Вы находитесь на странице: 1из 6

Available online at www.sciencedirect.

com

ScienceDirect
Procedia Engineering 86 (2014) 302 307

1st International Conference on Structural Integrity, ICONS-2014

Effect of Nitrogen Addition and Test Temperatures on Elastic-


Plastic Fracture Toughness of SS 316 LN
B. Shashank Dutt*, G. Shanthi, G. Sasikala, M. Nani Babu, S. Venugopal,
Shaju K Albert , A.K. Bhaduri and T. Jayakumar
Metallurgy and Materials Group, Indira Gandhi Centre for Atomic Research, Kalpakkam-603102, India
*
E-mail ID: shashank@igcar.gov.in

Abstract

In the present investigation J-R curves have been determined for SS 316 LN with nitrogen contents of 0.08, 0.14 and 0.22 wt% at
298, 653 and 823 K. Elastic-plastic fracture toughness (J0.2) corresponding to 0.2 mm crack extension, was determined from the
J-R curves. For all test temperatures, J0.2 values were the highest for the steel with 0.14 wt% nitrogen.
2014 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
2014 The Authors. Published by Elsevier Ltd.
(http://creativecommons.org/licenses/by-nc-nd/3.0/).
Peer-review under responsibility of the Indira Gandhi Centre for Atomic Research.
Peer-review under responsibility of the Indira Gandhi Centre for Atomic Research
Keywords: fracture toughness, J-R curve, DCPD, SS 316LN

1. Introduction

Austenitic stainless steel grade SS 316L(N) is used worldwide as high temperature structural material for
nuclear industry due to the combination of good high temperature mechanical properties, compatibility with coolant
liquid sodium, good weldability, and resistance to intergranular stress corrosion cracking (IGSCC) by reducing the
carbon content. SS 316L (N) has been chosen for the high temperature structural components of the 500 MWe
prototype fast breeder reactor (PFBR) which is in an advanced stage of construction at Kalpakkam. With the aim of
increasing the design life of the future reactors to 60 - 100 years, possibility of improving the creep life by
increasing the nitrogen content has been explored. Many of these components are subjected to cyclic thermo-
mechanical loading which induces different kinds of damage like creep, fatigue, creep-fatigue interaction etc.
Several studies on the effect of nitrogen content on the tensile [1], low cycle fatigue [2,3] and creep [4,5] behavior
have been undertaken; improved tensile, creep and stress corrosion behaviours have been observed in nitrogen
bearing steels. For the damage tolerant design of these components, inter alia the fatigue crack growth and fracture
toughness (J0.2) properties are necessary. This study addresses the determination of J-R curves of SS 316 LN at 298,

1877-7058 2014 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/3.0/).
Peer-review under responsibility of the Indira Gandhi Centre for Atomic Research
doi:10.1016/j.proeng.2014.11.042
B. Shashank Dutt et al. / Procedia Engineering 86 (2014) 302 307 303

653 and 823 K with nitrogen contents of 0.08, 0.14 and 0.22 wt%.

2. Experimental

SS316 LN plates (30 mm thick) were received from MIDHANI, India in solution-annealed (1323 K, I
hour) condition having chemical compositions, mentioned in table 1.

Table 1: Chemical composition (wt%) of SS316LN containing different nitrogen levels

C N Mn Cr Mo Ni Si S P Fe
0.027 0.08 1.7 17.53 2.49 12.2 0.22 0.0055 0.013 Bal.
0.025 0.14 1.74 17.57 2.53 12.15 0.2 0.0041 0.017 Bal.
0.028 0.22 1.7 17.57 2.54 12.36 0.2 0.0055 0.018 Bal.

The average grain sizes of the material in as-received condition were 88, 78 and 87 m respectively for
0.08, 0.14 and 0.22 % nitrogen contents. Compact tension (CT) specimens of 20 mm thickness were fabricated from
the plates as per ASTM E 399-06 [6] specifications. These were subjected to fatigue pre-cracking at room
temperature (298 K) to obtain a sharp fatigue crack ahead of the notch using a resonant fatigue machine. The crack
length at the end of pre-cracking (initial crack length a0 for the fracture test) was targeted between 25 and 30 mm
corresponding to an a0 W in the range of ~ 0.5-0.6, where W = Width of the specimen. The pre-cracking was
carried out in K (stress intensity factor) decreasing mode; as the crack propagated, K (stress intensity factor range =
0 .4 B y b 2
Kmax- Kmin) was gradually reduced such that the crack was grown at loads not exceeding Pm = as
( a 0 + 2W )
prescribed by ASTM E 1820 [7].

Then the test specimens were side-grooved to 20% net depth (2 mm deep groove on either surface along
crack growth path with 45 flank angle and 0.25 mm root radius) to restrict crack growth to the symmetry plane and
to minimize the thumbnail crack growth. J-tests were carried out at 298, 653 and 823 K using monotonic ramping at
a constant stroke rate of 0.01 mm/s in a servo-hydraulic machine, fully automated for test control and data
acquisition. The load-line displacement (LLD) was monitored using high resolution COD gages. A suitably
calibrated direct current potential drop (DCPD) device was used for online determination of crack length. The
limitations [8] of DCPD in differentiating potential drop due to material deformation and stable crack growth have
not been fully addressed. Chen et al [9] have discussed the need for correcting the crack length obtained using
DCPD in determination of J-R curves from unloading compliance tests for various materials to overcome these
limitations. High resolution load (P), LLD and crack length (a) data were recorded using the data acquisition
software integrated with the test application. The room temperature tested specimen was held at about 673 K for
about 30 minutes in order to heat-tint the fracture surfaces to identify the crack extension regime in the fracture test.
The samples were then pulled to fracture at a higher pulling rate at room temperature. The initial and final crack
lengths (ao and af) were optically determined using the nine-point average method prescribed in ASTM E1820.
From the tested fractographs, stretch zone width (SZW) measurements were done by SEM at a magnification of
50x. Averages of 20 areas per specimen were scanned for SZW measurements. A methodology for correcting the
crack length was developed [10] and applied to the present investigation and supplemented by use of SZW
measurements.

3. Results and Discussion

The tensile properties of SS 316 LN are mentioned in table 2.


From the P, LLD and a, J-R curves were determined as follows:
The J values were computed by the equations mentioned in section A2.4.2.2, ASTM E 1820 using eq.(1-3) as shown
304 B. Shashank Dutt et al. / Procedia Engineering 86 (2014) 302 307

below:
J = J el + J pl - (1)

K (i ) 2 1 2
J (i ) = +J - (2)
pl (i )
E
(i 1) A pl (i ) A pl (i 1) a (i ) a (i 1)
J pl (i ) = J pl (i 1) + 1 (i 1) - (3)
b(i 1) BN b(i 1)

where: Jel , Jpl are the elastic and plastic components of J, K = stress intensity factor, = Poissons ratio, Apl = plastic
area under LLD curve

b b
= 2.0 + 0.522 = 1.0 + 0.76
W W
b = uncracked ligament (W-a), W = Width of the specimen. The subscripts i and i-1 indicate the current and previous
measurements respectively.
For the initial crack blunting line, ASTM E1820 specifies, M = J a f = 2 , though for highly work
hardening materials, a value as high as 4 is allowed, provided it is justified by the data. Also, the French code for
design and construction of nuclear islands, RCC-MR [11] suggests a value of M = 4 for steels. ASTM E1820 also
mentions that any value (minimum M=2) could be taken based on test data, tensile properties and strain hardening
behavior of the material. Mills [12] recommended M=4 for 304 and 316 class of stainless steels and their welds. For
this type of materials [13-15] with low yield strength (YS) and high strain hardening exponents, the blunting line
equation can be considered as:
2 f
J = m YS (CTOD ) = 2( )(3a ) - (4)
3
In this study, we use the following equation:
J = 6 YS ( a ) - (5)
where YS = Yield strength (YS), f = flow stress =(YS+UTS)/2 , UTS= ultimate tensile strength

Table 2: Tensile properties of 316LN steels with different nitrogen levels at different temperatures

Nitrogen Test Yield Tensile Uniform Total Reduction in YS/U


content temperature strength strength elongation elongation area (%) TS
(wt%) (K) (MPa) (MPa) (%) (%)
298 274 595 53 66 37.5 0.46
0.08 653 147 450 27 34 40 0.32
823 136 417 24 32 44 0.32
298 328 670 40 57 39 0.48
0.14 653 164 500 31.5 40.5 42 0.32
823 150 463 27 35 43.5 0.32
298 369. 714 33 46 39 0.51
0.22 653 218 580 32 38 41 0.37
823 186 510 31 36 45 0.36

The J-R curves obtained using the correction procedures are shown in Fig.1 (a-c) for the three steels from
the tests at 298 K. The J-R curves determined at 653 and 823 K are shown in Fig.2 (a-b). The values of J0.2
determined as the intersection of 0.2 mm offset line (from blunting line) with the J-R curves are shown in Fig.3(a).
B. Shashank Dutt et al. / Procedia Engineering 86 (2014) 302 307 305

In Fig. 3(b) post-necking elongations are shown and it may be noted that J0.2 and post-necking elongations show
similar trends. For all test temperatures the J0.2 values were higher (Fig.3(a)) for 0.14 % nitrogen steel compared to
those with 0.08 % and 0.22 % N. This can be explained by considering the effect of nitrogen on strength and
ductility. At 298 K, strength (table 2) increased and total elongation (TE) decreased with nitrogen addition. The
increase in strength with nitrogen additions can be explained based on solid-solution contribution of nitrogen. The
other mechanisms by which nitrogen can contribute to strengthening include precipitation-strengthening, formation
of interstitial-solute-complexes, clustering, reduction in stacking-fault energy and order-strengthening [16]. Crack
blunting and stretch zone formation, which is dependent on the strain hardening behavior has a significant
contribution to the J0.2 values. The effects of nitrogen on the deformation characteristics, thus assume significance in
explaining the variations in fracture toughness. Thus the strain hardening parameters are expected to be strongly
influenced by nitrogen contents.

At 653 and 823 K, J0.2 values were lower than that at 298 K, in line with decrease in both strength (table 2)
and TE at these temperatures compared to 298 K. In this material dynamic strain ageing (DSA) occurs between 523
and 923 K [17]. One of the manifestations of DSA is the tendency for material to exhibit minimum ductility. Similar
influence of DSA on decrease in fracture toughness with temperatures was observed in SS 316L [18]. In a study of
low cycle fatigue behavior of the same steels as in the present work, the temperature range of occurrence of DSA
was also observed to be dependent on nitrogen contents [19]. In the present investigation at 653 and 823 K, DSA
can be expected to contribute to reduced fracture toughness. However, tests at a wider range of temperatures are
necessary to conclusively establish this.

2500
a b
2000
2000

1500 blunting line offset line


1500
-2

-2
J, kJ.m

1000
J, kJ.m

1000

500 500

0 0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
a, mm
a, mm

2000
c

1500
-2
J, kJ.m

1000

500

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
a, mm

Fig. 1 (a-c): J-R curves for SS 316 LN at 298 K containing (a) 0.08 wt% N (b) 0.14 % and (c) 0.22 % /N contents
306 B. Shashank Dutt et al. / Procedia Engineering 86 (2014) 302 307

1100
a 653 K 1000
b
823 K
2000
900
800

-2
1500 700

J, kJ.m
600
500
-2
J, kJ.m

1000 400
300
0.08 %
200
500 0.08 % N 0.14 %
100 0.22 %
0.14 % N
0.22 % N 0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5
0 a, mm
0 1 2 3 4 5 6 7
a, mm

Fig. 2 (a-b): J-R curves for SS 316 LN at (a) 653 and (b) 823 K having different nitrogen contents
1300 18
1200 a 298 K 17 298 K b
1100 653 K 16 653 K

Post necking elongation %


823 K 15 823 K
1000
14
900 13
800 12
700 11
-2
J0.2, kJ.m

600 10
9
500
8
400 7
300 6
200 5
100 4
0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20 0.22 0.24
0.06 0.08 0.10 0.12 0.14 0.16 0.18 0.20 0.22 0.24
Nitrogen content (wt%) Nitrogen content (wt%)

Fig. 3 (a-b): Variation of (a) fracture toughness results and (b) Post-necking elongations for SS
316 LN with nitrogen contents at different test temperatures

4. Conclusions

The elastic-plastic fracture behavior of SS 316 LN was studied with varying nitrogen additions at 298, 653
and 823 K. The contribution of crack tip deformation to the DCPD signal was accounted for in crack length
estimation. J-R curves were established using these corrected crack lengths. The J0.2 values were higher at 0.14 % at
all test temperatures. At 653 and 823 K decrease in J0.2 values occurred for all nitrogen additions, compared to 298
K.

Acknowledgements

The authors wish to acknowledge the contributions of Mr. Syed Meer Kaleem and Ms. Sreevidya for
experimental support.

References

1. Ganesan V, Mathew MD and Rao KBS, Mater Sci Technol 25 (2009) 614 -618.
2. Reddy GVP, Sandhya R and Rao KBS, Procedia Eng 2 (2010) 2181 -2188.
B. Shashank Dutt et al. / Procedia Engineering 86 (2014) 302 307 307

3. Mathew MD and Kim DW, Mater Sci and Eng A 474 (2008) 247 -253.
4. Sasikala G, Mathew MD and Rao KBS and Mannan SL, Metall Mater Trans 31A (2000) 1175 -1185.
5. Ganesan V, Mathew MD, Parameswaran P and Rao KBS, Trans Indian Inst Metals 63 (2010) 417 -421.
6. ASTM Standard Test Method for Linear-Elastic Plane-Strain Toughness KIC of Metallic Materials, ASTM
E399-06.
7. ASTM Standard Test Method for measurement of fracture toughness, ASTM E1820-06.
8. Lowes JM and Fearnehough GD, Eng Fract Mechanics 3 (1971) 103-108.
9. Chen X, Nanstad RK, Sokolov MA, in Transactions, SMIRT-22, California, USA, August 2013, Division-
II.
10. Shashank Dutt B and Sasikala G (unpublished results).
11. Design and construction rules for mechanical components for FBR nuclear islands, RCCMR (Edition
2002), Section Isubsection Z: Technical appendix A16, Chapter A16.9000 (Material properties) 185.
12. Mills WJ, Int Mater Review, No.2, 42 (1997) 45-82.
13. Mills WJ, J Test Eval, 9 (1981) 56-62.
14. Chipperfield CG, Int J Fract, 12 (1976) 873-886.
15. Paranjpe SA, Banerjee S, Eng Fract Mech, 11(1979) 43-53.
16. Mathew MD and Srinivasan VS, chapter 7, in High nitrogen steels and stainless steels (eds) Mudali UK,
High nitrogen steels, (ASM.) Narosa publishing house, New Delhi, India (2004) p 182.
17. Samuel KG, Mannan SL and Rodriguez P Acta Metall 36 (8) (1988) 2323-2327.
18. Samuel KG, Gossman O and Huthmann K, Int.J.Pres.Ves& Piping 41 (1990) 59-74.
19. Reddy GVP, Sandhya R, Rao KBS and Sankaran S, Procedia Engineering 2 (2010) 21812188.

Вам также может понравиться