Accepted Manuscript
Mechanical modeling of coupled plasticity and phase transformation effects in a martensitic high strength bearing steel Selcuk Hazar, Bo Alfredsson, Junbiao Lai PII: DOI: Reference:
S0167-6636(17)30149-7 10.1016/j.mechmat.2017.10.001 MECMAT 2802
To appear in:
Mechanics of Materials
Received date: Revised date: Accepted date:
24 February 2017 14 September 2017 2 October 2017
Please cite this article as: Selcuk Hazar, Bo Alfredsson, Junbiao Lai, Mechanical modeling of coupled plasticity and phase transformation effects in a martensitic high strength bearing steel, Mechanics of Materials (2017), doi: 10.1016/j.mechmat.2017.10.001
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT
Highlihgts • Both stress and strain induced phase change are taken into account in the definition of constitutive behavior of high strength steels.
CR IP T
• The phase transformation enhances and dominates the strength differential effect over plasticity effects.
• Strength differential effect simulated successfully by employing the phase transformation kinetics together with a pressure sensitive surface for plastic yielding.
AC
CE
PT
ED
M
AN US
• The nonlinearity of elastic strains are also taken into account in the material model.
1
ACCEPTED MANUSCRIPT
Mechanical modeling of coupled plasticity and phase transformation effects in a martensitic high strength bearing steel
CR IP T
Selcuk Hazara,∗, Bo Alfredssona , Junbiao Laib a Department
of Solid Mechanics, Royal institute of technology (KTH), 100 44 Stockholm, Sweden b SKF Engineering & Research Centre, P.O. Box 2350, 3430 DT Nieuwegein, The Netherlands
AN US
Abstract
CE
PT
ED
M
The stress and strain induced solid to solid phase transformation of retained austenite in a martensitic high strength bearing steel has been studied. Monotonic tension experiments that were carried out at different temperatures using this high strength steel showed that not only the strain induced but also the stress induced phase change plays a crucial role in the phase transformation of retained austenite to martensite. In the material model, plastic deformation was defined using the Drucker Prager yield surface through a nonassociated flow rule accompanied by nonlinear kinematic and isotropic hardening. The hardening was coupled with stress and strain induced phase transformations. A nonlinear elastic effect based on elastic dilation was included in the constitutive model by extending the bulk modulus with a second order term. For the finite element analysis, the material model was written as a user defined material subroutine (UMAT). The numerical simulations were done using ABAQUS and compared to monotonic tension, compression and cyclic experiments. The results showed that the strength differential effect and the volumetric change under loading are closely related to the transformation of retained austenite to martensite. At low temperatures the effect of stress induced phase transformation on yield strength was noticeable. It was concluded that at certain temperatures both strain and stress induced phase transformations significantly affect mechanical behavior of the high strength steel.
AC
Keywords: Transformation induced plasticity, Stress induced phase transformation, Strain induced phase transformation, Strength differential effect, Nonassociated flow rule
∗ Corresponding
author Email address:
[email protected] (Selcuk Hazar)
Preprint submitted to Mechanics of Materials
October 10, 2017
AC
CE
PT
ED
M
AN US
CR IP T
ACCEPTED MANUSCRIPT
2
AN US
CR IP T
ACCEPTED MANUSCRIPT
1. Introduction
4 5 6 7 8 9 10 11
M
CE
12
ED
3
At equilibrium conditions, materials tend to stay in their crystallographically more stable phase i.e. the atomic configuration having the lowest energy level. In high strength bearing steels and other materials undergoing phase change upon cooling, solid to solid phase transformation occurs from the parent phase (austenite), which is stable at high temperatures, to the product phase (martensite) that is stable at low temperatures. During quenching of the high strength steel, when the temperature is above the martensite finish temperature and the energy level required to complete the transformation is not reached, austenite cannot transform completely to martensite. As a result, a metastable phase named retained austenite forms. The amount of austenite retained after heat treatment is affected by carbon content, quenching temperature and alloy composition. Aside from temperature change, application of load and shear band intersection triggers the phase change of retained austenite to martensite. This mechanically induced phase change enhances ductility, formability, toughness and strength of such steels and makes them attractive for automotive, aerospace and variety of other engineering applications. The martensitic transformation of retained austenite upon loading can be either stress or strain induced (assisted). When the temperature is below transformation start temperature (Msd ), stress induced displacive martensitic transformation of the retained austenite to martensite takes place at nucleation sites in the parent phase. Up to Msd stress induced transformation takes place below the yield stress (σ Y ) of the material. In this temperature range, the transformation start stress increases linearly with the temperature (Bolling and Richman,
PT
1 2
13
14
15
AC
16
17
18
19
20
21 22 23
3
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23
ACCEPTED MANUSCRIPT
34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56
CE
57 58
59
60
61
AC
62
63
64
65
66
67 68 69
24 25 26 27 28 29
CR IP T
32 33
AN US
30 31
M
28 29
ED
26 27
1970a). Kinetics of stress induced transformation is described using either Gibbs or Helmholtz free energy potentials by considering the thermodynamic effect of the applied stress (Abeyaratne and Knowles, 1993; Bhattacharyya and Weng, 1994; Chan et al., 2012; Entchev and Lagoudas, 2004). Using Helmholtz free energy, Moumni et al. (2011) proposed a model for stress induced phase transformation based on the formalism of principle of generalized standard materials with internal constraints (Halphen and Nguyen, 1974; Moumni, 1995). The simulations that they performed using the thermomechanical model had good correlations with experimental results. In the literature, the constitutive models of stress induced transformation are studied not only for high strength steels but also for other alloys like Nitinol (Arghavani et al., 2010; Auricchio et al., 2007; Chan et al., 2012; Levitas and Ozsoy, 2009a; Moumni et al., 2011; Popov and Lagoudas, 2007; Saleeb et al., 2011; Zaki and Moumni, 2007b; Zhang et al., 2007). Strain induced (assisted) phase change initiates through formation of highly potent new sites via shear band interactions when the temperature is above Msd . In this case, depending on temperature, the transformation start stress can be higher than the yield stress. In previous studies, the strain induced transformation, that is commonly investigated in TRIP steels, was found to undergo irreversible martensitic transformation (Cherkaoui et al., 2000, 1998; Goel et al., 1985; Idesman et al., 2000; Iwamoto and Tsuta, 2000; Leblond et al., 1986; Olson and Azrin, 1978; Olson and Cohen, 1975; Sangal et al., 1985; Stringfellow et al., 1992; Taleb and Sidoroff, 2003; Tomita and Iwamoto, 1995). Olson and Cohen (1975) proposed a model that outlines the strain induced phase transformation behavior in austenite steel through interaction of shear bands. Their model was only able to capture uniaxial stress effects under isothermal conditions. Stringfellow et al. (1992) enhanced the model of Olson and Cohen (1975) by considering the stress state sensitivity, since the evolution of martensite at a material point is effected not only by the plastic strain and temperature but also by the stress state history. The model proposed by Stringfellow et al. (1992) was improved by taking into account of the influence of strain rate on toughness and ductility (Tomita and Iwamoto, 1995), thermocoupled effects (Iwamoto, 2002, 2004), deformation mode dependence (Iwamoto et al., 1998) and the effect of grain size (Iwamoto and Tsuta, 2000). Perlade et al. (2003) and Papatriantafillou et al. (2006) used the model of Stringfellow et al. (1992) to study the multiphase TRIP steels having different retained austenite levels. In a later study Sierra and Nemes (2008) modified the same model to study the interaction between embedded retained austenite islands and the surrounding matrix. In the early attempts, formulation of transformation kinetics in TRIP steels is based on either large scale orientation of the transformation strain in the martensite phase (Magee, 1966) or the micromechanical plastic strain accommodation that emerges from martensitic phase transformation and arises in the parent phase (Greenwood and Johnson, 1965). Based on these two principles, most of the constitutive models for TRIP steels are derived using either macro scale (Cherkaoui et al., 2000; Fischer et al., 1998, 2000; Leblond et al., 1989;
PT
24 25
4
30
31
32
33
34
35
36
37 38 39
40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69
ACCEPTED MANUSCRIPT
80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102
CE
103 104
105
106 107
AC
108 109 110 111 112 113 114 115
70 71 72 73 74 75
CR IP T
78 79
AN US
76 77
M
74 75
ED
72 73
Levitas, 1998; Moumni et al., 2011; Perlade et al., 2003; Tanaka et al., 1996) or micro scale approaches (Levitas and Ozsoy, 2009a,b; Ma and Hartmaier, 2015). In the vast majority of constitutive models developed for materials undergoing ˙ is prescribed as phase transformation, the rate of martensite volume fraction (ζ) the main internal state variable. Previous researches have documented that the rate of martensite transformation is affected by Ms temperature (Inoue et al., 1984; Neu and Sehitoglu, 1992), hydrostatic stress (Jacques et al., 2001; Neu and Sehitoglu, 1992), carbon content (Neu and Sehitoglu, 1992; Seong et al., 2004), quenching temperature and rate, plastic strains (Seong et al., 2004) and percentage of retained austenite (Neu and Sehitoglu, 1992). The study by Young (1988) showed that evolution of martensite fraction with respect to plastic strain during stress induced transformation is more linear compared to the case of strain induced phase change. Not only the transformation kinetics, but also the morphology of the martensite microstructure formed via stress or strain assisted transformation differs. Strain induced phase change results in fine lath martensite phase formed at the shear band intersections, whereas plate martensite formation is observed when the transformation is induced by stress (Young, 1988). Phase transformation in TRIP steels results in a strength differential effect (SDE), which is known as the asymmetry between tensile and compressive yield stresses, because the retained austenite is more susceptible to transform to martensite under tension than in compression. The stress induced transformation is affected by temperature, for this reason SDE is found to be higher when the temperature is decreased (Singh et al., 2000). It is observed that SDE depends not only on temperature but also on the hydrostatic stress, therefore Drucker Prager yield surface has to be used to define the plastic flow (Casey and Sullivan, 1985; Spitzig and Richmond, 1984). In addition, in high strength steels the associated flow rule would result in a plasticity induced volume change which is different from the experimental findings, for this reason a nonassociated flow rule has to be considered in the material model (Linares Arregui and Alfredsson, 2010; Stoughton and Yoon, 2004). Not only the phase transformation but also microcracking, residual stresses, volume expansion and solute dislocation interaction are some of the factors that can give rise to SDE. For instance Linares Arregui and Alfredsson (2010) observed SDE in high strength bainitic steel which did not undergo any mechanically induced phase transformation. Previously, the stress dependency of elastic modulus for high strength steels is discussed by some researchers (Cleveland and Ghosh, 2002; Lee et al., 2013; Linares Arregui and Alfredsson, 2013; Sun and Wagoner, 2011; Yoshida et al., 2002; Zavattieri et al., 2009). According to the results of push–pull cyclic experiments by Sommer et al. (1991) on SAE 52100 high strength steel, when the linear elastic part is subtracted, a sickle shaped asymmetric hysteresis loop is obtained in stress vs. plastic strain graphs. In a recent work, Linares Arregui and Alfredsson (2013) showed that the nonlinear elastic material behavior should be considered in the material model of the high strength bainitic roller bearing steels. Moreover, Linares Arregui and Alfredsson (2010) used combined nonlinear kinematic and nonlinear isotropic behaviors to precisely define the
PT
70 71
5
76
77
78
79
80
81
82
83
84
85 86 87 88 89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115
CR IP T
Stress (σ)
ACCEPTED MANUSCRIPT
σY
σs
Temperature (T)
AN US
Msd
Stress Induced Transformation Coupled Transformation
Strain Induced Transformation
Figure 1: The effect of temperature on the start of strain and stress induced transformations. Yellow region represents where the austenite is stable, blue and gray zones represents the stress induced and coupled transformation zones respectively.
119 120 121 122 123 124 125 126
M
CE
127
ED
118
plastic material response in their experimental measurements. Figure 1 illustrates the temperature effect on the start of strain and stress induced transformations. At a certain critical temperature stress induced transformation starts at a certain stress level, σs , which can be below yield stress (Bolling and Richman, 1970a; Olson and Cohen , 1972). Therefore coupled transformation behavior takes place after plastic yielding. Strain induced transformation on the other hand starts at the onset of yielding. Above Msd the shape of yield stress curve representing the case of coupled transformation depends on material properties but increases up to a certain temperature above which no phase change is observed (Young, 1988). Therefore, coupled transformation behaviour takes place after plastic yielding. According to Neu and Sehitoglu (1991) it is difficult to distinguish stress induced from strain induced transformation when the deformation increases beyond the yield point. Recently Ma and Hartmaier (2015) proposed a constitutive model considering strain induced martensite nucleation and stress assisted martensite growth using Helmholtz free energy potential. Although they stated that their numerical calculations fit to the experimental results of Lebedev and Kosarchuk (2000), no direct comparison were presented between the numerical results and the corresponding measurements. In the present research, stress and strain induced phase transformation mechanisms were coupled with plastic deformation that was described using a Drucker Prager yield surface with a nonassociated flow rule accompanied with combining nonlinear kinematic and isotropic hardening. For this purpose the model represented by Stringfellow
PT
116 117
128 129 130
AC
131 132 133 134 135 136 137 138
6
116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138
ACCEPTED MANUSCRIPT
160
2. Material model
143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158
161 162 163 164
AN US
142
M
141
This section presents the constitutive model of the investigated martensitic high strength bearing steel (DIN 100CrMnMoSi8), with the chemical composition given in Table 1. Figure 2 shows the microstructure of the quenched and tempered martensitic steel with retained austenite content.
ED
140
139 140 141 142 143 144
CR IP T
159
et al. (1992) was extended by including the stress induced transformation effect using Nguyen's formalism of standard materials, which was applied to TRIP steels by Moumni et al. (2011). The temperature dependence of the onset of nonlinear stress strain behavior was the main motivation for implementing coupled stress and strain induced phase change. Compared to the models existing the literature this model couples stress and strain induced phase changes with a nonassociated flow rule and nonlinear elasticity. The predictions were in a good agreement with the experimental findings. Since the morphology and mechanisms of the stress and strain induced martensite phases differ, the proposed model includes the contribution of each phase separately and in combination to the other mechanical properties of the high strength steel, such as nonlinear elasticity, nonlinear hardening and pressure dependence of the yield surface. The paper is structured in the following way: First the material model is described in section two. The first subsection introduces the free energy potential and the effect of nonlinear elasticity. The second subsection describes the nonlinear kinematic and isotropic hardening rules together with nonassociated flow rule. The third subsection presents the strain induced phase transformation mechanism and combines it with the stress induced transformation mechanism. In section three experiments and determination material parameters are discussed. The section 4 describes the method and procedure of numerical simulations. Finally, conclusions are drawn.
139
145 146 147 148 149 150 151 152 153 154 155 156 157 158 159
160
161 162 163 164
PT
Table 1: Chemical composition (wt %) of 100CrMnMoSi8 (Lai et al., 2016).
Cr 1.86
Mn 0.81
Mo 0.59
Si 0.21
CE
C 0.91
165
167 168 169 170 171 172
Cu 0.19
Fe balance
2.1. Definition of free energy potential and nonlinear elasticity Patel and Cohen (1953) were among the first researchers to investigate the contribution of the applied loading to phase transformation through a free energy formulation. Using a similar approach, Olson and Cohen (1982) represented the stress state sensitivity of the transformation and the toughening effect of stress assisted transformation. In this part, the thermodynamic force conjugate to evolution of martensite fraction was computed using the formalism of generalized standard materials with the internal constraints that were derived by
AC
166
Ni 0.18
7
165 166 167 168 169 170 171 172
AN US
CR IP T
ACCEPTED MANUSCRIPT
Figure 2: Microstructure of martensitic high strength steel DIN 100CrMnMoSi8 (Lai et al., 2016), in this work some of the retained austenite islands are marked inside red circles.
174 175 176
Halphen and Nguyen (1974). A similar aproach was implemented by Moumni et al. (2011). In this study Reuss scheme was used to define the macroscopic strain tensor which is an average over the representative volume element and given as (1 −
178 179
+
ζ>0
182
184
185
− εij = 0,
(1 − ζ) > 0.
and
ν1 > 0,
ν1 ζ = 0, ν2 > 0,
ν2 (1 − ζ) = 0.
174 175 176
177 178 179
(2) 180 181
182 183
(4)
By considering the contribution of internal variables of plastic deformation, the free energy potential can be written as 1 m m m 1 a a a εij Eijkl εkl + ζ εij Eijkl εkl + C(T ) W = (1 − ζ) 2 2 (5) 3 X 1 1 1 pl m m m el 3 pl + C χij χij + K1 εkk + Q εeq + exp(−bεeq ) . 3 m=1 3 b 8
173
(1)
where Λij , ν1 and ν2 are the Lagrange multipliers. ν1 and ν2 are associated with unilateral constraints and
AC
183
+
εpl ij
Using Eqs (1) and (2), the constraints potential, Wl , in the form of a Lagrangian formulation, was given by h i pl tr Wl = −Λij (1 − ζ)εaij + ζεm (3) ij + εij + εij − εij − ν1 ζ − ν2 (1 − ζ),
CE
181
+
εtr ij
PT
180
ζεm ij
where the terms εaij and εm ij are elastic strains of austenite and martensite phases, respectively. Martensite volume fraction, ζ, that is bounded in the interval [0, 1] can be stated in inequality form as
ED
177
ζ)εaij
M
173
184 185
ACCEPTED MANUSCRIPT
186 187 188
In Eq. (5), E a and E m are the linear elasticity tensors for martensite and austenite phases. C(T ) is the phase change heat density that is assumed to be a linear function of temperature (Zaki and Moumni, 2007b) C(T ) = ρ(T − Ms ) + κ.
191 192
193 194 195 196 197
In Eq. (5), C is the slope between the back stress Xij , where X˙ ij = 32 C ε˙pl ij − pl ˙ γXij λ, and the plastic strain rate ε˙ij when γ = 0, C/γ defines the saturated value of evolution of the back stress. χij is the conjugate of back stress Xij and defined as 3 X 3 m m χ˙ ij = C Xij (7) 2 m=1
In Eq. (5), Q is the limiting value of the change in isotropic yield surface size 3 and b is the rate of hardening before reaching the stress limit. The term K1 εel kk represents the nonlinear elastic potential with the same nonlinear bulk modulus parameter K1 for both phases. εel ij is the elastic strain tensor that is composed of the elastic strains of the martensite and austenite phases, a m εel ij = (1 − ζ)εij + ζεij .
200
190
191 192
193
194 195
196 197
Given the expressions of the free energy density W and of the constraints potential Wl in (3) and Eqs (5), the Lagrangian L = W + Wl was constructed to define the generalized forces (Halphen and Nguyen, 1974; Zaki and Moumni, 2007b) 1 a a a 1 m m m L(εij , T, εaij , εm , ζ) = (1 − ζ) ε E ε + ζ ε E ε + C(T ) ij 2 ij ijkl kl 2 ij ijkl kl 3 1 1 X m m m 1 3 pl pl + K1 εel + C χ χ + Q ε + exp(−bε ) kk ij ij eq eq 3 3 m=1 b h i pl tr − Λij (1 − ζ)(εaij ) + ζ(εm ij ) + εij + εij − εij
198 199 200 201
PT
ED
201
189
(8)
M
198 199
188
CR IP T
190
187
(6)
AN US
189
186
− ν1 ζ − ν2 (1 − ζ),
(9)
where Eq. (4) must be verified. The state equations can be written as
CE
202
AC
∂L,εij = σij
⇒
−∂L,εaij = 0
⇒
−∂L,εm =0 ij
⇒
−∂L,ζ = Cζ
⇒
202
Λij − σij = 0, 2 K1 εel kk δij
(10)
a + εakl Eijkl − Λij = 0,
(11)
m + εm kl Eijkl − Λij = 0, 1 a a a m m εkl Eijkl εkl − εm kl Eijkl εkl 2
(12)
pl tr (1 − ζ)(εaij ) + ζ(εm ij ) + εij + εij − εij = 0.
(14)
2 K1 εel kk δij
(13)
2
a m − C(T ) − (Λij − K1 εel ij )(εij − εij ) = Cζ ,
−∂L,Λij = 0
⇒
9
ACCEPTED MANUSCRIPT
203 204
in the equations above Cζ is the only nonzero thermodynamic force. Using the Eqs (10), (11), (12) and (14) the constitutive equation became 2
tr el σij = Eijkl (εkl − εpl kl − εkl ) + K1 εkk δij ,
where Eijkl is
204
(15) 205
CR IP T
205
203
a m −1 −1 Eijkl = (1 − ζ)(Eijkl . )−1 + ζ(Eijkl )
(16)
At undeformed state, assuming the same Poisson's ratio ν0 for both phases, the equivalent Young's, E0eq , bulk, K0eq , and shear, Geq 0 moduli were expressed as
207 208 209 210
(18) (19)
In the work of Linares Arregui and Alfredsson (2013) a small but significant nonlinear elastic behavior was detected during uniaxial cyclic push pull experiments on high strength bainitic steel. The nonlinear elastic behavior was traced to the volumetric bulk behavior of the material and defined using the bulk modulus as a function of elastic strains, K(εel kk ) and a constant shear modulus, G. Using Eq. (15) eq el K(εel (20) kk ) = K0 + K1 εkk .
ED
211
(17)
M
206
−1 1−ζ ζ + , E0a E0m E0eq , K0eq = 3(1 − 2ν0 ) E0eq Geq . 0 = 2(1 + ν0 )
AN US
E0eq =
206 207 208 209 210 211
CE
PT
The nonlinearity of Young's modulus and Poisson's ratio could be written in terms of their relation to K(εel kk ) as (21)
ν(εel kk ) =
eq 3K(εel kk ) − 2G0 eq . 2(3K(εel kk ) + G0 )
(22)
Using Eq. (15) the hydrostatic stress, σ h = σii /3, was stated as 2
eq el el el σ h = K(εel kk )εkk = K0 εkk + K1 εkk .
AC
212
el 9Geq 0 K(εkk ) eq , 3K(εel kk ) + G0
E(εel kk ) =
The uniaxial elastic strain can be calculated using q eq 2 eq 4 σ 3 σ11 K1 + (K0 ) − K0 11 el ε11 = + . 3Geq 6K1 0
213 214
212
(23)
(24)
Appendix A shows the derivation of uniaxial elastic strain, εel 11 , at tension or compression using Eqs (21) and (23). 10
213 214
ACCEPTED MANUSCRIPT
1.02 1.00
CR IP T
0.96 0.94 0.92 0.90 0.88 0.86 −0.004
ζ =0.99 ζ =0.80 ζ =0.60 −0.002
AN US
WNL /WL
0.98
0.000
0.002
0.004
0.006
el εkk /3
Figure 3: W L is the linear elastic strain energy and W N L is the nonlinear elastic strain energy.
The strain energy function of the nonlinear elastic material became 1 2 a NL a a 2 a a a W =(1 − ζ) K0 − G0 (εkk ) + G0 εij εij + 2 3 1 2 1 el 3 m 2 m m m ζ K0m − Gm (ε ) + G ε ε kk 0 ij ij + K1 εkk , 2 3 0 3
217 218 219 220
222
223
224
225
AC
226 227
228 229
230 231 232 233
(25)
3
CE
221
215
L which can be stated alternatively as W N L = W L + 13 K1 εel repkk , where W resents sum of the linear elastic strain energies of both phases. The change of W N L / W L with respect to hydrostatic strains is plotted in Fig. 3 for ζ = 0.60, 0.80, 0.99 to see the effect of phase change on nonlinear elasticity. Fig. 3 shows that strain energy is effected by nonlinearity when the strains are high and the nonlinear elasticity is more noticeable at high ζ when the retained austenite level is low. The stress strain graph of a push pull experiment under a cyclic load with a compressive mean load, Pm = −20.8 kN and amplitude of Pa = 48.5 kN performed by Linares Arregui and Alfredsson (2010) (test no: 13832) is given in Fig. 4(a). The figure shows the 1st and 10th cycles of the experiment. The effect of nonlinear elasticity on the inelastic stress strain cycle is given in Fig. 4(b). The inelastic curves are derived either by subtracting linear elastic (L), εin 11 = el = ε − ε and Eq. (24) ε11 − σ/E eq , or nonlinear elastic (NL) strains, using εin 11 11 11 from the experimental results in Fig. 4(a). When the two cyclic responses shown in Fig. 4(b) are compared, it is observed that nonlinear elastic strains exist during unloading, since steeper response is obtained for the NL curves. The vertical unloading response from both tension and compression that was derived
PT
216
ED
M
215
11
216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233
1500
1500
1000
1000
500
500 0
σ MPa
)
0
−500
−1000
−1000
−1500
−1500
−2000
−2000
−2500 −0.020
−0.015
−0.010
ǫ
−0.005
0.000
0.005
0.010
−2500 −0.0050
CR IP T
(
−500
(
σ MPa
)
ACCEPTED MANUSCRIPT
NL-cycle 10 L-cycle 10 NL-cycle 1 L-cycle 1
−0.0045
(a)
−0.0040
ǫin
−0.0035
−0.0030
−0.0025
−0.0020
(b)
AN US
Figure 4: a. Load controlled push-pull experiment with compressive mean load. b. Plastic hysteresis loops using linear (L) and nonlinear (NL) elasticity.
240
2.2. Nonlinear kinematic and isotropic hardening
237 238
241 242 243 244 245
246 247
By virtue of the SDE observed in the push pull experiments conducted by Linares Arregui and Alfredsson (2010), it was found that a Drucker Prager type (Drucker and Prager, 1952) yield surface with a linear dependency on the hydrostatic stress had to be used to define the plastic behavior. Hence, the yield surface was defined as f = σeq + 3aσ h − σ Y , (26) where a is the slope of the offset yield stress with respect to hydrostatic stress σ h = σii /3. σ Y defines the yield stress which could be expressed using nonlinear isotropic hardening behavior (Chaboche, 1991)
CE
248
249
251 252
σ Y = σ Y 0 + R.
˙ R˙ = b (Q − R) λ.
where λ˙ represents the plastic multiplier, λ˙ = loading, R in Eq. (28) can be integrated to
237 238 239
241 242 243 244 245
246 247 248
249 250
(28) q
2 pl pl 3 ε˙ij ε˙ij
R = Q(1 − exp(−bεpl eq )). 12
235 236
(27)
In the equation above, σ Y 0 represents the virgin yield stress at zero hydrostatic ˙ was defined as stress. The evolution of the yield surface radius, R,
AC
250
234
240
ED
236
PT
235
M
239
using NL illustrates the presence of nonlinear elastic strains. Hence, the cyclic response must be studied in order to separate nonlinear elasticity from plasticity. The martensite fraction ζ at start of the first cycle was taken as ≈ 0.78 and at 10th cycle it was ≈ 0.85. These values, ζ = 0.78, 0.85, were measured using X-ray diffraction on an undeformed reference configuration and on the specimen midsection where it had broken after 10 cycles.
234
= ε˙pl eq . For uniaxial
251 252
(29)
ACCEPTED MANUSCRIPT
The equivalent stress, σeq , used in Eq. (26) was defined as
253
σeq =
256 257
258 259
260
,
253
(30)
d d where σij is the deviatoric stress, σij = σij −σh δij , and δij is the Kronecker delta. In Eq. (30), the nonlinear kinematic hardening is introduced using the back stress tensor, described by Chaboche (1986, 2008); Frederick and Armstrong (2007), 2 ˙ − γXij λ. (31) X˙ ij = C ε˙pl 3 ij In the definition by Chaboche (1986) three sets of kinematic hardening parameters are used 2 m m m˙ X˙ ij = C m ε˙pl (32) ij − γ Xij λ, 3 with 3 X m Xij = Xij . (33)
254
CR IP T
255
d 3 d σij − Xij σij − Xij 2
1/2
AN US
254
255 256 257
258 259
260
m=1
262 263 264 265
Experimental measurements of the volume change due to the dislocation motion in austenitic steels undergoing phase change was found to be different from the predictions of the associated flow rule (Garofalo and Wriedt, 1962; Linares Arregui and Alfredsson, 2010; Rauch et al., 1975; Spitzig and Richmond, 1984), for this reason a nonassociated flow rule was used,
M
261
267 268
˙ ε˙pl ij = λ
270 271
AC
272
273
274 275 276
δg . δσij
R−
3 P
m=1
m 3 γ m m 2 C m αij αij
≤
σY 0 + R
263 264 265
266 267 268
(35)
a∗ ≤ 1, a
269 270 271 272
(36)
in order to satisfy the dissipation equation. The derivation is stated in Appendix B. Using the same procedure as (Linares Arregui and Alfredsson, 2010) the value of a∗ was continuously checked during the simulations to satisfy the condition in Eq. (36) when the plastic strains developed. 13
262
(34)
The nonassociated flow ruled used to define plastic material behavior might result in a physically instabil solution and violence of principal of maximum dissipation. For this reason, the bounds on the pressure dependence in nonassociated flow rule were defined by the
CE
269
3 X 3 γm m m X X = constant, 4 C m ij ij m=1
where a∗ was used to predict plastic dilatation ∆V /V0 , which is known to be very small compared to volume expansion due to phase change (Neu and Sehitoglu, 1992). Using the nonassociated flow rule the plastic strain was defined as
PT
266
ED
g = σeq + 3a∗ σ h − σ Y +
261
273 274 275 276
ACCEPTED MANUSCRIPT
278 279 280 281 282
2.3. Strain induced transformation coupled with stress induced phase transformation The strain induced phase change, that occurs after plastic yielding, see Fig. 1, was introduced by implementing the extended version of Olson and Cohen's model (Olson and Cohen, 1975) which was proposed by Stringfellow et al. (1992). The evolution of strain induced martensite fraction, ζ˙strain , was formulated as ζ˙strain = (1 −
284 285 286
˙ + Bζ Σ),
Aζ given in Eq. (37) was defined as
287
Aζ = αβ(1 − fsb )(fsb )r−1 P ,
291 292 293 294 295 296
where εeq is the equivalent plastic strain of the austenite phase. It was calculated using Eq. (35) assuming the same hardening properties for both phases (Zaki et al., 2010). P in Eq. (39) represents the probability function for transformation start, which was determined by assuming a Gaussian distribution of the potency of the shear bands to activate the nucleation sites that initiate transformation (Stringfellow et al., 1992). The probability function was defined as " 2 # ZΓ 1 1 Γ0 − g¯ P =√ exp − dΓ0 , (41) 2 sg 2π
CE
297
where α, β and r are the material parameters. The volume fraction of shear bands, fsb , is a function of equivalent plastic strain. The equation that describes fsb was written as fsb = 1 − exp −αεpl(a) , (40) eq pl(a)
298
299
301
302
303
281 282
283
284 285 286
288 289 290
291 292 293 294 295 296 297
−∞
where g¯ is the mean value of the distribution and sg is the standard deviation. Γ depends on temperature and stress triaxiality through a dimensionless driving force function for martensite nucleation
AC
300
280
(39)
M
290
279
287
ED
289
PT
288
278
(37)
where ζ is the overall martensite volume fraction and Σ represents the stress ii , which has an effect not only on the nucleation of voids triaxiality, Σ = σσeq (Rice and Tracey, 1969) but also on nucleation of martensite sites (Jacques et al., 2001; McClintock, 1968). Σ˙ was calculated as σ˙ ii σ˙ eq ˙ Σ=Σ − . (38) σii σeq
AN US
283
ζ)(Aζ ε˙pl eq
277
CR IP T
277
Γ = g0 − g1 Θ + g2 Σ,
299 300
(42)
in which g0 is a material parameter, g1 and g2 are the temperature and the linear triaxiality dependency coefficients. Θ is the normalized temperature defined as ˙ = 0, P˙ could be derived as Θ = MTs . When Θ " 2 # g2 1 Γ − g¯ ˙ ˙ P =√ exp − Σ. (43) 2 sg 2πsg 14
298
301 302 303
ACCEPTED MANUSCRIPT
307 308 309
310 311 312 313 314
In Eq (45) ∆v is the relative volume change due to martensite transformation (∆v ≈ 0.02 − 0.05 (Leal, 1984)). N is defined as N = using σeq A = A0 + A1 , sa
318 319 320
1 ˙ (Cζ )2 ξ1 (ξ2 − ζ) + Cζ (Aζ ε˙pl eq + Bζ Σ). 2
322
323
325
326
327
328
329 330 331
307
308
310 311 312 313 314
(47)
where ξ1 and ξ2 are material parameters and Cζ is given by Eq. (13). In Eq. (13) the term multiplied by Cζ is responsible for stress induced phase transformation and determines the value of evolution of stress induced martensite fraction, ζstress . In order to satisfy the dissipation inequality the requirement of Cζ > 0 is discussed in the Appendix C. 3. Experiments and Material parameters To examine stress and strain induced transformations and determine the parameters related to the stress and strain induced transformation mechanisms, monotonic tests were performed at different temperatures. For this purpose, tension experiments were carried out at 22◦ C (RT), 100◦ C and 150◦ C using servo-hydraulic test machines and the specimens shown in Fig. 5. The thin steel specimens with rectangular cross section used in this study were wire electro-discharged from forged rod specimens which had been hardened together with the specimens used in the experiments conducted by Linares Arregui and Alfredsson (2010). Thus, the material and heat treatment were 15
306
(46)
Therefore, the evolution of martensite fraction can be found using the function ∂Φ ˙ ζ˙ = = Cζ ξ1 (ξ2 − ζ) + Aζ ε˙pl (48) eq + Bζ Σ, ∂Cζ
AC
324
305
309
where A0 and A1 are material parameters and sa is the reference austenite hardness (Stringfellow et al., 1992). Finally, the stress and strain transformations were united through the pseudo potential for dissipation of the phase transformation proposed by Moumni et al. (2011). It was here modified as
CE
321
and A was calculated
ED
317
d 3σij 2σeq
PT
316
(44)
where H() is the Heaviside function. Using the evolution of martensite in Eq. (37), the transformation strain became ˙ √1 AN + 1 ∆v δij . ε˙tr = ζ (45) ij 3 2
Φ= 315
304
CR IP T
306
AN US
305
The parameter Bζ in Eq. (37) was (Stringfellow et al., 1992) " 2 # g2 1 Γ − g¯ r ˙ Bζ = √ β(fsb ) exp − H(Σ), 2 sg 2πsg
M
304
315 316
317 318 319 320 321
322
323 324 325 326 327 328 329 330 331
CR IP T
ACCEPTED MANUSCRIPT
Figure 5: Tensile test specimen, thickness 0.70 – 0.76 mm. Scale is cm with mm divisions. 4000 3500
AN US
3000
σ (MPa)
2500
σ (MPa)
1800 1600 1400 1200 1000 800 600 400 200 0 0.0
2000
Exp. RT Exp. 75 ◦ C Exp. 100 ◦ C Exp. 150 ◦ C 0.2
0.4
0.6
ε
0.8
1.0
1000
Exp. Tension Exp. Compression
500
0 0.0
1.4 1e−2
0.5
1.0
1.5 2.0 ε
2.5
3.0
3.5 4.0 1e−2
(b)
M
(a)
1.2
1500
333 334 335 336 337
CE
338
identical for all specimens used in the present and earlier experiments by Linares Arregui and Alfredsson (2010). The specimens with the aforementioned steel grade DIN 100CrMnMoSi8 had been heat treated to martensitic microstructure with a retained austenite content. The monotonic test setup was placed in an oven to maintain constant temperature during the experiment and the axial strains were measured using extensometers. To avoid assembly stresses, the fixtures were carefully aligned with the load line using metal shims. In addition to these experiments, the results of tension, compression and cyclic experiments performed at 75◦ C were taken from the study of Linares Arregui and Alfredsson (2010). The stress strain curves in monotonic tension, compression and cyclic experiments are presented in Figs 6 and 7(a). In this work, the plastic and nonlinear elastic material parameters given by Linares Arregui and Alfredsson (2010) and Linares Arregui and Alfredsson (2013) were used as starting values to determine the material parameters of the corresponding model that considers phase transformation. In their work the phase transformation was not explicitly modeled, instead all the inelastic strains were included in the plastic description. The material properties for nonlinear kinematic and isotropic hardening given in Table 2 are based on the
PT
332
ED
Figure 6: a Tension experiments to specimen rupture b. Tension and compression experiments at 75◦
339 340 341
AC
342
343
344 345 346 347 348 349
16
332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349
ACCEPTED MANUSCRIPT
0.95 1000
−1000 −2000
0.85 0.80
Exp. 75 ◦ C Cyclic
−0.015 −0.010 −0.005 0.000 0.005 0.010 0.015
CR IP T
ζ
σ (MPa)
0.90 0
FE 75 ◦ C Tension FE 75 ◦ C Cyclic Exp. 75 ◦ C Tension Exp. 75 ◦ C Cyclic
−0.010 −0.005
ε
(a)
0.000 ε
0.005
0.010
(b)
353 354 355 356 357 358 359 360 361 362 363 364 365 366 367
CE
368
M
352
studies by Linares Arregui and Alfredsson (2010). Since the present investigation focused on and modeled the phase change taking into acount both stress and strain induced transformation, some parameters in the plastic description that are given in Table 2 were adjusted from the values in the reference (σ Y 0 , a, a∗ , b, Q, C1 , C2 , C3 , γ1 , γ2 , γ3 ). The plasticity parameters were updated iteratively using inverse modeling of the stress strain curves with respect to the experimental results in Figs 6 and 7(a). The parameters quantify the effects of the different mechanical phenomena described in Eqs (1) to (48). They are grouped accordingly in Table 2. In previous studies it is reported that the elastic modulus of martensite is less than that of austenite (Liu and Shan, 2011; Mendiguren et al., 2015) therefore different elastic moduli have to be defined for each phase. The equivalent elasticity tensor should vary during the transformation with respect to the change of martensite fraction as given in Eq. (16). Here the Poisson's ratio and nonlinear bulk modulus K(εel kk ) were taken from Linares Arregui and Alfredsson (2013). The same values were used for both phases. The Young’s moduli for each phase were then determined by inverse modeling of the monotonic experiments presented in Fig. 6, the curves describing the nonlinear elastic behavior given in Fig. 4(b) and using Eqs (16) and (20). The X–ray diffraction (XRD) measurements were performed using a Bruker D8 Discover instrument and Co K–alpha radiation. A 1D silicon strip detector with 192 strips were used and a programmable divergence slit maintained the sample illumination at 5 mm throughout the measurements. The instrument was operated at 40 kV and 35 mA, and the step size used was 0.03 deg. The counting time was varied depending on the cross-section area of the sample to assure proper counting statistics. The accuracy of 2.2-3 % was maintained in the date analysis of single martensite fraction values. The measured X-ray diffraction values at different temperatures are shown in Fig. 8. The parameters affecting the kinetics of the phase transformation were set from the retained austenite levels measured after deformation. The parameter
ED
351
PT
350
AN US
Figure 7: a. Cyclic loading, experimental data taken from Linares Arregui and Alfredsson (2010) (test no: 13838). b. Uniaxial strain vs. martensite fraction.
369
370 371
AC
372
373
374 375 376 377
378 379
17
350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379
ACCEPTED MANUSCRIPT
Table 2: Material parameters used to fit the numerical predictions to experimental data.
Value
Parameter
920 MPa
a
76 × 10−6 Q
135 GPa 700 GPa 50 GPa
γ1 γ2 γ3
5.2 57 220◦ C 3.05 420 MPa 350
Stress Induced Transformation κ ξ1
0.054 MPa 0.010 (MPa·s)−1
ρ ξ2
7.8 MPa/◦ C 1.02
195 GPa 0.27
E0a K1
220 GPa -4000 GPa
M
g2 g¯ Ms β sa sg
ED
382
383
384
PT
β in Eqs (39) and (44) depends on ratio of the average volume of martensite and shear band intersection. In this study, the start value for β was based on values in the literature (Papatriantafillou, 2005; Stringfellow et al., 1992). The value of β = 3.05 that fits the transformation kinetics given in Figs 8(a) and 8(b) was finally determined again using inverse modeling with the FEM subroutine. The exponent r in Eqs (39) and (44) is commonly taken as 4 − 4.5 for martensitic TRIP steels (Iwamoto, 2002; Papatriantafillou, 2005; Stringfellow et al., 1992). Here r = 4 was found to give the best fit to the experimental data represented in Figs 8(a) and 8(b). In general α in Eqs (40) and (39), which controls the rate of shear band formation based on a stacking fault energy, is temperature dependent (Sierra and Nemes, 2008; Tomita and Iwamoto, 1995). Since the temperature degree of freedom was satisfied through the functions Γ(Θ, σ) and C(T ) in Eqs (42), (6), α and β were here kept temperature in-
AC
385
386
387 388
389 390 391
392
950 500 50
0.015 0.00623 800 4 0.0371 60 176000
CE
381
180 MPa
Strain Induced Transformation A0 A1 α r ∆v g0 g1
Nonlinear elasticity E0m ν0
380
0.04
AN US
100
Value
CR IP T
Parameter Yield Surface σY 0 Flow potential a∗ Isotropic hardening b Kinematic hardening C1 C2 C3
18
380 381 382 383 384 385 386 387 388 389 390 391 392
ACCEPTED MANUSCRIPT
0.86
ζ
0.84 0.82 0.80
0.792 0.790 0.788 0.786 0.784 0.782
FE 75 ◦ C Exp. 75 ◦ C
0.780
0.78 0.000 0.002 0.004 0.006 0.008 0.010 0.012
0.778 0.00
ε
(a)
CR IP T
0.88
0.794
FE RT FE 75 ◦ C FE 100 ◦ C FE 150 ◦ C Exp. RT Exp. 75 ◦ C Exp. 100 ◦ C Exp. 150 ◦ C
ζ
0.90
0.01
0.02
ε
0.03
0.04
0.05
(b)
396 397 398 399 400 401 402 403 404 405 406 407 408 409
CE
410
M
395
dependent. The parameters responsible for temperature dependence (g1 , ρ, κ) were obtained using the results of X-ray diffraction measurements presented in Figs 7(b) and 8(b). The g1 value was fitted to the measurements representing the change of ζ at high temperatures (75◦ C, 100◦ C and 150◦ C) where strain induced transformation was dominant. The parameters ρ and κ on the other hand changes the transformation start stresses in Eq. 6. Therefore they were determined considering the yield strength at RT to 75◦ C, where stress induced transformation starts before yielding, see Fig. 8(a). Parameters g0 , g2 , g¯ and sg in the probability function are determined through their fit to the X–ray diffraction measurements given in Fig. 8(a), especially at high temperatures where the strain induced phase transformation is dominant. The dimensionless parameters A0 , A1 and sa given in Eq. (46) were used to calculate A, which account for the strains related to shape change. These parameters were determined from the stress strain curves in Fig. 6. The starting values were based on the values by Iwamoto (2002); Papatriantafillou (2005); Stringfellow et al. (1992). According to Stringfellow et al. (1992) the difference in kinetics and morphology of stress and strain induced phase transformation affect the magnitude of transformation strains. However, in this study the same A was used to define the evolution of all transformation strains. Figure 2 shows the retained austenite as islands at the prior austenite grain boundary intersections. Some islands of retained austenite are circled in red in Fig. 2 which represent small inclusions constrained in martensite matrix that effect phase change and redistributes the stress field in the transformation zone. Therefore the volume expansion of retained austenite is not free and it is mainly constricted by elastic deformation of the martensite matrix. Using the empirical equation given by Moyer and Ansell (1975) the relative volume expansion, ∆v in Eq. (45), during martensitic transformation of the iron-carbon alloy with 0.91 wt % C was calculated as 0.0371. The determined value was in the range of 0.2 % to 0.4 %, that is reported by Leal (1984) and commonly used by previous
ED
394
PT
393
AN US
Figure 8: a. X-ray measurements of transformation of retained austenite to martensite in tension specimens are compared to model predictions. b. Simulated transformation kinetics during compression and X-ray measurement at experiment termination.
411
412
413
AC
414 415 416 417
418
419 420 421
19
393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421
ACCEPTED MANUSCRIPT
40 mm
1.5 mm 4 mm
CR IP T
P
r = 4 mm
Figure 9: FE model, thickness 0.70 – 0.76 mm. The triangles represents the symmetry conditions.
424 425
researchers. The relation between the parameters which were only related to the kinetics of the stress induced phase change, ξ1 and ξ2 , was determined using Eq. (48) (Moumni et al., 2011)
AN US
422 423
∂ζ ∂T ζ˙stress = = Cζ ξ1 (ξ2 − ζ). ∂T ∂t 426
Therefore
ζ1
T Rf
M
ξ1 = T˙
Rζ2
dζ (ξ2 −ζ)
422
423 424 425
(49) 426
(50)
Cζ dT
Ms
429 430 431 432 433 434
CE
435
ED
428
where Tf = 165◦ C representing the temperature at the end of heat treatment when ζ2 = 0.78 and Ms = 220◦ C (Ovako, 2016) at the martensite start where ζ1 = 0. Using ξ2 = 1, which is defined as the maximum martensite fraction that can be obtained as a result of phase transformation as stated by Moumni et al. (2011), and Eq. (50) ξ1 , was calculated. Table 3 lists the minimum number of experiments required to determine the model parameters, the number of experiments used here for parameter determination, the additional number of experiments used for model verification and which parameters are fitted to the corresponding curves. Some of the calculated parameters indicated in Table 3 were fitted independently (σ Y 0 , a, a∗ , b, Q, Ms , g1 , κ, ρ, ξ1 , ξ2 ) since they define a unique characteristic. Some parameters (C1 , C2 , C3 , γ1 , γ2 , γ3 , A0 , A1 , g2 , g¯, α, β, sa , sg , g0 , E0m , E0a , ν0 , K1 ) required number of iterations. The other parameters are either calculated through expressions existing in the literature or taken directly from the literature (∆v , ξ1 , ξ2 ,r). The experiments were monotonic tension to specimen rupture or a predefined strain, monotonic compression to a predefined strain and cyclic pushpull to rupture. Also, the tension experiments must be performed at different temperatures in the transition temperature range, see Fig. 1.
PT
427
436 437
438
AC
439
440 441
442 443 444
20
427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444
ACCEPTED MANUSCRIPT
Table 3: Number of experiments used
Used
Verification
Tensile
3
6
2
Compression
1
1
1
Cyclic
1
1
1
X-ray diff.
8
8
2
4. Methods of Numerical simulations
Parameters σ Y 0 , a, b, Q, C1 , C2 , C3 , γ1 , γ2 , γ3 A0 , A1 , g2 , α, sa , E0m , E0a , ν0 , K1 A0 , A1 , g2 b, Q, C1 , C2 , C3 , γ1 , γ2 , γ3 α, sa , E0m , E0a , ν0 , K1 A0 , A1 , g2 , g¯, r, β, sg , g0 , g1 , κ, ρ
CR IP T
Min.
AN US
445
Test
461
5. Results
449 450 451 452 453 454 455 456 457 458 459
462
464
465
466
As seen from true stress true strain curves in Fig. 10, the finite element results were in good agreement with the tension experiments performed at different temperatures. Figure 6(a) indicates that the yield stress increases with the increase of temperature from RT to 75◦ C, but over 75◦ C only a small difference is observed. According to Neu and Sehitoglu (1992) during stress induced transformation, the yield strength in tension increases considerably with temperature. Therefore for TRIP steels, the change of yield point can be used to identify whether stress induced transformation takes place before strain induced transformation or not. To see the temperature dependence of stress induced transformation and the change in transformation start stress, stress strain curves representing the data obtained from the tensile experiments to rupture and those calculated using FE are plotted separately in Figs 6(a) and 11. In the FE computations
AC
467 468 469 470
471
472 473 474
446 447 448 449 450 451 452 453 454 455 456 457 458 459 460
461
CE
463
ED
448
PT
447
M
460
FE calculations were done using ABAQUS through a user defined material subroutine (UMAT) that followed Eqs (26) to (48) with the material properties given in Table 2. Figure 9 shows the FE model used in the analysis. A 3D mesh with 67872 C3D8 elements was used after a mesh size and convergence study. The total CPU time was 13.1 × 103 s and the ratio of CPU time with and without considering phase transformation was 1.2. In the UMAT, stress and strain induced phase transformations were coupled with nonlinear elasticity, nonlinear kinematic and isotropic hardening behaviors and formulated using fully implicit integration of the constitutive relations. In the numerical solution, initially stress induced transformation was considered, the evolution of martensite fraction and transformation strains were calculated from Eqs (48) and (45). During dislocation movement, strain induced transformation takes place at shear band intersections along with stress induced phase change. Together they yield the evolution of martensite fraction. The flowchart of the UMAT subroutine is summarized in Algorithm 1.
446
445
21
462 463 464 465 466 467 468 469 470 471 472 473 474
1600
1400
1400
1200
1200 1000
σ (MPa)
σ (MPa)
1600
800 600 400
FE RT Exp. RT
200
200
500
ε
(b)
1500
σ (MPa)
ED
1000
FE 75 ◦ C Exp. 75 ◦ C
0 0.000 0.002 0.004 0.006 0.008 0.010 0.012 0.014
M
σ (MPa)
1500
600
400
0 0.000 0.002 0.004 0.006 0.008 0.010 0.012 0.014 ε
(a)
800
AN US
1000
CR IP T
ACCEPTED MANUSCRIPT
1000
500
FE 100 ◦ C Exp. 100 ◦ C
0 0.000 0.002 0.004 0.006 0.008 0.010 0.012 0.014 ε
PT
0 0.000 0.002 0.004 0.006 0.008 0.010 0.012 0.014 ε
FE 150 ◦ C Exp. 150 ◦ C
(d)
CE
(c)
AC
Figure 10: Tensile experiments and finite element results at a. RT b. 75◦ C conducted by Linares Arregui and Alfredsson (2010) (test no: 13823) c. 100◦ C d. 150◦ C
22
ACCEPTED MANUSCRIPT
CR IP T
Algorithm 1 UMAT subroutine for plastic deformation, stress and strain induced phase change. Initial Conditions pl 0 0 ζ0 = 0.78, [εtr ij ]0 = 0, [εij ]0 = 0, k:iteration number, n:increment number, [∆A] = [A]n+1 − [A]n . 1. Trial Stress pl 0 pl 0 0 0 tr 0 Set : k = 0, ζn+1 = ζn0 , [εtr ij ]n+1 = [εij ]n , [εij ]n+1 = [εij ]n pl
k [∆εpl ij ] =
d k m k 3 [σij ]n+1 −[Xij ]n+1 k Y 2 [σ ]n+1
AN US
Calculate trial stress: [σij ]0n+1 = [Eijkl ]0n+1 ([εkl ]0n+1 −[εkl ]0n+1 −[εkl tr ]0n+1 )+ 2 0 K1 [(εel kk ) ]n+1 δij ) 2. Check stress induced transformation and yield function k if (( Φ > 0 & ζ˙ > 0 ⇒ Cζ > 0) & fn+1 < 0) then only stress induced transformation occurs [∆ζ]k+1 = Cζ ξ1 (ξ2 − [ζ]k+1 n+1 ) k+1 [εtr ] calculated using Eq. (45) ij n+1 update stresses using Eq. (15) and go to step 6. k else if (Cζ < 0 and fn+1 > 0) then yielding occurs before stress induced transformation [f ]kn+1 = 0 gives ∆λ plastic strains are calculated using: + 3a∗ [σ h ]kn+1 [∆λ]k
AC
CE
PT
ED
M
k k k [∆ζstrain ]k = (1 − ζn+1 )(Aζ [∆εpl eq ] + Bζ [∆Σ] ) k else if Cζ > 0 and ζ < 1 and fn+1 > 0 then yielding, strain and stress induced phase changes occur k k k k [∆ζ]k = (1 − ζn+1 )(Aζ [∆εpl eq ] + Bζ [∆Σ] ) + Cζ ξ1 (ξ2 − ζn+1 ) k [εtr ij ]n+1 calculated using Eq. (45) else Elastic deformation and go to step 6. end if pl k+1 k+1 tr k+1 3. Update [ζ]k+1 n+1 ,[εij ]n+1 , [εij ]n+1 and [Eijkl ]n+1 . 4. Calculate stress using : pl k+1 k+1 k tr k+1 el 2 k+1 [σij ]k+1 n+1 = [Eijkl ]]n+1 ([εkl ]n+1 − [εkl ]n+1 − [εkl ]n+1 ) + K1 [(εkk ) ]n+1 δij ) 5. Set k=k+1 and go to step (2). 6. Set n = n + 1 and go to step (1). 7. When the calculation finishes at all the integration points ABAQUS checks the force equilibrium.
475
476 477 478 479 480 481
rupture of the specimens was not modeled. Instead all simulations continued to the total strain when the last experiment ruptured. During the experiments it was observed that the stress was homogeneously distributed and the specimens deformed without any necking. Therefore the localization of inhomogeneous stresses was not considered in the numerical simulations. Figure 6(a) shows that the start of stress induced phase change increases with the temperature up to a certain value above which no phase change occurs prior to yielding. Thus 23
475 476 477 478 479 480 481
σ (MPa)
1800 1600 1400 1200 1000 800 600 400 200 0 0.0
CR IP T
ACCEPTED MANUSCRIPT
FE RT FE 75 ◦ C FE 100 ◦ C FE 150 ◦ C 0.4
0.6
ε
0.8
1.0
1.2
1.4 1e−2
AN US
0.2
Figure 11: FE calculations
485 486 487 488 489 490 491 492 493 494 495 496
CE
497
M
484
the results obtained were compatible with the findings of Neu and Sehitoglu (1992). The results by Neu and Sehitoglu (1992) indicated that it is difficult to separate the stress and strain induce transformations by only analyzing monotonic loading experiments at one specific temperature, especially if the material yields before transformation. To illustrate the transformation kinetics distinctly, the conditions considering only stress induced or strain induced transformations were modeled and are plotted in Fig. 12 at two different temperatures. At RT in Fig. 12(a), the dashed green curve, representing the case where only stress induced transformation is active, coincides with black curve, including both transformation mechanics, till the start of strain induced transformation. Therefore, it is seen that for low temperature the stress induced transformation initiates before yielding and is a major part of the total transformation. At 75◦ C strain induced phase change becomes more dominant and the stress induced phase change takes place after yielding as is represented in Fig. 12(b). For visual representation of the dependence of yield strength on transformation behavior, the uniaxial stress versus martensite fraction is plotted in Fig. 13(a). By comparing the onset of nonlinearity in Fig. 6(a) with transformation start stress in Fig. 13(a) it can be clearly seen that the yield strength decreases with decreasing temperature as the result of early transformation start at low temperatures of stress induced transformation. On the other hand at higher temperatures the effect of temperature on yield strength is less because no stress induced transformation occurs before yielding. Moreover, Fig. 13(b) supports the claim that at RT phase change starts before yielding since εtr 11 starts evolving before εpl . It was concluded that by performing a series of different eq monotonic tension experiments at different temperatures it was possible to separate between stress and strain induced transformation. Figure 1 illustrates
ED
483
PT
482
498 499 500
AC
501 502 503 504
505 506
507 508
24
482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508
ACCEPTED MANUSCRIPT
σ (MPa)
1500
σ (MPa)
1500
1000
500 0 0.0
0.2
0.4
500
Stress & Strain induced Strain induced Stress induced 0.6 0.8 1.0 1.2 1.4 1e−2 ε
0 0.0
0.2
0.4
(a)
CR IP T
1000
Stress & Strain induced Strain induced Stress induced 0.6 0.8 1.0 1.2 1.4 1e−2 ε
(b)
AN US
Figure 12: Stress strain curve obtained from FEA, illustrating the effects and difference between stress and strain induced phase change at two different temperatures a. RT b. 75◦ C
2000
0.16
0.14
σ (MPa)
1000
FE RT FE 75 ◦ C FE 100 ◦ C FE 150 ◦ C
0.82
0.84
ζ
0.86
0.88
0.90
ED
0.80
0.10
FE RT FE 75 ◦ C FE 100 ◦ C FE 150 ◦ C
0.08
0.06
0.04
M
500 0 0.78
pl /ε11 εeq
0.12
1500
(a)
0.02 0.00 0.0
0.1
0.2
0.3
tr /ε11 ε11
0.4
0.5
(b)
509 510
for low temperatures that the decrease of inelastic deformation with decreasing temperature shows that stress induced transformation is present and predominating. It was then detected separately from strain induced transformation, see the schematic start stress below yielding for stress induced transformation in Fig. 1. Strain induced transformation starts after yielding which in the present temperature range is relatively constant. It was then found that after yielding, strain induced transformation was more pronounced than stress induced. The XRD measurements are shown in Fig. 8(a) together with the ζ curves obtained from FE using the calibrated material properties. When all data points were considered together with the model prediction in the figure, the measurements supports each other and deviation from the model is less than 0.5 % for each retained austenite value. Hence, the overall accuracy of multiple measurements were much better than the single point estimate suggests. Figures 14(a) and 14(b) represent the contour plot of martensite fraction at
CE
511
PT
Figure 13: a. Effect of martensite fraction on strength b. evolution of nonlinear strains at monotonic tension.
512 513 514
AC
515
516
517 518 519 520 521
522
25
509 510 511 512 513 514 515 516 517 518 519 520 521 522
ACCEPTED MANUSCRIPT
(a)
AN US
+8.609e−01 +8.609e−01 +8.493e−01 +8.378e−01 +8.262e−01 +8.147e−01 +8.031e−01 +7.916e−01 +7.800e−01
CR IP T
+8.609e−01 +8.493e−01 +8.378e−01 +8.262e−01 +8.147e−01 +8.031e−01 +7.916e−01 +7.800e−01
M
(b)
Figure 14: a. ζ contour plot, when a. ε11 = 0.0095 b. ε11 = 0.0112 (strains at notch root).
525 526 527 528 529
CE
530
the early transformation stage ( ε11 = 0.00275 in the test part with homogeneous strain and maximum ε11 = 0.0095 in the notch root) and at the specimen rupture under tensile loading at 75◦ C, see Fig. 6(a). During phase change that occurs more predominantly under tension than compression, the volume increase induces an increase in strain, resulting in a nonlinear stress-strain behavior in advance and/or during dislocation motion. For this reason, in high strength steels with retained austenite undergoing phase change, the asymmetry between compressive and tensile yield stresses seen in Fig. 6(b) is largely a result of phase transformation. The triaxiality parameter Σ in Eq. (38) affects the probability of martensite formation at the dislocation sites during strain induced phase change through the parameter Γ in Eq. (42) in the probability function P , Eqs (37), (41) and (44). The strain induced model of Stringfellow et al. (1992) that fitted to the martensite fraction evolution data of Young (1988) showed that martensite formation decreases when the triaxility is negative as in the case of compression. Additionally, Young (1988) reported that transformation curves for stress induced transformation exhibit linear behavior with respect to axial strains, whereas the curves for strain induced transformation was found to be nonlinear. The comparison of tension–compression tests and the calculated FE results together with X-ray
ED
524
PT
523
531 532 533
AC
534
535 536 537 538 539 540 541
26
523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541
ACCEPTED MANUSCRIPT
0
CR IP T
σ (MPa)
1000
−1000
Exp. 75 ◦ C Cyclic FE 75 ◦ C Cyclic
−2000
AN US
−0.015 −0.010 −0.005 0.000 0.005 0.010 0.015 ε
Figure 15: Cyclic loading, experimental data taken from Linares Arregui and Alfredsson (2010). (test no: 13838)
545 546 547 548 549 550 551 552 553
CE
554
M
544
diffraction measurements given in Fig. 8(b) pointed out that under compression a small amount of transformation occurs. The simulated behavior of ζ versus axial strain plotted in Fig. 8(b) differs quite markedly to the tensile loading case in Fig. 8(a). In compression, the transformation is dominated by the stress induced transformation that develops at the early stages of the deformation. This part of the transformation is furthermore linear with the strain, which agree with the findings of Young (1988). The last part of the curve includes small strain induced transformation but at this stage the rate of transformation ζ˙ is much lower than the early part. The strength differential effect was also examined for cyclic load and compared to the experiments performed by Linares Arregui and Alfredsson (2010) (test no:13889) in Fig. 7. The XRD results represented in Fig. 7(b) are in agreement with the simulations where the dotted the black line represents the martensite fraction under monotonic tensile loading and dashed blue curve is the evolution of martensite in the first cycle. The results confirm that the phase transformation predominantly develops during tension. Note that the retained austenite measurement for the cyclic curve in Fig. 7(b) was not directly used in determining the phase transformation parameters. Hence, it gave an indication on the model prediction. For high strength steels, the slope of the yield surface in the meridian plane is a ≈ 0.02 (Allen and Wilson, 2004; Guo et al., 2008; Spitzig and Richmond, 1979), which is smaller than the value determined by Linares Arregui and Alfredsson (2010), a ≈ 0.25. However, this value included all nonelastic effects. The effect of the pressure sensitive plasticity parameter, a on yield strength and
ED
543
PT
542
555 556 557
AC
558 559 560
561
562
563 564 565
27
542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565
1600
0.90
1400
0.88
1200
600 400 200 0.2
0.4
0.6
ε
0.8
1.0
1.2
0.84 0.82 0.80
0.78 0.000 0.002 0.004 0.006 0.008 0.010 0.012
1.4 1e−2
ε
AN US
0 0.0
0.86
a=0.04, 75 ◦ C a=0.08, 75 ◦ C a=0.10, 75 ◦ C a=0.12, 75 ◦ C a=0.04, RT a=0.08, RT a=0.10, RT a=0.12, RT
800
ζ
σ (MPa)
1000
Exp. RT Exp. 75 ◦ C a=0.04, 75 ◦ C a=0.08, 75 ◦ C a=0.10, 75 ◦ C a=0.12, 75 ◦ C a=0.04, RT a=0.08, RT a=0.10, RT a=0.12, RT
CR IP T
ACCEPTED MANUSCRIPT
(a)
(b)
AC
2
3
|
|
PT 1
4
inelastic strain
5
6
1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0
SDE= σC -σT (GPa)
Full Drucker & Strain Drucker & Stress Drucker Experiment
CE
|
|
SDE= σC -σT (GPa)
1.6 1.4 1.2 1.0 0.8 0.6 0.4 0.2 0.0 0
ED
M
Figure 16: The effect of parameter a a. on constitutive behavior b. on martensite transformation.
7 1e−3
(a)
Full Drucker & Strain Drucker & Stress Drucker
1
2
3
4
inelastic strain (b)
Figure 17: Strength differential effect a. at 75◦ C b. at RT.
28
5
6
7 1e−3
ACCEPTED MANUSCRIPT
571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598
CE
599 600 601 602 603
AC
604
605
606 607 608 609
566 567 568 569 570 571
CR IP T
570
AN US
569
M
568
phase transformation was investigated by using four different a values in tensile simulations. The parameter sensitivity study represented in Fig. 16(a) shows that the decrease in yield stress for increasing a is more significant at 75◦ C than RT. This result is mainly due to the stress induced transformation that occurs at lower temperatures before yielding. Therefore, the dependency of hydrostatic stress is influenced more by stress induced phase change, where the transformation is initiated by elastic stresses, compared to strain induced phase transformation. To determine the value of parameter a, not only the stress strain response but also the effect of a on phase transformation kinetics was studied. From Fig. 16(b) it is seen that decreasing a from a = 0.12 to a = 0.04 decreases the predicted martensite fraction at a certain strain value. Experimental measurements on martensite fraction ζ in Fig. 16(b), yield stress in Fig. 16(a) and the tensile and compressive stress strain responses in Fig. 6(b) were used to establish a = 0.04 as the best yield stress slope for the present material. During strain induced phase change, which dominates for elevated temperatures, the evolution of martensite fraction occurs at dislocation sites and the effect on the yield surface slope is illustrated in Fig. 16(b). Therefore the experimented evolution of ζ at elevated temperatures were also considered. The effect of phase change on SDE was analyzed for two different temperatures through simulations of monotonic loading where the different parts of phase changes were executed. The evolution of such simulated SDE is plotted against inelastic strains in Fig. 17. Figure 17(a) shows that for increased temperatures, here exemplified with 75◦ C, strain induced transformation together with Drucker Prager plasticity coincides with the curve representing the full model and they yield closer results to the experimental measurements of Linares Arregui and Alfredsson (2010) performed at 75◦ C. The difference in SDE between experimental measurements and full model predictions in Fig. 17(a) is explained in Fig. 6(b) by the corresponding difference in the compression curves at moderate strain. Figure 6(b) also illustrates that in absolute stress values the model deviation is small. Figures 17(a) and 17(b) includes simulated curves that only contains pressure sensitive plasticity, using Drucker Prager, i.e. the phase transformation was turned off. Comparing these curves with the full model prediction clearly illustrates that phase transformation dominates SDE effect for the current material structure. The result also show that stress induced phase change has only a small effect on SDE for high temperatures. Besides, when the temperature decreases to a level where the stress induced phase transformation is more effective on yield strength, e.g. RT in this case, the stress induced transformation is mainly responsible for SDE especially at low inelastic strains, see Fig. 17(b). Figure 18 compares the calculated volume changes during tensile and compressive loading to experimental measurements of Linares Arregui and Alfredsson (2010). It was observed that the computed volume change during compression is less than that during tensile loading which agree with the results by Neu and Sehitoglu (1992).
ED
567
PT
566
29
572
573 574 575 576 577 578 579 580 581 582
583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609
ACCEPTED MANUSCRIPT
0.6
0.8
1.0
1.2
1.4 1e−1
AN US
0.4
CR IP T
FE-tension FE-compression Exp-Compression Exp-Tension
∆V/V
9 1e−3 8 7 6 5 4 3 2 1 0 0.0 0.2
accum. plastic strain
Figure 18: Volumetric change vs. accumulated plastic strain
613 614 615 616 617 618 619 620 621
CE
622
The present study couples the plastic deformation obeying nonlinear hardening rules and nonassociated flow rule with stress and strain induced phase change. Good agreement was obtained with the experimental results. In addition to phase transformation kinetics, the nonlinearity of elastic strains were taken into account in the material model while scrutiny of the plastic response cycles from cyclic experiments clearly showed that nonlinear elastic effects do have influences for the high elastic strains that can develop in this type of high strength bearing steel. Although the full material model includes a relatively large amount of material parameters, a limited number of well planned and well instrumented experiments were sufficient to calibrate the model. The required series include monotonic tensile experiments to some different strains and at some different temperatures together with at least one compression and one cyclic experiment. The data captured should include retained austenite levels and densities at test termination as well as the stress strain curves. The experiments showed that the solid to solid phase transformation of retained austenite to martensite affects the material behavior before and after plastic deformation through stress and strain induced phase transformations and changes in the yield strength of the material. Results indicated that the change of the yield strength with temperature is the main indicator of stress induced transformation that takes place before plasticity and defines start of the measured yielding at RT. At low temperatures, approximately below 75◦ C for the present material, stress induced transformation results in a reversed tem-
M
612
ED
611
6. Conclusions
PT
610
623 624 625
AC
626
627 628 629 630 631
632 633
30
610
611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633
ACCEPTED MANUSCRIPT
675
Acknowledgment
639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666
CE
667
AN US
638
M
637
ED
636
PT
635
668 669 670 671
AC
672 673
676 677
634 635 636 637 638 639
CR IP T
674
perature relation with decreasing yield point for decreasing temperature. Also at these temperatures phase transformation was only stress induced, i.e. no strain induced transformation existed. At 75◦ C and for higher temperatures up to the martensite start temperature plastic yielding develop before phase change. Here strain induced transformation dominates the transformation, although some stress induced transformation continues to develop. By utilizing the understanding on how the yield point changes with temperature and the corresponding measured data points, stress and strain induced phase transformation were successfully separated from each other, their magnitude measured and the model parameter values were determined. Thus, the proposed model summarized in Eq. (48) separates stress and strain induced transformation. It was concluded that the stress and strain induced phase transformation was active in the present material and that the processes can be separated by coupled modeling with a series of different monotonic experiments at different temperatures where both the stress strain curves were captured and the retained austenite levels after experiment termination was determined. Using the proposed model not only the coupled behavior but also distinctly stress and strain induced phases can be identified. Therefore, using this modeling technique, for complex loading condition like the stress distribution around a crack tip or at contact surfaces, contour plot of strain and stress induced martensite transformation zones can be plotted separately and their contribution to the mechanical behavior can be calculated independently. According to the monotonic experiments and X-ray diffraction measurements it was concluded that the amount of retained austenite that transforms to martensite is higher under tension than compression. The calculations for compressive loads at 75◦ C showed that a small stress induced phase transformation at the early stage of compression accompanied by strain induced phase change which is very small in total compared to the change under tensile loading. The analysis showed that phase transformation enhances and dominates the SDE over plasticity effects and that SDE can be simulated successfully by employing the coupled stress and strain induced phase transformation kinetics together with a pressure sensitive surface for plastic yielding and flow as is proposed in the current study. In the light of the aforementioned reasons it is claimed that both stress and strain induced phase change should be taken into account in the definition of constitutive behavior of the present class of high strength steels. Therefore, the material model that was capable of capturing nonlinear hardening behaviors, pressure sensitive plastic yielding and nonlinear elasticity together with stress and strain induced phase change was required to simulate the mechanical material behavior including the SDE in the high strength martensitic bearing steel accurately.
634
640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655
656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674
675
This work was conducted in cooperation with SKF ERC. The authors would ¨ like to thank Dr P. Hedstr¨ om and Mr M. Oberg for providing measuring and 31
676 677
ACCEPTED MANUSCRIPT
679
experimental expertise. The financial support from SKF ERC is much appreciated.
680
References
685
686 687
688 689 690
691 692 693
694 695 696
697 698
699 700
701 702
Achenbach, M., 1989. A Model for an Alloy With Shape Memory. Int. J. Plast. 5, 371–395.
Allen, P.A., Wilson, C.D., 2004. Hydrostatic Stress Effect on the Yield Behavior of Inconel 100. J. Mech. Behav. Mater. 15, 27–48.
Arghavani, J., Auricchio, F., Naghdabadi, R., Reali, A., Sohrabpour, S., 2010. A 3-D phenomenological constitutive model for shape memory alloys under multiaxial loadings. Int. J. Plast. 26, 976–991. Auricchio, F., Reali, A., Stefanelli, U., 2007. A three-dimensional model describing stress-induced Solid phase transformation with permanent inelasticity. Int. J. Plast. 23, 207–226. Bhattacharyya, A., Weng, G.J., 1994. An energy criterion for the stress-induced martensitic transformation in a ductile system. J. Mech. Phys. Solids 42, 1699–1724. Bolling, G.F., Richman, R.H., 1970a. The influence of stress on martensite-start temperatures in Fe-Ni-C alloys. Scr. Metall. 4, 539–543. Bolling, G.F., Richman, R.H., 1970b. The plastic deformation-transformation of paramagnetic f.c.c. FeNiC alloys. Acta Metall. 18, 673–681. Boyd, J.G., Lagoudas, D.C., 1996. A Thermodynamical Constitutive Model for Shape Memory Materials. Part I. the Monolithic Shape Memory Alloy. Int. J. Plast. 12, 805–842. Berveiller, M., Fischer, F. D., 1997. Mechanics of Solids with Phase Changes. Int. Cent. Mech. Sci. 368.
CE
703
704
705
706
Brocca, M., Brinson, L.C., Baˇzant, Z.P., 2002. Three-Dimensional Constitutive Model for Shape Memory Alloys Based on Microplane Model. J. Mech. Phys. Solids 50, 1051–1077.
AC
707 708
709
710
711 712
679
CR IP T
684
AN US
683
Abeyaratne, R., Knowles, J.K., 1993. A continuum model of a thermoelastic solid capable of undergoing phase transitions. J. Mech. Phys. Solids 41, 541– 571.
M
682
ED
681
678
680
PT
678
Br¨ unig, M., 1999. Numerical simulation of the large elastic-plastic deformation behavior of hydrostatic stress-sensitive solids. Int. J. Plast. 15, 1237–1264. Casey, J., Sullivan, T.D., 1985. Pressure dependency, strength-differential effect, and plastic volume expansion in metals. Int. J. Plast. 1, 39–61.
32
681 682 683
684 685
686 687
688 689 690
691 692 693
694 695 696
697 698
699 700
701 702 703
704 705
706 707 708
709 710
711 712
ACCEPTED MANUSCRIPT
718
719 720
721 722 723 724
725 726 727
728 729 730
731 732 733
734 735
736 737 738
739
Chait, R., 1972. Factors influencing the strength differential of high strength steels. Metall. Trans. 3, 369–375.
Chan, C.W., Chan, S.H.J., Man, H.C., Ji, P., 2012. 1-D Constitutive Model for Evolution of Stress-Induced R-Phase and Localized L¨ uders-Like StressInduced Martensitic Transformation of Super-Elastic NiTi Wires. Int. J. Plast. 32-33, 85–105. Cherkaoui, M., Berveiller, M., Lemoine, X., 2000. Couplings between plasticity and martensitic phase transformation: overall behavior of polycrystalline TRIP steels. Int. J. Plast. 16, 1215–1241.
Cherkaoui, M., Berveiller, M., Sabar, H., 1998. Micromechanical modeling of martensitic transformation induced plasticity (TRIP) in austenitic single crystals. Int. J. Plast. 14, 597–626. Christ, H.J., Sommer, C., Mughrabi, H., Voskamp, A.P., Beswick, J.M., Hengerer, F., 1992. Fatigue Behaviour of Three Variants of the Roller Bearing Steel Sae 52100. Fatigue Fract. Eng. Mater. Struct. 15, 855–870. Cleveland, R.M., Ghosh, A.K., 2002. Inelastic effects on springback in metals. Int. J. Plast. 18, 769–785. Denis, S., Simon, A., Beck, G., 1982. Estimation of the Effect of Stress/ Phase Internal Transformation Quenching Interaction of Steel Internal Stress during Martensitic. Trans. Iron Steel Inst. Japan 22, 504–513. Diani, J.M., Sabar, H., Berveiller, M., 1995. Micromechanical modelling of the transformation induced plasticity (TRIP) phenomenon in steels. Int. J. Eng. Sci. 33, 1921–1934.
CE
740
Chaboche, J.L., 2008. A review of some plasticity and viscoplasticity constitutive theories. Int. J. Plast. 24, 1642–1693.
741
742
Drucker, D.C., Prager, W., 1952. Soil Mechanics and Plastic Analysis or Limit Design.
AC
743
744
745 746
747 748 749
713 714
715 716
CR IP T
717
Chaboche, J.L., 1991. On some modifications of kinematic hardening to improve the description of ratchetting effects. Int. J. Plast. 7, 661–678.
AN US
716
M
715
Chaboche, J.L., 1986. Time-independent constitutive theories for cyclic plasticity. Int. J. Plast. 2, 149–188.
ED
714
PT
713
Entchev, P.B., Lagoudas, D.C., 2004. Modeling of transformation-induced plasticity and its effect on the behavior of porous shape memory alloys. Part II: porous SMA response. Mech. Mater. 36, 893–913. Falk, F., Konopka, P., 1990. Three-Dimensional Landau Theory Describing the Martensitic Phase Transformation of Shape-Memory Alloys. J. Phys. Condens. Matter 2, 61–77. 33
717 718
719 720
721 722 723 724
725 726 727
728 729 730
731 732 733
734 735
736 737 738
739 740 741
742 743
744 745 746
747 748 749
ACCEPTED MANUSCRIPT
755
756 757 758
759 760
761 762
763 764
765 766 767
768 769
770 771
772 773 774
775 776
Fischer, F.D., Reisner, G., Werner, E., Tanaka, K., Cailletaud, G., Antretter, T., 2000. New view on transformation induced plasticity (TRIP). Int. J. Plast. 16, 723–748. Frederick, C. O., Armstrong, P. J., 2007. A mathematical representation of the multiaxial Bauschinger effect. Mater. High Temp. 24, 1–26. Gall, K., Sehitoglu, H., 1999. The Role of Texture in Tension-compression Asymmetry in Polycrystalline NiTi. Int. J. Plast. 15, 69–92.
Ganghoffer, J.F., Simonsson, K., 1998. A micromechanical model of the martensitic transformation. Mech. Mater. 27, 125–144. Gao, X., Huang, M., Brinson, L.C., 2000. A Multivariant Micromechanical Model for SMAs Part 1. Crystallographic Issues for Single Crystal Model. Int. J. Plast. 16, 1345–1369. Garofalo, F., Wriedt, H., 1962. Density change in an austenitic stainless steel deformed in tension or compression. Acta Metall. 10, 1007–1012. Giusti, J., Radenkovic, D., 1983. A model of thermo-mechanical behaviour of metals with phase changes. Mech. Res. Commun. 10, 163–169. Goel, N.C., Sangal, S., Tangri, K., 1985. A theoretical model for the flow behavior of commercial dual-phase steels containing metastable retained austenite: Part I. derivation of flow curve equations. Metall. Trans. A 16, 2013–2021. Greenwood, G., Johnson, R., 1965. The deformation of metals under small stresses during phase transformations, in: Proc. R. Soc. London. Ser. A, pp. 403–422.
CE
777
Fischer, F.D., Oberaigner, E.R., Tanaka, K., Nishimura, F., 1998. Transformation induced plasticity revised an updated formulation. Int. J. Solids Struct. 35, 2209–2227.
778
779
Guo, T.F., Faleskog, J., Shih, C.F., 2008. Continuum modeling of a porous solid with pressure-sensitive dilatant matrix. J. Mech. Phys. Solids 56, 2188–2212. Halphen, B., Nguyen, Q.S., 1974. Plastic and Visco-plastic Materials with Generalized Potential. Mech. Res. Commun. 1, 43–47.
AC
780
781
782
783 784
750 751 752
753 754
CR IP T
754
AN US
753
M
752
Fei, D., Hodgson, P., 2006. Experimental and numerical studies of springback in air v-bending process for cold rolled TRIP steels. Nucl. Eng. Des. 236, 1847–1851.
ED
751
PT
750
Idesman, A., Levitas, V., Stein, E., 2000. Structural changes in elastoplastic material: a unified finite-element approach to phase transformation, twinning and fracture. Int. J. Plast. 16, 893–949.
34
755
756 757 758
759 760
761 762
763 764
765 766 767
768 769
770 771
772 773 774
775 776 777
778 779
780 781
782 783 784
ACCEPTED MANUSCRIPT
790
791 792 793
794 795 796
797 798 799
800 801 802
803 804 805
806 807
808 809
810 811
Iwamoto, T., 2004. Multiscale computational simulation of deformation behavior of TRIP steel with growth of martensitic particles in unit cell by asymptotic homogenization method. Int. J. Plast. 20, 841–869.
Iwamoto, T., Tsuta, T., 2000. Computational simulation of the dependence of the austenitic grain size on the deformation behavior of TRIP steels. Int. J. Plast. 16, 791–804. Iwamoto, T., Tsuta, T., Tomita, Y., 1998. Investigation on deformation mode dependence of strain-induced martensitic transformation in trip steels and modelling of transformation kinetics. Int. J. Mech. Sci. 40, 173–182. Jacques, P., Furn´emont, Q., Pardoen, T., Delannay, F., 2001. On the role of martensitic transformation on damage and cracking resistance in TRIPassisted multiphase steels. Acta Mater. 49, 139–152. Lai, J., Huang, H., Buising, W., 2016. Effects of microstructure and surface roughness on the fatigue strength of high-strength steels. Procedia Struct. Integr. 2, 1213–1220. Leal, R., 1984. Transformation toughening of metastable austenitic steels. Ph.D. thesis. MIT. Lebedev, A., Kosarchuk, V., 2000. Influence of phase transformations on the mechanical properties of austenitic stainless steels. Int. J. Plast. 16, 749–767. Leblond, J., Mottet, G., Devaux, J., 1986. A theoretical and numerical approach to the plastic behaviour of steels during phase transformations-I. Derivation of general relations. J. Mech. Phys. Solids 34.
CE
812
Iwamoto, T., 2002. Computational simulation on deformation behavior of CT specimens of TRIP steel under mode I loading for evaluation of fracture toughness. Int. J. Plast. 18, 1583–1606.
813
814
Leblond, J.B., 1989. Mathematical modelling of transformation plasticity in steels II: Coupling with strain hardening phenomena. Int. J. Plast. 5, 573– 591.
AC
815
816
817 818
819
820 821 822
785 786 787
788 789
CR IP T
789
AN US
788
M
787
Inoue, T., Yamaguchi, T., Wang, Z., 1984. Stresses and Phase Transformations Occuring in Quenching of Carburized Steel Gear Wheel. Mater. Sci. Technol. 1, 872–876.
ED
786
PT
785
Leblond, J.B., Devaux, J., Devaux, J.C., 1989. Mathematical modelling of transformation plasticity in steels I: Case of ideal-plastic phases. Int. J. Plast. 5, 551–572. Lee, J., Lee, J.Y., Barlat, F., Wagoner, R., Chung, K., Lee, M.G., 2013. Extension of quasi-plastic-elastic approach to incorporate complex plastic flow behavior - application to springback of advanced high-strength steels. Int. J. Plast. 45, 140–159. 35
790
791 792 793
794 795 796
797 798 799
800 801 802
803 804 805
806 807
808 809
810 811 812
813 814 815
816 817 818
819 820 821 822
ACCEPTED MANUSCRIPT
828 829
830 831 832
833 834 835
836 837
838 839 840
841 842
843 844
845 846
847 848
Levitas, V.I., Ozsoy, I.B., 2009b. Micromechanical modeling of stress-induced phase transformations. Part 2. Computational algorithms and examples. Int. J. Plast. 25, 546–583. Liang, C., Rogers, C.A., 1990. One-dimensional Thermomechanical Constitutive Relations for Shape Memory Materials. J. Intell. Mater. Syst. Struct. 1, 207– 234.
Liang, C., Rogers, C.A., 1992. A Multi-dimensional Constitutive Model for Shape Memory Alloys. J. Eng. Math. 26, 429–443. Linares Arregui, I., Alfredsson, B., 2010. Elastic-plastic characterization of a high strength bainitic roller bearing steel-experiments and modelling. Int. J. Mech. Sci. 52, 1254–1268. Linares Arregui, I., Alfredsson, B., 2013. Non-linear elastic characterisation of a high strength bainitic roller bearing steel. Int. J. Mech. Sci. 68, 1–15. Liu, L., Shan, T.K., 2011. Effect of Transformation on Springback for TRIP Steel Stamping. Adv. Mater. Res. 221, 405–410. Loret, B., Prevost, J.H., 1986. Accurate numerical solutions for drucker-prager elastic-plastic models. Comput. Methods Appl. Mech. Eng. 54, 259–277. Ma, A., Hartmaier, A., 2015. A study of deformation and phase transformation coupling for TRIP-assisted steels. Int. J. Plast. 64, 40–55. Magee, C., 1966. Transformation kinetics, micro-plasticity and aging of martensite in Fe-31Ni. Ph.D. thesis. Carnegie Inst. of Technology.
CE
849
Levitas, V.I., Ozsoy, I.B., 2009a. Micromechanical modeling of stress-induced phase transformations. Part 1. Thermodynamics and kinetics of coupled interface propagation and reorientation. Int. J. Plast. 25, 239–280.
850
851
McClintock, F.A., 1968. A Criterion for Ductile Fracture by the Growth of Holes. J. Appl. Mech. 35, 363.
AC
852
853
854 855
856
857
823 824
825 826
CR IP T
827
Levitas, V.I., 1998. Thermomechanical theory of martensitic phase transformations in inelastic materials. Int. J. Solids Struct. 35, 889–940.
AN US
826
M
825
Lee, J.H., 1988. Research note on a simple model for pressure-sensitive strainhardening materials. Int. J. Plast. 4, 265–278.
ED
824
PT
823
Mendiguren, J., Cort´es, F., G´ omez, X., Galdos, L., 2015. Elastic behaviour characterisation of TRIP 700 steel by means of loading-unloading tests. Mater. Sci. Eng. A 634, 147–152. Morin, C., Moumni, Z., Zaki, W., 2011a. A constitutive model for shape memory alloys accounting for thermomechanical coupling. Int. J. Plast. 27, 748–767.
36
827 828 829
830 831 832
833 834 835
836 837
838 839 840
841 842
843 844
845 846
847 848
849 850
851 852
853 854 855
856 857
ACCEPTED MANUSCRIPT
863
864 865 866
867 868 869
870 871 872
873 874
875 876
877 878
879 880
881 882
883
Moumni, Z., Roger, F., Trinh, N.T., 2011. Theoretical and numerical modeling of the thermomechanical and metallurgical behavior of steel. Int. J. Plast. 27, 414–439. Moumni, Z., Zaki, W., Nguyen, Q.S., 2008. Theoretical and Numerical Modeling of Solid-solid Phase Change: Application to the Description of the Thermomechanical Behavior of Shape Memory Alloys. Int. J. Plast. 24, 614–645.
Moyer, J.M., Ansell, G.S., 1975. The volume expansion accompanying the martensite transformation in iron-carbon alloys. Metall. Trans. A 6, 1785– 1791. Neu, R.W., Sehitoglu, H., 1992. Stress-induced transformation in a carburized steelExperiments and analysis. Acta Metall. Mater. 40, 2257–2268. Neu, R.W., Sehitoglu, H., 1991. Transformation of retained austenite in carburized 4320 steel. Metall. Trans. A 22, 1491–1500. Olson, G.B., Azrin, M., 1978. Transformation behavior of TRIP steels. Metall. Mater. Trans. A 9A, 713–721. Olson, G., Cohen, M., 1972. A mechanism for the strain-induced nucleation of martensitic transformations. J. Less-Common Metals 28, 107–118. Olson, G.B., Cohen, M., 1975. Kinetics of strain-induced martensitic nucleation. Metall. Trans. A 6, 791–795. Olson, G.B., Cohen, M., 1982. Stress-assisted isothermal martensitic transformation: Application to TRIP steels. Metall. Trans. A 13, 1907–1914.
CE
884
Moumni, Z., 1995. Sur la Modelisation du Changement de Phase Solide : Application aux Materiaux a Memorie de forme et a l’endommagement fragile partiel. Ph.D. thesis.
885
886
Ottosen, N.S., Ristinma, M., 2005 The Mechanics of Constitutive Modeling . Ovako, 2016. Steel Navigator. http//www.ovako.com/steel navigator/pdf/Ovako 827 - 100CrMnMoSi8-4-6 v3.pdf, access date 2016-09-01 .
AC
887
888
889
890
891 892
858 859 860
861 862
CR IP T
862
AN US
861
M
860
Morin, C., Moumni, Z., Zaki, W., 2011b. Thermomechanical Coupling in Shape Memory Alloys under Cyclic Loadings: Experimental Analysis and Constitutive Modeling. Int. J. Plast. 27, 1959–1980.
ED
859
PT
858
Papatriantafillou, I., 2005. TRIP Steels: Constitutive modeling and computational issues. Ph.D. thesis. University of Thessaly. Papatriantafillou, I., Agoras, M., Aravas, N., Haidemenopoulos, G., 2006. Constitutive modeling and finite element methods for TRIP steels. Comput. Methods Appl. Mech. Eng. 195, 5094–5114.
37
863
864 865 866
867 868 869
870 871 872
873 874
875 876
877 878
879 880
881 882
883 884
885
886 887
888 889
890 891 892
ACCEPTED MANUSCRIPT
898
899 900
901 902 903
904 905 906
907 908
909 910
911 912 913 914
915 916 917
918 919
Perlade, A., Bouaziz, O., Furn´emont, Q., 2003. A physically based model for TRIP-aided carbon steels behaviour. Mater. Sci. Eng. A 356, 145–152.
Popov, P., Lagoudas, D.C., 2007. A 3-D constitutive model for shape memory alloys incorporating pseudoelasticity and detwinning of self-accommodated martensite. Int. J. Plast. 23, 1679–1720.
Rauch, G.C., Daga, R.L., Radcliffe, S.V., Sober, R.J., Leslie, W.C., 1975. Volume expansion, pressure effects, and the strength differential in as-quenched iron-carbon martensite. Metall. Mater. Trans. A 6, 2279–2287. Rauch, G.C., Leslie, W.C., 1972. The extent and nature of the strength differential effect in steel. Metall. Trans. 3, 373–385. Rice, J., Tracey, D., 1969. On the ductile enlargement of voids in triaxial stress fields. J. Mech. Phys. Solids 17, 201–217. Saleeb, A., Padula, S., Kumar, A., 2011. A Multi-axial, Multimechanism Based Constitutive Model for the Comprehensive Representation of the Evolutionary Response of Smas under General Thermomechanical Loading Conditions. Int. J. Plast. 27, 655–687. Sangal, S., Goel, N.C., Tangri, K., 1985. A theoretical model for the flow behavior of commercial dual-phase steels containing metastable retained austenite: Part II. calculation of flow curves. Metall. Trans. A 16, 2023–2029. Seong, B.S., Shin, E.J., Han, Y.S., Lee, C.H., Kim, Y.J., Kim, S.J., 2004. Effect of retained austenite and solute carbon on the mechanical properties in TRIP steels. Phys. B Condens. Matter 350.
CE
920
Paul, S.K., 2013. Effect of martensite volume fraction on stress triaxiality and deformation behavior of dual phase steel. Mater. Des. 50, 782–789.
921
922
Shan, T., Li, S., Zhang, W., Xu, Z., 2008. Prediction of martensitic transformation and deformation behavior in the TRIP steel sheet forming. Mater. Des. 29, 1810–1816.
AC
923
924
925
926
927
893 894
895 896
CR IP T
897
Patoor, E., Eberhardt, A., Berveiller, M., 1996. Micromechanical Modelling of Superelasticity in Shape Memory Alloys. Le J. Phys. IV 06, C1–277–C1–292.
AN US
896
M
895
Patel, J., Cohen, M., 1953. Criterion for the action of applied stress in the martensitic transformation. Acta Metall. 1, 531–538.
ED
894
PT
893
Sierra, R., Nemes, J., 2008. Investigation of the mechanical behaviour of multiphase TRIP steels using finite element methods. Int. J. Mech. Sci. 50, 649–665. Singh, A.P., Padmanabhan, K.A., Pandey, G.N., Murty, G.M.D., Jha, S., 2000. Strength differential effect in four. J. Mater. Sci. 35, 1379–1388.
38
897 898
899 900
901 902 903
904 905 906
907 908
909 910
911 912 913 914
915 916 917
918 919 920
921 922 923
924 925
926 927
ACCEPTED MANUSCRIPT
933
934 935
936 937 938
939 940 941
942 943
944 945 946
947 948 949
950 951
952 953 954
Spitzig, W.A., Richmond, O., 1984. The effect of pressure on the flow stress of metals. Acta Metall. 32, 457–463.
Spitzig, W.A., Sober, R., Richmond, O., 1975. Pressure dependence of yielding and associated volume expansion in tempered martensite. Acta Metall. 23, 885–893. Spitzig, W.A., Sober, R.J., Richmond, O., 1976. The effect of hydrostatic pressure on the deformation behavior of maraging and HY-80 steels and its implications for plasticity theory. Metall. Trans. A 7, 1703–1710. Stoughton, T.B., Yoon, J.W., 2004. A pressure-sensitive yield criterion under a non-associated flow rule for sheet metal forming. Int. J. Plast. 20, 705–731. Stringfellow, R., Parks, D., Olson, G.B., 1992. A constitutive model for transformation plasticity accompanying strain-induced martensitic transformations in metastable austenitic steels. Acta Metall. Mater. 40, 1703–1716. Sun, L., Wagoner, R., 2011. Complex unloading behavior: Nature of the deformation and its consistent constitutive representation. Int. J. Plast. 27, 1126–1144. Sun, Q., 1991. A micromechanics constitutive model of transformation plasticity with shear and dilatation effect. J. Mech. Phys. Solids 39, 507–524. Taleb, L., Sidoroff, F., 2003. A micromechanical modeling of the GreenwoodJohnson mechanism in transformation induced plasticity. Int. J. Plast. 19, 1821–1842. Tanaka, K., Nagaki, S., 1982. A Thermomechanical Description of Materials with Internal Variables in the Process of Phase Transitions. Arch. Appl. Mech. 51, 287–299.
CE
955
Spitzig, W.A., Richmond, O., 1979. Effect of Hydrostatic Pressure on the Deformation Behavior of Polyethylene and Polycarbonate in Tension and in Compression. Polym. Eng. Sci. 19, 1129–1139.
956 957
958
Tanaka, K., Oberaigner, E.R., Fischer, F.D., 1996. Kinetics on the micro- and macro-levels in polycrystalline alloy materials during martensitic transformation.
AC
959 960
961
962
963 964 965
928 929 930
931 932
CR IP T
932
AN US
931
M
930
Sommer, C., Christ, H.J., Mughrabi, H., 1991. Non-linear elastic behaviour of the roller bearing steel SAE 52100 during cyclic loading. Acta Metall. Mater. 39, 1177–1187.
ED
929
PT
928
Tanaka, K., Sato, Y., 1985. A mechanical view of transformation-induced plasticity. Ingenieur-Archiv 55, 147–155. Thibaud, S., Boudeau, N., Gelin, J.C., 2006. TRIP steel: Plastic behaviour modelling and influence on functional behaviour. J. Mater. Process. Technol. 177, 433–438. 39
933
934 935
936 937 938
939 940 941
942 943
944 945 946
947 948 949
950 951
952 953 954
955 956 957
958 959 960
961 962
963 964 965
ACCEPTED MANUSCRIPT
971
972 973
974 975 976
977 978
979 980
981 982
983 984
985 986 987
988 989 990
Turteltaub, S., Suiker, A.S.J., 2006. A multiscale thermomechanical model for cubic to tetragonal martensitic phase transformations. Int. J. Solids Struct. 43, 4509–4545.
Watt, D., Jain, M., 1984. Effect of martensite morphology on the strength differential effect in dual phase steels. Scr. Metall. 18, 1379–1382.
Yoshida, F., Uemori, T., Fujiwara, K., 2002. Elastic-plastic behavior of steel sheets under in-plane cyclic tension-compression at large strain. Int. J. Plast. 18, 633–659. Young, C.C., 1988. Transformation toughening in phosphocarbide-strengthened austenitic steels. Ph.D. thesis. MIT. Zaki, W., 2010. An Approach to Modeling Tensile-compressive Asymmetry for Martensitic Shape Memory Alloys. Smart Mater. Struct. 19, 025009–025026. Zaki, W., Moumni, Z., 2007a. A 3D Model of the Cyclic Thermomechanical Behavior of Shape Memory Alloys. J. Mech. Phys. Solids 55, 2427–2454. Zaki, W., Moumni, Z., 2007b. A three-dimensional model of the thermomechanical behavior of shape memory alloys. J. Mech. Phys. Solids 55, 2455–2490. Zaki, W., Zamfir, S., Moumni, Z., 2010. An extension of the ZM model for shape memory alloys accounting for plastic deformation. Mech. Mater. 42, 266–274. Zavattieri, P., Savic, V., Hector Jr., L., Fekete, J., Tong, W., Xuan, Y., 2009. Spatio-temporal characteristics of the Portevin-Le Chˆ atelier effect in austenitic steel with twinning induced plasticity. Int. J. Plast. 25, 2298–2330.
994
Zhang, Y., Cheng, Y.T., Grummon, D.S., 2007. Finite Element Modeling of Indentation-Induced Superelastic Effect Using a Three-dimensional Constitutive Model for Shape Memory Materials with Plasticity. J. Appl. Phys. 101, 053507–053513.
995
Appendix A. Elastic strain for monotonic tension-compression
991 992
AC
CE
993
996
966 967 968
969 970
CR IP T
970
AN US
969
M
968
Tomita, Y., Iwamoto, T., 1995. Constitutive modeling of trip steel and its application to the improvement of mechanical properties. Int. J. Mech. Sci. 37, 1295–1305.
ED
967
PT
966
The equation
971
972 973
974 975 976
977 978
979 980
981 982
983 984
985 986 987
988 989 990
991 992 993 994
995
996
el 9Geq 0 K(εkk ) E(εel eq , kk ) = 3K(εel kk ) + G0
40
(51)
ACCEPTED MANUSCRIPT
2
997 998
el and σ h = K0eq εel kk + K1 εkk was evaluated for uniaxial tension and compression. The uniaxial stress was then calculated through
1000
el Equation (22) together with εel kk = (1 − 2v)ε11 yields 3(K0eq + K1 εel el kk ) − 2G εel εkk = 1 − eq 11 . 3(K0eq + K1 εel kk ) + G0
el Using Eq. (53), εel kk can be defined in terms of ε11 as p eq el 2 3(K0eq + Geq (3(K0eq + Geq el 0 ) 0 ) + 36K1 G0 ε11 εkk = − + . 6K1 6K1
1001
1005 1006
1011
1012
1013
(54)
1001 1002
1004
The dissipation potential of the system defined through free energy potential by Eq. (5) was
1006
3 X
m=1
m m Xij χ˙ ij − Rκ˙ R + Cζ ζ˙ > 0
PT
˙ ∂g = 3 sij − αij χ˙ m ij = −λ m ∂Xij 2 σeq
1005
1007 1008
(57)
∂g κ˙ R = −λ˙ (58) ∂R When a nonassociative flow rule is employed in order to compute the bounds for the influence of pressure that satisfies the dissipation equation Eq. (56) the following analysis has to be done. Equations (57), (58) and (35) were used in Eq. (56) and the stresses were written in deviatoric and dillatation components so that ! 3 X 3 γm m m ∗ h D = σeq + 3a σ + X Xij − R + Cζ ζ˙ ≥ 0 (59) m ij 2 C m=1 41
1003
(56)
m The corresponding work conjugates to the thermodynamic forces Xij and R can be defined as
AC
1009
1010
1000
Appendix B. Criterion for pressure dependence in the nonassociated flow rule
CE
1008
(53)
el When the Eq. (54) is used in Eq. (52), noting that εel kk = 0 for ε11 = 0, el = 0 when σ11 = 0 and solved for ε11 the following equation was obtained q eq 2 eq 4 σ11 3 σ11 K1 + (K0 ) − K0 . (55) εel = + 11 3Geq 6K1 0
D = σij ε˙pl ij −
1007
999
M
1003 1004
εel 11
ED
1002
(52)
AN US
999
+ K1 εel kk ) εel 11 . K1 εel ) + G kk
998
CR IP T
σ11 =
3(K0eq 3G 3(K0eq +
997
1009 1010 1011 1012 1013
ACCEPTED MANUSCRIPT
1015 1016
The positivity of Cζ ζ˙ is discussed in appendix C. Therefore for the rest of the analysis the method used by Linares Arregui and Alfredsson (2010) is followed. Since λ˙ > 0 and f = 0 during plastic deformation, σY 0 − 3σ h (a − a∗ ) +
1018 1019 1020
a
R−
3 P
m=1
m 3 γ m m 2 C m Xij Xij
σY 0 + R
≤ a∗
1016
(60)
This condition should be satisfied for large plastic deformations when the kinematic hardening is zero, Xij = 0. Hence a ≥ a∗ . For this conditions the hydrosatic tension can be formulated as 3σ h = σY 0a+R which yielded the equation
AN US
1017
3 X 3 γm m m X X ≥0 2 C m ij ij m=1
1014 1015
CR IP T
1014
1017
1018
1019
1020
(61)
1024
Linares Arregui and Alfredsson (2010) states that in the material model that considers the nonassociated flow rule with pressure dependent plastic flow rule, the chosen parameters a∗ and a has to be controlled throughout the simulations to verify the stability of the solution and satisfaction of dissipation equation.
1024
1025
Appendix C. Dissipation Potential
1025
1023
The pseudopotential of dissipation was
1026
1029
where both terms in this Cζ = 0. This yields the condition Cζ ≥ division are positive and numerator is zero when there is no plastic deformation. It is known that stress induced phase change starts before strain induced phase change below a certain temperature and the evolution of stress induced martensite fraction defined in Eq (49), ζstress ≥ 0. In addition to that the term ˙ Aζ ε˙pl eq + Bζ Σ is positive during plastic deformation and zero otherwise. Therefore here, as a transformation function the condition of Cζ ≥ 0 should be used. This conditon satisfies positivity of the term responsible from the dissipation during phase change Cζ ζ˙ in dissipation equation.
CE
1030
where ( ∂Φ Cζ )0 = 0 since there is no phase change when the termodynamic force
1031
1032
1033
AC
1034
˙ Aζ ε˙pl eq +Bζ Σ , − ξ1 (ξ 2 −ζ)
1035
1036
1037
1038
42
1023
(62)
According to Ottosen and Ristinmaa (2005) the conditions of convex potential analysis ensures that Φ(Cζ ) − Φ(0) ∂Φ ≥ , (63) (Cζ − 0) ∂Cζ 0
PT
1027 1028
1 ˙ (Cζ )2 ξ1 (ξ2 − ζ) + Cζ (Aζ ε˙pl eq + Bζ Σ) 2
1021 1022
1026
ED
Φ(Cζ ) =
M
1021 1022
1027 1028
1029
1030 1031 1032 1033 1034 1035 1036 1037 1038