe or Fe-Ni matrix can react to the instable intermediate carbide M3C which decomposes to metal particles and carbon, in case of Ni-base materials a supersaturated solid solution of carbon is the intermediate.  相似文献   

17.
The extended use of powder metal components can be improved by the use of welding joining methods. This work investigates the weldability of iron-based powder metal alloys (Fe–Ni, Fe–Ni–P alloys) using the pulsed gas tungsten arc welding process (GTAW) with three different filler metals (AWS R 70S-6, AWS R 309L, AWS R Fe–Ni). Results revealed that the Fe–Ni powder metal alloy does not present any metallurgical difficulty concerning the weldability for all types of filler metal studied. The Fe–Ni–P powder metal alloy, microstructural examinations showed that, despite its high content of phosphorus (0.25 wt%), the utilization of pulsed GTAW process with stainless steel 309L filler metal resulted in welds free of porosities and solidification cracks. Metallographics examinations suggest that the absence of solidification cracks in this alloy can be mainly attributed to the presence of delta ferrite in the stainless steel weld metal which absorbed part of the phosphorus and significantly reduced the formation of the Fe3P low-melting eutectic in the weld pool during cooling. In contrast, solidification cracks were observed when joining the Fe–Ni–P powder metal alloy using RFe–NI and R70S-6 filler metals. Hardness tests carried out indicated a heat affected zone (HAZ) with no excessive hardening for all alloys studied. Furthermore, tensile tests showed that the fractures always occurred in the base metal with tensile strength slightly superior to the value of unwelded samples. As a result, this investigation showed the feasibility of joining iron-based powder metal alloys by the pulsed GTAW process since a rigid control of the heat input is implemented together with an adequate choice of the filler metal, especially when welding the Fe–Ni–P alloy.  相似文献   

18.
Abstract

In order to clarify the effect of solidification velocity on the weld solidification process of alloy tool steel during the welding, the information about microstructure evolution was obtained by the concurrent experiments of liquid tin quenching and time resolved X-ray diffraction technique using intense synchrotron radiation. It was found from the experiments that the solidification mode was transferred from an FA to an A mode at the high solidification velocity. The effect of solidification velocity on the phase selection during solidification between the primary δ-ferrite and γ-austenite was theoretically proved by the Kurz, Giovanola and Trivedi (KGT) model. It is thus explained that the solidification cracking susceptibility of the weld metal of alloy tool steel was enhanced due to the δ to γ transition of the primary phase.  相似文献   

19.
The effect of water vapor on the transition from internal to external oxidation of austenitic alloys has been conducted at 1,073 K under the equilibrium oxygen partial pressure for the coexistence of Fe and FeO. Critical Cr concentrations in the Fe–Cr–30Ni (at.%) austenitic alloys were determined to be 30 at.% in dry atmosphere and 37 at.% in humid atmosphere. Thus, water vapor significantly affected the transition from internal to external oxidation of austenitic alloys. Two oxides of Cr2O3 and FeCr2O4 precipitated in the Fe–5Cr–30Ni (at.%) alloy and solid state reaction for the formation of FeCr2O4 may be influenced by water vapor. Oxygen permeability, which was estimated by considering the effective stoichiometric ratio, was also enhanced by water vapor.  相似文献   

20.
Metal dusting of Fe–Ni alloys was investigated in a CO–H2–H2O–Ar gas corresponding to a C = 19.6 at 650 °C. Thermogravimetric analysis showed that increasing the nickel content in the alloy decreased the initial rate of carbon uptake. A uniform Fe3C scale formed on pure iron, a layer with mixed structures of Fe3C, γ and α-Fe developed on ferritic Fe–5Ni, and small amounts of Fe3C developed at the surface of an austenite layer grown on two-phase (α + γ) Fe–10Ni. At nickel levels above 10%, no carbide appeared. These observations are shown to be broadly consistent with local equilibrium according to the Fe–Ni–C phase diagram. However, the failure of higher nickel austenitic alloys to form the (Fe,Ni)3C expected at high carbon activities indicates a barrier to nucleation and growth of this phase. Graphite deposition was catalysed by (Fe,Ni)3C on ferritics and by the metal itself on austenitics. The rates of carbon deposition on Fe–60Ni corresponded to the existence of three parallel and independent paths: the synthesis gas, the Boudouard and the carbon methanation reactions.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号