Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/237358396

Effect of Heat Treatment on the Mechanical Properties of ASTM A 890 Gr6A


Super Duplex Stainless Steel

Article  in  Journal of ASTM International · January 2005


DOI: 10.1520/JAI13037

CITATIONS READS

26 3,058

3 authors, including:

Marcelo Martins L.C. Casteletti


Foundry & Metallurgy Consultant University of São Paulo
31 PUBLICATIONS   264 CITATIONS    114 PUBLICATIONS   1,049 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Development and Characterization of white cast iron Alloyed with Mo and/or Nb View project

Projeto de Doutorado - Production and heat treatment of cast graphitic steels with additions of niobium View project

All content following this page was uploaded by Marcelo Martins on 28 May 2014.

The user has requested enhancement of the downloaded file.


Journal of ASTM International, January 2005, Vol. 2, No. 1
Paper ID JAI13037
Available online at www.astm.org

Marcelo Martins1 and Luiz Carlos Casteletti2

Effect of Heat Treatment on the Mechanical Properties of


ASTM A 890 Gr6A Super Duplex Stainless Steel

ABSTRACT: Duplex stainless steels were initially developed for the manufacture of
thermomechanically formed products. However, the increasing demand for casting components of these
materials has led to the application of a widely developed technology for this forming process. After
undergoing a solution annealing heat treatment, these materials become thermodynamically metastable
systems, since the concentration of solute atoms in solid solution is so high that they become saturated,
causing them to seek a lower free energy state when exposed to different temperatures. These systems
reach a more stable thermodynamic condition by the precipitation of various intermetallic phases,
depending on the temperatures to which they are exposed. Thermal energy serves as a catalyst to
“overcome” the energy barrier that separates the metastable and stable phases. The objective of this work
was to determine the influence of several heat treatment temperatures on the microstructure and
mechanical properties of an ASTM A 890/A 890M Gr6A super duplex stainless steel. The increase in
hardness and the decrease in impact toughness of these materials in impact tests were found to be directly
correlated with the increase in sigma phase concentration in their microstructure, which tended to
precipitate into ferrite/austenite interfaces. When the sigma phase was completely dissolved by the heat
treatment, the material’s hardness was determined by the volumetric concentration of ferrite and austenite
in the microstructure, and the energy absorbed in the impact test reached approximately 220 J at room
temperature.
KEYWORDS: sigma phase, microstructure, super duplex stainless steel

Introduction
Duplex stainless steels are widely used in “offshore” platforms, since they offer good
mechanical properties and excellent pitting corrosion resistance.
Despite the greatly varying concentrations of salt, hydrogen sulfide (H2S), carbon dioxide
(CO2), sulfide ions such as (HS-) and (S-2), and oxygen (O2) contents in seawater, duplex
stainless steels are still a good option for such applications [1].
Duplex stainless steels include a particular category called super duplex stainless steels
(SDSS) whose microstructure is composed of ferrite and austenite, each present in a volumetric
concentration of about 50 ± 5 %. SDSSs present a Pitting Resistance Equivalent number (PREN)
higher than 40 [2], a property which is theoretically expressed by Eq 1 [3].
PREN = {%Cr +[(3,3).( %Mo)] + 16(%N)} (1)
Duplex and super duplex stainless steels are Fe-Cr-Ni-Mo-N alloys containing up to 0,3wt.%
of nitrogen in atomic form and displaying a similar corrosion resistance and a superior
mechanical performance to that of copper alloys [2].

Manuscript received 18 August 2004; accepted for publication 15 October 2004; published January 2005.
1
M.Sc., Industrial Manager, SULZER Brasil S/A - Fundinox Division, Av. Eng. João F. Gimenez Molina, 905 -
Jundiaí - SP - Brazil, CEP 13213-080 - Cx Postal 2114. Professor of Centro Universitário Salesiano (Unisal).
2
Ph.D., Associate Professor, Department of Materials, Aeronautics and Automotive Engineering, University of São
Paulo’s São Carlos School of Engineering.

Copyright © 2005 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.
2 JOURNAL OF ASTM INTERNATIONAL

The main properties of duplex and super duplex stainless steels represent a combination of
the best characteristics of their two phases, with the austenite contributing to their impact
toughness and the stronger ferrite improving their tensile strength and weldability. These steels
offer advantages over austenitic stainless steels, displaying a higher stress corrosion resistance
and a greater intercrystalline corrosion resistance, as well as superior mechanical properties.
Their mechanical properties depend on the ferrite content in the microstructure, whose normal
concentration varies from 60 % to 40 % in volume. Higher ferrite (δ) contents increase the
mechanical strength but favor the precipitation of sigma phase during solidification cooling
through a eutectoid reaction of the type:
δ ⇌σ+γ (2)

This reaction is unbalanced toward the right, i.e., in the direction of sigma phase formation,
as the concentration of ferrite-stabilizing elements, particularly chromium and molybdenum,
increases [2].
It is practically impossible to prevent the formation of sigma phase during solidification
cooling, because the compositional ranges favor its precipitation; however, the volumetric
fraction of that phase can be minimized by adjusting the chemical composition and cooling rate
during the solidification process.
In addition to the sigma phase, complex chromium and molybdenum carbides, R phase, G
phase, ε phase (in copper containing steels), and chromium nitrides, among others, can also
appear in these materials’ microstructure.
Due to these steels’ strong tendency for secondary phase precipitation, it is essential to study
the temperatures at which these materials become sensitive.
The “incubation” time for precipitation of the σ phase is 5 min between 850°C and 900°C.
The material’s ductility is strongly reduced by σ phase precipitation. The effect of this
phenomenon increases as the volumetric fraction of σ phase increases. The σ phase formation
mechanism occurs through controlled nucleation and subsequent, relatively fast growth. The σ
phase is the most exhaustively studied of all the intermetallic phases that appear in these steels’
microstructure, for, in addition to appearing in a higher proportion, it drastically affects the
material’s impact toughness and corrosion resistance. Therefore, we have attempted to
investigate the influence of the several heat treatment temperatures on the phase proportions,
hardness, and energy absorbed in impact tests.

Experimental Procedure
The test specimens were cast in vacuum induction degassing (VID) furnaces operating at the
electric supply system’s frequency (60 Hz) and at a maximum power of 400 KW.
The first stage consisted of the casting design of the test specimens, having a 25 mm
diameter and 300 mm length, using AUTOCAD 2000 equipped with MECHANICAL
DESKTOP software, followed by a simulation of the solidification process using a specific
software program called SOLSTAR.
The chemical composition of the material was analyzed via optical emission spectrometry.
Based on the chemical composition, the values of Crequivalent and Niequivalent were calculated
according to the following expressions [4]:
MARTINS AND CASTELETTI ON HEAT TREATMENT OF STEEL 3

Creq.= (%)Cr + [(1,5). (%)Si] + [(1,4). (%)Mo] + (%)Nb – 4,99 (3)

Nieq. = (%)Ni + [(30). (%)C] + [(0,5). (%)Mn] + [26(%)N - 0,02)] + 2,77 (4)

The solution annealing heat treatment recommended for this material was performed in an
electric furnace with a heating capacity of 1300ºC.
The solution annealing heat treatment at 1160°C was carried out to completely dissolve all
the precipitates in the ferritic matrix, forming an unsaturated solid solution. After this solution
annealing heat treatment, followed by water quenching, the samples were aged at various
temperatures ranging from 520–1180°C, increasing the treatment temperatures in 20°C steps, to
check the influence of these temperatures on the phase concentrations and on the energy
absorbed in the impact test. All the samples were kept for 2 h at the aging temperatures.
For the quantitative analysis, the samples were etched with three different etchants: Behara
II, Murakami, and a 50–50 % solution of HCl and HNO3, using full immersion. The percentage
of each phase in the microstructure was determined by the grid point counting method. The
method adopted here involved the use of a grid containing 100 points and 200 times
magnification, based on the ASTM E 562 standard.
A Stereoscan 440– LEO scanning electron microscope (SEM) was also used, and the images
by secondary electrons were taken in digital form. This device was equipped with a 3PC
Microspec-type EDS (Energy Dispersive Spectrometer) detector.
After being heat treated, the samples were subjected to a hardness test to ascertain the
influence of the quantities of phases on this property of the material. The samples were also
subjected to an impact test at room temperature.

Results and Discussion


The chemical compositions of the samples studied here are shown in Table 1. After the
chemical analysis, the Creq/Nieq ratio obtained from Eqs 3 and 4 resulted in 1.71, corresponding
to a theoretical ferrite percentage of approximately 48.2 % and, hence, to 51.8 % of austenite,
according to the ASTM A 800/A 800M standard [4].

TABLE 1—Concentration of chemical elements (wt %) indicated by optical emission


spectrometry.
C (%) Cr(%) Mo(%) Ni(%) Si(%) Mn(%) Cu(%) W(%) N(%) P(%)
0,016 25,69 3,80 7,18 0,74 0,52 0,716 0,736 0,22 0,027
Nb(%) Ti(%) Al(%) V(%) Zr(%) Co(%) Sn(%) Pb(%) S(%) Fe(%)
0,014 0,005 0,016 0,049 0,065 0,055 0,0069 0,0018 0,008 Rem.

The delta ferrite content, indicated by quantitative metallography using the grid point
technique, was 43 % in volume, representing a downward variation of 10.8 %, according to the
value forecasted by ASTM A 800. The delta ferrite content in the microstructure depends on the
heat treatment temperature, since the volumetric concentration of ferrite increases as this variable
rises. The volumetric fraction of delta ferrite in the microstructure also depends on the
correlation between the ferrite and austenite stabilizing elements and on the heat treatment
temperature. The figures below depict the materials’ microstructures under different heat
treatment conditions after solution annealing at 1160°C followed by water quenching.
4 JOURNAL OF ASTM INTERNATIONAL

The sample solution, annealed at 1160°C, quenched, and treated at 520°C for 2 h, displays a
microstructure composed of only two phases: the ferritic matrix and the austenite precipitated in
the form of “islands,” as illustrated in Fig. 1. No intermetallic phases are precipitated at the δ/γ
interfaces or at the ferrite grain boundaries. The sample annealed at 1160°C and treated at 820°C
for 2 h shows the presence of austenite, sigma phase, and ferrite dispersed in the sigma phase and
austenite mixture, as depicted in Fig. 2.

FIG. 1—Scanning electron micrograph of the sample treated at 520°C for 2 h.

σ+δ+γ

FIG. 2—Scanning electron micrograph of the sample treated at 820°C for 2 h.

Ferrite is difficult to identify because at that temperature, its volumetric concentration is very
low, i.e., about 3.0 %. However, as can be noted, the austenite appears on a plane below that of
the sigma phase plane because the latter is harder than the austenite and during mechanical
MARTINS AND CASTELETTI ON HEAT TREATMENT OF STEEL 5

polishing, it stands out in high relief since it becomes less worn. The austenite obtained from the
eutectoid reaction is mixed with the ferrite, but it is difficult to identify each of these phases at
this level of magnification.
In the sample treated at 920°C (Fig. 3), the visible sigma phase is clearly outlined and in
relief for the aforementioned reason. At that temperature, this phase clearly precipitates in a
continuous network. This morphology favors the propagation of cracks at the interfaces between
the phases. The treatment at 940°C (Fig. 4) produced two types of morphologies in the sigma
phase: an elongated one in relief, with well-defined boundaries, and the second, lacy one
encompassing austenite and delta ferrite, which provided chromium and molybdenum for its
precipitation.

σ+γ

FIG. 3—Scanning electron micrograph of the sample treated at 920°C for 2 h.

δ+σ+γ

FIG. 4—Scanning electron micrograph of the sample treated at 940°C for 2 h.


6 JOURNAL OF ASTM INTERNATIONAL

The major difficulty involved in the manufacture of these steels is to obtain the maximum
PREN value without impairing the microstructural balance between ferrite and austenite. When
super duplex stainless steels are exposed to temperatures below the solution annealing
temperature, the metastable thermodynamic balance is disturbed, causing the system (material)
to seek a more stable thermodynamic state through the precipitation of intermetallic phases,
carbonic phases, and a microstructural imbalance between ferrite and austenite. These
intermetallic phases include the sigma phase, which is very rich in chromium and molybdenum.
Figure 5 shows the correlation between the proportions of the phases and the diverse heat
treatment temperatures.

Austenite Ferrite Sigma Phase

70
Phase Volumetric Concentrations

60

50

40
(%)

30

20

10

0
500 600 700 800 900 1000 1100 1200
Heat Treatm ent Tem peratures (°C)

FIG. 5—Proportions of the phases versus heat treatment temperature.

As can be seen, the sigma phase begins to precipitate at 720°C, and its concentration rises
sharply up to 800°C, when it reaches about 45 % in volume. The concentration peak of this
phase occurs close to 880°C, with a volumetric concentration of 50 %. Between 880°C and
900°C, the percentage of sigma phase drops by 10 % in volume. From 900–1000°C, the
volumetric concentration of the sigma phase gradually decreases, displaying a sharp drop
between 1000°C and 1020°C. The sigma phase gradually disappears from the material’s
microstructure at temperatures above 1060°C. The curve indicating the variation in volumetric
concentration of ferrite displays an inverse behavior to that representing the concentration of
sigma phase.
An increasing concentration of sigma phase causes the concentration of ferrite to decrease
because the elements forming the intermetallic phase (Cr and Mo) are rejected from the ferrite,
where they are more soluble. This causes a depletion of these elements in the ferrite and a
considerable decrease in the Crequiv/Niequiv ratio, favoring its transformation into secondary
austenite. Thus, the eutectoid reaction: δ ⇌ σ + γ2 is unbalanced toward the right.
When the heat treatment temperature rises from 1000°C to 1020°C, the volumetric
concentration of ferrite increases abruptly from 5 % to 38 %, then stabilizes at around 46 %
between 1140°C and 1180°C.
MARTINS AND CASTELETTI ON HEAT TREATMENT OF STEEL 7

A practical way to detect the presence of sigma phase in the microstructure is to measure the
material’s hardness, as can be seen by a comparison of Figs. 5 and 6. Figure 6 shows that the
steel’s hardness begins to increase effectively from 720°C on, displaying a peak in the order of
400 Brinell at 800°C, which remains high up to treatments at 840°C, thereafter diminishing as
the temperature rises. Figure 6 also reveals that the hardness peak at 800°C does not coincide
with the sigma phase concentration peak, which occurs at 880°C. A possible reason for this is
the fact that, with the eutectoid decomposition of the delta ferrite into sigma phase and secondary
austenite, the ferritic phase disappears completely, in detriment to the two new phases. Thus, as
a result of the significant increase in austenitic phase, the hardness decreases considerably. At
temperatures exceeding 1060°C, the material’s hardness tends to stabilize between 250 and 260
Brinell, which represents hardness typical of a super duplex stainless steel [7].

450

400
Brinell Hardness (HB)

350

300

250

200
500 600 700 800 900 1000 1100 1200
Heat Treatment Temperatures (°C)

FIG. 6—Hardness of the material as a function of heat treatment temperatures.

Figure 7 illustrates the material’s hardness as a function of the volumetric percentage of


sigma phase. As the volume of sigma phase increases, the material’s hardness increases with an
approximately parabolic tendency. This intermetallic precipitate is hard, leading to a significant
increase in this property from the macroscopic standpoint.
Figure 08 shows the impact toughness values measured through a Charpy test conducted at
room temperature on a V-notched test specimen. Note that the energy absorbed in the test drops
to 10% of the original value, with a sigma phase volume of only 3% in the microstructure.
Figure 09 illustrates the behavior of impact toughness as a function of the heat treatment
temperature. As can be seen, the energy absorbed in the test drops abruptly from 540oC on,
reaches a value of 20J at 640oC, remains at extremely low values (below 10J) up to 1040oC, then
“leaps” rapidly to the level of 200J from 1060oC on, remaining at around that value at higher
temperatures. Moreover, even without the presence of the sigma phase, i.e., with a volumetric
percentage equal to zero, there is a visible drop in absorbed energy during the impact test. This
effect was not detected either in the hardness versus heat treatment temperature curve through an
increase in this property, or in the micrographs obtained by optical microscopy.
8 JOURNAL OF ASTM INTERNATIONAL

BRINELL HARDNESS VS SIGMA PHASE CONCENTRATION


450

400
BRINELL HARDNESS

350

300

250

200
0 10 20 30 40 50 60
VOLUMETRIC CONCENTRATION OF SIGMA PHASE (%)

FIG. 7—Correlation between Brinell hardness and sigma phase volume.

250
ABSORBED ENERGY (J)

200

150

100

50

0
0 10 20 30 40 50 60
VOLUMETRIC CONCENTRATION OF SIGMA PHASE (%)

FIG. 8—Energy absorbed in Charpy Test versus sigma phase volumetric percentage.
MARTINS AND CASTELETTI ON HEAT TREATMENT OF STEEL 9

250

200
Absorbed Energy (J)

150

100

50

0
500 600 700 800 900 1000 1100 1200
Heat Treatment Temperature (°C)

FIG. 9—Absorbed energy versus heat treatment temperature.

According to Fig. 5, the sigma phase begins precipitating from 720°C on, so the drop in
absorbed energy at temperatures below 720°C (see Fig. 9) cannot be attributed to the
precipitation of this intermetallic phase. The analyses of the fractured surfaces of samples
treated at 520°C, 980°C, and 1160°C are shown in Figs. 10–15.
Figures 10 and 11 show the appearance of the fractures in samples treated at 520°C. Note the
fibrous aspect typical of highly tough materials, which indicates that the heat treatment
temperature did not modify the mechanical behavior in terms of impact toughness. This finding
is also confirmed by an analysis of the absorbed energy versus temperature curve (Fig. 9). This
fracture surface displays included particles originating from the deoxidization process, as well as
voids around the particles. Figure 12 shows the peaks revealed by a qualitative analysis of the
particle obtained by EDS detector. As can be seen, this is an inclusionary particle of a complex
silicon, zirconium, and aluminum oxide originating from the deoxidization process.
Figure 13 depicts the appearance of the fracture in the sample treated at 980°C for 2 h. At
this temperature, the material displays a brittle behavior due to the presence of a large amount
(approximately 31 %) of sigma phase in the microstructure. The appearance of the fracture
differs totally from the previous case, lacking the presence of fibrous structures. Note the
appearance of an intergranular crack likely due to the precipitation of sigma phase in that region.
The appearance of the fracture differs slightly from the typical pattern for brittle materials since,
in this case, there is a high volumetric fraction of precipitated austenite (58 %), in detriment to
the reduction of the volumetric fraction of ferrite. This high content of a ductile and tough phase
alters the fracture appearance pattern for brittle materials.
Figures 14 and 15 show aspects of the fractures in samples treated at 1160°C. Note the
presence of a fibrous structure typical of ductile material at that treatment temperature, indicating
the absence of intermetallic phases negatively influencing the impact toughness. The presence of
tough phases in the microstructure indicates that part of the energy required to break the test
specimen is absorbed, plastically straining the material and hindering crack propagation.
10 JOURNAL OF ASTM INTERNATIONAL

FIG. 10—Appearance of the fracture in the sample treated at 520°C for 2 h.

FIG. 11—Appearance of the fracture in the sample treated at 520°C for 2 h.


MARTINS AND CASTELETTI ON HEAT TREATMENT OF STEEL 11

FIG. 12—Qualitative analysis of the particle shown in Fig. 11.

FIG. 13—Appearance of the fracture of the sample treated at 980°C for 2 h.


12 JOURNAL OF ASTM INTERNATIONAL

FIG. 14—Appearance of the fracture in the sample treated at 1160°C.

FIG. 15—Appearance of the fracture in the sample treated at 1160°C.

Conclusions
The sigma phase begins precipitating at heat treatment temperatures in the order of 720°C,
dissolving completely from solution annealing above 1060°C, followed by water quenching.
The sigma phase precipitates preferentially at the ferrite/austenite interface, growing toward
ferrite, which supplies elements such as chromium and molybdenum for its stabilization. At the
heat treatment temperature of 880°C, the volumetric concentration of sigma phase reaches the
maximum value of 50 % in volume.
MARTINS AND CASTELETTI ON HEAT TREATMENT OF STEEL 13

The volumetric fraction of ferrite drops to very low values, gradually disappearing at heat
treatment temperatures in the 840°C to 880°C interval. The material’s hardness begins to
increase significantly from 720°C up, reaching its highest values at treatment temperatures of
800–820°C. This peak does not coincide exactly with the peak of maximum volumetric
concentration of sigma phase, which occurs at 880°C.
At high treatment temperatures exceeding 1060°C, the material’s hardness is affected only by
the ferrite and austenite phases. The material’s hardness is strongly influenced by the volumetric
fraction of sigma phase in the microstructure. The energy absorbed in the impact test is strongly
reduced by the presence of sigma phase in the microstructure. A mere 3 % in volume of sigma
phase in the microstructure reduces the absorbed energy to 10 % of its original value, i.e., when
this intermetallic phase is absent from the microstructure. After heat treatments at high
temperatures (above 1060°C), the energy absorbed in the impact test returns to the material’s
original levels.

References
[1] Duplex Stainless Steels ’91, Charles, J. and Bernhardsson, S., Eds., Beaune, Les Ulis,
France, Les Éditions of Physique, Vol. 1, 1991, pp. 3–48.
[2] Weber, J., “Materials for Seawater Pumps and Related Systems,” Sulzer Brothers Limited,
Winterthur, Switzerland, pp. 1–11.
[3] ASTM Standard A 890/A 890M–91, “Ferrous Castings; Ferralloys,” Annual Book of ASTM
Standards, Vol. 01.02, ASTM International, West Conshohocken, PA, 1999, pp. 556–569.
[4] ASTM Standard A 800/A 800M–91, “Ferrous Castings; Ferralloys,” Annual Book of ASTM
Standards, Vol. 01.02, ASTM International, West Conshohocken, PA, 1999, pp. 458–463.
[5] ASTM Standard A 781/A 781M–91, “Ferrous Castings; Ferralloys,” Annual Book of ASTM
Standards, Vol. 01.02, ASTM International, West Conshohocken, PA, 1999, pp. 440–443.
[6] ASTM Standard A 370–97a, “Ferrous Castings; Ferralloys,” Annual Book of ASTM
Standards, Vol. 01.02, ASTM International, West Conshohocken, PA, 2001, pp. 148–200.
[7] Weber, J. and Schlapfer, H. W., “Austenitic-Ferritic Duplex Steels,” Sulzer Brothers
Limited, Winterthur, Switzerland, pp. 1–10.
[8] Maehara, Y., Ohmori, Y., Murayama, J., Fujino, N., and Kunitake, T., “Effects of Alloying
Elements on σ Phase Precipitation in δ-γ Duplex Stainless Steels,” Metal Science, Vol. 17,
November 1983, pp. 541–547.
[9] Li, J., Wu, T., Riquier, Y., “σ Phase Precipitation and Its Effect on the Mechanical
Proprieties of the Super Duplex Stainless Steel,” Materials Science and Engineering, 1994,
Vol. 174A, pp. 149–156.
[10] Anson, D. R., Pomfret, R. J.,and Hendry, T., “Prediction of the Solubility of Nitrogen in
Molten Duplex Stainless Steel,” ISIJ International, Vol. 36, No. 7, 1996, pp. 750–758.
[11] Goldstein, J. I., et al., “Scanning Electron Microscopy and X-Ray Microanalysis,” 2nd ed.,
New York, Plenum Press, Cap. 3, 1992, pp. 69–147.
[12] Rossitti, S. M., “Effect of Niobium in the Microstructure and in Mechanical Properties of
the Molted Super Duplex Stainless Steel SEW 410 W. Nr. 1.4517,” Thesis, Universidade de
São Paulo, Interunidade, EESC-IFSC-IQSC, 2000, pp. 49–53.
[13] Weiss, B. and Stickler, R., “Phase Instabilities at High Temperature Exposure of 316
Austenitic Stainless Steel,” Metallurgical Transaction, Vol. 3, April 1972, pp. 851–864.
14 JOURNAL OF ASTM INTERNATIONAL

[14] Pohl, M., “The Ferrite/Austenite Ratio of Duplex Stainless Steels,” Z. Metallkd., Vol. 86,
No. 2, May 1995, pp. 97–101.
[15] Siewert, T. T., McCowan, C. N., and Olson, D. L., “Ferrite Number Prediction to 100 FN in
Stainless Steel Weld Metal,” Welding Research Supplement, December 1988, pp. 289-s–
297-s.

View publication stats

You might also like