You are on page 1of 12

Materials Science and Engineering C 51 (2015) 87–98

Contents lists available at ScienceDirect

Materials Science and Engineering C


journal homepage: www.elsevier.com/locate/msec

Analytical modeling of the thermomechanical behavior of ASTM F-1586


high nitrogen austenitic stainless steel used as a biomaterial under
multipass deformation
Fabiano R. Bernardes a, Samuel F. Rodrigues a,⁎, Eden S. Silva a, Gedeon S. Reis a, Mariana B.R. Silva b,
Alberto M.J. Junior b, Oscar Balancin b
a
Department of Mechanic and Materials, Federal Institute of Education, Science and Technology of Maranhão — IFMA, Av. Getúlio Vargas, 4, Monte Castelo, CEP 65030-005 São Luís, MA, Brazil
b
Department of Materials Engineering, Federal University of São Carlos — UFSCar, Rod. Washington Luiz, Km 235, CEP 13565-905 São Carlos, SP, Brazil

a r t i c l e i n f o a b s t r a c t

Article history: Precipitation–recrystallization interactions in ASTM F-1586 austenitic stainless steel were studied by means of
Received 22 September 2014 hot torsion tests with multipass deformation under continuous cooling, simulating an industrial laminating pro-
Received in revised form 14 January 2015 cess. Samples were deformed at 0.2 and 0.3 at a strain rate of 1.0 s−1, in a temperature range of 900 to 1200 °C and
Accepted 23 February 2015
interpass times varying from 5 to 80 s. The tests indicate that the stress level depends on deformation tempera-
Available online 25 February 2015
ture and the slope of the equivalent mean stress (EMS) vs. 1/T presents two distinct behaviors, with a transition at
Keywords:
around 1100 °C, the non-recrystallization temperature (Tnr). Below the Tnr, strain-induced precipitation of Z-
Biomaterial phase (NbCrN) occurs in short interpass times (tpass b 30 s), inhibiting recrystallization and promoting stepwise
Recrystallization stress build-up with strong recovery, which is responsible for increasing the Tnr. At interpass times longer than
Precipitates 30 s, the coalescence and dissolution of precipitates promote a decrease in the Tnr and favor the formation of re-
Z-phase crystallized grains. Based on this evidence, the physical simulation of controlled processing allows for a domain
refined grain with better mechanical properties.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction the friction, lubrication and heat transfer conditions are different, in-
cluding the plastic behavior determined in the simulation of the hot
Austenitic stainless steels are being used as a material for orthopedic conformation physical equations of each process leading to optimal re-
implants. ASTM F-138 steel is currently the one most widely used by sults. Thus, it is possible to evaluate the role of the Z-phase and the soft-
Brazil's National Health Service (SUS), but it presents unsatisfactory me- ening mechanisms that occur during microstructural evolution, and to
chanical and corrosion properties. Moreover, this material has a high design and investigate the sequence of passes that improve the proper-
nickel content (N12%) that causes allergic reactions in many patients ties of materials and that can be implemented in industrial processing
[1]. Steels containing Mn, N and a low concentration of Ni, such as routes.
ASTM F-1586, have been developed to mitigate these problems. This In the manufacture of orthopedic prostheses, e.g., the austenitic
material consists of an austenitic matrix rich in Nb, Cr and N, which stainless steels used in the manufacture of hip prosthesis, hot processing
can form second phase particles, Z-phase (NbCrN), with evidence of is carried out in two distinct stages: (i) manufacture of bars and (ii) forg-
precipitation within the industrial processing window [2,3]. However, ing of prostheses [4,5]. In the case of steels not subjected to phase trans-
the interaction between precipitation and recrystallization of this steel formation during cooling after hot forging, controlled processing is
is still not fully established. This characteristic makes it a promising sub- more restricted. One possibility is to promote repeated recrystallization
stitute of ASTM F-138 steel in more critical applications of more severe of the austenite and prevent grain growth in the intervals between
loading and longer periods within the human body. passes. A second alternative would be to accumulate strains below the
Physical simulations of thermomechanical processing by miniaturized non-recrystallization temperature (Tnr), thereby increasing the number
devices allow for the execution of designs with multipass deformation ap- of nucleation sites. Tnr is influenced by the type and amount of
proximate to industrial conditions to investigate the mechanisms that op- microalloying element. Different microalloying elements have different
erate during deformation and in the intervals between passes. However, effects on Tnr. The microalloying element present in the composition of
the ASTM F-1586 steel under study is niobium (Nb), which is the most
effective element for increasing the Tnr. This behavior is related to its
⁎ Corresponding author. precipitation capacity through its chemical potential and to the process-
E-mail address: samuel.filgueiras@ifma.edu.br (S.F. Rodrigues). ing conditions.

http://dx.doi.org/10.1016/j.msec.2015.02.040
0928-4931/© 2015 Elsevier B.V. All rights reserved.
88 F.R. Bernardes et al. / Materials Science and Engineering C 51 (2015) 87–98

When stainless steel is reheated at 1250 °C, compounds that are This work investigates the thermomechanical behavior of ASTM F-
present as precipitates dissolve as the temperature rises, until the mate- 1586 steel based on hot torsion tests with multipass deformation
rial becomes fully monophasic. Particularly in the case of the ASTM F- under continuous cooling. The aim is to determine the Tnr and its depen-
1586 steel, Z-phase is present at room temperature and is dissolved dence on the applied deformation per pass and the waiting time, ana-
partially or totally during reheating at high temperatures. This phase lyzing the action of softening mechanisms and the occurrence of
is a little known complex nitride, whose tetragonal crystal structure precipitation, as well as its role in the shape of the curves.
was first determined by JACK [6], by means of stoichiometric NbCrN in
an alloy of similar composition. 2. Material and method
The Z-phase precipitation potential is determined from the solubility
product by calculating the supersaturation of the microalloying ele- The austenitic stainless steel studied here is ASTM F-1586 used in
ments at a given temperature, comparing the amount of dissolved the manufacture of orthopedic implants. This steel is manufactured by
microalloying element during reheating with the amount in solution, Villares Metals S/A Brazil in the form of 20 mm diameter laminated
whose formation reaction is [7]: bars annealed at 1030 °C for 60 min and cooled in water. This material
is a Cr–Ni–Mn–Mo–Nb based composition containing added N. Its Cr,
Nb2 þ 2Cr þ N2 ↔2NbCrN: ð1Þ Mo and N contents ensure higher corrosion resistance than that of
ASTM F-138 steel. The beneficial effects afforded by the presence of
The equilibrium constant (K) for the reaction is given by: the elements C, Nb and Mo are enhanced corrosion resistance and me-
chanical and fatigue strength. In the solubilized condition, the plastic
½aNbCrN 2 flow of ASTM F-1586 steel is two-fold higher than that of ASTM F-138
K¼ ð2Þ
2hNb hCr hN steel. Table 1 describes the composition of this steel.
The technique developed by Boratto et al. [12] was used here to an-
where the aNbCrN activity is equal to unity for the pure compound and alyze the precipitation–recrystallization interaction under multipass
the [hCr], [hNb] and [hN] activities are considered proportional to the di- deformation. The test specimens were induction heated to a tempera-
luted concentrations of elements in solution. The solubility product of Z- ture of 1250 °C at a heating rate of 5.5 °C/s and held at that temperature
phase can be generated by taking into account the equilibrium constant for 300 s. They were then subjected to consecutive deformation passes
of reaction (K) and relating it with the system's free energy (ΔGo), en- of 0.2 or 0.3 at a constant strain rate of 1.0 s−1, while undergoing contin-
thalpy (ΔHo) and entropy (ΔSo), obtaining the equation of the solubility uous cooling at a controlled rate, αcool = 20 °C/tpass (interpass time).
product given by [8,9]: Seventeen deformation passes were applied in the temperature interval
o o o of 1250 and 900 °C, with interpass times (tpass) varying from 5 to 80 s.
ΔG ¼ ΔH −TΔS ¼ −RT lnK: ð3Þ Some of the tests were interrupted at different levels of deformation
and the material abruptly cooled in water to examine the microstruc-
Hence, for Z-phase precipitation: ture in the deformation stage. The temperature was controlled by
o o means of an optical pyrometer.
ΔS ΔH
ln f½hNb ½hCr ½hN g ¼ − : ð4Þ Based on the experimental data, the mean equivalent stress (EMS)
R RT was determined by trapezoidal numerical integration of the area
under the stress vs. strain curve at each of the deformations, using equa-
Rewriting the above equation in terms of the weight percentage of
tion [12]:
Cr, Nb and N in solution, one obtains the equation of the solubility prod-
uct, which, according to Giordani [10], is expressed by:
Zεb
1
12; 277 EMS ¼ σ eq dεeq ð7Þ
ln f½Nb½Cr½Ng ¼ 7:55− : ð5Þ εb −ε a
T εa

The precipitation potential based on the calculated supersaturation


at a given temperature can be determined by means of the solubility where EMS is the equivalent mean stress, σeq is the equivalent stress,
product. The supersaturation coefficient (Ks) at a given temperature and (εb–εa) is the equivalent stress at each pass of interest.
below the equilibrium temperature is defined as the ratio of the amount The samples were characterized by optical and scanning electron
of [Cr][Nb][N] at that temperature to the amount at the equilibrium microscopy, with the microstructure observed in the longitudinal
temperature: plane underneath the deformed surface. The samples were polished
and electrolytically etched in HNO3 solution (65% in water solution)
f½Cr½Nb½Ngsol 107:557−ð12;277=T sol Þ for microstructural analysis.
Ks ¼ 7:557−ð12;277=TÞ
¼ 7:557−ð12;277=T Þ ð6Þ The precipitate analysis was determined by scanning electron mi-
10 10 passe

croscopy (SEM) and transmission (TEM). During TEM, the technique


where Tsol and Tpass are the temperatures of solubilization and absolute of precipitate extraction by carbon replica was used for analysis of fine
deformation. The supersaturation coefficient (Ks) determines the driv- precipitates and the SEM was used for morphological analysis and
ing force for precipitation that occurs when the supersaturation reaches coarsening of precipitates.
a critical value necessary to compensate for the formation of interfaces Carbon replicas were prepared etching the matrix, following the
and the elastic distortions caused by the precipitates. same procedures as for optical microscopy and, after that, a thin carbon
Retardation of recrystallization occurs when the anchoring force de- layer was evaporated on the etched surface and then the carbon layer
veloped by the precipitates at the grain boundaries is greater than the containing precipitates was extracted from the surface by electrolytic
driving force for recrystallization [10,11]. This interaction is important etching and supported on a 300-mesh Cu grid. Thin foil samples were
because it determines the softening mechanisms that act during hot
processing, and therefore, the resulting microstructural evolution. Tak- Table 1
ing as reference Nb microalloyed steel, in which this interaction has Chemical composition of ASTM F-1586 steel (mass %).
been studied extensively, one can analyze this interaction through re- C Si Mn Ni Cr Mo S P N Cu Nb Fe
crystallization–precipitation–time–temperature (RPTT) diagrams with
0.035 0.37 4.04 10.6 20.3 2.47 0.001 0.022 0.36 0.06 0.29 bal.
determination of the Tnr.
F.R. Bernardes et al. / Materials Science and Engineering C 51 (2015) 87–98 89

Fig. 1. Plastic flow curves obtained under continuous cooling with multipass deformations of 0.2 and 0.3 and interpass times of 5 to 80 s.
90 F.R. Bernardes et al. / Materials Science and Engineering C 51 (2015) 87–98

300 also prepared by electrolytic polishing using a solution of acetic acid


ε = 0,2 - 1s
-1
5,0 s
Region I Region II (95%) and perchloric acid (5%), at room temperature (~25 °C).
30 s
50 s
250 80 s
3. Results and discussion
EMS (MPa)

200 3.1. Plastic flow curves under multipass deformation

Fig. 1 illustrates the plastic flow curves. As can be seen, the stress in-
150 creases with the evolution of sequential deformations, and this increase
is intensified at lower temperatures. Fig. 2 shows the evolution of equiv-
alent mean stress (EMS) as a function of deformation temperature. Note
100 that the curves are clearly separable into two regions with different
slopes in EMS. Also note that, in region I, the EMS value of some of the
first deformations deviates from the expected linear behavior.
50
-4 -4 -4 -4 -4 -4 -4
6.0x10 6.5x10 7.0x10 7.5x10 8.0x10 8.5x10 9.0x10
-1
3.2. Equivalent mean stress (EMS) vs. 1/T curves
1/T (K )
300 Upon reducing the deformation temperature, the EMS behavior
5,0 s -1
Region I Region II 0,3 - 1s shows a sharp increase in the EMS value required to deform the material
30 s
50 s (Fig. 2). This increase causes the slope of the EMS vs. 1/T curve to change,
250 80 s demarcating two distinct regions. The temperature at which the two line
segments intersect is identified as the non-recrystallization temperature
(Tnr) [11,12]. The steeper slope of the line in region II is an indication
200
that, in addition to the increased strength of the material in response to
EMS (MPa)

stress built up through the decrease in deformation temperature, another


150
hardening mechanism is activated simultaneously, leading to the onset of
strain-induced precipitation. Thus, Tnr represents the temperature at
which recrystallization is no longer completed in the intervals between
100 passes. Therefore, the change in inclination, and hence, in the level of
stress is associated with the microstructural evolution taking place during
and between multipass deformations.
50 The stress–strain behavior reveals that the level of stress increases in
-4 -4 -4 -4 -4 -4 -4
6.0x10 6.5x10 7.0x10 7.5x10 8.0x10 8.5x10 9.0x10 region I (high T), especially in the first four passes when the time is longer
1/T (K )
-1 than 30 s. This increase is due to the decrease in deformation temperature
and the longer time required for the softening phenomena to act when
Fig. 2. Evolution of equivalent mean stress (EMS) with inverse absolute temperature (1/T), strain exceeds the critical value, added to the contribution of strain-
and determination of Tnr. induced precipitation of Z-phase that may occur after recrystallization.

Fig. 3. Phase diagram of ASTM F-1586 steel calculated by FactSage 6.1.


F.R. Bernardes et al. / Materials Science and Engineering C 51 (2015) 87–98 91

Element Weight% Atomic%

CK 4.14 14.38

NK 12.48 37.23

VK 1.34 1.10

Cr K 22.39 17.99

Mn K 0.93 0.71

Fe K 6.96 5.21

Ni K 0.43 0.31

Nb L 51.33 23.08

Totals 100.00

Fig. 4. Energy dispersive X-ray spectroscopy (EDS) and SEM analysis.

In this region of temperatures above Tnr, the driving force for precip- explained by the increase in nucleation sites, with work hardened
itation kinetics (supersaturation constant, ks) decreases significantly grains replaced by new undeformed grains that migrate into the most
and metadynamic recrystallization (MDRX) predominates. This is deformed grains, characterizing the subsequent softening, with short
interpass times not affecting the EMS, which thus becomes only a func-
tion of temperature.
In the region of intermediate temperatures, between 1040 and
1125 °C, the Tnr range, note the abrupt change in the level of stress
along the passes, with deformation varying from 1.5 to 2.5, see Fig. 1.
In this condition, metadynamic recrystallization is the main softening
mechanism during multipass deformations above the critical deforma-
tion for the onset of dynamic recrystallization (εc = 0.15) with precip-
itation appearing subsequently and dependent on differences in
temperature, with both controlled by diffusion. Finally, in the low tem-
perature region below Tnr, note the stress build-up with intense recov-
ery, and partial interpass recrystallization with strong hardening (high
EMS) and high incidence of strain-induced precipitation, changing the
slope of the curve.
An estimate of the precipitates in ASTM F-1586 steel using the ther-
modynamic calculation program FactSage 6.1 predicts the formation of
four types of precipitates: Z-phase (NbCrN), FCC (NbN), M23C6 (possi-
bly Cr23C6) and σ phase (FeCrMoNi) (see Fig. 3). At high temperatures
the precipitates are more stable, such as Z-phase and NbN, which are the
ones most likely to interact with the recrystallization kinetics, where a
delay in the recrystallization kinetics with an increase in EMS, Fig. 2,
and elevation in stress build-up Δσ, Fig. 6, are observed.
Energy dispersive X-ray spectroscopy (EDS) and SEM analysis reveal
particles rich in Nb, Cr, Mo and N, confirming the abovementioned cal-
Fig. 5. Precipitates analysis by thin foils at 1000 °C for time between pass of 5.0 s. culations, see Fig. 4. The chemical composition of the Z-phase showed
92 F.R. Bernardes et al. / Materials Science and Engineering C 51 (2015) 87–98

(a) (b)
Fig. 6. Analysis by TEM of carbon replicas and thin foils under the conditions: a) 1200 °C with times 5.0 s and 50 s (third pass) and b) 1000 °C with times 5.0 s and 50 s (13th pass).

atomic percentages of approximately 50% Nb, 28% Cr and 7.5% N, which stored during deformation and the thermodynamic potential for recrys-
are similar to those reported in the literature [7]. tallization, delaying the nucleation and growth of recrystallized regions.
The analysis and identification of particles precipitated by TEM using Fig. 9 shows the dependence of non-recrystallization temperature
carbon replica techniques and thin foils, Fig. 5, confirm the presence of (Tnr) on interpass time (tpass) and applied strain. As can be seen, short
Z-phase (NbCrN) and niobium nitride (NbN), Fig. 6, confirmed by elec- interpass times of up to about 30 s cause the Tnr to increase. Thereafter,
tron diffraction, Fig. 7. Note the precipitates coarsening when the time the Tnr decreases as the interpass time increases. Also note that the Tnr in-
between passes is increased. creases as the applied strain increases. As can be seen, the Tnr reaches a
The increase in EMS in the initial deformation passes above Tnr at maximum at about 30 s, presenting basically two regions: (i) ascending
interpass times longer than 30 s (Fig. 2) is due to the decrease in tem- Tnr in short interpass times and (ii) descending Tnr in longer interpass
perature with conspicuous hardening, hindering the action of thermally times. The first region shows significant retardation of recrystallization
activated mechanisms. At temperatures below Tnr, recrystallization is due to strain-induced precipitation, which is time-dependent, such that
delayed by the presence of strain-induced precipitation of Z-phase par- the effect of Tnr retardation increases in these deformation conditions. In
ticles (Fig. 8; a–d), resulting from the reduction of free energy of the el- region II, at interpass times longer than 30 s, the Tnr reaches a peak, sug-
ements Nb, Cr and N, which obstruct grain boundary mobility, which gesting that at this point the effect of precipitation on recrystallization be-
explains the change in slope and fluctuations in EMS values. As the tem- comes less effective due to coalescence and dissolution of the precipitates,
perature decreases, the recrystallization phenomenon decreases and thus increasing the extent of recrystallization and reducing Tnr.
the effect of precipitation increases until it reaches the stage of coales- As for deformation, a slight change is visible in Tnr, particularly in
cence, temperature at 1000 °C and interpass times of 5.0 and 50.0 s long interpass times, with a decrease in Tnr as deformation increases,
(Fig. 8; e–f), leading to a local reduction in the level of stress in region i.e., deformation accelerates the coalescence of precipitates as the
II. It should be noted that the effect of solute (N, Mo, Nb) and stacking interpass time increases, see Fig. 8; e–f. Note that the time required
fault energy (SFE) on the mobility of dislocations increases the energy for the formation of recrystallization decreases continuously with
F.R. Bernardes et al. / Materials Science and Engineering C 51 (2015) 87–98 93

(a)

(b)
Fig. 7. Energy dispersive X-ray spectroscopy: a) particle 1 (Z phase) and b) particle 2 (NbN).

temperature, indicating a typical process of diffusion-controlled growth agreement with results reported by Giordani [2,7,10], who studied an
[17,18]. The recovery–precipitation interaction then occurs as the tem- alloy of similar composition, in which the fraction recrystallized below
perature decreases, with recovery delaying the progress of precipitation 1100 °C did not exceed 16%, even with long interpass times (tpass), indi-
by reducing the number of available nucleation sites. In addition, a sig- cating the effect of induced Z-phase precipitation and providing strong
nificant concentration of dislocations can trap fine dispersed precipi- evidence of the recovery of this steel with markedly elongated grains.
tates, thereby delaying the recovery process. Note that, unlike the data
obtained by previous methods, precipitation may occur at higher 3.3. Behavior of isothermally simulated curves
temperatures, in this case close to 1100 °C. Precipitation will probably
always occur below this temperature, as already observed in the previ- In Fig. 12, note that the flow curves in region I are rounded and well
ous analyses, even without considering the cooling rate and accumulat- delineated, with the first passes showing higher EMS values than those
ed energy. indicated by the straight lines, which represent the effect of decreasing
Fig. 10 illustrates the behavior of stress build-up (Δσ), determined temperature, particularly at longer interpass times. In these experi-
directly from the EMS vs. 1/T graph (see Fig. 2) as a function of interpass ments, the analysis of softening between passes in this region is quite
time and non-recrystallization temperature (Tnr). Considering that complex, since the tests were performed under continuous cooling. An
stress build-up (Δσ) describes the progress of hardening of the alloy alternative is to make corrections to the experimental curves, making
under multipass deformation at temperatures below the Tnr, it is clear them equivalent to tests performed isothermally. Initially, we deter-
that the higher the Tnr the higher the Δσ resulting from the increase mined the equations that describe the straight lines which represent
in Z-phase precipitation, without recrystallization, with elongated only the effect of temperature on the EMS in region I, which are
pancake-shaped grains, however, longer interpass times (tpass) will expressed as
lead to the formation of coarser precipitates, due to coalescence and dif-
fusion, with less effect on grain size in the softening kinetics and lower
stress build-up, Fig. 11. This high build-up of stress (Δσ ~ 140 MPa) is in σ ðTME; MPaÞ ¼ a þ bð1000=TðKÞÞ ð8Þ
94 F.R. Bernardes et al. / Materials Science and Engineering C 51 (2015) 87–98

(a) (b)

(c) (d)

(e) (f)
Fig. 8. Precipitated particles interacting with the matrix and grain boundaries of ASTM F-1586 steel: (a) and (b) precipitates in grain boundaries; (c) and (d) precipitates inside the grains;
(e) and (f) coalescence of precipitated Z-phase particles in multiple deformation tests at 1000 °C with interpass time between 5.0 and 50.0 s, respectively.

where a and b are constants; T is the absolute temperature; and σ is the onset of dynamic recrystallization (εc) [14–19], it can be expected that
equivalent mean stress. The equations presented in Table 2 were ob- softening between the first and second passes was generated by
tained for the experiments with deformations of 0.2 and 0.3 and metadynamic recrystallization, particularly in the three experiments
interpass times of 10 and 50 s. Using these equations, corrections in question. Note that the overlap of the plastic flow curve shows a char-
were made of all the stress values, transforming the curves into equiva- acteristic “single peak” curve, indicating the occurrence of dynamic re-
lents at 1200 °C. The result of these corrections is depicted in Fig. 12. crystallization. Starting from the second and third reloadings, there is
The curves shown in Fig. 12 reveal significant softening between the a new build-up of deformations from one pass to the next, which can
first and second deformations, suggesting that complete softening oc- be attributed to increased critical deformation as a function of test tem-
curred in the four conditions under study. In the second reloading, the perature. At higher deformations, it is expected that the softening pro-
shapes of the curves in the simulations show some differences; with cess is triggered by static recrystallization, except in the experiment
ε = 0.2 and (tpass) = 10 s, and the shape and level of stress in the second performed with ε = 0.3 and tpass = 50 s.
curve are similar to those of the first, while the level of stress and work Fig. 13 shows the microstructures of samples that were deformed
hardening rate were lower in the other three experiments. Considering under continuous cooling at ε = 0.3 and tpass = 5 s, with interrup-
that the deformations applied are greater than the critical strain for the tions of the tests in regions I and II (see Fig. 2). Fig. 13(a) shows a
F.R. Bernardes et al. / Materials Science and Engineering C 51 (2015) 87–98 95

ε
ε

(a)
Fig. 9. Dependence of non-recrystallization temperature (Tnr) on interpass time (tpass) and
applied strain per pass.

microstructure composed of recrystallized grains; this test was stopped in


the third pass, at 1200 °C, region I. The microstructure in Fig. 13(b) shows
elongated grains, indicating incomplete recrystallization between passes

(b)
Fig. 11. (a) Elongated grains pancake-shaped without recrystallization at 1000 °C and 5.0 s
and b) recrystallized grains at 1000 °C and 50.0 s.

with the test stopped in the 13th pass, at 1000 °C, in region II below the
Tnr, and intense stress build-up leading to a matrix of pancake-shaped
grains, which is detrimental to grain refinement. However, this does not
cause dynamic recrystallization (DRX) of this steel because of interpass
precipitation. Note the difference from the initial microstructure, showing
coarse grain boundaries after soaking and numerous annealing twins
with an average diameter of 86 μm, Fig. 14 [17]. Figs. 8 and 13 also
show numerous coalesced precipitates of various sizes, particularly inside
shapeless grains smaller than 100 nm, which did not dissolve completely
during reheating.
From the mechanical standpoint, a comparison of the multipass
curves corrected at constant temperature and those obtained from the
continuous isothermal runs (Fig. 12) indicates that the material softens
at each interruption and that softening is more intense in longer
interpass times. With the increase in deformation during reloading,
the curve reaches stress levels above the curves obtained in the contin-
uous isothermal runs [20]. In contrast with softening during long
interpass times, the material hardens during reloading, because after
complete softening the material behaves like it does in the first stage,
i.e., it hardens until the dislocation density reaches a critical value for
the onset of recrystallization.
When deformation is performed at 1200 °C, the size of the recrystal-
Fig. 10. Dependence of stress build-up (Δσ) on interpass time (tpass) and non-recrystalli- lized grain grows, as indicated in Fig. 8. In this region, also note the pres-
zation temperature (Tnr) in ASTM F-1586 steel. ence of Z-phase precipitates, with recrystallization beginning dynamically
96 F.R. Bernardes et al. / Materials Science and Engineering C 51 (2015) 87–98

Fig. 12. Plastic flow curves corrected to account for the effect of temperature drop, simulating tests at a constant temperature (1200 °C).

and developing statically at the end of deformation, and hence, without continuous isothermal hot torsion tests in the conditions of 900, 1000
the presence of stresses. Note the strong hardening that occurs in the and 1050 °C, with strain rates varying from 0.01–10 s −1, reported in
low temperature condition at 1000 °C in the 13th pass, causing the level previous studies [16].
of stress to increase and the temperature to decrease. As can be seen,
the microstructure processed with short interpass times shows finer 3.4. Calculation of the supersaturation through the solubility product of
grains and twinning. Both static (SRX) and dynamic recrystallizations Z-phase
(DRX) cause softening of the material. However, SRX produces new unde-
formed grains that continue to grow, while DRX produces new grains The critical supersaturation required for the nucleation of undeformed
whose dislocation density is preserved under the applied deformations, ASTM F-1586 austenitic steel (solubilized sample) is high, ks = 4.8, and is
without significant grain growth. the driving force or change in free enthalpy for precipitation to occur.
Tests with multipass deformations under continuous cooling indi- However, the stored energy increases with applied deformation, i.e., the
cate the temperature range in which ASTM F-1586 steel can be de- thermodynamic potential for recrystallization with an increase in the
formed without the occurrence of stress build-up from one pass to number of nucleation sites and Z-phase precipitation is observed at low
another. The non-recrystallization temperature (Tnr) indicates that supersaturation, ks = 1.3. These values are below those reported in the lit-
grain refinement can be controlled by static or metadynamic recrystal- erature for microalloyed Nb steels with precipitation of niobium
lization at high temperatures (T N 1100 °C) and by controlled rolling carbonitride Nb (CN), 5.0 b ks b 8.0 [13]. The reason for this is that solid
at temperatures below Tnr (T b 1050 °C). It should be noted that at state precipitation of Z-phase at temperatures below 1100 °C (close to
very low temperatures (T b 900 °C), continuous isothermal tests and Tnr) is favorable, because Cr and N solubility decreases significantly
tests under continuous cooling indicate that this material is not suffi- below this temperature, and the higher concentration of Cr is removed
ciently ductile to be subjected to a sequence of more severe passes. In from the matrix by the precipitates that are formed, decreasing its me-
other words, this material cannot withstand an increase in loading, chanical strength and corrosion resistance [17].
which causes it to fail, as revealed by the plastic flow curves from
4. Conclusions

Table 2
Correlation between EMS and 1/T when interpass softening is promoted by complete stat- – At high temperatures, this alloy undergoes metadynamic recrystalli-
ic interpass recrystallization. zation (MDRX) and static recrystallization (SRX) after deformation
under continuous cooling, with partial recrystallization at tempera-
Deformation (tpass) = 10 s (tpass) = 50 s
tures below 1100 °C, the range in which Z-phase precipitation occurs
0.2 σ ¼ −159:430 þ 382  ð1000=TÞ σ ¼ −299:760 þ 570  ð1000=TÞ upon cooling, as predicted by the phase diagram and microstructural
0.3 σ ¼ −185:715 þ 414  ð1000=TÞ σ ¼ −226:580 þ 462  ð1000=TÞ
analysis.
F.R. Bernardes et al. / Materials Science and Engineering C 51 (2015) 87–98 97

(a)

(b)
Fig. 13. Microstructures obtained under continuous cooling, with ε = 0.3 and (tpass) = 5.0 s, with testing interrupted in (a) region I (1200 °C) and (b) region II (1000 °C).

– The tests indicate that the degree of stress depends on the defor- Acknowledgments
mation temperature, and that the slope of the EMS vs. 1/T pre-
sents two distinct behaviors at decreasing temperature, with a The authors thank the Brazilian research funding agencies CAPES
transition at about 1100 °C, i.e., the non-recrystallization temper- (Federal Agency for the Support and Improvement of Higher Educa-
ature (Tnr). tion), CNPq (National Council for Scientific and Technological Develop-
– At temperatures below T nr, strain-induced precipitation of Z- ment) and FAPEMA (Maranhão Foundation for Scientific Research and
phase (NbCrN) occurs over a given time, inhibiting recrystalliza- Development) for their financial support of this work.
tion and causing stepwise stress build-up, which in turn is re-
sponsible for the formation of elongated pancake-shaped grains.
– The influence of interpass time tpass on Tnr presents two distinct References
behaviors. When the interpass time is short (less than 30 s), re-
[1] M. Navarro, A. Michiardi, O. Castano, J.A. Planell, Biomaterials in orthopaedics, J. R.
crystallization is delayed by the increase in the fraction of Z- Soc. Interface 5 (2008) 1137–1158.
phase precipitates, and Tnr increases. At interpass times of more [2] E.J. Giordani, A.M. Jorge Jr., O. Balancin, Proportion of recovery and recrystallization
during inter-pass times at high temperatures on a Nb- and N-bearing austenitic
than 30 s, the coalescence and dissolution of precipitates reduce
stainless steel biomaterial, Scr. Mater. 55 (2006) 743–746.
the Tnr. [3] C. Ornhagen, J.O. Nilsson, H. Vannevik, Characterization of a nitrogen-rich austenitic
stainless steel used for osteosynthesis devices, J. Biomed. Mater. Res. 31 (1996)
97–103.
[4] L.J. Cuddy, in: A.J. DeARDO (Ed.), Thermomechanical Processing of Microalloyed
Austenite, AIME, Warrendale, 1982, p. 129.
[5] S.S. Hansen, J.B. Vander Sande, M. Cohen, Niobiun carbonitride precipitation and
austenite recrystallization in hot-rolled microalloyed steels, Metall. Trans. 11A
(1980) 387–402.
[6] D.H. Jack, K.H. Jack, Invited review: carbides and nitrides in steel, Mater. Sci. Eng. 11
(1973) 1–27.
[7] E.J. Giordani, Evidence of strain-induced precipitation on a Nb- and N-bearing aus-
tenitic stainless steel biomaterial, Mater. Sci. Forum 500–501 (2005) 179–186.
[8] I. Andersen, O. Grong, Analytical modelling of grain growth in metals and alloys in
the presence of frowing and dissolving precipitates—I. Normal grain growth, Acta
Metall. Mater. 43 (7) (1995) 2673–2688.
[9] J. Ernrman, M. Schwind, P. Liu, O. Nilsson, H.O. Andren, J. Agren, Precipitation reac-
tions caused by nitrogen uptake during service at high temperatures of a niobium
stabilised austenitic stainless steel, Acta Mater. 52 (2004) 4337–4350.
[10] E.J. Giordani, Propriedades e mecanismos de nucleação de trincas por fadiga em
meio neutro e meio fisiológico artificial de dois aços inoxidáveis austeníticos
utilizados como biomateriais, FEM/UNICAMP, (Tese de Doutorado) Campinas, 2001.
[11] D.Q. Bai, S. Yue, W.P. Sun, J.J. Jonas, Effect deformation parameters on the no-
recrystallization temperature in Nb-beating steels, Metall. Trans. A 24A (1993)
2151–2159.
[12] F. Boratto, R. Barbosa, S. Yue, J.J. Jonas, Proc. Int. Conf. Physical Metallurgy of
Fig. 14. Initial microstructure heated to 1200 °C and held there for 300 s and after cooled in Thermomechanical Processing of Steels and Other Metals Thermec 88, ISIJ Interna-
water. tional, Tokyo, 1988. 383.
98 F.R. Bernardes et al. / Materials Science and Engineering C 51 (2015) 87–98

[13] S.F. Medina, The influence of niobium on the static recrystallization of hot deformed steel biomaterial: influence of strain rate and temperature, Mater. Sci. Eng. A 582
austenite and on strain induced precipitation kinetics, Scr. Metall. Mater. 32 (1) (2012) 96–107.
(1995) 43–48. [18] M. Gomes, L. Rancel, B.J. Fernandez, S.F. Medina, Evolution of austenite static recrys-
[14] E.I. Poliak, J.J. Jonas, A one-parameter approach to determining the critical condi- tallization and grain size during hot rolling of a V-microalloyed steel, Mater. Sci. Eng.
tions for the initiation of dynamic recrystallization, Acta Mater. 44 (n1) (1996) A 501 (2009) 188–196.
127–136. [19] T. Sakai, A. Belyakov, R. Kaibyshev, H. Miur, J.J. Jonas, Dynamic and post-dynamic re-
[15] S.H. Mousavi, S. Yue, The necessity of dynamic precipitation for the occurrence of crystallization under hot, cold and severe plastic deformation conditions, Prog.
no-recrystallization temperature in Nb-microalloyed steel, Mater. Sci. Eng. A 528 Mater. Sci. 60 (2014) 130–207.
(2011) 803–807. [20] S.F. Rodrigues, E.S. Silva, G.S. Reis, R.C. Sousa, O. Balancin, Prediction of hot flow plas-
[16] E.S. Silva, R.C. Sousa, A.M. Jorge Jr., O. Balancin, Hot deformation behavior of an Nb- and tic curves of ISO 5832-9 steel used as orthopedic implants, Mater. Res. 17 (2) (2014)
N-bearing austenitic stainless steel biomaterial, Mater. Sci. Eng. A 543 (2012) 69–75. 436–444.
[17] R.C. Sousa, E.S. Silva, A.M. Jorge Jr., J.M. Cabrera, O. Balancin, Dynamic recovery and
dynamic recrystallization competition on a Nb- and N-bearing austenitic stainless

You might also like