Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Mohsen Zadeh 2016

Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

Author’s Accepted Manuscript

On the Yield Point Phenomenon in Low-Carbon


Steels with Ferrite-Cementite Microstructure

M.S. Mohsenzadeh, M. Mazinani

www.elsevier.com/locate/msea

PII: S0921-5093(16)30789-4
DOI: http://dx.doi.org/10.1016/j.msea.2016.07.033
Reference: MSA33864
To appear in: Materials Science & Engineering A
Received date: 15 May 2016
Revised date: 4 July 2016
Accepted date: 9 July 2016
Cite this article as: M.S. Mohsenzadeh and M. Mazinani, On the Yield Point
Phenomenon in Low-Carbon Steels with Ferrite-Cementite Microstructure,
Materials Science & Engineering A,
http://dx.doi.org/10.1016/j.msea.2016.07.033
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 galley proof before it is published in its final citable 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.
On the Yield Point Phenomenon in Low-Carbon Steels with Ferrite-Cementite
Microstructure
M.S. Mohsenzadeh, M. Mazinani*
Department of Materials Engineering, Faculty of Engineering, Ferdowsi University of Mashhad,
P.O. Box 91775-1111, Mashhad, Iran
*
Corresponding author. mazinani@um.ac.ir

Abstract
Low-carbon steels with a ferrite-cementite microstructure represent a
discontinuous yielding resulting in several industrially undesirable problems.
Considering the thermal contraction mismatch between the ferrite and cementite
phases during the cooling stage to room temperature, which is comparable with the
volume expansion due to austenite to martensite transformation, one may expect
that these steels, similar to dual-phase steels, exhibit a continuous yielding
behavior. In order to examine this phenomenon, different microstructures including
ferrite-cementite, ferrite-cementite-martensite and ferrite-martensite were produced
using a low-carbon steel and yielding behavior of steel samples with these
microstructures were evaluated during uniaxial tensile loading. In addition, the
effect of volume expansion due to austenite to martensite transformation and
thermal contraction mismatch between ferrite and cementite phases on each
microstructure was also evaluated. It was found that for the case of ferrite-
cementite steels with very small cementite particles, continuous yielding cannot be
observed. Finally, yielding behavior of different steel samples were explained
based on the theoretical results obtained.
Keywords: yield point phenomenon; steel microstructure; dislocations; internal
stress induced yielding of ferrite.

Nomenclature

Ms Martensite start temperature

ΔVγ→α Difference between the specific volumes of austenite and martensite phases

Vγ Specific volume of austenite


C Carbon content of martensite in weight percent

εM* Transformation strain of martensite

εc Transformation strain of cementite

εc* Equivalent transformation strain of cementite

I Thermal expansion coefficients of cementite phase

M Thermal expansion coefficients of ferrite phase

ΔT Temperature change

CI Stiffness tensor of inclusion (cementite particle)

CM Stiffness tensor of matrix (ferrite phase)

S Eshelby tensor

δkl Kronecker delta

a Radius of second phase particle (martensite or cementite)

R Radius of plastic zone formed around the particle due to internal stress induced yielding
of matrix

ε(r) Equivalent plastic strain

E Young’s modulus

 Poisson’s ratio

Yo Yield strength of ferrite at the temperature of internal stress induced yielding

α Constant

E Energy of interaction of solute atoms and dislocations in Equation (10)

c Concentration of solute atoms

k Boltzmann’s constant

T Temperature

b Burgers vector

KHP Hall-Petch parameter


D Grain size of ferrite phase

M Taylor factor

G Shear modulus

d The diameter of the particle

L Mean particle spacing

f Volume fraction of particles

Y Yield strength of ferrite

ρGND Average density of geometrically necessary dislocations (GNDs)

γ Shear strain

ε Normal strain

fmin Minimum volume fraction of the second phase required for the occurrence of internal
stress induced yielding of ferrite throughout the microstructure

 Equivalent stress

σm Mean normal stress

Ho Initial strain hardening rate of the ferrite at internal stress induced yielding

σA Uniaxial tensile stress

A Weighted average yield stress in the internal stress induced yielded zone

θ The angle from the x1-x2 plane to the r direction with the assumption that the tensile
stress acts in the x3 direction

σs Stresses required for operating dislocation sources

σi Stresses required for overcoming obstacles to the motion of dislocations produced by the
sources

γ Amount of crystallographic slip in Equation (27) and (29) generated by the dislocations
moving a certain distance

ρ Dislocation density

l The distance moved by dislocations


fPZ Volume fraction of plastic zones generated by internal stress

1. Introduction
Due to the existence of solute interstitials such as carbon and nitrogen atoms, low
carbon steels like many metals and alloys exhibit discontinuous yielding
manifesting itself in the appearance of a sharp yield point and subsequent yield
plateau in the stress-strain curve. By diffusing to the dislocation sites, solute atoms
hinder the movement of dislocations. In order to unlock these dislocations, a
certain stress is required which is known as the upper yield stress. When the stress
level reaches the upper yield stress, dislocations will be released at the stress
concentration sites. After releasing, dislocations could move with a lower stress
level corresponding to the lower yield stress. The stress drop at the yield point is
accompanied by the nucleation of slip bands named Lüders bands at the points of
stress concentration. Since stress concentration is generated at the Lüders bands
front, dislocations are gradually unlocked in elastically deformed sections of the
specimen and propagation of the Lüders bands along the whole gauge length of the
specimen is enabled. Since Lüders bands propagate under a constant stress level
equals to the lower yield stress, a plateau is formed in the stress-strain curve called
the yield point elongation [1, 2].
Causing industrially important problems, the yield point phenomenon in carbon
steels has been investigated for several decades. Besides early studies on
explaining the mechanism of discontinuous yielding in such steels, the influences
of many factors such as ferrite grain size, type of the steel microstructure, strain
rate, deformation temperature and carbon content on discontinuous yielding, the
magnitude of Lüders strain and the velocity of Lüders bands front have been
studied [3-10].
Formation of a certain amount of martensite phase in the microstructure of low
carbon steels leads to a continuous yielding behavior. Elimination of the yield
point phenomenon has been attributed to introducing mobile dislocations into the
ferrite matrix as a result of volume expansion accompanying austenite to
martensite transformation. Internal stresses produced due to austenite to martensite
transformation have been also considered to be responsible for continuous
yielding behavior [11-13]. On the other hand, in the stress-strain curve of low
carbon steels with ferrite-cementite microstructure during uniaxial tensile loading,
yield point phenomenon is also observed. Considering the thermal contraction
mismatch between the ferrite and cementite phases generated upon cooling, which
is comparable to the volume expansion due to austenite to martensite
transformation, a very important question arises that why continuous yielding
behavior could not be observed in the stress-strain curve of these steels with
ferrite-cementite microstructure. The present research work is an effort to account
for this question. In order to achieve this goal, the yielding behavior of steels with
different microstructures including ferrite-cementite, ferrite-cementite-martensite,
and ferrite-martensite microstructures has been investigated theoretically.
2. Experimental procedure
A low carbon steel in the form of a sheet 4.6 mm in thick was used in this research
work. The chemical composition of the investigated steel is given in Table 1.
Samples with the size of 80 mm×15 mm were cut from the sheet and then, were
austenitized at 1000 ºC for 30 min followed by quenching in an ice brine solution.
The resulting martensitic steel was tempered at 650 ºC for 1 hour and then, was
cold rolled by a 80% reduction using a laboratory rolling mill with the roll
diameter of 57 cm. The cold rolled samples were tempered for another 2 hours at
650 ºC. This treatment resulted in a microstructure consisting of cementite
particles in a ferrite matrix. The samples with ferrite-cementite microstructure are
labeled “F” throughout the paper. In order to produce different volume fractions of
martensite in steel microstructures, some of the samples with ferrite-cementite
initial microstructure were intercritically annealed at 740 ºC for 30 seconds, 1 and
15 min followed by quenching into an ice brine solution. Consequently, two types
of microstructures including ferrite-cementite-martensite (labeled “B”) and ferrite-
martensite (labeled “C”) were produced. The microstructures of steel samples were
investigated using scanning electron microscope (SEM) (Leo 1450VP). Samples
for microstructural examination were etched with 3% Nital solution. The sizes of
martensite islands and cementite particles as well as the volume fractions of
martensite and cementite phases were measured using Clemex image analysis
software. In order to evaluate yielding behavior, tensile tests were conducted on
specimens prepared according to ASTM-E8 standard with a 25 mm gauge length
and the strain rate of 0.002 s-1 using Zwick Z250 universal tensile test machine. In
order to obtain stress-strain curve of ferrite, specimens of a steel containing 0.09
wt.% carbon were also tested.
3. Results of the tensile testing
The results of quantitative metallography for each specimen are presented in Table
2. The microstructures of these specimens are shown in Figure 1. Figure 2 (a to d)
illustrates the stress-strain curves of the specimens F (2.3% cementite), B1 (1.8%
cementite, 6.2% martensite), B2 (1.2% cementite, 13.1% martensite) and C (39%
martensite). In addition, the stress-strain curve of the steel containing 0.09 wt.%
carbon, which can be considered to be the stress-strain curve of ferritic steel [14] is
shown in Figure 2-e. A yield drop and the yield point elongation can be observed
in this mostly ferritic steel. In the specimen F with ferrite-cementite microstructure,
the yield drop is eliminated, however, the yield point elongation is still observed.
As can be seen in Figure 3, the magnitude of the yield point elongation is larger for
specimen F than that for the ferritic steel. The formation of 6.2% martensite phase
in steel B1 has eliminated the yield point elongation, however, a clear yield point
appears in the specimen B1 consisting of cementite particles (150 nm, 1.8%) and
martensite islands (1.763 m, 6.2%) in a ferrite matrix. As can be seen in the
stress-strain curve of specimen B1, the work hardening rate increases after a small
plastic deformation. To show more clearly the change of work-hardening rate at
the early stages of deformation for this steel specimen, the variation of work-
hardening rate as a function of applied true strain is shown in Figure 4. The work
hardening rate of specimen C (with coarse martensite islands, 4.15 m, 39%) is
also illustrated in this figure for comparison. As can be clearly observed, instead of
gradually decreasing trend as one can see for the case of specimen C, the work
hardening rate of specimen B1 increases after experiencing a certain amount of
plastic strain. Increasing the volume percent martensite to 13.1 (still having 1.2%
cementite) has resulted in an almost complete continuous yielding behavior in the
specimen B2 similar to the specimen C containing martensite islands as the only
second phase in the microstructure.
In the following sections, the effects of volume expansion due to austenite to
martensite transformation and thermal contraction mismatch between ferrite and
cementite phases on the microstructures will be evaluated and the experimental
observations will be explained in more details.
4. Misfit strains of martensite islands and cementite particles
Based on the study conducted by Moyer et al. [15], the volume expansion as a
result of austenite to martensite transformation at martensite start temperature (Ms)
is determined using the following equation:
 V   
   0.017  0.013C (1)
 V  M s

where ΔVγ→α is the difference between the specific volumes of austenite and
martensite phases, Vγ is the specific volume of austenite, and C is the carbon
content of martensite in weight percent.
Considering the transformation strain of martensite to be an isotropic expansion,
the magnitude of the transformation strain (εM*) can be calculated as follows [13]:
M
*
 0.0058  0.0045C (2)

According to the above relation, the transformation strain of martensite island


depends only on its carbon content. Since the carbon concentration in the
martensite islands as measured by the energy dispersive X-ray analyzer (EDS) for
specimens B1, B2 and C are 0.427 wt.%, 0.502 wt.% and 0.4 wt.%, respectively, the
transformation strains of martensite islands in these specimens are determined as
follows:
M
*
, ( B1 )  0.0077215 ,  M , ( B2 )  0.008059 ,  M ,(C )  0.0076
* *

On cooling the specimen F from 650 ºC to room temperature (27 ºC), both the
ferrite matrix and cementite particles shrink. However, due to the lower thermal
expansion coefficient of cementite phase, the amount of its shrinkage is less than
that of the ferrite phase generating a misfit between the cementite particles and
corresponding holes in the matrix from whence the cemented particles have came.
This misfit results in the internal stress set up in both the cementite particles and
ferrite phase similar to that created due to austenite to martensite transformation.
The magnitude of the transformation strain in this case is given by [16]:
 c  ( I   M )T (3)
where I and M are the thermal expansion coefficients of cementite and ferrite
phases, respectively, and ΔT the temperature change. The values of the thermal
expansion coefficients of cementite and ferrite phases are [17]:
 I  6 106  3 109 (T  273)  1011(T  273)2 (K 1)

 M  1.244 105 (K 1 )

For the temperature change from 650 ºC to room temperature (27 ºC), εc can be
obtained as:
 c  6.44 106  3 109 (T  273)  1011(T  273)2 dT

One may consider an equivalent transformation strain (εc*) for the calculation of
the internal stress. εc* is the transformation strain of an equivalent particle with a
stiffness equals to that of the matrix generating identical internal stress like the real
particle. εc* can be calculated using the following equation [16]:
 c*  [(CI  CM )S  CM ]1CI  c (4)

where, εc and εc* are, respectively, the transformation strain tensor and equivalent
transformation strain tensor, CI and CM are the stiffness tensors of inclusion
(cementite particle) and matrix (ferrite phase), respectively, and S is the Eshelby
tensor calculated using the particle’s aspect ratio and Poisson’s ratio. In the indicial
notation, the previous equation can be written as:
 cij
*

 (CI ijkl  CM ijkl )Sijkl  CM ijkl 
1
CI ijkl  ckl i, j, k  1,2,3 (5)

where  ckl is given by the following equation:


 ckl   c kl (6)

in which δkl is the kronecker delta, i.e.:


0 k l
 kl   (7)
1 k l
The components of the stiffness tensors of ferrite and cementite phases as well as
those of the Eshelby tensor for spherical particles used in the calculations are
reported in Table 3. From equation 5 and 6, εc* can be obtained as:

 c*  3.54 c  3.54   6.44 106  3 109 (T  273)  1011(T  273)2 dT 
For the cooling stage from 923 K (650 ºC) to 300 K (27 ºC):
300
 c*  
923
(2.306 105  8.708 109 T  3.54 1011T 2 )dT  0.00872

Comparing this transformation strain with the value of εM* for the case of
martensite phase in specimen B2, it is clearly found that the equivalent
transformation strain for the cementite particle is nearly equal to that of the
martensite island containing 0.5 wt.% carbon. Consequently, from the point of
view of their micromechanical effect on the matrix (ferrite phase), the cementite
particles could be considered to be martensite islands creating an approximately
similar volume expansion as that of the martensite islands.
5. Internal stress induced yielding of ferrite
Assuming a spherical inclusion, a martensite island or a cementite particle, with a
radius “a” embedded in a matrix (ferrite phase), internal stress generated due to the
volume expansion of martensite island or the thermal contraction mismatch
between ferrite and cementite phases could be calculated using the Eshelby’s
method. Considering the origin of the co-ordinate system at the center of the
particle, a strong hydrostatic stress acts on the particle in the range of r < a. On the
other hand, an equivalent stress acts on the ferrite matrix around the particle
decreasing with the distance from the inclusion-matrix interface. Just outside the
interface of particle-matrix, the magnitude of the equivalent stress is so high that it
causes the ferrite to yield within a radius R. The equivalent plastic strain, ε(r), in
the range of a ≤ r ≤ R can be calculated as follows [13]:
a3
 (r )  2 * (8)
r3
As can be seen, the value of the plastic strain is maximum at the interface of
particle-matrix and decreases rapidly with 1/r3. Figure 5 illustrates the distribution
of ε(r) for the investigated specimens. ε(r) depends on the size of the particle and
misfit strain. Misfit strain of cementite particles is nearly equal to that of
martensite islands. However, because of the smaller size of cementite particles,
although the same values are predictable immediately outside the cementite
particle and martensite island, ε(r) decreases more rapidly with the distance from
the interface of cementite particle-matrix.
Considering ε(R) = 0, the radius of the yielded zone (R) can be obtained as [13]:
Ea 3 *
R3  (9)
(1  )Yo

where E is the Young’s modulus,  the Poisson’s ratio and Yo the yield strength of
ferrite at the temperature at which the internal stress induced yielding occurs. In
the steel sample containing martensite islands, the assumption is that
transformation induced yielding occurs at Ms temperature. Therefore, the yield
strength of ferrite phase at Ms temperature must be determined. On the other hand,
in the steel with cementite particles, yielding of ferrite phase occurs in a
temperature range from 650 to 27 ºC. Consequently, the yield strength of ferrite
matrix as a function of temperature must be determined.
Combining the contributions of different strengthening mechanisms including the
Peierls-Nabarro stress, solid solution strengthening, grain boundary strengthening,
and precipitation strengthening (for steels containing cementite particles), the yield
strength of the ferrite matrix can be theoretically calculated. In the investigated
temperature range, athermal strengthening mechanisms including grain boundary
strengthening and precipitation strengthening dominate [18]. Thus, the yield
strength of ferrite does not decrease significantly with increasing temperature. The
slight decrease of yield strength with temperature is due to the decrease of shear
modulus and thermal components of stress, i.e. solid solution strengthening and
Peierls-Nabarro stress.
According to the theory presented by Suzuki for BCC metals, the temperature
dependence of solid solution strengthening is expressed as [19]:
E 2 c
 (10)
kTb3
where α is a constant, E energy of interaction of solute atoms and dislocations, c
concentration of solute atoms, k Boltzmann’s constant, T temperature, and b is the
burgers vector. Solid solution strengthening due to interstitial (mainly carbon,
0.0218 wt.% [20] and substitutional solute atoms (manganese and silicon, 1.15
and 0.411 wt.%, respectively, as can be seen in Table 1) at room temperature are
reported to be 57.22 and 178.72 MPa, respectively [2, 19, 21]. From Equation 10
and using the magnitude of solid solution strengthening at room temperature, the
following relation for the temperature dependence of solid solution strengthening
is obtained:
70782
 (MPa)
T
The grain boundary strengthening can be well described by the Hall-Petch
relationship [2], i.e.:
K HP
 gb  (11)
D

where KHP is the Hall-Petch parameter and D is the grain size. KHP is the only
temperature dependent factor showing a weak temperature dependency in the
investigated range [18]. Consequently, the temperature dependence of KHP is
assumed to be negligible and the value of grain boundary strengthening is
considered to be almost constant in the whole temperature range. By substituting
the measured value of ferrite grain size, D = 10 μm, and the value of Hall-Petch
parameter from the literature, KHP = 150 MPa(μm)1/2 [22], the grain boundary
strengthening component of the ferrite yield strength is obtained to be 47.43 MPa.
Assuming the cementite particles to be non-sharable incoherent particles, their
precipitation strengthening contribution could be described by Orowan relation, i.e.
[2]:
MGb
 p  (12)
Ld

where M is the Taylor factor, G shear modulus of the matrix, b Burgers vector of
the matrix, d the diameter of the particle, and L the mean particle spacing
calculated as:
1/ 2
 d2 
L 
 4f  (13)
 
in which f is the volume fraction of particles. The only temperature dependent
parameter in the above relation is shear modulus. The temperature dependency of
shear modulus of ferrite phase can be expressed as follows:
G  90.033  0.00267T (GPa) (14)

From the quantitative metallography conducted in this study, the diameter of


cementite particles was measured to around 150 nm and their volume percent in
specimens F, B1 and B2 were 2.3, 1.8 and 1.2%, respectively. Therefore, the
contribution of precipitation strengthening of ferrite phase in these specimens can
be obtained, i.e.:
 ( F )  98.59  0.0293T (MPa)

 ( B1 )  85.19  0.0253T (MPa)

 ( B2 )  67.35  0.02T (MPa)

Summing all the contributions of different strengthening mechanisms, the yield


strength of ferrite (Y) as a function of temperature can be calculated for each
specimen:
70782
Y (T )( F )   146.021  0.0293T (MPa) (15)
T
70782
Y (T )( B1 )   132.62  0.0253T (MPa) (16)
T
70782
Y (T )( B2 )   114.78  0.002T (MPa) (17)
T
70782
Y (T )(C )   47.43 (MPa) (18)
T
Using Equations 15 and 18, the yield strengths of the ferrite phase containing 2.3%
cementite particles and that without cementite particles at room temperature (27
ºC) are calculated to be 373.2 and 283.4 MPa, respectively. Comparing the
calculated yield strengths with the experimental ones obtained for specimen F and
the steel with 0.09 wt.% carbon (the ferritic steel), i.e. 390.6 MPa and 299 MPa
respectively, it is clearly seen that the model presents a relatively good estimation
of the yield strength. It should be mentioned that the higher value of experimental
yield strength compared with the calculated value is due to the triaxial stress state
developed at the Lüders bands front [6].
Now the values of R can be calculated using Equation 9. Since the carbon
concentration in the martensite islands in specimens B1, B2 and C are 0.427, 0.502
and 0.4 weight percent, respectively, the corresponding Ms temperatures are about
655.8, 631.9 and 665.2 K, respectively [23]. Using Equations 16 to 18, the yield
strength of ferrite matrix (Yo) in specimens B1, B2 and C at Ms temperature are
obtained as 223.93, 214.13, 153.83 MPa, respectively. Thus, the magnitude of R
for each specimen can be obtained as follow:
R( F )  160.35 (nm) , R( B1 )  1.884 (m) , R( B2 )  2.523 ( μm) , R(C )  5.096 ( μm)

From Equation 9, it can be seen that both the particle size (a) and the yield strength
of ferrite (Yo) at the temperature of internal stress induced yielding affect the value
of R. However, the influence of particle size is more significant than that of yield
strength. Hence, if Yo for the specimen F is considered to be equal to that of ferrite
without cementite precipitates, the increase of R will be 21.52 nm, whereas with
the assumption that the radius of cementite particle being equal to that of the
martensite island in specimen C (for the fixed value of Yo), the value of R is
determined to be 4.43 μm, i.e. approximately equal to that of specimen C.
Consequently, because of the smaller size of cementite particles (150 nm) than that
of martensite islands in steels B1, B2 and C, i.e. 1.73, 2.25 and 4.15 μm,
respectively, a plastic zone with a smaller radius is expected to form around
cementite particles.
Due to the plastic deformation of ferrite, a high density of unlocked dislocations
called geometrically necessary dislocations (GNDs) is formed around the
martensite island or cementite particle (inclusion) to maintain the lattice continuity
[11, 12]. Since the plastic strain at the interface between particle and matrix is
maximum and decreases with the distance from the interface, a non-uniform
distribution of dislocations the majority of them at the interface, would be
generated. For the investigated microstructures, the average density of GNDs
(ρGND) can be estimated using the following equation [24]:
 f  8  f  16
GND      (19)
d b d b
where d is the diameter of the second phase particles, f the volume fraction of the
second phase, γ the shear strain and ε the normal strain. The average values of
plastic strains for the specimens F, B1, B2 and C are 0.00599, 0.00517, 0.00519 and
0.00435, respectively, and using these values, ρGND are calculated to be 56.52,
11.43, 18.55 and 25.15 1/μm2, respectively, for the specimens F, B1, B2 and C. It
can be seen that possessing a higher value of f/d (equal to 0.153) due to the
smaller particle size, the specimen F has a greater value of ρGND than other
specimens having the values of f/d equal to 0.036, 0.058 and 0.093.
The minimum volume fraction of the second phase (fmin) required for the
occurrence of internal stress induced yielding of ferrite throughout the
microstructure can be obtained using the following equation [13]:
a3  (1  )Yo 
f min  100   * 
100 (20)
 E 
3
R

The magnitudes of fmin for the specimens F, B1, B2 and C are obtained to be 10.23,
9.66, 8.85 and 6.74%, respectively. Based on Equation 20, fmin depends on ε* and
Yo. Since ε* could be considered to be approximately the same for all specimens,
one may speculate that the value of Yo has a significant influence on fmin. Because
of the greatest contribution of precipitation strengthening and consequently, having
the highest Yo (255.8 MPa) compared with those of other specimens (223.9 MPa,
214.1 MPa and 153.8 MPa for specimens B1, B2 and C, respectively), the value of
fmin is maximum for the specimen F and decreases gradually with a decrease in the
volume fraction of cementite particles until reaches its minimum value in the
specimen C with a ferrite matrix which does not contain cementite particles.
Despite having a noticeable influence on R, it appears that the particle size does not
affect fmin. However, it can be said that the particle size indirectly affects fmin via
influencing on the precipitation strengthening component of Yo. Eliminating the
contribution of precipitation strengthening in calculating Yo, fmin is determined to be
7.01% for specimen F. Thus, for a specimen containing a greater volume fraction
of cementite particles, it is possible to find a particle size at which the ferrite
matrix yields throughout the microstructure.
Due to the yielding of ferrite, internal stresses generated in the particle and matrix
are partially relaxed. Equivalent stress (  ) and mean normal stress (σm) in the
particle and matrix can be calculated as follow [13]:
2 2 R3
 m (r )   [Yo  H o (a)]  hYo ln ,  (r )  0 (r  a) (21)
3 3 a3
1
h 1
 2(1  ) H o 
1  
 E 

2 R3
 m (r )   hYo ln ,  (r )  Yo  H o (r ) ( a  r  R) (22)
3 a3

R3
 m (r )  0 ,  (r )  Yo ( r  R) (23)
a3
where Ho is the initial strain hardening rate of the ferrite at internal stress induced
yielding. Since Hoε(r) is a negligible value (considering the value of Ho to be 1000
MPa, Hoε(r)≤17 MPa), the variation of Ho with temperature is neglected and its
magnitude is considered to be approximately 1000 MPa [13]. Thus, internal
stresses for the investigated specimens are obtained as follow:
0.00736 1.054
 ( F )  255.78  3
( a  r  R) ,  ( F )  ( r  R)
r r3
9.994 1499.211
 ( B1 )  223.93  3
(a  r  R) ,  ( B1 )  ( r  R)
r r3
22.949 3442.383
 ( B2 )  214.13  3
(a  r  R) ,  ( B2 )  ( r  R)
r r3
135.799 20369.911
 (C )  153.83  3
( a  r  R) ,  (C )  ( r  R)
r r3
Figure 6 illustrates the distribution of the equivalent stress (  ).
6. Evaluation of yielding of ferrite in the steel under uniaxial tension
Under a uniaxial tensile stress (σA), the ferrite phase in the steel microstructure
yields initially. The yield strength of ferrite in the internal stress induced plastic
zone is considered to be Y + Hoε(r), where Y is the yield strength of undeformed
ferrite at room temperature and Y ≥ Yo. Because of the work hardening caused by
the internal stress induced plastic deformation, the yield strength of ferrite in the
plastic zone is greater than Y. However, since Hoε(r) is less than Y, similar yield
strength of Y can be considered for the entire ferrite matrix.
Possessing maximum internal stress, plastic zones formed around the particles
yields initially and the yielded zones propagates then throughout the ferrite matrix.
Using the Von-Mises yielding criterion, the yield stress of ferrite in the elastic
deformation zones around the particles generated by the internal stress can be
expressed as [13]:
 2 2 
 3 sin   1 Yo R
2 3  3 sin 2   1 Yo R3   Yo R3  
A  Y        1 (r  R)
2 Y r 3  2 3 
Y r  Y r  3  (24)
  
 

with the assumption that the tensile stress acts in the x3 direction, θ is the angle
from the x1-x2 plane to the r direction (see Figure 7). The yield stress of ferrite in
the plastic zones around the particles generated by the internal stress can be
obtained as [13]:
 2 
 3 sin   1 Yo
2  3 sin 2   1 Yo   Yo 2 
A  Y        1 (a  r  R)
2 Y  2 Y  Y  (25)
   
 

and for the specimens investigated in the present study:


 2 
 3 sin   1
2  3 sin 2   1  
 A( F )  373.17 (0.68)   (0.68)   0.68  1
2
(a  r  R)
2  2 
   
 
 2 
 3 sin   1
2  3 sin 2   1  
 A( B )  360.95 (0.62)   (0.62)   0.62  1
2
(a  r  R)
2  2 
   
1

 

 2 
 3 sin   1
2  3 sin 2   1  
 A( B )  344.71 (0.621)   (0.621)  0.621  1
 2
(a  r  R)
2  2 
   
2

 
 2 
 3 sin   1
2  3 sin 2   1  
 A( C )  283.37 (0.54)   (0.54)  0.54  1
 2
(a  r  R)
2  2 
   
 

From Equation 25, it can be seen that the yield stress of ferrite in the plastic zone
depends on Y, Y/Yo and . In general, σA enhances with increasing  and a gradient
of yield stress is formed in the plastic zone. With an increase in the ratio of Y/Yo,
this gradient reduces and approaches zero. Figure 8 shows the yield stress in the
plastic zone (σA) as a function of θ. For all of the investigated specimens, the ratio
of Y/Yo is almost the same (see Figure 8), thus the gradient of σA in the plastic zone
is also the same. Having the highest value of Y (373.2 MPa), the specimen F
possesses the maximum value of σA. With decreasing the volume fraction of
cementite particles and consequently Y (see Equations 15-18, the values of Y are
361 MPa, 344.7 MPa and 283.4 MPa for specimens B1, B2 and C, respectively), the
magnitude of σA decreases.
The weighted average yield stress (  A ) in the internal stress induced yielded zone
can be calculated as follows [13]:
  2
1  3 sin 2   1 Yo   Yo  2
A   2
A 2R cos  d  Y        1. cos  d (a  r  R)
2 2
2R 2  2 Y  Y 
0 0
 
(26)
 A( F )  295.38 (MPa),  A( B1 )  299.87 (MPa),  A( B2 )  286.17 (MPa),  A(C )  247.92 (MPa)

 A depends only on the ratio of Y/Yo and Yo. Because of the presence of internal
stresses in the plastic zone, the average yield stress in this zone is considerably
lower than the yield strength of ferrite at room temperature. Since the ratio of Y/Yo
is the same for all the specimens,  A is mainly determined by Yo and therefore,  A
is maximum for the specimen F and decreases with a decrease in the volume
fraction of cementite particles in the steel.
7. Discussion
As mentioned previously, the plastic zone generated by the internal stress exhibits
a yield stress gradient. With increasing the applied load, yielding begins from some
regions around the second phase particles with the minimum yield stress. In order
to continue yielding process, the applied stress must increase. Increasing the
applied stress causes neighboring regions to yield and previously yielded regions
undergo work hardening. Consequently, if internal stress induced plastic zones
percolate into the whole microstructure, i.e. f ≥ fmin, the elimination of Lüders
bands and continuous yielding is expected.
The presence of a high density of unlocked dislocations in the plastic zone lowers
the yield stress. Considering the yield stress to be the sum of the stresses required
for operating dislocation sources (σs) and overcoming obstacles to the motion of
dislocations produced by the sources (σi) [1], the decrease of yield stress due to the
presence of unlocked dislocations could be attributed to the decrease of σs. Non-
uniform distribution of unlocked dislocations, i.e. reduction of ρGND with the
distance from the interface of particle-matrix, could be considered to promote
continuous yielding.
Based on the experimental results obtained from the tensile tests, yield drop is not
observed in the specimen with a ferrite-cementite microstructure. However, this
specimen exhibits a large yield point elongation. Thermal expansion mismatch
between the ferrite and cementite phases results in the formation of unlocked
dislocations (GNDs) distributed in the plastic zone formed around the cementite
particles. Due to the presence of unlocked dislocations, the elimination of yield
drop is expected. The volume percent of particles, fmin, required for the elimination
of yield drop in the specimen F is 10.23%. Considering the volume percent of
cementite particles in this specimen, i.e. 2.3%, one may conclude that the plastic
zones generated by internal stresses occupy only a small portion of the
microstructure which is not enough to prevent the formation of Lüders bands.
Consequently, under uniaxial tensile loading, yielding starts at some regions of the
microstructure and encompasses the entire gauge length of the specimen under a
relatively constant stress. Since unlocked dislocations are distributed only in the
plastic zones around the cementite particles, the main part of the specimen contains
locked dislocations which are released due to the simultaneous effects of stress
concentration produced by dislocation pile-up and the applied stress [1]. The
magnitude of Lüders strain is greater for the specimen F (3.48%) than that for the
ferritic steel (2%). During the growth of Lüders bands along the entire gauge
length of the specimen, the material does not work harden. Therefore, the density
of dislocations is constant and the amount of crystallographic slip (γ) generated by
the dislocation density of ρ moving the distance l can be obtained as follows [2]:
  bl (27)

where b is the burgers’ vector. One can consider the magnitude of the Lüders strain
to be proportional to γ. As a result, according to Equation 27, the Lüders strain
depends on both the dislocation density (ρ) and the distance moved by dislocations
(l). If “l” is considered to be the distance between dislocations and additionally,
distribution of dislocations is assumed to be uniform, the following relation can be
written:
1
l (28)

Using Equation 28, Equation 27 can be written as:


 b  (29)

As mentioned previously, because of the presence of GNDs, the density of


dislocations in the ferrite-cementite microstructure is greater than that in the ferritic
steel. Consequently, the Lüders strain is greater in the specimen F.
The formation of 6.2% martensite in the specimen B1 results in the elimination of
yield point elongation, although a clear yield point was observed in the stress-strain
curve (see Figure 2-b). The value of fmin for this specimen is 9.66%. Considering
the volume percent martensite in this specimen (6.2%), it is expected that the
internal stress induced plastic zones occupy a considerable portion of the
microstructure and unlocked dislocations corresponding to these zones are
distributed in a larger area of the microstructure. As a result, under the uniaxial
tensile loading, yielding begins in many regions of the microstructure and the
Lüders bands formation is highly prevented. However, because the plastic zones
have not been percolated to the whole microstructure, a clear yield point is still
observed in the stress-strain curve. If the Lüders bands formation is prevented, the
material work hardens after initial yielding and the density of dislocations
increases. In the specimen B1, yielding starts at internal stress induced plastic zones
having a lower yield stress and these zones of the microstructure will then work
harden. When the stress reaches the yield stress of other regions, these regions also
yield and subsequently work harden giving rise to an increase in the work
hardening rate after a small plastic strain, as can be seen in the stress- strain curve
of the specimen B1 in Figure 2-b and Figure 4.
In the specimens B2 and C, a continuous yielding phenomenon is observed. Since
the volume fractions of martensite in these specimens (13.1 and 39%, respectively)
are much greater than corresponding fmin values (8.85 and 6.74%, respectively), the
plastic zones generated by internal stresses are able to encompass the entire
microstructure. Observation of continuous yielding in the stress-strain curves of
specimens B2 and C can be attributed to the percolation of plastic zones to their
entire microstructures.
It can be concluded that the presence of plastic zones generated by internal stress
in the entire microstructure has a main role in continuous yielding of the steel and
thus, fmin is a very important parameter determining the extent to which the plastic
zones are developed in the microstructure.
It has been found that producing a bimodal grain size distribution in a low carbon
steel with ferrite-cementite microstructure results in continuous yielding behaviour
[25, 26]. According to the Hall-Petch relation, the variation of ferrite grain size
causes the yield strength to vary in different regions of the steel microstructure.
Consequently, the situation is similar to the formation of yield stress gradient in
internal stress induced plastic zones of inclusions and therefore, the microstructure
with different yield strength yields continuously. In addition, non-uniform
deformation of different regions of the microstructure gives rise to the formation of
GNDs assisting continuous yielding.
As mentioned previously, fmin depends on ε* and Yo (Equation 20). Via influencing
Yo (Equation 12), the size of cementite particles could indirectly affect fmin. Because
of precipitation hardening caused by small cementite particles, fmin is always
greater than the volume fraction of cementite particles (f) in steel with ferrite-
cementite microstructure. As a result, discontinuous yielding does not eliminate in
steels with ferrite-cementite microstructure. The volume fraction of plastic zones
(fPZ) generated by internal stress can be approximately obtained as:
3
R 3
E f
* 2
f PZ  3
100  3 (30)
L
 2 (1  )Y
o

As can be seen in Equation 30, fPZ depends on both Yo and f. For a constant volume
fraction of cementite particles (fixed f), the microstructural condition resulting in
the elimination of yield point elongation, e.g. the situation as for the specimen B1,
depends on Yo, and indirectly on the particle size. Thus, the presence of small
cementite particles can be considered to be the main reason that discontinuous
yielding does not eliminate in steels with ferrite-cementite microstructures. In
steels with higher volume fractions of cementite particles, it is possible to find a
particle size at which the yield point phenomenon does not occur.
8. Conclusion
In this study, the influence of volume expansion due to austenite to martensite
transformation and thermal contraction mismatch between ferrite and martensite
phases in different steel microstructures including ferrite-cementite, ferrite-
martensite-cementite and ferrite-martensite as well as yielding behavior of these
steels were investigated. The main results of the present study are summarized as
follows:
1. The magnitude of equivalent transformation strain for a cementite particle is
approximately equal to that of a martensite island. Therefore, one may speculate
that the cementite particle generates a volume expansion equal to that of the
martensite island. Because of much smaller size of the cementite particle (150 nm
compared with 1.73 to 4.15 m for the case of steels containing martensite phase
as well), the radius of the plastic zone (R) generated by the internal stress resulted
from the volume expansion/thermal contraction mismatch is much smaller for the
cementite particles than that for the martensite islands. Consequently, although the
equivalent plastic strain is the same at the interface of particle/island-matrix, it
decreases more rapidly with the distance from the interface for the case of
cementite particles. Because of the precipitation strengthening caused by small
particles, the steel containing cementite particles possesses a higher Yo. Thus, the
minimum volume fraction required for the percolation of the internal stress
induced plastic zones to the entire microstructure (fmin) is higher for the ferrite-
cementite steel when compared with the ferrite-martensite steel.
2. Due to the presence of GNDs, yield drop was not observed in the stress-strain
curve of the specimen consisting of cementite particles in a ferrite matrix, although
a fairly large yield point elongation was still present. In this specimen, the volume
fraction of cementite particles was 2.3% whereas fmin was determined by the
calculation to be 10.23%. The formation 6.2% martensite in the steel
microstructure resulted in the elimination of yield point elongation in the specimen
B1, although a clear yield point was observable. The value of fmin in this specimen
was calculated to be 9.667%. Because of having the martensite volume fractions
greater than corresponding fmin, the specimens B2 and C exhibited a completely
continuous yielding behavior. Consequently, it was concluded that the percolation
of internal stress induced plastic zone to the whole microstructure is a requisite for
the occurrence of continuous yielding behavior. To achieve this condition, the
volume fraction of second phase must be greater than fmin.
3. Since small cementite particles cause precipitation strengthening and
consequently, increase of Yo, fmin is always higher than the volume fraction of
cementite particles in steels. Thus, because cementite particles are very small,
continuous yielding cannot be observed in steels with ferrite-cementite
microstructure.
4. In steels with a greater volume fraction of cementite phase (f), it is possible to
find a particle size for which f ≥ fmin and consequently, it is expected that a
continuous yielding behavior occurs.
5. In low-carbon ferrite-cementite steels, the formation of a certain amount of
martensite phase in the microstructure besides the cementite phase could eliminate
the yield point phenomenon.
Reference
[1] G.E. Dieter, D. Bacon, Mechanical Metallurgy, McGraw-Hill, 1988.

[2] M.A. Meyers, K.K. Chawla, Mechanical metallurgy: principles and


applications, Prentice-Hall, 1984.
[3] D. Johnson, M. Edwards, P. Chard-Tuckey, Materials Science and Engineering:
A, 625 (2015) 36-45.

[4] S. Nagarajan, R. Narayanaswamy, V. Balasubramaniam, Journal of materials


engineering and performance, 22 (2013) 3085-3092.

[5] S. Nagarajan, N. Raghu, B. Venkatraman, Materials Science and Engineering:


A, 561 (2013) 203-211.

[6] R. Schwab, V. Ruff, Acta Materialia, 61 (2013) 1798-1808.


[7] N. Tsuchida, Y. Tomota, K. Nagai, K. Fukaura, Scripta materialia, 54 (2006)
57-60.

[8] M.R. Wenman, P.R. Chard-Tuckey, International Journal of Plasticity, 26


(2010) 1013-1028.
[9] F. Yoshida, Y. Kaneda, S. Yamamoto, International Journal of Plasticity, 24
(2008) 1792-1818.

[10] J. Zhang, Y. Jiang, International Journal of Plasticity, 21 (2005) 651-670.

[11] J. Kadkhodapour, S. Schmauder, D. Raabe, S. Ziaei-Rad, U. Weber, M.


Calcagnotto, Acta Materialia, 59 (2011) 4387-4394.
[12] A. Ramazani, K. Mukherjee, A. Schwedt, P. Goravanchi, U. Prahl, W. Bleck,
International Journal of Plasticity, 43 (2013) 128-152.
[13] T. Sakaki, K. Sugimoto, T. Fukuzato, Acta Metallurgica, 31 (1983) 1737-
1746.

[14] T. Gladman, I. McIvor, F. Pickering, J. Iron Steel Inst., 210 (1972) 916-930.

[15] J. Moyer, G. Ansell, Metallurgical Transactions A, 6 (1975) 1785-1791.


[16] T.W. Clyne, P.J. Withers, An Introduction to Metal Matrix Composites,
Cambridge University Press, 1995.
[17] C.G. de Andrés, F. Caballero, C. Capdevila, H. Bhadeshia, Scripta Materialia,
39 (1998) 791-796.

[18] J.H. Kim, T.S. Byun, D.T. Hoelzer, C.H. Park, J.T. Yeom, J.K. Hong,
Materials Science and Engineering: A, 559 (2013) 111-118.

[19] W.C. Leslie, Metallurgical transactions, 3 (1972) 5-26.


[20] W.C. Leslie, The physical metallurgy of steels, Hempisphere Pub. Corp.,
1981.
[21] S. Takeuchi, Journal of the Physical Society of Japan, 27 (1969) 929-940.

[22] M. Delincé, Y. Bréchet, J.D. Embury, M. Geers, P. Jacques, T. Pardoen, Acta


Materialia, 55 (2007) 2337-2350.

[23] G. Krauss, Steels: heat treatment and processing principles, ASM


International, 1990.
[24] A. Ohmori, S. Torizuka, K. Nagai, ISIJ international, 44 (2004) 1063-1071.
[25] H. Azizi-Alizamini, M. Militzer, W. Poole, Scripta Materialia, 57 (2007)
1065-1068.
[26] T.S. Wang, F.C. Zhang, M. Zhang, B. Lv, Materials Science and Engineering:
A, 485 (2008) 456-460.

[27] M. Nikolussi, Universitätsbibliothek der Universität Stuttgart, Stuttgart, 2009.

[28] D. Tjahjanto, S. Turteltaub, A. Suiker, S. Van der Zwaag, Modelling and


Simulation in Materials Science and Engineering, 14 (2006) 617.

Table 1 Chemical composition of the steel used in this study in weight percent.

Fe C M Si P S Cr Ni Mo Al Cu Co Ti Nb V W
n
Balanc 0.16 1.1 0.41 0.01 0.0 0.03 0.06 0.00 0.04 0.06 0.00 0.00 0.00 0.00 0.02
ed 5 5 1 8 1 5 6 8 2 2 1 1 1 1 8

Table 2 Martensite volume fraction (Vm), cementite volume fraction (Vc), martensite island size
(dm) and cementite particle size (dc) in the studied specimens.

Nomenclature Vm (%) Vc (%) dm (μm) dc (nm)


F - 2.3 - 150
B1 6.2 1.8 1.73 150
B2 13.1 1.2 2.25 150
C 39 - 4.15 -

Table 3 Stiffness constants of cementite (CI) and ferrite (CM) as well as Eshelby S tensor used in
the calculation of this study.

CI (GPa)a C1111=385, C2222=341, C3333=316, C2323=13, C1313=131, C1212=131


C1122=C2211=157
C1133=C3311=162
C2233=C3322=167
CM (GPa)b C1111=C2222=C3333=233.5
C2323=C1313=C1212=118
C1122=C2211=C1133=C3311=C2233=C3322=135.5
Sc S1111=S2222=S3333=0.523
S2323=S3232=S3223=S2332=S2323=S1313=S1331=S3113=S3131=S1212=S1221=S2112=S2121=0.238
S1122=S2211=S1133=S3311=S2233=S3322=0.0476
a
[27]
b
[28]
c
[16]
(a)

(b)

(c)
(d)

Figure 1. SEM micrographs of the (a) specimen F (ferrite-2.3% cementite), (b) specimen B1
(ferrite-1.8% cementite-6.2% martensite), (c) specimen B2 (ferrite-1.2% cementite-13.1
martensite) and (d) specimen C (ferrite-39% martensite). “F”, “M” and “C” represent ferrite,
martensite and cementite, respectively.

500
Engineering Stress (MPa)

400

300

200

100
Microstructure:
ferrite-2.3% cementite
0
0.00 0.05 0.10 0.15 0.20 0.25
Engineering Strain

(a)
600

Engineering Stress (MPa)


500

400

300

200

100 Microstructure:
ferrite-1.8% cementite-6.2% martensite
0
0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16
Engineering Strain

(b)

700

600
Engineering Stress (MPa)

500

400

300

200

Microstructure:
100
ferrite-1.2% cementite-13.1% martensite
0
0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16
Engineering Strain

(c)
1000

Engineering Stress (MPa)


800

600

400

200 Microstructure:
ferrite-39% martensite
0
0.00 0.02 0.04 0.06 0.08
Engineering Strain

(d)

450

400
Engineering Stress (MPa)

350

300

250

200

150

100

50 mostly ferritic steel


0
0.0 0.1 0.2 0.3 0.4
Engineering Strain

(e)

Figure 2. Results of the tensile tests on a) specimen F (ferrite-2.3% cementite), b) specimen B1


(ferrite-1.8% cementite-6.2% martensite), c) specimen B2 (ferrite-1.2% cementite-13.1%
martensite), d) specimen C (ferrite-39% martensite), and e) steel containing 0.09 %wt carbon
considered as ferrite.
500

Engineering Stress (MPa)


400

300

200

100
Sample F (ferrite-cementite microstructure)
Ferritic steel
0
0.000 0.005 0.010 0.015 0.020 0.025 0.030 0.035 0.040
Engineering Strain

Figure 3. Comparison of the extent of yield point elongation of specimen F (ferrite-2.3%


cementite) with ferrite.

50000
Sample C (ferrite-39% martensite)
45000 Sample B1 (ferrite-1.8% cementite-6.2% martensite)
Work Hardening Rate (MPa)

40000

35000

30000

25000

20000

15000

10000

5000

0
0.00 0.02 0.04 0.06 0.08 0.10 0.12
True Strain

Figure 4. Work hardening rate as a function of strain for the samples B1 (ferrite-1.8% cementite-
6.2% martensite) and C (ferrite-39% martensite). The stress-strain curves of these steels are
shown in Figures 2-b (for sample B1) and 2-d (for sample C).
0.020 aF aB 1
aB2 Sample F
aC Sample B1

Equivalent Plastic Strain


Sample B2
0.015
Sample C

0.010

0.005

0.000
0.00 0.05 0.10 0.15 1 2 3 4 5

r (m)

Figure 5. The distribution of the plastic strain (ε(r)) for the studied specimens.

aF
300 aB1 Sample F
aB2 Sample B1
aC
Sample B2
Equivalent Stress (MPa)

250
Sample C
200

150

100

50

0
0.0 0.1 0.2 0.3 0.4 0.5 1 2 3 4 5 6 7 8 9 10

r (m)

Figure . The distribution of the e uivalent stresses (σ) for the studied specimens.
X3
r

θ
R X2
particle

Figure 7. Schematic illustration showing the angle θ used in the calculations. with the
assumption that the tensile stress acts in the x3 direction, θ is the angle from the x1-x2 plane to the
r direction.

600 Sample F
Y/Y0=1.4
Sample B1 Y/Y0=1.6
500 Sample B2
Y/Y0=1.6
Sample C
Yield Stress (MPa)

400 Y/Y0=1.8

300

200

100

0
0 20 40 60 80 100

Theta (deg)

Figure 8. The yield stress of the plastic zone (σA) as a function of θ for the studied specimens.

You might also like