You are on page 1of 6

Conference proceedings, 7th International Conference on Creep, Fatigue and Creep-Fatigue Interaction,

19-22 January 2016, IGCAR, Kalpakkam, India, D.P.R. Palaparti et al.

CREEP FRACTURE CHARACTERISTICS OF P9 STEEL AT 873 K


D.P Rao Palaparti, E. Isaac Samuel*, P. Parameswaran and B.K. Choudhary
Metallurgy and Materials Group
Indira Gandhi Centre for Atomic Centre, Kalpakkam – 603102, TN, India
*
e-mail: isaac@igcar.gov.in

ABSTRACT

Fracture characteristic of grade P9 steel which had undergone creep deformation at 873 K has
been examined. Creep rupture tests conducted on the steel revealed that stress dependence of
minimum creep rate obey Norton’s power law with distinct low and high stress regimes.
Applied stress dependence of rupture life also conformed to the observed two-slope
behaviour. Fracture behaviour remained transgranular ductile characterized by dimples in
both stress regimes. The observed extensive tertiary creep, high creep ductility and
transgranular fracture indicate that the alloy is not prone to typical creep cavitation damage.
However, a few cracks arising from decohesion at the precipitate-matrix interface were
observed at low stresses. Secondary crack density near the fracture surface was significantly
higher than that observed at regions away from the fracture surface, at all stress conditions.

Key Words: 9Cr-1Mo Ferritic Steel, Creep, Fracture Characteristic

1.0 INTRODUCTION

In Cr-Mo family of steels, 9Cr-1Mo ferritic steel (P9 steel) has been an important creep
resistant material developed for steam generator applications in fossil fired and nuclear power
generating industries [1]. The low-chromium low-alloy ferritic steels, in general, suffer from
loss of creep ductility and typical inter-granular creep damage due to formation of precipitate
free zone (PFZ) along the prior austenite gain boundaries [2,3]. The PFZ facilitates strain
concentration resulting from zone shear and gain boundary sliding processes leading to
cavitation damage [2]. 9Cr–1Mo retains its creep ductility over prolonged thermal exposure
and display absence of a PFZ and transgranular fracture. Also, comparable relative strength
of matrix and grain boundaries in 9Cr–1Mo steel results in high degree of grain boundary
cohesion. This inhibits accumulation of stress concentration at the grain boundaries and
subsequent development of cavitation damage [2,3]. In addition to its superior creep strength,
choice of as P9 steel for steam generator application is guided by its low coefficient of
thermal expansion and high resistance to stress corrosion cracking.

Creep deformation and damage contribute significantly to the degradation in strength of this
important class of steel during service. Creep deformation in 9Cr-1Mo family of steels had
been a subject of continued interest in view of their importance in application to power
industry. Typically, the variation of minimum creep rate and rupture life with stress in these
creep resistant steels are known to exhibit two slope behaviour with two distinct values of
stress exponents [4-10]. The different values of stress exponent were also associated with the
distinct values of apparent activation energy in the low and high stress regimes [6]. By
invoking the concept of back/resisting stress, the two stress regimes for creep deformation
were rationalised, and it was suggested that the creep deformation in both the stress regimes
are controlled by dislocation climb process [6]. Further, from the similar values of stress
exponents for the stress dependence of minimum creep rate and rupture life in the respective
stress regimes, it had been suggested that the controlling mechanisms of deformation and
fracture are the same in the 9Cr–1Mo steel [5,6].

1
Conference proceedings, 7th International Conference on Creep, Fatigue and Creep-Fatigue Interaction,
19-22 January 2016, IGCAR, Kalpakkam, India, D.P.R. Palaparti et al.

Prolonged exposure of the 9Cr-1Mo steel to elevated temperatures during service lead to
adverse microstructural changes, such as precipitation of secondary phases and their
coarsening and formation of intermetallic Laves phase, which had been reported to
deteriorate the mechanical properties of the steel [1,3,5,8]. This paper highlights the influence
of creep exposure at 873 K on fracture characteristics of the 9Cr-1Mo steel plate material.

2.0 EXPERIMENAL

Creep rupture tests were carried out on 9Cr-1Mo steel plate material of the chemical
composition (wt. %): C-0.10, Si-0.75, Mn-0.63, S-0.001, P-0.02, Cr-9.27, Mo-1.05, N-190
ppm and balance Fe. The 20 mm plates were normalised at 1223 K for 15 min. and tempered
at 1053 K for 2 h. Microstructure of the P9 steel in normalized and tempered condition is
shown in Fig. 1, which is composed of tempered lath martensite with precipitates along the
prior austenite grain and martensite lath boundaries. The prior austenite grain size was ~ 25
m. Uniaxial creep rupture tests were performed on cylindrical creep specimens of 10 mm
gauge diameter and 50 mm gauge length, machined out of P9 steel plates in the normalised
and tempered condition. The creep tests were conducted at 873 K employing stresses ranging
from 50 to 175 MPa. Optical and scanning electron microscopy (SEM) examinations on
creep tested specimens were carried out for obtaining detailed information on the
microstructure and fracture behaviour of the steel. Typically creep deformation resulted in
ductile fracture with a noticeable region of necking, as shown in Fig. 2 for 125 MPa at 873 K.
Detailed microstructural examinations were carried out at the creep deformed regions close to
the fracture surface (zone-1 in Fig.2) and regions away from fracture (zone-2 in Fig.2).

Fig. 1. Microstructure of P9 steel in Fig. 2. Macrostructure of creep failed


normalized and tempered condition specimen at 125 MPa / 873 K

3.0 RESULTS AND DISCUSSIONS


3.1 Creep Rupture Behaviour

Figure 3 shows the typical creep curve observed for the P9 steel at 873 K and 80 MPa. Creep
deformation of P9 steel plate material in normalized and tempered condition at all test
conditions is characterised by a small amount of instantaneous strain followed by well-
defined primary (or transient), secondary and tertiary creep stages. The instantaneous strain
on loading was subtracted from the total strain and secondary stage creep deformation was
characterized by minimum creep rate. At all stress levels, creep strain accumulated in the
tertiary stage is larger than those accumulated over transient and secondary stages of creep.
Stress dependence of minimum creep rate obeyed Norton’s power law with two distinct
slopes for low and high regimes. Figure 4 shows the variation of creep life with applied stress

2
Conference proceedings, 7th International Conference on Creep, Fatigue and Creep-Fatigue Interaction,
19-22 January 2016, IGCAR, Kalpakkam, India, D.P.R. Palaparti et al.

obtained for the P9 steel which also exhibit the two distinct, low and high, stress regimes with
respective exponents of 4.1 and 8.6. The observed two-slope behaviour and the values of the
exponents agree well with those reported for the 9Cr-1Mo family of steels [4-6,9,10]. For all
stress conditions at 873 K, the creep rate decreased continuously, in the transient stage, with
increasing creep time and creep strain to reach a minimum value in the secondary stage. In
the tertiary stage, the creep rate increased sharply with increasing creep time and creep strain
before fracture. The observed stages of creep deformation in the P9 steel plates are in
agreement with those reported for P9 thick section material, P91 and P92 steels [5,9,10].

Fig. 3. Creep curve for P9 steel at 873 K Fig. 4. Variation of rupture time with stress
and 80 MPa for P9 steel at 873 K

The prolonged tertiary creep in terms of both the large time spent and accumulation of large
creep strain observed is consistent with those reported for plain and modified versions of
9Cr–1Mo steels [5,8-10]. In Cr-Mo-V steel, Williams and Wilshire [11] had attributed the
enhanced tertiary creep to precipitate coarsening and the associated reduction in back stress
for dislocation movement and thereby increasing the effective stress with time. Increased
creep rate in the tertiary had been ascribed to reduction in creep strength due to recovery of
dislocations, the agglomeration and growth of carbides and subgrain coarsening [5,12]. The
subgrain coarsening was higher in tungsten free 9Cr steel compared to tungsten bearing 9Cr
steels [12]. In view of the above investigations, the progressive reduction in creep strength
can be attributed to microstructural degradation during creep exposure leading to enhanced
tertiary creep in the 9Cr–Mo steel.

3.2 Fracture Behaviour

SEM images of fractured surfaces of creep failed specimen revealed transgranular ductile
failure characterized by dimples resulting from coalescence of micro voids (Fig. 5). At the
low stress level of 60 MPa, the dimples were of mixed sizes with higher fraction of large
sized dimples. At high stress levels, the fracture surface revealed higher fraction of fine
dimples with few large sized dimples. The steel retained its high creep ductility and loss of
ductility is not observed at the longer test durations. Extensive necking and the occurrence of
cup and cone fracture has been observed at all the test conditions (e.g., Fig. 2).

The fracture characteristic of P9 steel in the low stress and high stress regimes has been
examined on longitudinally sectioned specimen at regions close to fracture surface and
regions away from the fracture surface. Figures 6a-6d show the optical microscope (OM) and
SEM image of the microstructure observed close to fracture surface (Zone-1) at 60 and
125 MPa. Grain flow along the direction of stress can be seen at the high stress of 125 MPa
than that observed at 60 MPa (Fig. 6a and 6c). Further, in the region close to fracture,

3
Conference proceedings, 7th International Conference on Creep, Fatigue and Creep-Fatigue Interaction,
19-22 January 2016, IGCAR, Kalpakkam, India, D.P.R. Palaparti et al.

relatively higher number of large sized cavities can be seen at 60 MPa than at 125 MPa, in
the SEM micrographs (Figs. 6b and 6d). Figures 7a and 7b show the SEM micrographs in the
regions away from the fracture surface for 60 and 125 MPa, respectively. Equi-axed grains
with lath and grain boundaries decorated with coarse precipitates can be seen in the
micrograph. At low stresses, transformation of lath structure giving rise to relatively larger
precipitates in the matrix regions can be seen in Fig. 7a.

Fig. 5. Fracture surface of creep tested specimen of P9 steel at 873 K for the applied
stress of (a) 60 and (b) 125 MPa.

Fig. 6. Microstructure at regions close to fracture surface (e.g. Zone-1 in Fig.2) of the
creep tested specimen at 873 K for the applied stress of (6a and 6b) 60 and (6c and
6d) 125 MPa for P9 steel.

In the vicinity of Zone-1, the triaxial stress would be high and the microstructure reflects the
excess strain accumulated during post necking regime in the tertiary stage of creep. Hence the
microstructure observed may be attributed to the coarsened precipitates and presence of large
voids nucleating from decohesion of precipitate-matrix interface. From the tensile fracture

4
Conference proceedings, 7th International Conference on Creep, Fatigue and Creep-Fatigue Interaction,
19-22 January 2016, IGCAR, Kalpakkam, India, D.P.R. Palaparti et al.

analysis of thermally exposed 9Cr-1Mo steel, Blach et al [13] has observed increasing size of
secondary precipitates and decohesion between precipitate/matrix. The large sized cavities
observed at the low stress of 60 MPa (Fig. 6b) result from voids nucleating from decohesion
of coarse precipitates. The secondary crack density near the fracture surface (Zone-1) was
higher than that observed at regions away from the fracture surface (Zone-2), at all stress
conditions. This is consistent with the fracture behaviour reported by Baker et al [4] in case
of smooth and double-edge-notched test on 9Cr-1Mo steel. In the regions away from the
fracture surface (Zone-2), observed microstructure is primarily signifies the damage
accumulated due to coarsening of precipitation. It is known that the coarsening of the
precipitates such as M23C6 and Laves phase and sub-grain coarsening result in degradation of
creep strength with increasing creep strain in 9Cr-1Mo steel [5,12]. The observed high creep
ductility, grain flow, transgranular ductile fracture and significant coarsening of precipitates
suggest that the creep behaviour of P9 steel is controlled by the domination of microstructural
degradation.

Fig. 7. Microstructure at regions away from fracture surface (e.g. Zone-2 in Fig.2) of the
creep tested specimen at 873 K for the applied stress of (a) 60 and (b) 125 MPa for
P9 steel.

4.0 CONCLUSIONS

Creep rupture tests carried out on 9Cr-1Mo steel in normalised and tempered condition
revealed that the stress dependence of minimum creep rate and rupture life exhibited two
slopes behaviour with distinct stress exponents at 873 K. Extensive tertiary creep and
transgranular ductile fracture observed suggest that the creep deformation of P9 steel is
governed by microstructural degradation, in both the stress regimes. Microstructure close to
fracture surface exhibited excessive grain flow, coarse precipitates and large number of
secondary cracks originating from triaxiality developed during necking. In regions away from
the fracture surface, coarsened precipitates and negligible secondary cracks have been
observed.

REFERENCES

1. R.L.Klueh, Int. Mater. Rev., 50 (2005) p.287.


2. B.J. Cane, Acta Metall. 29 (1981) p.1581.
3. S.J. Sanderson: Ferritic Steels for High Temperature Applications,Ashok K. Khare, ed.,
ASM, Metals Park, OH, 1981, pp. 85-99.
4. E. Barker, G.J. Lloyd, and R. Pilkington: Mater. Sci. Eng., 84,(1986) pp.49-64.
5. B.K. Choudhary, S. Saroja, K.B.S. Rao and S.L. Mannan, Metall. Trans., 30 (1999) pp.
2825-2834.

5
Conference proceedings, 7th International Conference on Creep, Fatigue and Creep-Fatigue Interaction,
19-22 January 2016, IGCAR, Kalpakkam, India, D.P.R. Palaparti et al.

6. B.K. Choudhary, K.B.S. Rao and S.L. Mannan, Trans. Indian Inst. Met., 52 (1999) pp.
327-336.
7. E.M. Haney, F. Dalle, M. Sauzay, L. Vincent, I. Tournié, L. Allais, B. Fournier, Mater.
Sci. and Eng. A 510–511 (2009) pp.99–103
8. B.K. Choudhary, Mater. Sci. Eng. A585 (2013) pp.1–9
9. D.P. Rao Palaparti, E. Isaac Samuel, B.K. Choudhary and M.D. Mathew, Procedia Eng.
55 (2013) pp.70–77
10. E. Isaac Samuel, B.K. Choudhary, D.P. Rao Palaparti and M.D. Mathew, Procedia Eng.
55(2013) pp.64–69
11. K.R. Williams, B. Wilshire, Mater. Sci. Eng. 28 (1977) p.189.
12. F. Abe, S. Nakazawa, Metall. Trans. A 23 (1992) p.3025
13. J. Blach, L. Falat, P. Sevc, Eng. Failure Analysis. 16 (2009) pp.1397-1403.

You might also like