Next Article in Journal
Characteristics of the Dissimilar AA7075 and Q235 Steel Joints Fabricated by Friction Stir Welding
Previous Article in Journal
Prediction of Spherical Sheet Springback Based on a Sparrow-Search-Algorithm-Optimized BP Neural Network
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phase Transformations of 5Cr-0.5Mo-0.1C Steel after Heat Treatment and Isothermal Exposure

by
Maribel L. Saucedo-Muñoz
1,*,
Victor M. Lopez-Hirata
1,
Hector J. Dorantes-Rosales
1,
Jose D. Villegas-Cardenas
2,
Diego I. Rivas-Lopez
1,
Manuel Beltran-Zuñiga
1,
Carlos Ferreira-Palma
3 and
Joel Moreno-Palmerin
4
1
Instituto Politecnico Nacional (ESIQIE), Ciudad de Mexico 07300, Mexico
2
Universidad Politecnica del Valle de Mexico, Tultitlán 54900, Mexico
3
Universidad Veracruzana (FCQ), Boca del Río 94294, Mexico
4
Departamento de Ingeniería en Minas, Metalurgia y Geología, Universidad de Guanajuato, Guanajuato 36000, Mexico
*
Author to whom correspondence should be addressed.
Metals 2022, 12(8), 1378; https://doi.org/10.3390/met12081378
Submission received: 20 June 2022 / Revised: 27 July 2022 / Accepted: 12 August 2022 / Published: 19 August 2022

Abstract

:
This study consists of the experimental and numerical analysis of the phase transformations of 5Cr-0.5Mo.0.1C steel after heat treatment. The microstructure of the as-received steel comprised ferrite and bainite, which is in agreement with the microconstituents predicted by the Calphad-calculated TTT diagram. Calphad-based precipitation calculations show that the cooling stage of normalizing treatment did not cause carbide formation. In contrast, tempering at 700 °C for 15 min promotes the intergranular precipitation of Fe3C, M7C3 and M23C6 carbides, which is consistent with experimental results. Aging at 600 °C for short periods caused the precipitation of both M7C3 and M23C6 carbides; however, M23C6 is the dominant phase after prolonged aging. This is in agreement with experimental results. A rapid decrease in the steel hardness was observed after short aging, which is attributable to bainite transformation. Further reduction in hardness is associated with the diffusion-controlled coarsening of M23C6 carbide.

1. Introduction

Low-alloyed ferritic steels are widely used for the fabrication of equipment for petroleum refineries and carbon power plants [1,2]. This type of steel exhibits good mechanical properties, including good tensile strength and toughness at room temperature and creep resistance at temperatures up to 600 °C. This type of steel is designated as heat-resistant steel [3]. The low-C Cr-Mo ferritic steels are an example of heat-resistant steels. Higher Cr and Mo contents increase the mechanical strength at high temperatures [4]. Besides solid solution strengthening, these steels also present carbide precipitation hardening. The precipitation hardening treatment of low-carbon 1Cr-0.5 Mo steel relies on the formation of M3C, M7C3 and M23C6 carbides [5]. In contrast, the precipitation hardening of low-carbon 2.25Cr-1Mo steel originates from the presence of M23C6 and Mo2C [6]. The precipitation of carbides is relevant for improving the creep strength, since the alloyed carbides constitute barriers that impede dislocation movement and grain boundary sliding during creep at high temperatures [7,8]. The ferritic low-carbon 5Cr-0.5 Mo also presents good creep strength at temperatures up to 600 °C [8]. These characteristics are achieved by heat treatments and precipitation hardening, which contribute to maintaining their properties at high temperatures. There are few studies on phase transformations during heat treatments and heating at operating temperatures. Therefore, it is relevant to carry out this type of study to better understand the relationship between the microstructure and properties of this steel.
The heat treatment of these steels consists of normalizing and tempering, which allows achieving the required tensile strength and promotes the formation of precipitates to present good creep strength during service operation [1].
Calphad-based software [8,9,10,11,12,13] is beneficial for understanding the phase stability and the precipitation of carbides during the isothermal exposure of low-alloyed steels. This software allows us to estimate the growth kinetics of precipitation and coarsening of precipitates, which are important in the deterioration of creep properties during its service operation at high temperatures in industrial components.
We performed numerical and experimental analyses of the phase transformations after heat treating and the precipitation process after isothermal aging at 600 °C for a low carbon 5Cr-0.5 Mo-0.1C steel to understand their effect on mechanical properties.

2. Materials and Methods

2.1. Numerical Method

With the Calphad-based tools, Thermo-Calc (TC) and TC-Prisma [11,12,13,14,15,16], we established the phase stability using the steel composition of Table 1. The time–temperature–transformation (TTT) diagram and precipitation growth kinetics during the heat-treatment process and isothermal aging at 600 °C were also calculated. The TCFe11 and MobFe6 databases (Thermo-Calc, Stockholm, Sweden) were employed [17], assuming an austenite grain size of 30 µm and grain boundary precipitation. This software was also used to determine the interfacial free energy of the precipitate/matrix.

2.2. Experimental Method

Table 1 presents the chemical composition of the steel. Specimens were cut to 10 mm × 10 mm × 10 mm dimensions. The initial condition of the steel, referred to as “as-received”, was obtained after normalizing at 950 °C for 30 min and tempering at 715 °C for 15 min. Additionally, samples were aged at 600 °C for periods between 0 and 7000 h. We used a scanning electron microscope (SEM, JEOL Ltd., Tokyo, Japan) equipped with EDX for the microstructural characterization of as-received and aged samples. Electrolytic dissolution of the ferrite phase with 2 vol % HNO3 in CH3-OH at 4 V (dc) enabled the extraction of the precipitates from the ferrite matrix. The residues were collected, rinsed with water and subsequently dried. The X-ray diffraction (XRD, JEOL LTD., Tokyo, Japan) analysis of these residues with a monochromated Cu Kα radiation allowed us to determine the carbide type. These residues were mounted on carbon-coated 300 mesh copper grits, and subsequently analyzed by TEM equipped with EDX at 200 kV. Rockwell “B” hardness of aged specimens was measured according to the ASTM E-18 standard [18]. Grain size was determined using the intercept procedure according to the ASTM E-112 standard [19]. The volume fraction of precipitates was quantified using SEM micrographs at 30,000×, based on a point-counting method consistent with ASTM E-562 standard [20].

3. Results

3.1. Phase Stability and TTT Diagram

Figure 1 shows the variation in the volume fraction of all equilibrium phases against temperature, calculated by Calphad-based software [17]. This figure indicates the formation of austenite from the liquid during cooling. TiC precipitates from the austenite matrix at temperatures below 915 °C. This carbide also contains a small amount of Cr and Mo. These precipitates contribute to control grain size, as presented in subsequent sections. As temperature decreases, austenite transforms into ferrite around 827 °C, the A3 intercritical temperature. These calculations also suggest that the precipitation of Cr-rich M23C6 is possible by aging at temperatures below approximately 810 °C.
The time–temperature–transformation diagram, TTT, calculated by a Calphad-based steel transformation software [17], is shown in Figure 2. The precipitation of ferrite from austenite below the A3 temperature was determined using Calphad-based precipitation software [17]. This diagram suggests that the microstructure in the normalized condition, at a cooling rate close to 1 °C/s, consists of ferrite and bainite. This means that the first phase transformation expected to occur during the cooling is the ferrite precipitation in the austenite matrix at temperatures lower than 820 °C (Figure 1). Ferrite is the dominant microconstituent; however, the formation of some bainite fraction seems possible by the end of the cooling process. Water-quenching with a cooling rate of about 1000 °C/s may produce martensite formation according to the calculated TTT diagram.

3.2. Microstructure Characterization

Figure 3 shows a light microscope LM micrograph of the as-received steel, which indicates that the most evident microconstituents correspond to equiaxed ferrite grains with a size of approximately 30 ± 5 µm. These fine ferrite grains can be attributed to the presence of a small amount of TiC carbides. Some elongated ferrite plates are observed as part of the bainite microconstituent. No pearlite colonies seem to be present in the micrograph. This suggests that the cooling rate of about 1 °C/s was fast enough to avoid the formation of pearlite during the normalizing treatment, as indicated qualitatively in the Calphad-calculated TTT diagram. The continuous cooling transformation (CCT) diagram is more adequate for continuous cooling treatments; nevertheless, the TTT diagram was used because this diagram shows more clearly the bainitic transformation, which occurred during the cooling of the steel. It is important to note that the calculation of TTT utilizes thermodynamic data, contrary to other commercial software [15].
The carbide precipitation is not clearly observable in the optical micrograph of Figure 3. Figure 4a–f presents SEM micrographs for the steel in the as-received condition and after aging at 600 °C for 100, 300, 1000, 6000 and 7000 h. The precipitation of both intragranular and intergranular carbides is evident in the aged steel, as indicated by arrows (Figure 4b,e). This precipitation is more notorious as the aging process advances. The coarsening of carbides is evident for aging times longer than 1000 h.
Figure 5a–c presents the electrolytically extracted precipitates for the specimens after aging at 600 °C for 0 h, 300 and 7000 h, respectively. The carbide size increases with time, and the particle morphology changes from irregular to polygonal. The EDX-SEM analysis indicates that the main component for all the extracted particles is Cr.
The XRD patterns corresponding to the as-received steel and those aged at 600 °C for 100, 300, 1000 and 7000 h are presented in Figure 6. The XRD peaks of extracted precipitates from the as-received specimen are mainly of Cr-rich M7C3 carbides. The M7C3 carbide belongs to the Pmcn (62) space group [21], and the crystals present an irregular morphology, as previously presented in Figure 5a. The XRD peaks of residues from specimens aged for 300 h correspond to a mixture of M7C3 and M23C6 carbides. In the case of steel aged for periods longer than 300 h, the XRD patterns correspond to the M23C6 carbide with the Fm3 m (225) space group [22]. These carbides are faceted crystals. This suggests that the aging process promotes the change from M7C3 to M23C6 carbides. No Fe3C XRD peaks were detected in the diffractogram corresponding to the as-received steel, which confirms its absence, as shown in Figure 3.
Figure 7 and Figure 8 present the TEM micrographs and electron diffraction patterns of a precipitate extracted from the as-received steel and the steel aged for 7000 h, respectively. Figure 7a verifies the irregular morphology of M7C3 carbide, while the electron diffraction pattern, Figure 7b, corresponds to an orthorrombic crystalline structure, Pmcn, with a zone axis ZA [ 2 6 ¯ 1 ] . Conversely, Figure 8a shows the faceted shape morphology of M23C6 carbides, with fcc crystalline structure, Fm3m, indicated by the electron diffraction pattern with a zone axis [ 1 ¯ 11 ] in Figure 8b.

4. Discussion

4.1. Characterization of Precipitation Evolution

The Calphad-based precipitation kinetics calculation for the normalizing process from 950 °C at a cooling rate of about 1 °C/s estimated a maximum volume fraction of 1.5 × 10−10 for the intergranular precipitation of M7C3 carbide after the normalizing process. Neither Fe3C nor M23C6 precipitates were present after normalizing. The calculated precipitation kinetics of the tempering process at 715 °C for 900 s (15 min) is presented in Figure 9. The plot of volume fraction for precipitation clearly shows that the precipitation sequence starts with Fe3C formation, then M7C3 precipitation, and finally, M23C6 precipitation. This indicates that Fe3C is formed and subsequently disappears. This could explain the absence of cementite and the presence of M7C3 and M23C6 carbides in the as-received steel, as observed in Figure 6 and Figure 7. This suggests that the precipitation reaction during tempering is as follows:
αsss → α + Fe3C → α + M7C3 → α + M23C6
On the other hand, the Calphad-calculated precipitation kinetic behavior for aging at 600 °C exhibits that the precipitation of M23C6 carbide is dominant, as shown in Figure 10. A very low volume fraction of M7C3 is also present during short periods, of about 1000 s, and subsequently disappears. This result agrees with the precipitated phases detected in the aged specimens by XRD and TEM in Figure 6 and Figure 8. These results indicate that the precipitation reaction during aging at 600 °C follows the sequence:
αsss → M7C3 → α + M23C6
Figure 11 shows the Calphad-calculated precipitate mean radius against time plot with experimental values of the equivalent radius for M23C6 precipitates. There is a good agreement between the calculated and experimental data. The growth kinetics of precipitation for times longer than 300 h (1.08 × 106 s) correspond to the diffusion-controlled coarsening stage. The calculated interfacial energy was 0.1 Jm−2 between the M23C6 precipitate and ferrite, and 0.28 Jm−2 between the M7C3 precipitate and ferrite. The former value suggests a coherent interface for the M23C6/ferrite interface [23].
This suggests that creep strength is based on the intergranular and intragranular precipitation of Cr-rich M23C6, which impedes the dislocation motion and grain boundary sliding at high temperatures [1,2,3].

4.2. Mechanical Characterization

Figure 12 presents the Rockwell “B” hardness variation with aging time. There is a decreasing tendency as time proceeds. The normalizing treatment is mainly responsible for the initial hardness of 90.5 HRB due to the presence of the ferrite and bainite microconstituents.
The fast decrease in hardness from 90.5 to 86 HRB for short times can be related to the transformation of bainite into the equilibrium phases: ferrite and M23C6. The decrease from 86 to 83.5 HRB after prolonged periods of aging is related to the diffusion-controlled coarsening of M23C6 carbide, even though the precipitate size is still of nanometric order. The Calphad-calculated coarsening constant, k, for M7C3 carbide at 400 and 600 °C is 8.1 × 10−36 and 1.52 × 10−30 m3 s−1, while for M23C6 carbide, it is 3.8 × 10−36 and 1.01 × 10−30 m3 s−1, respectively. These values suggest that the coarsening resistance is higher for the M23C6 carbide, which is the most predominant carbide at temperatures between 400 and 600 °C, as shown by Calphad calculations. This behavior seems to be caused by the lower interfacial energy. This also indicates that the mechanical strength may remain for long aging times at these temperatures.

5. Conclusions

A study of phase transformations of 5Cr-0.5Mo-0.1C steel after heat treating yielded the following conclusions:
  • The Calphad-calculated TTT diagram of the steel precisely predicted the microconstituents observed in the as-received steel.
  • The cooling stage of the normalizing treatment promoted neither intergranular nor intragranular precipitation of carbides.
  • The tempering process at 700 °C originated the precipitation of M7C3 and M23C6 carbides.
  • Aging at 600 °C produced the precipitation of M23C6 carbide, which is responsible for the mechanical strength at high temperatures.
  • The reduction in steel hardness can be attributed to the bainite transformation and diffusion-controlled coarsening of M23C6 carbide.

Author Contributions

Conceptualization, V.M.L.-H. and M.L.S.-M.; methodology, M.L.S.-M.; software, V.M.L.-H.; validation, H.J.D.-R., M.B.-Z. and D.I.R.-L.; formal analysis, M.L.S.-M. and J.M.-P.; investigation, V.M.L.-H. and M.L.S.-M.; writing—original draft preparation, V.M.L.-H., J.D.V.-C. and C.F.-P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Conacyt, grant number A1-S-9682.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon request.

Acknowledgments

The authors acknowledge the financial support from Conacyt A1-S-9682, and SIP-IPN.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gneezy, V.C.; Ugwuegbu, C.C.; Mark, U. Physical metallurgy of modern creep-resistant steel for steam power plants: Microstructure and phase transformations. J. Metall. 2016, 3, 5468292. [Google Scholar] [CrossRef]
  2. Bhadeshia, H.-K.D.H.; Strange, A.; Gooch, D.J. Ferritic power plant steels: Remanent life assessment and approach to equilibrium. Inter. Mater. Rev. 1998, 43, 45–69. [Google Scholar] [CrossRef]
  3. Davis, J.R. ASM Specialty Handbook: Heat-Resistant Materials; ASM International: Materials Park, OH, USA, 1997; pp. 50–75. [Google Scholar]
  4. Bhadeshia, H.K.D.H. Design of ferritic creep-resistant steels. ISIJ Int. 2001, 41, 626–640. [Google Scholar] [CrossRef]
  5. Viswanathan, R. Effect of stress and temperature on the creep and rupture behavior of a 1.25 Pct.Chromium-0.5 Pct. molybdenum steel. Met. Trans. 1977, 8, 877–884. [Google Scholar] [CrossRef]
  6. Saucedo-Muñoz, M. Precipitation kinetics of carbides during cyclical and isothermal aging of 2.25Cr–1Mo steel and its effect on mechanical properties. J. Irons Steel Res. Int. 2021, 28, 1282–1290. [Google Scholar] [CrossRef]
  7. Das, S.; Vora, J.; Patel, V.; Andersson, J.; Pimenov, D.Y.; Giasin, K. Elucidating the effect of step cooling heat treatment on the properties of 2.25 Cr–1.0 Mo steel welded with a combination of GMAW techniques incorporating metal-cored wires. Materials 2021, 14, 14206033. [Google Scholar] [CrossRef] [PubMed]
  8. Matsuyama, F. History of power plants and progress in heat resistant steels. ISIJ Int. 2001, 41, 612–625. [Google Scholar] [CrossRef]
  9. Chen, Q.; Wu, K.; Sterner, G.; Mason, P. Modeling Precipitation Kinetics During Heat Treatment with Calphad-Based Tools. J. Mater. Eng. Perform. 2014, 23, 4193–4196. [Google Scholar] [CrossRef]
  10. Costa e Silva, A. Challenges and opportunities in thermodynamic and kinetic modeling microalloyed HSLA steels using computational thermodynamics. Calphad 2020, 68, 101720. [Google Scholar] [CrossRef]
  11. Saucedo-Muñoz, M.; Ortiz-Mariscal, A.; Lopez-Hirata, V.; Villegas-Cardenas, J.; Soriano-Vargas, O.; Avila-Davila, E.O. Precipitation analysis of as-cast HK40 steel after isothermal aging. Int. J. Min. Metall. Mater. 2017, 24, 1125–1133. [Google Scholar] [CrossRef]
  12. Lopez-Hirata, V.M.; Hernandez-Santiago, F.; Saucedo-Muñoz, M.L.; Dorantes-Rosales, H.J.; Paniagua-Mercado, A.M. Analysis of β´(Cu4Ti) Precipitation during isothermal aging of a Cu-4wt.%Ti alloy. Mater. Res. 2018, 21, 1–7. [Google Scholar] [CrossRef]
  13. Thermo-Calc. Thermo-Calc Prisma Software TCFE 11 and MOBFE6 Steels/ Fe Alloys Databases. Available online: https://thermocalc.com/content/uploads/Documentation/Databases/Mobility/mobfe7-technical-info.pdf (accessed on 21 June 2022).
  14. Yan, J.; Agren, J.; Jeppsson, J. Pearlite in multicomponent steels: Phenomenological steady-state modeling. Met. Mater. Trans. 2020, 51, 1978–2001. [Google Scholar] [CrossRef]
  15. Stormvinter, A.; Borgenstam, A.; Agren, J. Thermodynamically based prediction of the martensite start temperature for commercial steels. Met. Mater. Trans. 2012, 43, 3870–3879. [Google Scholar] [CrossRef]
  16. Leach, L.; Kolmskog, P.; Hönglund, L.; Hillert, M.; Borgenstam, A. Critical driving forces for formation of bainite. Met. Mater. Trans. 2018, 49, 4509–4520. [Google Scholar] [CrossRef]
  17. Thermo-Calc. Thermo-Calc Software TCFE Steels 11/ Fe Alloys Database. Available online: https://thermocalc.com/content/uploads/Documentation/Databases/Thermodynamic/tcfe12-technical-info.pdf (accessed on 21 June 2022).
  18. ASTM Standard E 18-15; Standard Test Methods for Rockwell Hardness of Metallic Materials. ASTM: West Conshohocken, PA, USA, 2016.
  19. ASTM E 112-19; Standard Test Methods for Determining Average Grain Size. ASTM: West Conshohocken, PA, USA, 2019.
  20. ASTM E 562-19; Standard Test Method for Determining Volume Fraction by Systematic Manual Point Count. ASTM: West Conshohocken, PA, USA, 2019.
  21. Rouault, M.A.; Herpin, P.; Fruchart, M.R. Etude cristallographique des carbures Cr7 C3 et Mn7 C3. Ann. Chim. 1970, 5, 461–470. [Google Scholar]
  22. Bowman, A.L.; Arnold, G.P.; Storms, E.K.; Nereson, N.G. The crystal structure of Cr23 C6. Acta Cryst. B 1972, 28, 3102–3103. [Google Scholar] [CrossRef]
  23. Kostorz, K. Phase Transformations in Materials, 2nd ed.; Wiley-VCH: Weinheim, Germany, 2001; pp. 309–407. [Google Scholar]
Figure 1. Plot of volume fraction of equilibrium phases against temperature.
Figure 1. Plot of volume fraction of equilibrium phases against temperature.
Metals 12 01378 g001
Figure 2. TTT diagram of the steel.
Figure 2. TTT diagram of the steel.
Metals 12 01378 g002
Figure 3. LM micrograph of the as-received steel.
Figure 3. LM micrograph of the as-received steel.
Metals 12 01378 g003
Figure 4. SEM micrographs of (a) the as-received steel and after aging at 600 °C for (b) 100, (c) 300, (d) 1000, (e) 6000 and (f) 7000 h.
Figure 4. SEM micrographs of (a) the as-received steel and after aging at 600 °C for (b) 100, (c) 300, (d) 1000, (e) 6000 and (f) 7000 h.
Metals 12 01378 g004
Figure 5. SEM micrograph of extracted residues of the steel after aging at 600 °C for (a) 0, (b) 300 and (c) 7000 h.
Figure 5. SEM micrograph of extracted residues of the steel after aging at 600 °C for (a) 0, (b) 300 and (c) 7000 h.
Metals 12 01378 g005
Figure 6. XRD patterns of the extracted residues from the steel specimen after aging at 600 °C for 0 (as-received), 100, 300 and 7000 h.
Figure 6. XRD patterns of the extracted residues from the steel specimen after aging at 600 °C for 0 (as-received), 100, 300 and 7000 h.
Metals 12 01378 g006
Figure 7. (a) TEM micrograph of M7C3 carbide with (b) electron diffraction pattern with ZA = [ 1 ¯ 11 ] .
Figure 7. (a) TEM micrograph of M7C3 carbide with (b) electron diffraction pattern with ZA = [ 1 ¯ 11 ] .
Metals 12 01378 g007
Figure 8. (a) TEM micrograph of M23C6 carbide with (b) electron diffraction pattern with ZA =   [ 2 6 ¯ 1 ] .
Figure 8. (a) TEM micrograph of M23C6 carbide with (b) electron diffraction pattern with ZA =   [ 2 6 ¯ 1 ] .
Metals 12 01378 g008
Figure 9. Calphad-calculated volume fraction of precipitates vs. time during aging at 715 °C.
Figure 9. Calphad-calculated volume fraction of precipitates vs. time during aging at 715 °C.
Metals 12 01378 g009
Figure 10. Calphad-calculated volume fraction of precipitates vs. time during aging at 600 °C.
Figure 10. Calphad-calculated volume fraction of precipitates vs. time during aging at 600 °C.
Metals 12 01378 g010
Figure 11. Calphad-calculated mean radius of precipitates vs. time during aging at 600 °C.
Figure 11. Calphad-calculated mean radius of precipitates vs. time during aging at 600 °C.
Metals 12 01378 g011
Figure 12. Plot of Rockwell “B” hardness vs. time for aging at 600 °C.
Figure 12. Plot of Rockwell “B” hardness vs. time for aging at 600 °C.
Metals 12 01378 g012
Table 1. Chemical composition (wt %) of the present work’s steel.
Table 1. Chemical composition (wt %) of the present work’s steel.
CCrMoMnSiCuNiTi
0.104.500.460.360.350.050.120.002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Saucedo-Muñoz, M.L.; Lopez-Hirata, V.M.; Dorantes-Rosales, H.J.; Villegas-Cardenas, J.D.; Rivas-Lopez, D.I.; Beltran-Zuñiga, M.; Ferreira-Palma, C.; Moreno-Palmerin, J. Phase Transformations of 5Cr-0.5Mo-0.1C Steel after Heat Treatment and Isothermal Exposure. Metals 2022, 12, 1378. https://doi.org/10.3390/met12081378

AMA Style

Saucedo-Muñoz ML, Lopez-Hirata VM, Dorantes-Rosales HJ, Villegas-Cardenas JD, Rivas-Lopez DI, Beltran-Zuñiga M, Ferreira-Palma C, Moreno-Palmerin J. Phase Transformations of 5Cr-0.5Mo-0.1C Steel after Heat Treatment and Isothermal Exposure. Metals. 2022; 12(8):1378. https://doi.org/10.3390/met12081378

Chicago/Turabian Style

Saucedo-Muñoz, Maribel L., Victor M. Lopez-Hirata, Hector J. Dorantes-Rosales, Jose D. Villegas-Cardenas, Diego I. Rivas-Lopez, Manuel Beltran-Zuñiga, Carlos Ferreira-Palma, and Joel Moreno-Palmerin. 2022. "Phase Transformations of 5Cr-0.5Mo-0.1C Steel after Heat Treatment and Isothermal Exposure" Metals 12, no. 8: 1378. https://doi.org/10.3390/met12081378

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