You are on page 1of 8

Materials Science & Engineering A 575 (2013) 119–126

Contents lists available at SciVerse ScienceDirect

Materials Science & Engineering A


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

Applicability of the one-internal-variable Kocks–Mecking approach for tensile


flow and work hardening behaviour of modified 9Cr–1Mo steel
J. Christopher, B.K. Choudhary n, M.D. Mathew, T. Jayakumar
Metallurgy and Materials Group, Indira Gandhi Centre for Atomic Research, Kalpakkam 603102, Tamil Nadu, India

art ic l e i nf o a b s t r a c t

Article history: Tensile flow and work hardening behaviour of modified 9Cr–1Mo steel in normalised and tempered
Received 17 January 2013 (N+T) and additional three different simulated post-weld heat treatments (PWHTs of 993, 1013 and
Accepted 15 March 2013 1033 K for 1 h) conditions have been examined in the framework of the one-internal-variable Kocks–
Available online 28 March 2013
Mecking (K–M) approach in the temperature range 300–873 K. The steel exhibited two-stage work
Keywords: hardening behaviour characterised by a rapid decrease in instantaneous work hardening rate, θ (θ ¼dsp/
Modified 9Cr–1Mo steel dεp, where sp is the plastic flow stress component and εp is the true plastic strain) with sp at low stresses
Post-weld heat treatment (transient stage) followed by a gradual and linear decrease at high stresses (stage-III). The variations in
Kocks–Mecking approach work hardening parameters with temperature exhibited three distinct temperature regimes. The
Work hardening behaviour
influence of PWHT is reflected in systematic variations in work hardening parameters with increasing
PWHT temperature from 993 to 1033 K due to microstructural softening.
& 2013 Elsevier B.V. All rights reserved.

1. Introduction one-parameter model of the Kocks–Mecking approach [7,8], the


mean free path of gliding dislocations is assumed to be solely
Appropriate description of tensile flow and work hardening determined by localised obstacles associated with dislocations and
behaviour of metals and alloys is one of the most important areas evolution of average dislocation density to be the governing
of materials research in view of improving the conditions for parameter defining the course of deformation. In the K–M
material processing and for ensuring safe performance during approach, the evolution of the average dislocation density (ρ) is
service. Tensile flow and work hardening behaviour of metals and expressed as the competition between a storage term and an
alloys has been described using one of the several empirical annihilation term in the following equation:
relations proposed in the literatures such as Hollomon [1], Ludwik  þ 
dρ dρ dρ−
[2], Swift [3], Voce [4,5] and Ludwigson [6] relationships. In these ¼M − ; ð1Þ
dεp dεp dεp
relationships, the flow stress is related to the plastic strain εp
(which is not a physical state variable) as an external variable and where M is the Taylor factor. Eq. (1) can also be expressed using
involving relevant additional term defining the work hardening dislocation accumulation or hardening parameter, k1 (the disloca-
characteristics of the materials. Among these flow relationships, tion accumulation term dρ+/dεp is given as dρ+/dεp ¼k1ρ1/2)
Voce equation [4,5] has attracted more attention in view of and dislocation annihilation due to dynamic recovery, k2 para-
the prediction of saturation stress close to the experimentally meter (the dislocation annihilation term dρ−/dεp is given as
observed value of ultimate tensile strength from some initial stress dρ−/dεp ¼k2ρ) as
value and the sound interpretation subsequently provided by the dρ
Kocks–Mecking approach [7–10]. In the Kocks–Mecking (K–M) ¼ M ðk1 ρ1=2 −k2 ρÞ: ð2Þ
dεp
phenomenological approach [7,8], the evolution of dislocation
structure with strain at constant strain rate is assumed as a single In the K–M approach, the plastic flow stress component (sp)
structural parameter responsible for plastic flow. The work hard- is linked to the dislocation density [11–13] following Taylor
ening is controlled by the competition between storage and equation as
annihilation (rearrangement) of dislocations, which are assumed sp ¼ Mαμbρ1=2 ; ð3Þ
to superimpose in an additive manner. In this simplified
where α is a constant, μ is the shear modulus and b is Burger's
vector. The plastic contribution of the flow stress is experimentally
n
Corresponding author. Tel.: +91 44 27480118; fax: +91 44 27480075,
defined as sp ¼s−sY, where s is the total flow stress and sY is
+91 44 27480118. yield stress component of the flow stress. The work hardening rate
E-mail address: bkc@igcar.gov.in (B.K. Choudhary). due to the evolution of dislocation density with strain can be

0921-5093/$ - see front matter & 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.msea.2013.03.044
120 J. Christopher et al. / Materials Science & Engineering A 575 (2013) 119–126

derived from the differentiation of Eq. (3), and by the substitution Detailed investigation on the influence of PWHT on room and
of Eq. (2), work hardening relationship is obtained as elevated temperatures tensile properties of modified 9Cr–1Mo
steel base metal in normalised and tempered (N+T) condition
dsp M 2 αμbk1 Mk2 sp
¼θ¼ − ; ð4Þ indicated a systematic decrease in both yield and ultimate tensile
dεp 2 2
strength values with increasing PWHT temperature [18]. An
where θ ¼dsp/dεp is the work hardening rate. On rearrangement examination on the applicability of appropriate flow relationship
and by substituting the plastic component of saturation stress, i.e., in the normalised and tempered and with additional PWHT
sp,sat ¼Mαμbk1/k2, the work hardening relation describing stage-III conditions suggested that a combination of Ludwigson and Hollo-
hardening is expressed as mon relations describe true stress–true plastic strain behaviour
  most accurately [19]. The Ludwigson equation follows closely s–ε
sp
θ ¼ θ0 1− ; ð5Þ data at room and intermediate temperatures, whereas at high
sp;sat
temperatures, Ludwigson equation reduces to Hollomon equation.
where θ0 ¼ M2αμbk1/2 is the initial work hardening rate at plastic Alternately, Voce equation as a single relationship can provide an
flow stress, sp ¼s−sY ¼ 0. The applicability of Eq. (5) has been adequate description of s–ε behaviour for the temperature range
successfully demonstrated for understanding the deformation and of interest, i.e., 300–873 K [19,20]. In this paper, work hardening
work hardening characteristics of materials ranging from single behaviour of modified 9Cr–1Mo steel in N+T and PWHT conditions
and polycrystalline pure face-centred cubic metals and alloys has been examined in terms of the variation of instantaneous
[7–10,13] to complex ferritic/martensitic 7–9% Cr steels [11,12,14]. work hardening rate (θ) with total true stress (s) and plastic flow
Modified 9Cr–1Mo steel is an important high temperature stress (sp) in the framework of the one-internal-variable Kock–
material for steam generator (SG) applications in fossil-fired Mecking approach [7–10] in the temperature range 300–873 K.
thermal and nuclear power plants. The choice of modified 9Cr– The influence of additional three different PWHTs and tempera-
1Mo steel for steam generator applications in sodium cooled fast tures on the flow and work hardening parameters associated with
reactors (SFRs) is based on the low thermal expansion coefficient the K–M approach has also been discussed.
and high resistance to stress corrosion cracking in water-steam
systems compared to austenitic stainless steels, in addition to
better mechanical properties at elevated temperatures than the
alternate 2.25Cr–1Mo and plain 9Cr–1Mo steels [15,16]. The steel
also offers a good combination of high creep strength and ductility 2. Experimental details and analysis of data
over long exposures at high temperatures. The good weldability
and microstructural stability over long exposures at elevated 2.1. Material, heat treatment and tensile testing
temperatures are other attractive features for the steel. Further-
more, the fabrication of steam generator in monometallic material Modified 9Cr–1Mo steel of 12 mm thick plates in normalised
enhances the reliability of the components including critical tube and tempered condition was used in this investigation. The
to tube plate welds. Welding, which remains a major joining chemical composition of the steel is given in Table 1. The plates
process for the manufacture of steam generator components of were normalised at 1333 K for 25 min and tempered at 1023 K for
SFRs, strongly influences the microstructure and properties of the 1 h. In order to simulate PWHTs, additional heat treatments of
weld joint. During weld thermal cycle, a narrow heat affected zone 993 K/1 h, 1013 K/1 h and 1033 K/1 h were given to 12  12 
(HAZ) consisting of coarse grain, fine grain, partially transformed 60 mm3 specimen blanks machined from the plate with stress
or intercritical HAZ and over tempered regions are formed in base axis along the rolling direction. The rates of heating and cooling of
metal on either side of the weld metal due to large temperature 200 K/1 h were used during all the post-weld heat treatments. The
gradients. This heterogeneous distribution of microstructure in the Vickers hardness measurements were performed on N+T and
HAZ leads to large gradient in the strength values across the weld PWHT specimens employing a load value of 10 kg. Microstructural
joints. The presence of a soft intercritical HAZ is often regarded as examinations were performed on N+T and PWHT specimens
a weak link and a life limiting region of a weld joint due to its prepared by the standard metallographic technique and by etching
inferior creep strength [17]. The inferior creep strength of the in Villela's reagent (1 g picric acid+5 ml HCl+100 ml ethyl alcohol)
intercritical HAZ results from the preferential accumulation of using a Philips XL30 scanning electron microscope (SEM).
creep deformation coupled with extensive creep cavitation due to Button head cylindrical specimens of 26 mm gauge length and
micro-mechanistic effects of constrained deformation and devel- 4 mm gauge diameter were machined from the specimen blanks
opment of damage commonly known as type IV cracking [17]. in N+T and PWHT conditions. Tensile tests were carried out in air
Apart from the heterogeneous microstructure across HAZ, a large using a floor model Instron 1195 universal testing machine
amount of residual stresses generated due to the weld thermal equipped with a three-zone temperature control furnace and a
cycle adds to strength gradient in the weld joint. In order to relieve stepped-load suppression unit. Tests were performed on N+T and
the residual stresses and minimise the strength gradient across PWHT specimens over a temperature range 300-873 K at a
the weld joint, a suitable post-weld heat treatment (PWHT) is nominal strain rate of 1.26  10−3 s−1. The temperatures in all the
prescribed and practised. It is known that the PWHT modifies the tests were controlled within 72 K. Load–elongation curves were
microstructure, which in turn affects the mechanical properties of recorded using the Instron autographic recorder. Use of suitable
the base metal that is previously unaffected by the weld chart speed and stepped zero suppression gave strain resolution of
thermal cycle. 7.5  10−4 and stress resolution of 0.80 MPa, respectively.

Table 1
Chemical composition (wt%) of modified 9Cr–1Mo steel.

Element C Cr Mo Si Mn P S Nb V N Fe

Amount 0.097 9.29 0.92 0.31 0.37 0.018 0.005 0.08 0.26 0.057 Balance
J. Christopher et al. / Materials Science & Engineering A 575 (2013) 119–126 121

2.2. Analysis of data 3. Results and discussion

The load–elongation curves at room and high temperatures 3.1. Microstructure and hardness
were smooth, whereas at intermediate temperatures (523–673 K)
jerky flow was observed in N+T and PWHT conditions. An The microstructures of modified 9Cr–1Mo steel in N+T and
examination of the jerky flow at different temperatures in the different PWHT conditions are shown in Fig. 1. The microstructure
range 523–673 K revealed that these serrations could be classified in N+T condition is composed of tempered lath martensite with
into types A–C serrations [18]. A uniform approach based on the fine precipitates of M23C6 carbides decorated along prior austenite
classification of types of serrations was adopted for measuring grain and martensite lath boundaries. The matrix regions con-
load values from the load–elongation curves showing serrated tained fine MX type V(C,N) and Nb(C,N) precipitates (Fig. 1a).
flow [18,21]. For type A serrations, which show an abrupt rise in PWHT specimens exhibited coarse precipitates both at the grain
the loads followed by discontinuous drops to or below the general and lath boundaries and in the matrix regions (Fig. 1b–d).
level of load–elongation curves, general levels of load values were A systematic coarsening of the precipitates resulting in the
considered. For type B serrations that oscillated about the general decrease in number density of precipitates with increasing PWHT
level of load values, mean load values were taken. For type C temperatures can be seen in Fig. 1. The variation in hardness with
serrations characterised by load drops always below the general the introduction of PWHT is shown in Fig. 2. A rapid decrease in
level of load values, envelope load values were taken for stress– the hardness can be seen for the specimen subjected to PWHT of
strain data. Typical averaging scheme of load values in the load– 993 K/1 h followed by a marginal decrease with increasing PWHT
elongation curves exhibiting types A–C serrations can be seen in temperature from 993 to 1033 K. The observed decrease in the
Refs. [22,23] for plain 9Cr–1Mo steel as an example. The true hardness following PWHT results from the microstructural soft-
stress–true plastic strain data were obtained using a computer ening in terms of precipitate coarsening and overall decrease in
programme from the digitised load–elongation data up to the the number density of the precipitates, as observed in Fig. 1.
maximum load values from the load–elongation curves. Since no
strain gauge was employed, the cross head displacement was 3.2. Influence of test temperature and PWHT on tensile flow and
taken as the specimen extension. The linear elastic portion of the work hardening behaviour
load–elongation data was contributed by the specimen, machine
frame, and load–train assembly. The combination of these elastic Typical variations in flow stress as a function of true plastic
elongations was subtracted from the total elongation for the strain at different temperatures as double logarithmic plots are
estimation of plastic strain. Stress and plastic strain data were shown in Fig. 3 for modified 9Cr–1Mo steel subjected to PWHT of
used to determine true stress (s)–true plastic strain (εp). The 993 K for 1 h as an example. At all the strains, a general decrease in
instantaneous work hardening rate with flow stress or plastic flow stress with increase in temperature in the range 300–873 K
component of flow stress was calculated by centred difference has been observed for the steel in N+T and PWHT conditions.
method as A substantial decrease in the flow stress is seen with increase in
    temperature from 300 to 473 K. At intermediate temperatures
sðiþ1Þ −sðiÞ sðiÞ −sði−1Þ spðiþ1Þ −spðiÞ spðiÞ −spði−1Þ (473–723 K), marginal decrease in flow stress with increasing
θ ¼ 1=2 þ ¼ 1=2 þ ;
εpðiþ1Þ −εpðiÞ εpðiÞ −εpði−1Þ εpðiþ1Þ −εpðiÞ εpðiÞ −εpði−1Þ temperature is observed at all strains, and the s–εp data appear
ð6Þ to lie in a narrow band. At high temperatures (above 723 K), flow
stress decreases rapidly with increasing temperature. Apart from
where subscript i refers to particular data point of flow stress (s) the decrease in flow stress, a decrease in the true uniform plastic
or plastic component of flow stress (sp) and plastic strain (εp). strain with increasing temperature from 300 to 473 K followed by

Fig. 1. Microstructures of modified 9Cr–1Mo steel in (a) N+T condition and additional PWHTs of (b) 993 K/1 h, (c) 1013 K/1 h and (d) 1033 K/1 h.
122 J. Christopher et al. / Materials Science & Engineering A 575 (2013) 119–126

Fig. 4. Variations of instantaneous work hardening rate (θ) as a function of true


stress (s) at different temperatures for post-weld heat treatment of 993 K for 1 h in
Fig. 2. Variations in hardness with respect to post-weld heat treatments of 993 K/1 h, modified 9Cr–1Mo steel.
1013 K/1 h and 1033 K/1 h.

Fig. 3. Double logarithmic plots of true stress (s)–true plastic strain (εp) at different
temperatures for post-weld heat treatment of 993 K for 1 h in modified 9Cr– Fig. 5. Double logarithmic plots of true stress (s) vs. true plastic strain (εp) at 523 K
1Mo steel. for N+T and PWHT conditions in modified 9Cr–1Mo steel.

a marginal decrease in the values at intermediate temperatures towards origin with increase in temperature is observed. Like s vs. εp,
and a drastic reduction at high temperatures are observed. The θ vs. s plots also exhibited three distinct temperature regimes of
variations in the flow stress and the true uniform plastic strain room, intermediate and high temperatures. At intermediate tem-
values with respect to temperature explicitly suggest three distinct peratures (473–723 K), θ vs. s data points fall into narrow band and
temperature regimes of room, intermediate and high tempera- only marginal variations with respect to temperature are seen. At
tures for s–εp behaviour in modified 9Cr–1Mo steel. high temperatures (T4723 K), rapid shift in θ vs. s plots to lower
Following the K–M approach, work hardening behaviour of stresses is observed (Fig. 4). Further, the influence of increasing
modified 9Cr–1Mo steel in N+T and PWHT conditions has been temperature is reflected in a general increase in the rates at which θ
examined in terms of the variations of instantaneous work hardening decreases with increasing s particularly in stage-III. This can be
rate, θ (θ¼ds/dε) with true stress (s) at different temperatures. observed from the increasing slopes of θ–s plots in stage-III with
Representative θ–s plots at different temperatures are shown in increasing temperature (Fig. 4). The observed two-stage work hard-
Fig. 4 as an example for the steel post weld heat treated at 993 K for ening behaviour in θ–s is in agreement with those reported for 9% Cr
1 h. In all the heat treatment conditions, θ vs. s exhibited two-stage steels [22–26]. Further, three distinct temperature regimes observed
work hardening behaviour characterised by a rapid decrease in θ in s–εp and θ–s behaviour in modified 9Cr–1Mo steel compare
with s at low stresses followed by a gradual decrease in θ with s at favourably with those observed for plain 9Cr–1Mo steel [22,23,26].
high stresses. The two-stage work hardening is shown for 300 K in The influence of PWHTs on flow and work hardening behaviour
Fig. 4. The rapid decrease in θ at low stresses is designated as of N+T modified 9Cr–1Mo steel has been examined in the
transient stage (TS) and the gradual linear decrease in θ at high temperature range 300–873 K. At all the temperatures, a systema-
stresses represents typical stage-III work hardening governed by the tic decrease in flow stress values with increasing PWHT tempera-
K–M approach. The observed transient stage is different from stage-I ture has been observed. The variations of true stress–true plastic
in single crystal, in which the value of θ is much lower than that in strain with respect to PWHT temperature at 523 K is shown as an
stage-III. A general decrease in θ vs. s to low stresses, i.e., shift example in Fig. 5. At all the strains, true stress values exhibited a
J. Christopher et al. / Materials Science & Engineering A 575 (2013) 119–126 123

decrease with increasing PWHT temperature in the range


993–1033 K compared to that observed for N+T condition. At all
the temperatures, like s vs. εp, the influence of PWHT is also
reflected in a systematic shift in θ vs. s towards lower stresses with
increasing PWHT temperature. The θ vs. s plots at 523 K in N+T
and PWHT conditions exhibiting two-stage work hardening beha-
viour and a shift in θ–s towards low stresses with increasing
PWHT temperature is shown in Fig. 6 as an example. The
decreased flow stress values due to microstructural softening with
increasing PWHT temperatures (Fig. 1) is reflected in a shift in θ–s
towards lower stresses.

3.3. Variations of work hardening parameters associated with the


K–M approach with temperature and PWHT

In order to investigate the influence of decrease in flow stress


associated with microstructural softening due to PWHT on the
work hardening behaviour of modified 9Cr–1Mo steel, analysis of
work hardening in the framework of the K–M approach [7–10,13] Fig. 7. Variations of instantaneous work hardening rate (θ) as a function of plastic
was performed in terms of variations in θ with plastic flow stress component of flow stress (sp) at 473 K in modified 9Cr–1Mo steel for PWHT of
(sp) component defined as sp ¼s−sY, where sY is the true yield 993 K for 1 h. Best fit line for stage-III hardening obeying the K–M approach has
been shown as broken line.
stress at offset plastic strain of ε ¼0.001. The values of sY have been
determined either from the experimental s–ε data or from the
extrapolated flow stress values at ε ¼ 0.001 at all the test condi-
tions. Representative θ–sp plot at 473 K for PWHT of 993 K for 1 h
and fitted with the K–M equation, i.e., Eq. (5) is shown in Fig. 7 as
an example. A good fit of the K–M work hardening relation for
stage-III is shown by broken line in Fig. 7. The initial work
hardening rate θ0 is obtained as the intercept on the y-axis at
sp ¼s−sY ¼0 and the plastic component of saturation stress sp,sat is
evaluated as the intercept on the x-axis at θ¼ 0. The slope of θ–sp
plot in stage-III denoted by nV,K–M defines the rate at which work
hardening rate decreases with increasing stress. The slope of θ–sp
plot in stage-III can also be expressed in terms of dynamic
recovery k2 parameter as nV,K–M ¼−Mk2/2.
The temperature dependence of true yield stress sY for mod-
ified 9Cr–1Mo steel in N+T and PWHT conditions is shown in
Fig. 8. In order to bring out the influence of temperature alone on
the yield stress, the shear modulus (μ) compensated true yield
stress, i.e. sY/μ has been used. The shear modulus is obtained as
μ¼E/2(1+υ), where υ is Poisson's ratio and temperature dependent
of Young's modulus (E) values have been taken from French
nuclear design code RCC-MR for the steel [27]. A marginal Fig. 8. Variations of normalised true yield strength (sY/μ) with temperature for N
+T and PWHT conditions in modified 9Cr–1Mo steel.

decrease in sY/μ from 300 K to 473 K followed by well defined


plateaus at intermediate temperatures and a rapid decrease of sY/μ
at high temperatures can be seen in Fig. 8 for N+T and PWHT
conditions. At all the temperatures, a systematic decrease in true
yield stress values has been observed with increasing PHWT
temperatures of 993, 1013 and 1033 K. The reduction in true yield
stress is more pronounced for higher PWHT temperatures of
1013 K and 1033 K compared to those observed for N+T and
993 K/1 h PWHT conditions. The variations of modulus compen-
sated plastic flow stress sp,sat/μ with temperature for N+T and
PWHT conditions are shown in Fig. 9. Like sY/μ, the variations in
sp,sat/μ with temperature also exhibited three distinct temperature
regimes as a gradual decrease from 300 to 473 K followed by
plateaus at intermediate temperatures (523–673 K) and a rapid
decrease at high temperatures. Fig. 9 also demonstrates a sys-
tematic increase in plastic flow stress sp,sat/μ in the temperature
range 300–873 K with the introduction of increasing PWHT
Fig. 6. Variations of instantaneous work hardening rate (θ) as a function of true temperature. The increase in sp,sat/μ with increasing PWHT tem-
stress (s) at 523 K for N+T and PWHT conditions in modified 9Cr–1Mo steel. perature can arise from the dominance of improved contribution
124 J. Christopher et al. / Materials Science & Engineering A 575 (2013) 119–126

Fig. 9. Variations of normalised plastic component of saturation stress (sp,sat/μ) Fig. 11. Variations of θ0/μ with temperature for N+T and PWHT conditions in
with temperature for N+T and PWHT conditions in modified 9Cr–1Mo steel. modified 9Cr–1Mo steel.

Fig. 10. Variations of nV,K–M with temperature for N+T and PWHT conditions in Fig. 12. Variations of K2/K1 with temperature for N+T and PWHT conditions in
modified 9Cr–1Mo steel.
modified 9Cr–1Mo steel.

of tensile work hardening and gain in the plastic flow stress due to marginal with increasing PWHT temperature. At high tempera-
microstructural softening in PWHT conditions. tures, significant and systematic decrease in nV,K–M and θ0/μ values
The slope of θ–sp plot in stage-III, i.e., nV,K–M (proportional to with increasing PWHT temperature is seen.
dynamic recovery k2 parameter) exhibited a marginal increase in The occurrence of saturation stress or steady state stress has
nV,K–M from room to intermediate temperatures followed by a been interpreted as a state of dynamic equilibrium between
rapid increase at high temperatures in N+T and PWHT conditions generation of dislocations and their annihilation by dynamic
(Fig. 10). Like nV,K–M, modulus compensated initial work hardening recovery during deformation process [7–10]. A decrease in satura-
rate, i.e., θ0/μ (proportional to hardening parameter k1) also tion stress with increasing temperature implies that steady state is
exhibited a marginal increase in θ0/μ from room to intermediate reached at lower stresses due to the dominance of dynamic
temperatures followed by a rapid increase at high temperatures recovery. The occurrence of plateaus in true yield stress (sY/μ),
(Fig. 11). The variations in the ratio of dynamic recovery parameter saturation stress (sp,sat/μ) and the ratio of dynamic recovery
and hardening parameter, k2/k1 (evaluated as k2/k1 ¼Mαμb/sp,sat) parameter and hardening parameter (k2/k1) and the marginal
with temperature displayed a marginal increase in the value from variations in nV,K–M and θ0/μ along with the grouping of s–ε and
300 to 473 K followed by plateaus at intermediate temperatures θ–s data indicate anomalous flow and work hardening behaviour
and rapid increase at high temperatures (Fig. 12). A marginal at intermediate temperatures. This anomalous behaviour can be
decrease in k2/k1 with respect to increasing PWHT temperature is attributed to dynamic strain ageing (DSA) as manifested by the
noticed in the temperature range 300–873 K. Rapid increase in occurrence of serrated flow at intermediate temperatures in
k2/k1 with increasing temperature indicates the dominance of the steel [18]. Based on the measurement of activation energy
dynamic recovery at high temperatures. This is also reflected in for the onset of serrated flow, it has been suggested that diffusion
the rapid increase in the rate parameters nV,K–M and θ0/μ at high of interstitial solutes such as carbon to be responsible for DSA in
temperatures (Figs. 10 and 11). The influence of PWHT on nV,K–M 9% Cr steels [18,28–31]. It has been demonstrated that DSA causes
and θ0/μ at room and intermediate temperatures is observed to be an increased rate of multiplication of dislocations and delay in
J. Christopher et al. / Materials Science & Engineering A 575 (2013) 119–126 125

recovery of dislocation structure due to reduced dislocation microstructural softening associated with coarsening of precipi-
mobility and reduced propensity to crossslip [29,32–36]. These tates resulting in the overall reduction in the number density of
investigations indicate reduced dynamic recovery in the DSA precipitates with increasing PWHT temperature (Fig. 1). In addi-
temperature regime. Contrary to this, rapid decrease in sY/μ and tion to this, recovery of dislocation substructure with increasing
sp,sat/μ and a rapid increase in absolute nV,K–M, θ0/μ and k2/k1 with PWHT temperature from 993 to 1033 K is expected to add into the
increasing temperature indicate acceleration of recovery processes reduction in strength values of the steel. Contrary to this, the
at high temperatures. It has been suggested that the dynamic plastic flow stress sp,sat exhibited an increase in the values with
recovery parameter k2 or nV,K–M determines the rate at which increasing PWHT temperature (Fig. 9) due to increased work
stress tends to attain a steady state value [9]. Therefore, the value hardening contribution in PWHT specimens. The improved work
of nV,K–M is invariably associated with rate controlling deformation hardening in PWHT conditions is derived from its ability to achieve
mechanism. It has been also suggested that at room and inter- higher work hardening due to softened microstructures. Higher
mediate temperatures, where crossslip is dominant, nV,K–M is work hardening due to reduced dynamic recovery in the softened
expected to have a low absolute value. On the other hand, a microstructures in PWHT conditions is reflected in the lower k2/k1
change in controlling mechanism from cross-slip to dislocation ratio than that observed for N+T condition (Fig. 12). In N+T
climb and subboundary migration results in high nV,K–M value at condition, the dominance of dynamic recovery as indicated by
high temperatures. A rapid increase in nV,K–M at high temperatures higher k2/k1 ratio leads to lower sp,sat values than those in PWHT
observed for N+T and PWHT conditions indicates the dominance conditions at all temperatures. Since, θ0 is inversely related to the
of dislocation climb and subboundary migration during high mean free path of dislocations [9], the observed decrease in θ0/μ
temperature deformation in modified 9Cr–1Mo steel. The dom- with increasing PWHT temperature can be ascribed to the
inance of dynamic recovery is also reflected in the rapid increase decrease in number density of precipitates and dislocation density
in k2/k1 at high temperatures. A comment on the variations in with increasing PWHT temperatures. The dynamic recovery in FCC
work hardening parameters at room temperature and in the DSA metals has been interpreted in terms of the components of
temperature regime is in order. Marginal difference in the respec- dynamic recovery parameters arising from spontaneous annihila-
tive work hardening parameters with increasing temperature tion of dipoles or by the annihilation arising from dipole climb and
obtained at room and intermediate temperatures suggests that its direct proportionality to θ0 [37]. Since, θ0 and nV,K–M are
the dislocation substructural behaviour in its totality is not very interrelated, the decrease in θ0/μ is reflected in a systematic
different in the two temperature regimes. Therefore, as suggested decrease in nV,K–M with increasing PWHT temperature. All these
[25,26], long range stress field due to multiplication of dislocations observations indicate that the work hardening parameters asso-
and their interaction is expected to play a major role at room and ciated with the K–M approach such as θ0/μ, nV,K–M, k2/k1 and sp,sat/μ
intermediate temperatures. This is also reflected in the normalised are sensitive to initial microstructures.
plots of θ/μ vs. sp/sp,sat, i.e., θ/μ vs. (s−sY)/sp,sat at different The influence of improved work hardening in PWHT conditions
temperatures in modified 9Cr–1Mo steel post-weld heat treated due to microstructural softening is reflected in a systematic
at 993 K for 1 h as shown in Fig. 13. θ/μ vs. sp/sp,sat data displayed increase in plastic flow component of saturation stress sp,sat with
two different narrow bands, one for room and intermediate increasing PWHT temperature. In this context, it is interesting to
temperatures (300–673 K) and the other for high temperatures examine the variations of flow stress in its totality with respect to
(723–873 K). The two different groupings in θ/μ vs. sp/sp,sat clearly increasing PWHT temperature in the temperature range 300–
suggest the dominance of two different recovery mechanisms, i.e., 873 K. The total flow stress in terms of saturation stress obtained
cross-slip of dislocations at room and intermediate temperatures as sS ¼sY+sp,sat exhibited distinct three temperature regimes
and dislocation climb and subboundary migration at high temp- (Fig. 14) similar to those shown by the variations of yield stress
eratures. and plastic flow component of saturation stress with respect to
The large reduction in true yield stress with increasing PWHT temperature (Figs. 8 and 9). In addition to this, saturation stress
temperature (Fig. 8) observed for the steel can be attributed to shows a systematic decrease with increasing PWHT temperature.
However, the influence of increasing PWHT temperature on
decreasing saturation stress values is less pronounced compared

Fig. 13. Variations of normalised instantaneous work hardening rate (θ/μ) as a


function of normalised plastic component of flow stress (sp/sp,sat) at different Fig. 14. Variations of normalised total saturation stress (sS/μ) with temperature for
temperatures in modified 9Cr–1Mo steel for PWHT of 993 K for 1 h. N+T and PWHT conditions in modified 9Cr–1Mo steel.
126 J. Christopher et al. / Materials Science & Engineering A 575 (2013) 119–126

the improved work hardening and this is reflected in a systematic


increase in plastic flow component of saturation stress (sp,sat) with
increasing PWHT temperature compared to that in N+T condition. An
examination in terms of the influence of PWHTs on total saturation
stress, i.e., sS ¼sY+sp,sat indicated a marginal decrease in sS with
increasing PWHT temperature due to larger gain in plastic flow
component of saturation stress. It is observed that the work hard-
ening parameters are sensitive to initial microstructures. Excellent
correlation in terms of equivalence between saturation flow stress
(sS) and true ultimate tensile strength (sU) has been observed for the
steel in N+T and PWHT conditions.

References

[1] J.H. Hollomon, Trans. AIME 162 (1945) 268–290.


[2] P. Ludwik, Elements der Technologischen Mechanik, vol. 32, Verlag Von Julius
Springer, Leipzig, 1909.
[3] H.W. Swift, J. Mech. Phys. Solids 1 (1952) 1–18.
[4] E. Voce, J. Inst. Met. 74 (1948) 537–562.
[5] E. Voce, Metallurgia 51 (1955) 219–226.
Fig. 15. Variations of saturation stress (sS) with true ultimate tensile strength (sU) [6] D.C. Ludwigson, Metall. Trans. 2 (1971) 2825–2828.
obeying sS EsU for different temperatures and heat treatment conditions in [7] U.F. Kocks, J. Eng. Mater. Technol. 98 (1976) 76–85.
modified 9Cr–1Mo steel. [8] H. Mecking, U.F. Kocks, Acta Metall. Mater. 29 (1981) 1865–1875.
[9] Y. Estrin, H. Mecking, Acta Metall. Mater. 32 (1984) 57–70.
[10] U.F. Kocks, H. Mecking, Prog. Mater. Sci. 48 (2003) 171–273.
to that in yield strength due to improved work hardening causing [11] P. Spatig, G.R. Odette, E. Donahue, G.E. Lucas, J. Nucl. Mater. 283–287 (2000)
721–726.
larger gain in plastic flow component of saturation stress. Fig. 15 [12] R. Bonade, P. Spatig, M. Victoria, T. Yamamoto, G.R. Odette, J. Nucl. Mater. 329–
shows a comparison between saturation stress and true ultimate 333 (2004) 278–282.
tensile strength (sU) in the temperature range 300–873 K for N+T [13] Y. Estrin, K. Rhee, R. Lapovok, P.F. Thomson, J. Eng. Mater. Technol.—Trans.
ASME (2007)380–389.
and PWHT conditions. Theoretical sS ¼sU line is shown as solid
[14] P. Spatig, R. Schaublin, M. Victoria, Material instabilities and patterning in
line in Fig. 15. Excellent correlation observed between saturation metals, in: H.M. Zbib, G.H. Campbell, M. Victoria, D.A. Hughes, L.E. Levine
flow stress (sS) and true ultimate tensile strength (sU) for the steel (Eds.), Materials Research Society Symposium Proceedings, vol. 683E, San
in N+T and PWHT conditions is not surprising and is in order. Francisco, 2001, pp. BB1.10.1–BB1.10.6.
[15] V.K. Sikka, in: J.W. Davies, D.J. Michael (Eds.), Ferritic Alloys for Use in Nuclear
Following Considere's criterion for necking, sU is determined Energy Technologies, TMS-AIMEPA, Warrendale, 1984, pp. 317–327.
as s¼ sU at θ¼ s. Following the K–M approach, sS,K–M is obtained [16] S.L. Mannan, S.C. Chetal, Baldev Raj, S.B. Bhoje, Trans. Indian Inst. Met. 56
as the sum of sY+sp,sat, where sp,sat is determined as s¼ sp,sat at (2003) 155–178.
[17] Baldev Raj, B.K. Choudhary, Trans. Indian Inst. Met. 63 (2010) 75–84.
θ¼0 from the θ–sp plots in stage-III. Therefore, sS,K–M is expected [18] E. Isaac Samuel, B.K. Choudhary, K.B.S. Rao, Mater. Sci. Technol. 23 (2007)
to be higher than sU as demonstrated in Fig. 15. 992–999.
[19] B.K. Choudhary, J. Christopher, E. Isaac Samuel, D.P. Rao Palaparti, M.D.
Mathew, communicated to Mater. Des., 2013.
[20] D.P. Rao Palaparti, B.K. Choudhary, J. Christopher, E. Isaac Samuel, M.D.
4. Conclusions
communicated to Mater. Des., 2013.
[21] E. Pink, A. Grinberg, Mater. Sci. Eng. 51 (1981) 1–8.
Tensile flow and work hardening behaviour of modified 9Cr–1Mo [22] B.K. Choudhary, D.P. Rao Palaparti, J. Nucl. Mater. 430 (2012) 72–81.
steel in N+T and PWHT conditions have been adequately described [23] B.K. Choudhary, D.P. Rao Palaparti, E. Isaac Samuel, Metall. Mater. Trans. A 44
(2013) 212–223.
using the one-internal-variable Kocks–Mecking approach. At all the [24] P. Spatig, N. Baluc, M. Victoria, Mater. Sci. Eng. A 309–310 (2001) 425–429.
heat treated and test temperature conditions, θ–sp plots exhibited [25] R. Bonade, P. Spatig, Mater. Sci. Eng. A 400–401 (2005) 234–240.
two-stage work hardening behaviour characterised by a rapid [26] J. Christopher, B.K. Choudhary, E. Isaac Samuel, V.S. Srinivasan, M.D. Mathew,
Mater. Sci. Eng. A 528 (2011) 6589–6595.
decrease in instantaneous work hardening rate (θ) with plastic [27] Design and Construction Rules for Mechanical Components of FBR Nuclear
component of flow stress (sp) at low stresses (transient stage) Islands, RCC-MR, Section 1, Subsection Z, Appendix A3.18S.22, 2002.
followed by a gradual decrease at high stresses (stage-III). The [28] B.K. Choudhary, K. Bhanu Sankara Rao, S.L. Mannan, B.P. Kashyap, Mater. Sci.
Technol. 15 (1999) 791–797.
variations in work hardening parameters associated with the K–M [29] A.K. Roy, P. Kumar, D. Maitra, Mater. Sci. Eng. A 499 (2009) 379–386.
approach with temperature exhibited distinct three temperature [30] B.K. Choudhary, V.S. Srinivasan, M.D. Mathew, Mater. High Temp. 28 (2011)
regimes along with signatures of dynamic strain ageing at inter- 155–161.
[31] B.K. Choudhary, Mater. Sci. Eng. A 564 (2013) 303–309.
mediate temperatures and dominance of dynamic recovery at high
[32] D.J. Dingley, D. McLean, Acta Metall. 15 (1967) 885–901.
temperatures. At all the test temperatures, a systematic decrease in [33] A.M. Garde, A.T. Santhanam, R.E. Reed-Hill, Acta Metall. Mater. 20 (1972)
true yield stress (sY) values has been observed with increasing 215–220.
microstructural softening due to increasing PHWT temperatures in [34] D.J. Michel, J. Moteff, A.J. Lovell, Acta Metall. 21 (1973) 1269–1277.
[35] J.G. Morris, Mater. Sci. Eng. 13 (1974) 101–108.
the range 993–1033 K. On the contrary, the microstructural softening [36] B.P. Kashyap, K. McTaggart, K. Tangri, Philos. Mag. 57 (1988) 97–114.
due to additional PWHTs in modified 9Cr–1Mo steel has resulted in [37] E. Nes, K. Marthinsen, Y. Brechet, Scr. Mater. 47 (2002) 607–611.

You might also like