You are on page 1of 6

Corrosion Science 66 (2013) 211–216

Contents lists available at SciVerse ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

Evaluation of the effect of grain size on chromium carbide precipitation and


intergranular corrosion of 316L stainless steel
Shu-Xin Li ⇑, Yan-Ni He, Shu-Rong Yu, Peng-Yi Zhang
School of Petrochemical Engineering, Lanzhou University of Technology, Lanzhou 730050, China

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

Article history: The effect of grain size on chromium carbide precipitation and sensitization of 316L stainless steel was
Received 27 July 2012 investigated based on inspection of microstructures and electrochemical potentiokinetic reactivation
Accepted 18 September 2012 test. Various grain size samples were produced by heat treating the base material at 1100 °C for different
Available online 27 September 2012
durations. The result showed that chromium carbide precipitations were much delayed in larger grains.
The degree of sensitization decreases with increasing grain size. The diffusion bonded joint has good
Keywords: resistance to intergranular corrosion due to the coarsened grains and the increased percentage of low
A. Stainless steel
energy coincident site lattice, while the grain coarsening is the main contribution.
B. Polarization
C. Intergranular corrosion
Ó 2012 Elsevier Ltd. All rights reserved.

1. Introduction c) fine-grained microstructure, which promotes a very rapid pre-


cipitation of carbides. While this 2-phase microstructure is gener-
Failures due to intergranular corrosion and intergranular stress ally absent in 316 stainless steel.
corrosion cracking in structures and components of austenitic Any processing that changes the reactivity of a surface will
stainless steel materials have always been a tough problem in engi- cause a change in corrosion response of austenitic stainless
neering practice. Effort has been made by reducing carbon content, steels. In diffusion bonding of 316L stainless steel, the bonded
changing chemical composition, increasing grain size and intro- joint was kept in the furnace at 1100 °C under 10 MPa for 3 h.
ducing appropriate strain to shift the time–temperature–sensitiza- The long exposure at high temperature promoted severe grain
tion curve to the right, thus increasing the time of sensitization and coarsening in the diffusion bonded joint. The joint experienced
desensitization process and leading to an increase in intergranular the temperature range 900–450 °C for around 45 min during
corrosion resistance. Researchers have been extensively conducted the cooling process, which could increase the sensitization sus-
on sensitization of austenitic stainless steel in terms of testing ceptibility for austenitic stainless steels and thus the corrosion
methods [1–10], influences of strain and strain state [11–14] and response would be changed correspondingly. However, the pre-
grain size [15–17] on carbide precipitation, aging condition on vious work [28] indicated that the diffusion bonded joint has
chromium carbide precipitation [18] and models for prediction of much less susceptibility to intergranular corrosion than the base
chromium depletion [19,20]. Also, increasing coincident site lattice material although suffered long exposure to the sensitization
(CSL) in austenitic stainless steels by grain boundary engineering temperature. No chromium carbide precipitation was observed
has been an effective way to enhance resistance to intergranular in the joint even after 100 h0 treatment at 650 °C. It showed very
corrosion [21–27]. Among these studies with regard to the effect low susceptibility to intergranular corrosion in electrochemical
of grain size on sensitization of austenitic stainless steels, most potentiokinetic reactivation tests due to coarsened grains and in-
of them dealt with 304 stainless steel. Only a few talks about the creased percentage of twin-induced low CSL boundaries in the
effect of grain size on intergranular corrosion in 316 stainless steel. joint. However, the degree of sensitization in electrochemical
It has been proved [17] that the carbide precipitation and sensiti- test is the result of combined effect of microstructural texture,
zation in the deformed 304 stainless steel are different from those grain size and electrochemical parameters. In a given electro-
in the deformed 316 stainless steel because a high volume fraction chemical condition, how the grain size affects the electrochemi-
of strain-induced, a0 -martensite in 304 can nucleate a 2-phase (a0 / cal behavior of intergranular corrosion and to what extent the
sensitization susceptibility changes with the grain size have
not been clearly understood for 316L stainless steel.
⇑ Corresponding author. School of Petrochemical Engineering, Lanzhou University
Therefore, in the present study, the effect of the grain size on
of Technology, 287 Langongping Road, Lanzhou 730050, China. Tel.: +86 931
2973647; fax: +86 931 2973648. susceptibility to intergranular corrosion of 316L stainless steel
E-mail address: li_shuxin@163.com (S.-X. Li). was studied. Samples with various grain sizes were produced by

0010-938X/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.corsci.2012.09.022
212 S.-X. Li et al. / Corrosion Science 66 (2013) 211–216

heat-heating the base material at 1100 °C for different durations 3. Results and discussion
and then sensitized at 650 °C. Double loop electrochemical poten-
tiokinetic reactivation test (DLEPR) has been successfully em- 3.1. Microstructures of samples with various grain sizes
ployed to characterize the degree of sensitization of austenitic
stainless steel [29–36]. Thus, in this study, the DLEPR test was con- Fig. 1 shows the microstructures of samples with various grain
ducted on samples of various grain sizes of 316L stainless steel and sizes. Clearly, the grain size increased as the heat-treatment dura-
the diffusion bonded joint to evaluate their intergranular corrosion tion increased from 0.5 to 3 h. The maximum and the average grain
behavior. Microscopic structures were inspected after sensitization sizes were measured using an image analysis software and are
treatment and DLEPR tests. listed in Table 1. The grains grew initially from 46 to 55 lm at
0.5 h and then increased dramatically after 1 h duration and the
average grain size doubled at 1.5 h. As the duration increased to
2. Experimental procedure 3 h, the average grain size reached 173 lm, four times larger than
the base material. It can also be seen from the microstructures that
2.1. Diffusion bonding of 316L austenitic stainless steel the longer the duration, the more unevenly distributed the grain
size. The huge and the tiny grains coexist in the sample, Fig. 1(d).
As received cold-worked 316L stainless steel with the following Also included in Table 1 is the grain size of the diffusion bonded
chemical composition in wt.%: C-0.01, Si-0.41, Mn-1.41, P-0.036, S- joint. It should be noted that, different from the base material heat-
0.006, Ni-12.43, Cr-17.84, Mo-2.16, Fe-balance was used in this treated at 1100 °C for 3 h, the diffusion bonded joint experienced
study. 316L stainless steel was diffusion bonded at temperature 1100 °C for 3 h but with 10 MPa pressure applied on the joint.
of 1100 °C under pressure of 10 MPa for holding time of 3 h in vac- Grains in the diffusion bonded joint were dramatically coarsened
uum of 0.00133 Pa. The whole bonding process was conducted un- with the maximum grain size of 605 lm and average grain size
der vacuum and the joint was cooled down in the furnace from of 220 lm, larger than the base material heat-treated for 3 h with
1100 °C to the room temperature after 3 h holding time. The joint the average grain size of 173 lm. In addition, a large number of
experienced the temperature range of 850–450 °C for around twins were found in both heat-treated samples and the diffusion
45 min during the cooling process. The inspection of the interface bonded joint.
and the mechanical properties of the bonded joint can be seen in
the previous study [37].
3.2. Effect of the grain size on chromium carbide precipitations at grain
boundaries
2.2. Preparation of various grain sizes samples
The author’s previous study on the sensitization of the base
Grain size variation was achieved by heat-treating the 316L material and the diffusion bonded joint of 316L stainless steel
stainless steel base material at 1100 °C for various durations rang- has showed [28] that chromium carbide precipitations could be
ing from 0.5 to 3 h, then aged at the sensitization temperature of seen clearly in the base material after 8 h treatment at 650 °C while
650 °C for durations of 2, 8, 30, 50 and 100 h. Samples were almost no precipitations were observed in the diffusion bonded
grounded and polished and then etched with chloroazotic acid to joint after 100 h. One of the main reasons contributed to the great
reveal the microstructure. Detailed microstructure inspection improvement of intergranular corrosion resistance of the joint is
was conducted on these samples with optical microscope. the coarsened grains. Fig. 2 shows the microstructures of sensitized
samples of various grain sizes. A number of chromium carbides
precipitation were found in the sample of the average grain size
2.3. DLEPR test procedure
of 55 lm for 20 h durations, Fig. 2(a), but no precipitations were
observed at 8 h. As mentioned above that the base material of
10  10  10 mm samples were cut from the diffusion bonded
the average grain size of 46 lm sensitized at 8 h, having less sen-
joint and the heat treated samples for DLEPR tests. Samples were
sitization time than the sample of 55 lm. When the grain size in-
mounted in epoxy resin with brass rod welded to the surface and
creased to 77 lm, chromium carbides started to precipitate after
then successively ground to 1000 grit abrasive paper. The working
20 h and were clearly seen at 30 h, shown in Fig. 2(b). But for the
solution was 1.0 mol/L H2SO4 + 0.003 mol/L Na2S4O6. The electro-
samples with the average grain sizes of 89 and 173 lm in Fig. 2
chemical testing system consists of three electrodes (saturated cal-
(c) and (d) respectively, only a few chromium carbide precipitates
omel electrode (SCE) as reference electrode and graphite electrode
were found at 50 and 100 h durations. The grain size affects the
as auxiliary electrode) and Powersuite software which was used to
time to start the sensitization and to reach the complete sensitiza-
process electrochemical data.
tion. Increasing the grain size can dramatically delay the onset of
Specimen was immersed into the solution for 5–10 min to ob-
sensitization as larger grains have much wider chromium de-
tain open circuit potential (about 430 mV (SCE)). The specimen
pleted-zones and lower grain boundary chromium concentrations.
was started to polarize from a potential of 500 mV (SCE) and
The study on measuring the sensitization rates and M23C6 precip-
kept at this potential for 2 min, followed by anodic polarization
itation behavior over a range of grain sizes from 15 to 150 lm in
from 500 mV (SCE) to the reverse potential 400 mV (SCE). Then
304 stainless steel also showed [13] that the sensitization process
the specimen was cathodically polarized to the open circuit
is further accelerated as the grain size decreases. The precipitation
potential. The degree of sensitization can be characterized by
of Cr-rich carbides and distribution of chromium concentration by
determining the reactivation ratio Rr (ir/ia) in polarization curve.
cellular automaton simulation indicated [38] that the precipitation
ir is the maximum current density of the reverse scan (cathodic)
of Cr-rich carbides of the large grain microstructure is less than
and ia is the maximum current density of the forward (anodic)
that of the small grain microstructure.
scan. The higher the ratio, the higher the degree of sensitization.
The test was performed with the reverse potential of 400 mV
(SCE), scan velocity of 1.111 mV/s, solution temperature of 3.3. Effect of the grain size on the electrochemical behavior
40 °C and 1.0 mol/L H2SO4 + 0.003 mol/L Na2S4O6. Rr was the
average of results of three samples which were tested using Typical DLEPR polarization curves for samples of various grain
the same electrode. sizes were presented for 20 h in Fig. 3(a) and 50 h in Fig. 3(b).
S.-X. Li et al. / Corrosion Science 66 (2013) 211–216 213

Fig. 1. Microstructures of samples with various grain sizes produced by heat treating the base material at 1100 °C for various durations (a) Base material (0 h) (b) 0.5 h (c) 2 h
(d) 3 h.

Table 1
The maximum and the average grain sizes of samples.

No. BM 1 2 3 4 5 6 DBJ
Durations at 1100 °C (h) 0 0.5 1.0 1.5 2.0 2.5 3.0 3.0
Max grain size Dmax (lm) 187 202 285 296 356 389 407 605
Average grain size Dav (lm) 46 55 77 89 110 145 173 220

Note: BM represents the base material; DBJ represents the diffusion bonded joint.

Fig. 2. Microstructures of samples with various grain sizes sensitized at 650 °C for various durations of (a) 55 lm, 20 h (b) 77 lm, 30 h, (c) 89 lm, 50 h (d) 173 lm, 100 h.
214 S.-X. Li et al. / Corrosion Science 66 (2013) 211–216

Fig. 4. The change of Rr with the average grain size for various samples sensitized at
650 °C for 100 h, 50 h, 20 h and 2 h.

electrochemical response of the sample, so the bigger the grains


in the sample, the smaller Rr is, causing some error in testing data.
An attempt was made to set up a relationship between the grain
size and the reactivation ratio Rr at the given condition. But it is dif-
ficult to find as the Rr is the result of combined effect of the testing
parameters (such as the exposure time, temperature and scan
velocity, etc.), the grain size and the texture of the material. A small
variation in these parameters may result in a big change in Rr,
especially the reactivation potential and the solution temperature
[28]. Also, the Rr varies with the sensitization durations as shown
in Fig. 4. In fact, effort has been made to develop a relationship be-
tween the grain size and the corrosion resistance [15], but hard to
follow a specific relationship. Ralston and Birbilis [39] intended to
set up a Hall–Petch relationship between grain size and corrosion
response but failed due to the inherent difficulty of measuring
Fig. 3. Comparison of DLEPR polarization curves for various grain sizes (a) Grain the true grain size from two dimensional images and the effect
sizes of 77 lm and 173 lm sensitized at 650 °C for 20 h (b) Grain sizes of 55 lm and
of grain refinement processing on corrosion response.
145 lm sensitized at 650 °C for 50 h.
Comparing the sample heat-treated at 1100 °C for 3 h (the aver-
age grain size of 173 lm) with the diffusion bonded joint of
220 lm, it can be seen that the latter has smaller Rr than the for-
The reactivation ratio Rr decreased from 0.1114 to 0.0821 with
mer with the exception of 100 h sensitized sample. Also a few pre-
increasing grain size from 77 to173 lm in Fig. 3(a), indicating that
cipitations were found in the 3 h heat-treated sample in Fig. 2(d),
the bigger the grain size, the less the susceptibility to intergranular
while almost no chromium carbides precipitated at grain bound-
corrosion is. The same tendency of Rr exists in Fig. 3(b) for 50 h, the
aries in the diffusion bonded joint after 100 h, indicating that the
reactivation ratio Rr decreased from 0.1936 to 0.0988 with increas-
diffusion bonded joint has less susceptibility to intergranular cor-
ing grain size from 55 to145 lm. The DLEPR results are consistent
rosion than the 3 h heat-treated base material at 1100 °C.
with the microstructure inspections.
Fig. 4 also tells the effect of the sensitization duration on the de-
The change of Rr with the average grain size was plotted in
gree of sensitization. As the sensitization time increases, Rr in-
Fig. 4. As can be seen that the biggest Rr always occurred in the
creases correspondingly. This is easily understood as the longer
smallest grain size at all heat-treat durations. The samples sensi-
exposure to the sensitization temperature, the more chromium
tized at 2 and 20 h follow the trend of Rr decreasing with increasing
carbides precipitate.
grain size except for the grain of 110 lm at 2 h. But as the grain
size increased, the scatter exits in the data and it is hard to follow
this trend. This is especially true in the 100 h’ sensitized samples. 3.4. Effect of the texture on intergranular corrosion
The biggest and the smallest Rr happen at the grains of 55 and
77 lm respectively, while the rest of Rr scatter around 0.1530. This It is generally accepted that deformation has a big effect on
can be explained that as the grain size is greater than 100 lm, it is electrochemical response of austenitic stainless steels. Although
getting harder for the chromium carbide to precipitate on the grain the stain in the 316L stainless steel diffusion bonded joint is no
boundaries and more time is needed for the precipitation, as more than 0.05% [37], the texture of the joint has been changed
shown in Fig. 2(c) and (d) in which there are only a few precipita- due to the deformation in the diffusion bonding process. As illus-
tions observed on the samples. On the other hand, the grains are trated above that the sample heat-treated at 1100 °C for 3 h (0%
becoming even more unevenly distributed as the grain size ex- strain) has less resistance to intergranular corrosion than the diffu-
ceeds 100 lm, which leads to a big difference in grain size distribu- sion bonded joint on which 10 MPa pressure was applied in diffu-
tion in the electrochemical sample. Each Rr reflects the sion bonding process. It is difficult to isolate the grain size effect
S.-X. Li et al. / Corrosion Science 66 (2013) 211–216 215

Fig. 5. Microstructures of samples after the DLEPR test for various grain sizes (a) 55 lm (b) 77 lm (c) 110 lm (d) 173 lm.

from other micro-structural changes which are caused by the ther- sample of the average grain size of 55 lm in Fig. 5(a), that it was
momechanical processing used to achieve variation in grain size almost entirely sensitized with large amount of chromium carbides
[15]. Compared to the strain of ranging from 5% up to 20% which precipitated at grain boundaries. As the grain size increased, the
greatly changes the texture of the austenitic stainless steels precipitations could be clearly seen but only parts of grains were
[12,13,16], the small strain of 0.05% in the diffusion bonded joint sensitized in the sample of 77 lm in Fig. 5(b) and slight intergran-
can be neglected. But this small deformation which was to acceler- ular corrosion was observed in the sample of 110 lm, Fig. 5(c).
ate the diffusion of atoms did promote grain coarsening and an in- While no intergranular corrosion occurred in the sample of
P
crease of twin boundaries proportion. The frequency of 3 CSL 173 lm, Fig. 5(d). The microstructure analysis is consistent with
boundaries of 316L diffusion bonded joint is 59.4%, higher than the reactivation result. The DLEPR results provide the further evi-
P
that of the base material of 36% and the 3 CSL boundaries dom- dence that the bigger the grain size of 316L stainless steel, the less
inated the whole joint with about 96.5% [28]. The fraction of low- susceptibility to the intergranular corrosion resistance.
CSL boundaries of 316 stainless steel for 1100 °C solution treat-
ment is only 28.8% in Yu and Chen’s study [38], much less than that
of the diffusion bonded joint as their sample was only solution 4. Conclusions
treated without applying mechanical force, in which the grain size
increased but with no apparent twins produced. The effect of the grain size on the intergranular corrosion of
Therefore, it can be concluded that both the grain coarsening 316L stainless steel was investigated. Both the DLEPR tests and
and the increased amount of twin-induced grain boundaries are microstructure inspections showed that the 316L stainless steel
contributed to the increase of intergranular corrosion resistance has less susceptibility to intergranular corrosion as the grain size
of the 316L diffusion bonded joint. But compared with the in- increased. Samples with the average grain size above 89 lm were
creased amount of twin-induced grain boundaries, the grain coars- found to have only a few chromium carbide precipitations on grain
ening is the main contribution. boundaries. The long duration at high temperature and the small
Many studies focused on improving the intergranular corrosion strain caused the grain coarsening and the increase in the amount
resistance of austenitic stainless steels by changing the texture of of the twin boundaries, leading to the great improvement of inter-
austenitic stainless steel through grain boundary engineering granular corrosion resistance of the 316L diffusion bonded joint.
[21–27], such as introducing the strain to increase the percentage But the grain coarsening is the main contribution. The study in
of low energy grain boundary and twin boundaries. However, the the present paper suggests that increasing grain size at an opti-
study in the present paper indicates that increasing grain size at mum level could be an effective way to increase the intergranular
an optimum level could also be a way to reduce the susceptibility corrosion resistance of 316L stainless steel, but it is a big challenge
to intergranular corrosion resistance for 316L stainless steel. But to obtain optimized grain size for improvement of corrosion resis-
increasing the grain size will lead to a decrease in ductility. It is tance without losing its good mechanical properties.
a big challenge to obtain optimized grain size for improvement
of intergranular corrosion resistance without losing its good
mechanical properties. Acknowledgements

3.5. Microstructures of samples of various grain sizes after DLEPR test The authors are grateful for the supports provided by China
Natural Science Foundation (No. 50805072) and PHD foundation
Fig. 5 presents the microstructures of various grain sizes sensi- of Lanzhou University of Technology (SB05200801) for which due
tized at 650 °C for 50 h after the DLEPR test. It can be seen from the acknowledgement is given.
216 S.-X. Li et al. / Corrosion Science 66 (2013) 211–216

References [19] M.K. Samal, A. Abhishek, Proc. Inst. Mech. Eng. Part C. 225 (2011) 809–815.
[20] H. Sahaloui, H. Sidhom, J. Philibert, Acta Mater. 50 (2002) 1383–1392.
[21] M. Michiuchi, Z.J. Wang, Acta Mater. 54 (2006) 5179–5184.
[1] A.H. Tuthill, Weld. J. 5 (2005) 36–40.
[22] H.Y. Bi, H. Kokawa, Z.J. Wang, Scr. Mater. 49 (2003) 219–223.
[2] W.E. White, Mater. Charact. 28 (1992) 349–358.
[23] S. Tsurekawa, S. Nakamichi, T. Watanabe, Acta Mater. 54 (2006) 3617–3626.
[3] M. Matula, L. Hyspecka, M. Svoboda, Mater. Charact. 46 (2001) 203–210.
[24] E.M. Lehockey, G. Palumbo, P. Lin, Scr. Mater. 36 (1997) 1211–1218.
[4] H. Shaikh, N. Sivaibharasi, B. Sasi, T. Anita, R. Amirthalingam, Corros. Sci. 48
[25] R. Jones, V. Randle, Mater. Sci. Eng., A 527 (2010) 4275–4280.
(2006) 1462–1482.
[26] T. Sadahiro, N. Shinya, W. Tadao, Acta Mater. 54 (2006) 3617–3626.
[5] M.G. Pujar, N. Parvathavarthini, R.K. Dayal, S. Thirunavukkarasu, Corros. Sci. 51
[27] H. Kokawa, M. Shimada, M. Michiuchi, Z.J. Wang, Y.S. Sato, Acta Mater. 55
(2009) 1707–1713.
(2007) 5401–5407.
[6] H.E. Bühler, L. Gerlach, O. Greven, Corros. Sci. 45 (2003) 2325–2336.
[28] S.X. Li, L. Li, S.R. Yu, R. Akid, H.B. Xia, Corros. Sci. 53 (2011) 99–104.
[7] B. Deng, Y. Jiang, J. Xu, T. Sun, J. Gao, Corros. Sci. 52 (2008) 969–977.
[29] G.C. Sandip, S. Raghuvir, Scr. Mater. 58 (2008) 1102–1105.
[8] J. Gong, Y.M. Jiang, B. Deng, J.L. Xu, J.P. Hu, Electrochim. Acta 55 (2010) 5077–
[30] G.H. Aydoğdu, M.K. Aydinol, Corros. Sci. 48 (2006) 3565–3583.
5083.
[31] A.S. Lima, A.M. Nascimento, H.F.G. Abreu, P. De Lima-Neto, J. Mater. Sci. 40
[9] C. Garcia, M.P. de Tiedra, Y. Blanco, O. Martin, F. Martin, Corros. Sci. 50 (2008)
(2005) 139–144.
2390–2397.
[32] P. De Lima-Neto, J.P. Farias, L.F.G. Herculano, H.C. de Miranda, et al., Corros. Sci.
[10] Y. Cui, C.D. Lundin, Mater. Lett. 59 (2005) 1542–1546.
50 (2008) 1149–1155.
[11] E.A. Trillo, L.E. Murr, Acta Mater. 47 (1998) 235–245.
[33] A. Arutunow, K. Darowicki, Electrochim. Acta 54 (2009) 1034–1041.
[12] L.E. Murr, A. Advani, S. Shankar, D.G. Atteridge, Mater. Charact. 39 (1997) 575–
[34] S.S.M. Tavares, V.F. Terra, P. De Lima-Neto, J. Mater. Sci. 40 (2005) 4025–4028.
598.
[35] M. Dadfar, M.H. Fathi, F. Karimzadeh, M.R. Dadfar, A. Saatchi, Mater. Lett. 61
[13] E.A. Trillo, R. Beltran, J.G. Maldonado, Mater. Charact. 35 (1995) 99–112.
(2007) 2343–2346.
[14] Y. Fu, X. Wu, E. Han, Electrochim. Acta 55 (2009) 1618–1629.
[36] K.H. Lo, C.T. Kwok, W.K. Chan, Corros. Sci. 53 (2011) 3697–3703.
[15] K.D. Ralston, N. Birbilis, Corrosion 66 (2010) 1–13.
[37] S.X. Li, F.Z. Xuan, S.T. Tu, Mater. Sci. Eng., A 480 (2008) 125–129.
[16] N. Parvathavarthini, S. Mulki, P.K. Dayal, Corros. Sci. 51 (2009) 2144–2150.
[38] X. Yu, S. Chen, Y. Liu, F. Ren, Corros. Sci. 52 (2010) 1939–1947.
[17] R. Beltran, J.G. Maldonado, Acta Mater. 45 (1997) 4351–4360.
[39] K.D. Ralston, D. Fabijanic, N. Birbilis, Electrochim. Acta 56 (2011) 1729–1736.
[18] H. Sahlaoui, K. Makhlouf, H. Sidhom, J. Philibert, Mater. Sci. Eng., A 372 (2004)
98–108.

You might also like