You are on page 1of 11

Materials and Design 89 (2016) 759–769

Contents lists available at ScienceDirect

Materials and Design

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

Influence of thermo-mechanical embrittlement processing on


microstructure and mechanical behavior of a pressure vessel steel
Xue Bai a,b, Sujun Wu a,⁎, Peter K. Liaw b
a
School of Materials Science and Engineering, Beihang University, Beijing 100191, P.R. China
b
Department of Materials Science and Engineering, The University of Tennessee, Knoxville, TN 37996, USA

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

Article history: A thermo-mechanical embrittlement processing (TMEP) consisting of thermal aging and cold strain could cause
Received 11 June 2015 the deterioration of reactor pressure vessel (RPV) steels in the form of an increase in the ductile to brittle transi-
Received in revised form 30 September 2015 tion temperature (DBTT) and a decrease in the upper-shelf energy (USE). In this study, the TMEP was employed
Accepted 5 October 2015
to investigate the microstructure and the evolution of mechanical behavior of the SA508-IV pressure vessel steel.
Available online 9 October 2015
In the microstructure of the as-received state, Cr and Mn atoms replace Fe atoms and form alloying cementites,
Keywords:
(Fe,Cr)3C and (Fe,Mn)3C, through in-situ nucleation. Due to the slower diffusion coefficient, Cr precipitates in the
SA508-IV outer layer of the Mn clusters. In the subsequent embrittlement process, needle-shaped Mo2C, fine copper-rich
Thermo-mechanical embrittlement processing precipitates (CRPs) and P-rich precipitates are formed, which play a great role in the mechanical behavior evolu-
(TMEP) tion. Mechanical test results show that a series of changes in mechanical behavior occurred. It has been found that
Precipitates DBTT fits a linear function of the square root of embrittling time at 520 °C (t 1/2) from 10 h to 90 h degradation and
Reactor pressure vessel steel the degree of embrittlement reaches saturation after 90 h.
Ductile to brittle transition temperature (DBTT) © 2015 Elsevier Ltd. All rights reserved.

1. Introduction toughness and the premature failure of steels [8,9], which is a non-
hardening embrittlement caused by the grain-boundary segregation of
A main objective for the nuclear industry is to enhance the safety impurity elements, mainly phosphorus [2].
and the reliability of nuclear power plants and to extend their operation Currently, the nature and mechanisms of irradiation embrittlement,
lifetime. The reactor pressure vessel (RPV) steel plays the most impor- especially the changes in the microstructure and mechanical behavior of
tant role in safety rankings of the whole nuclear power plant system. materials, are the subjects of major studies. However, as neutron irradi-
The SA508-IV Ni–Cr–Mo low alloy steel has excellent mechanical be- ation is harmful, the thermo-mechanical process, thermal aging and
havior [1], which possesses higher strength and fracture toughness by cold strain were employed to deteriorate the RPVs to simulate the effect
adding Ni and Cr elements [2]. Neutron irradiation has been considered of neutron irradiation. To be specific, the temper-embrittlement process
as the main factor responsible for the embrittlement of RPVs [3]. As the could be combined with cold strain, which is called the TMEP procedure.
experiments under neutron irradiation are not easy due to the radioac- The heat treatment can produce a larger concentration of vacancies and
tivity, the thermo-mechanical embrittlement processing (TMEP) meth- vacancy clusters that are supposed to accelerate the formation of CRPs
od has been chosen to study the performance of RPVs. based on a solute-vacancy exchange diffusion mechanism [10]. After
Neutron irradiation is generally known as a major reason for the em- that, the cold strain may cause the interaction of dislocations and pre-
brittlement of RPV steels because of the damage produced in collision cipitates, and then the steel could be embrittled by the large numbers
cascades consisting of self-interstitial atoms and vacancies. The neutron of dislocation loops forming behind them.
irradiation embrittlement of RPV steels is known to be related to the In the previous study, the pre-strain followed by aging process was
presence of residual elements [4], such as Phosphorus, which can also applied to study the microstructure evolution of RPV steels. Ghosh
be obtained from long-term tempering [5,6]. Copper-rich precipitates et al. [11] have found that the 50% pre-strain prior to aging could trigger
(CRPs) are also considered as the reason for irradiation embrittlement the precipitation by decreasing the activation energy values for Cu pre-
[7]. In the temperature range of 300–500 °C, temper embrittlement cipitation from 182 to 126 kJ/mol. Kamada et al. have conducted a series
has been considered as the main factor responsible for the loss of of experiments to simulate the irradiation embrittlement of nuclear RPV
steels [12]. The Fe-1 weight percent (wt.%) Cu alloys with and without
⁎ Corresponding author. pre-deformation in the solid-state solution were thermally aged at
E-mail address: wusj@buaa.edu.cn (S. Wu). 773 K (500 °C) for various times. After that, the evolution of hardness,

http://dx.doi.org/10.1016/j.matdes.2015.10.024
0264-1275/© 2015 Elsevier Ltd. All rights reserved.
760 X. Bai et al. / Materials and Design 89 (2016) 759–769

conductivity, and microstructure were investigated. It has been found


that the pre-deformation enhanced Cu precipitation and caused precip-
itation at dislocations.
Although the microstructure evolution of the above methods was
close to the results of the neutron irradiation state, the mechanical
tests could not obviously show the embrittlement of RPVs. Wu et al.
[13] studied the TMEP process, applying the 8% uniform cold strain fol-
lowing the heat treatment, to simulate the effect of neutron irradiation
on the mechanical properties of a pressure vessel weld metal. The frac-
ture tests were carried out using blunt notch four-point-bend speci-
mens in slow bend over a range of temperatures. The ductile–brittle
transition temperature (DBTT) was shown to increase by approximately
100 °C as a result of the degradation.
In this study, the TMEP was applied to the new type of pressure ves- Fig. 1. Schematic of the degradation process.
sel steel A508-IV. It has been stated by Fatehi et al. [14] that the precip-
itation of Cu in steels is usually observed in the temperature range of
400–650 °C. Thus, the current work was conducted at 520 °C using the 150 kg at 9 positions, and the results are averaged to assure the
TMEP deterioration process. It was focused on studying the embrittle- repeatability.
ment effect of the hardening and grain-boundary P segregation, to com-
pare with the in-service SA508-III and the neutron irradiation 2.3. Microstructure analysis
embrittlement. The microstructure evolution, which leads to the em-
brittlement, will be studied in details and the resultant degradation of After standard grinding and polishing, the samples were etched by 5
the mechanical behavior will also be discussed. volume percent (vol.%) nital. The microstructures were examined using
the optical microscopy (OM) and scanning electron microscopy (SEM).
Transmission electron microscope (TEM) was employed to examine
2. Experimental procedure the fine details of the microstructure and to obtain the diffraction pat-
tern of the selected phases. The TEM specimens were prepared using a
2.1. Testing materials and specimens double-jet electrolytic thinning technique. The images were taken
under both TEM and scanning transmission electron microscope
The chemical composition of the SA508-IV Ni–Cr–Mo low-alloy steel (STEM) mode.
is listed in Table 1. As shown in Fig. 1, the specimens were subjected to a
heat treatment of 1.5 h at 920 °C, followed by water quenching (WQ) to 3. Results and discussion
room temperature (RT), re-heating to 650 °C for 30 h and furnace
cooling to RT. Specimens without the further TMEP treatment are 3.1. Microstructural observation
termed as “the as-received state”. The degradation procedure has been
worked out with aging, followed by the 10% uniform cold strain (Fig. 3.1.1. Undegraded state
1). The cold strain is applied by the compressive stress at the speed of Figs. 3(a) and (b) are SEM micrographs of the undegraded (as-re-
0.5 mm/min using the 100 t computer-controlled electro-hydraulic ceived) samples of SA508-IV. Fig. 3(a) shows that it has an overall mar-
servo universal testing machine. The corresponding specimens are tensitic structure. In Fig. 3(b), it could be found that there are some
termed as “the degraded state”. coarse carbides precipitated along grain boundaries. According to the
The above two states of materials are all fabricated into tensile research of Lee et al. [15], they should be the M3C-type carbides. Then
specimens and Charpy-impact test specimens. The tensile specimens based on the study of Hashimoto et al. [16], the carbides distributed
are bars with a gage length of 25 mm and gage diameter of 5 mm. The di- along grain and/or lath boundaries could be defined as M23C6 and Cr-,
mension of Charpy-impact test specimens is 10 mm × 10 mm × 55 mm Ni-rich M6C. The fine M2C-type carbides are also dispersed inside
with the notch direction transverse to acting surfaces applied by the laths, similar to the results mentioned in Refs. [13] and [14]. In Fig. 4,
compressive stress (Fig. 2). it can be seen that the plate-shaped Cr-rich M23C6 carbides have a diam-
eter of about 142 nm [Fig. 4(a)], and the rod-shaped M6C is about
129 nm in length [Fig. 4(b)]. Besides, a few MC precipitates mainly dis-
2.2. Mechanical tests tribute along dislocations (see Fig. 5).
In the tempering process, the alloying elements diffuse quickly and
Tensile tests at room temperature were carried out using a re-locate inside α-Fe and Fe3C phases. It can be obviously seen from
computer-controlled 50 kN testing machine with a loading rate of 0.5 the STEM distribution diagram of elements (Fig. 6) of the undegraded
mm/min.
Charpy-impact tests were conducted using a 300 J pendulum impact
test machine in a temperature range from −160 °C to RT. The test re-
sults were used to plot ductile to brittle transition temperature (DBTT)
curves.
In order to assess the age-hardening effect, Rockwell C Hardness
(HRC) values were measured for each state of samples under a load of

Table 1
Chemical composition of SA508-IV Ni–Cr–Mo alloy steel in wt.%.

C Si Mn P S Ni Cr Cu Mo Al Fe

0.15 0.36 0.34 0.011 0.008 3.26 1.66 0.041 0.46 0.005 Balance
Fig. 2. Position relationship of the compressive stress and notch location.
X. Bai et al. / Materials and Design 89 (2016) 759–769 761

Fig. 3. (a) SEM micrographs of the undegraded state of the SA508-IV RPV steel, showing a martensitic structure; (b) amplified photo of (a). Nital etched.

RPV steel that the precipitates are full of Cr and Mn elements but lack of smaller, the tiny M2C grows up a little [Fig. 7(e) and (f)]. After degraded
Fe and contain a mild quantity of Mo atoms. for 90 h [Fig. 7(f) and (h)], the coarse M3C carbides join together along
In Fig. 6, the diameter of Cr element is measured as 148.24 nm and grain boundaries. Besides, the size of M23C6 and M6C carbides goes up,
that of Mn is 91.67 nm (as arrows indicated in Fig. 6). Thus it could be and the density of M2C carbides increases. Obviously, the size of the
considered that Cr elements form the outer layer surrounding the Mn needle-like carbides Mo2C inside martensitic laths begins to grow up
clusters, which means that Mn atoms move faster than Cr atoms. The re- [see in Fig. 7(h)].
sults are consistent with the study of Zhang et al. [17], the diffusion co- Figs. 8, 9, 10, and 11 represent the STEM images of 10 h, 30 h, 90 h,
efficient of Mn is greater than that of Cr, DMn N DCr (the element Cr is one and 180 h-degraded state samples, respectively. In Fig. 8, the Cr-Mn
order of magnitude slower than Mn). After Cr and Mn gather near Fe3C, rich precipitates in the 10 h state could be found to be smaller compared
they will replace Fe atoms to alloy cementites (Fe,Cr)3C and (Fe,Mn)3C with Fig. 6, which represents the undegraded state. It can also be seen
in the manner of the in-situ nucleation. Therefore, there are no Fe from Fig. 8 that the elements Ni and Mo have precipitated together,
atoms in the positions of Cr and Mn atoms. Furthermore, it can be and the distributions of them are the same as phosphorus in 10 h-
seen from Fig. 6 that Mo atoms disperse in the area and do not form degraded specimens. Kameda [18] holds the opinion that the presence
clusters. of Mo plays a key part in impeding intergranular P segregation and frac-
ture probably due to the attractive Mo-P interactions in the grain matrix
3.1.2. Degraded state and a grain-boundary toughening effect of the segregated Mo. As the
The microstructure of the degraded state samples could be found in aging time grows to 30 h, the attractive Mo-P interactions are dispersed
Fig. 7. They correspond to the 10 h, 30 h, 90 h, and 180 h-degraded state homogeneously (see Fig. 9). However, the beneficial effect of Mo is di-
samples, respectively. Fig. 7(b), (d), (f), and (h) are the enlarged ones of minished in long-term aging where the dissolved Mo content decreases
(a), (c), (e), and (g), respectively. Comparing Fig. 7(a), (c), (e), and (g), as a result of the precipitation of fine Mo-rich carbides [18], Mo2C, which
the grain size of these degraded state samples do not change significant- are the tiny needle-like carbides precipitate inside the martensitic laths
ly. However, the size of carbides increases as the thermal aging in- [19], seen in Fig. 12. However, in Fig. 9, Cr atoms are found to segregate
creases from 10 h to 180 h. Unlike Fig. 3, there are more tiny M2C and form clusters, so Mn atoms precipitate outside.
precipitates inside the martensite laths in the 10 h-degraded state sam- Mo2C is a member of the interstitial compounds where the transition
ple. In the 30 h-degraded state sample [Fig. 7(d)], the needle-like car- metal atoms form a close-packed-hexagonal arrangement, and the car-
bides began to precipitate inside the martensitic laths, which is bon atoms are found in the octahedral interstices, which has relatively
supposed to be Mo2C [18]. In 90 h-degraded state sample, the size of higher hardness, melting point, and dissolving temperature. Mo2C usu-
the coarse precipitates on grain boundaries and lath boundaries become ally locates inside the martensite lath or segregates at the sub-

Fig. 4. M23C6 precipitates and M6C precipitates on grain and/or lath boundaries of the undegraded state, (a) M23C6 precipitates; (b) M6C precipitates.
762 X. Bai et al. / Materials and Design 89 (2016) 759–769

The alloying elements Ni and Cr promote P segregation through the


mechanisms of co-segregation and a reduced carbon activity in alloys,
and Ni conversely improves the low temperature cleavage toughness
by decreasing the energy barrier of kink formation [18]. It has been
shown that the addition of Mn and/or Si influences the intergranular
embrittlement in iron, Cr-Mo and Ni–Cr–Mo–V steels with different
plasticity [18]. The content of P affects irradiation hardening more sig-
nificantly than that of Cu due to the distinct formation of P-rich precip-
itates arising from the stabilization of vacancies [23]. Therefore, it can be
expected that the TMEP hardening will be affected by the phosphorus
concentration on the grain boundary.
In Fig. 10, it could be found that there are no phosphorus precipi-
tates. Thus, it could be concluded that the phosphorus has completely
re-dissolved back into the matrix [22] after long-term degradation,
which is different from the evolution of copper precipitates. What's
more, there is no large difference between 180 h's aging and 90 h's
aging (Figs. 10 and 11).

3.2. Mechanical behavior

3.2.1. Tensile tests


Tensile test results are listed in Table 2, which has been plotted in
Fig. 13(a). It could be found that the yield strengths (σy) and tensile
Fig. 5. MC precipitates on dislocations of the undegraded state.
strengths (σb) firstly increase and then decrease slightly with the degra-
dation time, showing peak values of 803 MPa and 866 MPa at 90 h's
aging, respectively. As for the elongation, the undegraded-state samples
possess good elongation of over 20%. However, with the degradation
boundary. It forms through a nucleation and growth process in a charac- time increases, elongation values fluctuate between 11.3%–14.7%. The
teristic needle shape [19], which is one of the main strengthening phase strain-hardening exponent, n, tends to decrease suddenly from the
of steels in the temperature range of 500–650 °C. undegraded state to the 10 h-degraded state, then rise up to a relatively
The equilibrium solid solution of Cu in α-iron at 500 °C can be deter- higher level at 30 h, and drop gradually from 30 h to 180 h.
mined as 0.0328 wt.% extrapolated from the empirical relationship pro- As for the shift of strength, it can be seen in Fig. 13(a) that the yield
posed by Salje and Feller-Kniepmeier [20]. The bulk content of Cu in this strength increases and the ductility decreases with the increase of deg-
experimental steel is 0.041 wt.%. Therefore, the supersaturated Cu radation time from 10 h to 90 h in the degradation process. This trend
would precipitate in the form of the CRPs. The calculation results obtain- can be attributed to the dislocation cells and distorted martensitic lath
ed by Becquart [21] indicate that a strong vacancy affinity for copper in the degraded material, the aging, as well as the cold-worked material
atoms, and slightly faster diffusion of the Cu impurities compared [24]. From the degradation of 10 h to 30 h, there are large numbers of
with Fe. With the aging time goes by, the number of vacancies and nanoscale precipitates in the specimens, such as CRPs, phosphorus-
vacancy-clusters will increase rapidly. Diffusion of Cu will be accelerat- rich precipitates, and Mo2Cs. According to the formula,
ed by a large number of single vacancies and vacancy-clusters based on
the solute-vacancy exchange diffusion mechanism to form copper-rich τ ¼ Gb=L ð1Þ
clusters [16]. It can be seen in Fig. 8 that there are plenty of CRPs
(CRPs with a size of 5–10 nm) in the 30 h-degraded RPV steel. where τ is the shear stress of particles placed upon dislocations, G is
When the aging time rises up to 90 h (Fig. 10), CRPs grow larger, and shear modulus, b is Burgers vector, L is the distance between particles,
the density of CRPs decreases. This phenomenon is consistent with the the material could be strengthened because of the close distance be-
results observed by Valo et al. [22]. The reason is that during annealing, tween precipitates.
copper precipitates do not massively re-dissolve back into the matrix. It should be noticed that there is a turning point at 90 h in the tensile
However, they grow larger with the decrease of density. strength versus degradation time curve, as seen in Fig. 13(a). The

Fig. 6. Elemental mapping of the undegraded RPV steel characterized by STEM.


X. Bai et al. / Materials and Design 89 (2016) 759–769 763

Fig. 7. The microstructure of the 10 h, 30 h, 90 h and 180 h degraded state samples.

specimens degraded for 90 h possess the highest yield strength, partial annihilation of dislocations produced during quenching and/or
803 MPa, and relatively higher ductility, 14.7% elongation (Table 2). cold straining and tempering of martensite laths (the decomposition
After being degraded for 180 h, the precipitates become coarse, which of hard martensites), which over-balances the hardening feature due
would weaken the strengthening effect. Otherwise, the softening be- to the Cu precipitation, as well as fine M2C carbides and the needle-
havior of the 180 h-degraded samples could also be attributed to the like Mo2C carbides that precipitate inside martensite laths. Besides,
764 X. Bai et al. / Materials and Design 89 (2016) 759–769

Fig. 8. Elemental mapping of the 10 h-degraded RPV steel characterized by STEM.

Fig. 9. Elemental mapping of the 30 h-degraded RPV steel characterized by STEM.

the coherent to incoherent transformation, the coarsening characteris- dislocations because of the mutual elastic interactions. The dislocations
tics of the nanoscale particles (mainly the CRPs) [14], M23C6 and M6C could be seen in the 90 h-degraded state specimens (Fig. 14). After
carbides, which distributed along grain and/or lath boundaries, the coa- being degraded for more than 90 h, there is enough time for the defects
lescence of the coarse M3C precipitates along grain boundaries, can all to diffuse. In this case, the stress field of a dislocation that initially exists
lead to the slightly drop of the strength of the RPV steel. It has been re- in the RPV steel would be compensated for by the stress field of these
ported that [12] nano-size defect clusters can accumulate near defects. Therefore, the strength would decrease after 90 h.

Fig. 10. Elemental mapping of the 90 h-degraded RPV steel characterized by STEM.
X. Bai et al. / Materials and Design 89 (2016) 759–769 765

Fig. 11. Elemental mapping of the 180 h-degraded RPV steel characterized by STEM.

It has been found that the strengthening process is highly complex, Correspondingly, in the TMEP, the dislocation density and carbides
and there are various sources of strengthening combined to produce dispersion have changed. Specifically, the heat treatment produced
the total flow stress, σy [25], some precipitates, such as ε-Cu and Mo2C precipitates. The cold-strain
process forces the dislocations to move and then form many dislocation
loops behind, which leads to the embrittlement of the specimens. The
σ y ¼ σ 0 þ Δσ ss þ Δ σ disl þ Δσ gr þ Δ σ part ð2Þ
TMEP in this study contributes to the changes of the parameters
Δσpart and Δσdisl in Eq. 3. Therefore, it can be used to forecast the perfor-
where the quantity, σ0, is the temperature and strain-rate-sensitive fric- mance of RPV steels under real neutron irradiation and screen the ones
tional stress from the inherent lattice resistance to slip, Δσss is the solid- which possess the superior neutron irradiation embrittlement
solution-hardening effect increasing with the content of dissolved for- resistance.
eign atoms, Δσdisl represents the effect due to the long-range stress
fields of the dislocations and is proportional to the square root of the 3.2.2. Hardness tests
dislocation density, Δσgr is the well-known Hall–Petch grain strength- The HRC hardness test results are listed in Table 3. The variation
ening, and Δσpart is the contribution of particles, precipitates, or similar trend can be seen in Fig. 13(b). It is evident that the hardness values
structural defects. The last one mainly results from the carbides disper- are in the direct proportion to the duration of the degradation time at
sion. The irradiation does not change the inherent lattice resistance, the first, and then saturate gently after 30 h.
grain size, the dislocation density, and the carbide dispersion. It prefer- The first rapid increase stage of hardness values between 0 h-30 h
entially produces a new dispersion of dislocation obstacles, copper-rich degraded states [see Fig. 13(b)] is caused by several reasons [24],
precipitates, which are termed as irradiation defects. Therefore, the last (a) The gradual precipitation of nano-scale CRPs and needle-like
parameter Δσpart should be paid attention to in the neutron irradiation Mo2Cs increases the hardness. As previously mentioned, Mo2C is one
embrittlement study. of the mainly strengthening phases in steels in the temperature range
of 500–650 °C [19].
(b) It can be stated that the interaction of P with vacancies, and the
P-rich precipitation or clustering are responsible for the precipitation
hardening in the P-doped steels.
(c) The hardening behavior may arise from the increased density of
dislocation-pinning sites, which is the same as in the highly irradiated
steel.
(d) It is well known that Ni has a strong effect on hardening, arising
from the formation of Cu-rich clusters made up of a Ni shell and Cu core.
In addition, Ni increases the density of interstitial-type dislocation
loops, which increases the hardness.

Table 2
Tensile test results of the different degradation state of SA508-IV steel.

Sample Yield Tensile Elongation, n


condition strength, strength, δ (%)
σy (MPa) σb (MPa)

0 h (as-received) 509 622 22.2 0.161/0.165


10 h 743 777 12.9 0.142/0.148
30 h 769 812 11.3 0.159/0.154
90 h 803 866 14.7 0.144/0.133
180 h 797 854 13.0 0.084/0.100
Fig. 12. Mo2C precipitates in the 30 h-degraded RPV steel.
766 X. Bai et al. / Materials and Design 89 (2016) 759–769

Table 3
Mechanical test results of the different degradation state of the SA508-IV steel.

Samples σy/MPa Δσy/MPa DBTT41J/°C ΔDBTT41J/°C HRC ΔHRC

0 h (as-received) 509 – −81 – 30.93 –


10 h 743 234 −68 13 35.05 4.12
30 h 769 260 −53 28 40.67 9.74
90 h 803 294 −42.5 38.5 41.13 10.2
180 h 797 288 −42 39 41.71 10.78

where Δσy = σy (TMEP) − σy (as-received).


ΔDBTT 41J = DBTT 41J (TMEP) − DBTT 41J (as-received).
ΔHRC = HRC (TMEP) − HRC (as-received).

deformation and tempering of martensite laths (the decomposition of


hard martensites); (b) aging hardening due to Cu [26], clustering of
Cu atoms and coarsening of Cu precipitates, respectively; (c) P-rich pre-
cipitates or clusters re-dissolve back into the matrix after aging for a
long time. Therefore, the hardness values would not go up dramatically
at the initial stage.

3.2.3. Charpy-impact tests


In this study, the degraded specimens, which are supposed to simu-
late the irradiation effect, have been used to conduct impact testing. The
testing results are listed in Table 3. Fig. 15 reveals the variation of DBTT
under different degradation states. The DBTT curve of the as-received
state samples can be expressed as,

Impact Absorbed Energy ¼ 92:79 þ 83:09


 tanh½ðT þ 43:54Þ=52:10: ð3Þ

In this study, the temperature corresponding to the impact test ener-


gy of 41 J has been used as the DBTT. From the intercepts of the DBTT
curves with the horizontal line, it can be concluded that the DBTT
Fig. 13. The variation of σy, HRC, DBTT and USE upon the degradation time.
keeps on increasing from the 0 h (the as-received state) to the 90 h-
degraded state. Fig. 16 shows the morphology of the fracture surfaces
The second gentle increase stage can be attributed to the combined ef- of both the undegraded state [Fig. 16 (a)] and 90 h-degraded state
fects of the concomitant softening (recovery) and age-hardening (pre- [Fig. 16 (b)] specimens. The impact absorbed energy at − 40 °C is
cipitation) reactions taking place during the aging process. The factors 103.39 J for the undegraded specimen and 36.09 J for the 90 h-
which cause this phenomenon are listed below: (a) the partial annihila- degraded specimen. Compared with Fig. 16 (a), there are much more in-
tion of dislocations produced during quenching and/or cold tergranular facets in Fig. 16 (b), which makes it brittle. However, the
DBTT curves of 90 h and 180 h almost overlap. The variation tendency
of DBTT upon the degradation time could be concluded more clearly
from Fig. 13(c), which increases continuously at the beginning and
then reaches a plateau after 90 h, corresponding to the saturation of

Fig. 14. Dislocations of 90 h-degraded state specimen. Fig. 15. The variation of DBTT under different degradation states.
X. Bai et al. / Materials and Design 89 (2016) 759–769 767

Fig. 16. Fracture surface of Charpy impact test specimens, (a) undegraded state, −40 °C, 103.39 J; (b) 90 h-degraded state, −40 °C, 36.09 J.

the embrittlement-hardening effect. The upper-shelf energy (USE) numbers of dislocation loops behind. Eventually, the specimens have
drops obviously after being degraded for 10 h, increases a little at 30 h been embrittled, resulting in the results of DBTT-upwards shift.
and decreases after 30 h. Then it keeps constant after 90 h [Fig. 13(d)]. Besides, in Ref. [24], it has been stated that the DBTT shift induced by
(1) Ductile–brittle transition temperature (DBTT). the real neutron irradiation (ΔT 41J) is linearly correlated with ΔCgbP,
Specifically, it can be seen from Fig. 15 that the shift of DBTT, which is that is,
termed as ΔDBTT 41J, is 39 °C after degradation for 180 h. The variation
tendency of DBTT [Fig. 13(c)] can be explained as follows. It is well ΔT 41 J ¼ 63 þ 530 ΔC gb P ð4Þ
known that the phosphorus segregation promotes the initiation of in-
tergranular cracking in the stress-intensified zone accompanied by the where ΔCgbP is the shift of the bulk content of phosphorus segregated at
cleavage and ductile crack propagation. According to the study by grain boundaries.
Druce [27], P segregation was shown to reduce the critical void- The segregation of P on prior austenite boundaries was proved to re-
nucleation strain giving rise to a higher proportion of the second- duce the fracture toughness of steels [28], and the enhanced segregation
phase particle distribution contributing to the fracture process and a of P was then found to be a linear function of the square root of the em-
consequential decrease in the average void diameter and toughness. brittling time at 520 °C, which means,
With the aging time goes from 0 h to 90 h, the dynamic interaction be-
tween vacancies and P near grain boundaries becomes more effective,
ΔC gb P ¼ C  t 1=2 ð5Þ
whereas their interaction in the grain interior is not greatly promoted.
Therefore, the DBTT rises gradually with the degree of the activation
of P. However, after 90 h, the grain boundary segregation of the solute where C is a constant, and t is the embrittling time at 520 °C.
elements reached maximum value and no more extra solute elements From Eqs. (4) and (5), it can be concluded that,
in the matrix can move to grain boundaries with longer tempering
time. Thus, there is no significant difference between the DBTT of 90 h ΔT 41 J ¼ 63 þ 530  C  t 1=2 ð6Þ
samples and that of 180 h. As for the effect of the cold strain of 10% ap-
plied on the specimens, the compressive stress forces the dislocations to Therefore, the DBTT shift induced by neutron irradiation (ΔT 41J) is
move and climb the carbides or precipitates and then leave large linearly correlated with the embrittling time at 520 °C.

Fig. 18. Relationship between ΔDBTT41J and Δσy of the RPV steel SA508-IV after different
Fig. 17. Relationship between DBTT 41J and the square root of embrittling time at 520 °C. degradation procedures.
768 X. Bai et al. / Materials and Design 89 (2016) 759–769

In the TMEP degraded samples, the DBTT41J values versus the square (3) DBTT fits a linear function of the square root of embrittling time
root of embrittling time at 520 °C (t 1/2) have been plotted in Fig. 17, at 520 °C (t 1/2) from 10 h to 90 h degradation.
from which it can be seen that DBTT fits a linear function of t 1/2 up to (4) There is a linear relationship between ΔDBTT 41J and Δσy for
90 h. The result above is consistent with the microstructure of 0 h to specimens after various degradation processes. Thus, the neutron irradi-
180 h-degraded state, in which case the phosphorus segregation in- ation embrittlement resistance of RPV steels could be studied by the
creases gradually at the beginning and reaches saturation at 90 h. There- TMEP instead of the real neutron irradiation experiments.
fore, a direct relationship could be deduced between the TMEP up to (5) Test results showed that SA508-IV possesses better embrittle-
90 h and the real neutron irradiation embrittlement. ment resistance than the A533B RPV steel.
The ΔDBTT 0→180 h of SA508-IV, 39 °C, can also be compared with the
result of that of A533B RPV steel, which is 47.2 °C, studied by Wu and
Cao [5]. The tempering-degradation process in Ref. [5] is that holding Acknowledgments
the specimens at 520 °C for 30, 90, and 180 h, respectively, and then
air cooling to room temperature. The fact that the ΔDBTT0→180h of This work is financially supported by the National Science and Tech-
A533B is higher than that of SA508-IV implies that SA508-IV possesses nology Key Specific Project: Life Management Technology of Nuclear
better embrittlement resistance than A533B. Power Plant (Grant No. 2011ZX06004-002).
(2) Upper-Shelf-Energy (USE)
The values of USE are supposed to keep on decreasing from 0 h References
to 180 h degradation process. However, in Fig. 13(d) it can be
seen that the USE drops obviously after degradation for 10 h, in- [1] S.G. Park, M.C. Kim, B.S. Lee, D.M. Wee, Evaluation of temper embrittlement effect
and segregation behaviors on Ni–Mo–Cr high strength low alloy RPV steels with
creases slightly at 30 h, and decreases after 30 h. Then it keeps con- changing P and Mn contents, Korean J. Met. Mater. 48 (2010) 122.
stant after 90 h. [2] S.G. Park, K.H. Lee, M.C. Kim, B.S. Lee, Effects of boundary characteristics on resis-
The 30 h degradation samples possess the relatively higher USE. It tance to temper embrittlement and segregation behavior of Ni–Cr–Mo low alloy
steel, Mater. Sci. Eng. A 561 (2013) 277–284.
may be because of the nano-size precipitates and Mo2C strengthening [3] P. Bowen, S.G. Druce, J.F. Knott, Effects of microstructure on cleavage fracture in
phase, unlike the coarse precipitates in 10 h, 90 h, and 180 h TMEP pressure vessel steel, Acta Metall. 34 (1986) 1121–1131.
states. [4] E. Meslin, B. Radiguet, M. Loyer-Prost, Radiation-induced precipitation in a ferritic
model alloy: an experimental and theoretical study, Acta Mater. 61 (2013)
(3) The relationship between ΔDBTT 41J and Δσy.
6246–6254.
It has been proved that the relationship between ΔDBTT 41J and Δσy [5] B. Yang, P.K. Liaw, H. Wang, L. Jiang, J.Y. Huang, R.C. Kuo, J.G. Huang, Thermographic
after the neutron-irradiation embrittlement follows the following equa- investigation of the fatigue behavior of reactor pressure vessel steels, Mater. Sci. Eng.
A 314 (2001) 131–139.
tion [29],
[6] S.J. Wu, L.W. Cao, Effect of intergranular failure on the critical fracture stress and the
fracture toughness of degraded reactor pressure vessel steel, Int. J. Press. Vessel. Pip.
ΔDBTT 41 J ¼ 0:65 Δσ y ð7Þ 101 (2013) 23–29.
[7] A.V. Barashev, S.I. Golubov, D.J. Bacon, P.E.J. Flewitt, T.A. Lewis, Copper precipitation
in Fe–Cu alloys under electron and neutron irradiation, Acta Mater. 52 (2004)
However, in the present research, the test results of ΔDBTT 41J 877–886.
against Δσy after TMEP have been plotted in Fig. 18, it fits the linear re- [8] C. Lea, M.P. Seah, Site competition in surface segregation, Surf. Sci. 53 (1975)
lationship as expressed by Eq. (8), 272–285.
[9] P.K. Liaw, H. Wang, L. Jiang, B. Yang, J.Y. Huang, R.C. Kuo, J.G. Huang, Thermographic
detection of fatigue damage of pressure vessel steels at 1,000 Hz and 20 Hz, Scr.
ΔDBTT 41 J ¼ −87 þ 0:43 Δσ y ð8Þ Mater. 42 (2000) 389–395.
[10] L.W. Cao, S.J. Wu, B. Liu, On the Cu precipitation behavior in the thermos-
mechanically embrittlement processed low copper reactor pressure vessel model
From which it can be seen that the ΔDBTT 41J of 10 h, 30 h, and 90 h- steel, Mater. Des. 47 (2013) 551–556.
degraded state specimens linearly increases with the Δσy, which is sim- [11] S.K. Ghosh, A. Haldar, P.P. Chattopadhyay, Effect of pre-strain on the ageing behavior
ilar to the neutron irradiated ones. Fig. 18 also shows that the difference of directly quenched copper containing micro-alloyed steel, Mater. Charact. 59
(2008) 1227–1233.
of the embrittlement results after 90 h and 180 h is insignificant which [12] Y. Kamada, S. Takahashi, H. Kikuchi, S. Kobayashi, K. Ara, J. Echigoya, Y. Tozawa, K.
indicates that the embrittlement effect is saturated after 90 h degrada- Watanabe, Effect of pre-deformation on the precipitation process and magnetic
tion and no longer time is necessary. properties of Fe–Cu model alloys, J. Mater. Sci. 44 (2009) 949–953.
[13] S.J. Wu, J.F. Knott, Effects of degradation on the mechanical properties and fracture
Based on the above analysis, it can be concluded that the mechanical
toughness of a steel pressure-vessel weld metal, Int. J. Press. Vessel. Pip. 80 (2003)
property change of the steel after TMEP from 10 h to 90 h is close to that 807–815.
of the steel resulted from the real neutron irradiation process. Thus, the [14] A. Fatehi, J. Calvo, A.M. Elwazri, S. Yue, Strengthening of HSLA steels by cool defor-
neutron irradiation embrittlement resistance of RPV steels could be mation, Mater. Sci. Eng. A 527 (2010) 4233–4240.
[15] S. Lee, S. Kim, B. Hwang, B.S. Lee, C.G. Lee, Effects of carbide distribution on the frac-
studied by the TMEP from 10 h to 90 h instead of the real neutron irra- ture toughness in the transition temperature region of an SA 508 steel, Acta Mater.
diation experiments. 50 (2002) 4755–4762.
[16] N. Hashimoto, R.L. Klueh, Microstructural evolution of nickel-doped 9Cr steels irra-
diated in HFIR, J. Nucl. Mater. 305 (2002) 153–158.
4. Conclusions [17] H. Zhang, J.W. Wu, X.B. Liu, A. Baker, Studies on elements diffusion of Mn/Co coated
ferritic stainless steel for solid oxide fuel cell interconnects application, Int. J.
Mechanical behavior of RPV steels under neutron irradiation embrit- Hydrog. Energy 38 (2013) 5075–5083.
[18] J. Kameda, Y. Nishiyama, Combined effects of phosphorus segregation and partial in-
tlement has been successfully simulated using this TMEP method. The tergranular fracture on the ductile–brittle transition temperature in structural alloy
mutual effect of precipitates, vacancies, and dislocations in the long- steels, Mater. Sci. Eng. A 528 (2011) 3705–3713.
time degradation process leads to the variation of the mechanical prop- [19] D.J. Dyson, S.R. Keown, D. Raynor, J.A. Whiteman, The orientation relationship and
growth direction of Mo2C in ferrite, Acta Metall. 14 (1966) 867–875.
erties of this steel. It can be concluded as follows: [20] G. Salje, M. Feller-Kniepmeier, The diffusion and solubility of copper in iron, J. Appl.
(1) The microstructure of the undegraded state samples is tempered Phys. 48 (1977) 1833–1839.
martensite with alloy cementites (Fe,Cr)3C and (Fe,Mn)3C in the man- [21] C.S. Becquart, RPV steel microstructure evolution under irradiation: a multiscale ap-
proach, Nucl. Instrum. Meth. B. 228 (2005) 111–121.
ner of in-situ nucleation.
[22] M. Valo, L. Debarberis, A. Kryukov, A. Chernobaeva, Copper and phosphorus effect on
(2) Needle-shaped Mo2C, Cu-rich precipitates and P-rich precipi- residual embrittlement of irradiated model alloys and RPV steels after annealing, Int.
tates play a key role in the mechanical properties evolution after the J. Press. Vessel. Pip. 85 (2008) 575–579.
degradation process. The coalescence of M3C and the coarsening behav- [23] Y. Nishiyama, K. Onizawa, M. Suzuki, J.W. Anderegg, Y. Nagai, T. Toyama, M.
Hasegawa, J. Kameda, Effects of neutron-irradiation-induced intergranular phos-
ior of M23C6 and M6C lead to the deterioration of the mechanical phorus segregation and hardening on embrittlement in reactor pressure vessel
properties. steels, Acta Mater. 56 (2008) 4510–4521.
X. Bai et al. / Materials and Design 89 (2016) 759–769 769

[24] S.K. Ghosh, A. Haldar, P.P. Chattopadhyay, Effect of pre-strain on the ageing behavior [27] S.G. Druce, Effects of austenitisation heat treatment on the fracture resistance and
of directly quenched copper containing micro-alloyed steel, Mater. Charact. 59 temper embrittlement of MnMoNi steels, Acta Metall. 34 (1986) 219–232.
(2008) 1227–1233. [28] P. Bowen, C.A. Hippsley, J.F. Knott, Effects of segregation on brittle fracture and fa-
[25] J. Bohmert, H.W. Viehrig, A. Ulbricht, Correlation between irradiation-induced tigue crack growth in coarse-grained, martensitic A533B pressure vessel steel,
changes of microstructural parameters and mechanical properties of RPV steels, J. Acta Metall. 32 (1984) 637–647.
Nucl. Mater. 334 (2004) 71–78. [29] G.R. Odette, P.M. Lomborzo, R.A. Wullaert, In: Garner FA, Perrin JS (Eds). Effects of
[26] S.K. Ghosh, A. Haldar, P.P. Chattopadhyay, On the Cu precipitation behavior in Radiation on Materials, 12th International Symposium, ASTM STP 870. West
thermomechanically processed low carbon microalloyed steels, Mater. Sci. Eng. A Conshohocken, PA: American Society for Testing and Materials; 1985.
519 (2009) 88–93.

You might also like