Sie sind auf Seite 1von 10

Available online at www.sciencedirect.

com
Available online at www.sciencedirect.com

ScienceDirect
ScienceDirect
Available
online
www.sciencedirect.com
Available
online
at at
www.sciencedirect.com
Structural Integrity Procedia 00 (2016) 000000

ScienceDirect
ScienceDirect

Structural Integrity Procedia 00 (2016) 000000

Procedia
Structural
(2016)
612621
Structural
IntegrityIntegrity
Procedia200
(2016)
000000

www.elsevier.com/locate/procedia
www.elsevier.com/locate/procedia

www.elsevier.com/locate/procedia

21st European Conference on Fracture, ECF21, 20-24 June 2016, Catania, Italy
21st European Conference on Fracture, ECF21, 20-24 June 2016, Catania, Italy

A three-dimensional model of non-slipping stress corrosion


cracking under low loads
cracking under low loads
Thermo-mechanical
modeling
of a high
Longkui
Zhu a, b *, Zhengcao
Li apressure
, Lijie Qiao bturbine blade of an
a, b
a
Longkui
Zhu *,
Zhengcao
Li ,engine
Lijie Qiao b
airplane
gas
turbine
State Key Laboratory of New Ceramics & Fine Processing, Key Laboratory for Advanced Materials of Ministry of Education, School of

XV Portuguese
Conference on Fracture,
PCFof
2016,
10-12 Februarystress
2016, Pao
de Arcos, Portugal
A three-dimensional
model
non-slipping
corrosion

a
a

Materials
Science
and Engineering,
Tsinghua
University,
Beijing
100084, Materials
Peoples Republic
of China
State Key Laboratory
of New
Ceramics
& Fine Processing,
Key
Laboratory
for Advanced
of Ministry
of Education, School of
a
b
c
Materials Science and Engineering, Tsinghua University, Beijing 100084, Peoples Republic of China

P. Brando , V. Infante , A.M. Deus *

a
Corrosion
and Protection
Center,
Key Laboratory
for environmental
fracture
of Ministry
Education,
SchoolPais,
of Materials
Science
and
Department
of Mechanical
Engineering,
Instituto
Superior Tcnico,
Universidade
de of
Lisboa,
Av. Rovisco
1, 1049-001
Lisboa,
University
of Science and
TechnologyPortugal
Beijing,
Beijing
100084,
Republic
Corrosion and Engineering,
Protection Center,
Key Laboratory
for environmental
fracture
of Ministry
of Peoples
Education,
School of
of China
Materials Science and
b
IDMEC, Department
of Mechanical
Engineering,
SuperiorBeijing,
Tcnico,
Universidade
Lisboa,Republic
Av. Rovisco
Pais, 1, 1049-001 Lisboa,
Engineering,
University
of Science Instituto
and Technology
Beijing
100084, de
Peoples
of China
Portugal
c
CeFEMA, Department of Mechanical Engineering, Instituto Superior Tcnico, Universidade de Lisboa, Av. Rovisco Pais, 1, 1049-001 Lisboa,
Portugal
Abstract
b
b

Abstract
Stress corrosion cracking (SCC) of 316L single-crystal austenitic stainless steel subjected to low loads (nom = 20-40 MPa)
Abstract
in a 45
% boiling
MgCl
was of
studied
synchrotron
X-ray stainless
computedsteel
tomography,
and so
on.
2 solution
MPa)
Stress
corrosion
cracking
(SCC)
316L using
single-crystal
austenitic
subjected finite
to lowelement
loads (analysis
nom = 20-40
Results
show
that
there
was
no
surface
slip
band
around
the
nucleation
sites
and
the
tips
of
short
cracks.
Three-dimensional
in
a
45
%
boiling
MgCl
solution
was
studied
using
synchrotron
X-ray
computed
tomography,
finite
element
analysis
and
so
on.
2
During their operation,
modern aircraft engine components are subjected to increasingly demanding operating conditions,
reconstruction
of
discontinuous
zig-zag
surface
SCC
crack
indicates
that
the
crack
was
continuous
inside
the
specimen.
The
Results
show
that
there
was
no
surface
slip
band
around
the
nucleation
sites
and
the
tips
of
short
cracks.
Three-dimensional
especially the high pressure turbine (HPT) blades. Such conditions cause these parts to undergo different types of time-dependent
obtaining
through
two-surface
analysis
manifests,
rather
than
{1 1that
1} slip
planes,
the continuous
cracks
extended
along
{1 able
0 0}toplanes
reconstruction
of
zig-zag
surface
SCCthe
crack
the (FEM)
crack was
inside
the
specimen.
The
degradation,
onediscontinuous
of which is trace
creep.
A model
using
finiteindicates
element
method
was
developed,
in
order
to be
predict
with
lowest
surface
energy.
Itblades.
is analysis
considered
that
and
dissolution
synergistically
led aalong
to commercial
the {1
SCC
advance,
obtaining
through
two-surface
trace
manifests,
rather than
{1
1 local
1}a slip
planes,
the
cracks
extended
0 0}
planes
thethe
creep
behaviour
of HPT
Flight
datamicrocleavage
records
(FDR)
for
specific
aircraft,
provided
by
aviation
and
microshear
was
alsoenergy.
ofItthe
primarily
SCCfor
mechanisms
underflight
the synergistically
high
load.
three-dimensional
with
the lowest
surface
is
considered
that microcleavage
and
local
dissolution
led
the SCC
company,
were
used
toone
obtain
thermal
and microscopic
mechanical
data
three
different
cycles.
In The
order
totocreate
theadvance,
3DSCC
model
model
wasfor
created
atalso
the
lowofstress
level,
where
a main
grew
MPD
due toThe
anodic
dissolution,
and
a
and
microshear
was
one
the
primarily
microscopic
SCC crack
mechanisms
underthe
thecomposition
high load.
three-dimensional
SCC
needed
the
FEM
analysis,
a HPT
blade
scrap
was SCC
scanned,
and
its along
chemical
and
material
properties
were
microdefect
was data
formed
onwas
thegathered
crack
the
microdefect
enlarged
toalong
a critical
microcracks
obtained.
The
wasWhen
fed
into
FEM
model
and
different
simulations
weretorun,
first dissolution,
with a nucleated
simplified
model
was created
atthat
the
low
stressfront.
level,
where
a the
main
SCC
crack
grew
the size,
MPDsecondary
due
anodic
and ata3D
rectangular
block
shape,
toshoulders.
betterWhen
establish
model,
and
thenpropagated
with
realto3D
mesh
obtained
from
thethrough
blade
scrap.
The
the
stress-concentrated
microdefect
Then,
the
microcracks
two
sides
of the
MPD
anodic
microdefect
was
formed
onintheorder
crack
front.
the the
microdefect
enlarged
to athe
critical
size,
secondary
microcracks
nucleated
at
expected
behaviour
termsshoulders.
of displacement
was
observed,
in the
particular
at the
trailing
edge
thediscontinuous
blade.through
Therefore
such a
dissolution,
microcleavage
or in
mciroshear,
resulting
in the
of
river-like
fractography
and
the
surface
theoverall
stress-concentrated
microdefect
Then,
theformation
microcracks
propagated
to
two sides
ofofthe
MPD
anodic
model
canmicrocleavage
be useful
in theor
goal
ofslipping.
predicting
turbine
blade
life, given
setriver-like
of FDR data.
SCC
cracks
with
or without
surface
dissolution,
mciroshear,
resulting
in the
formation
of athe
fractography and the discontinuous surface
2016
The with
Authors.
Published
by Elsevier
SCC
cracks
or without
surface
slipping.B.V.
Copyright
The
2016
The Authors.
Published
by Elsevier
B.V. This is an open access article under the CC BY-NC-ND license

2016
Authors.
Published
B.V.
Peer-review
under
responsibility
ofby
theElsevier
Scientific
Committee of ECF21.

2016
The
Authors.
Published
by
Elsevier
B.V.
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
Peer-review
under
responsibility
of
the
Scientific
Committee
PCF 2016.
Peer-review
under
responsibility
of
the
Scientific
Committee
of of
ECF21.
Peer-review under responsibility of the Scientific Committee of ECF21.
Keywords: Single-crystal stainless steel; Stress corrosion cracking; Three-dimensional model; Slip band; Discontinuous cracks
Keywords:
High Pressure
Turbine
Blade;
Creep;
Finitecracking;
Element Method;
3D Model;model;
Simulation.
Keywords:
Single-crystal
stainless
steel;
Stress
corrosion
Three-dimensional
Slip band; Discontinuous cracks

* Corresponding author. Tel.: +86-10-62772513; fax: +86-10-62772513.


address:author.
l.k.zhu@163.com
* E-mail
Corresponding
Tel.: +86-10-62772513; fax: +86-10-62772513.

E-mail address: l.k.zhu@163.com


2452-3216 2016 The Authors. Published by Elsevier B.V.
Peer-review
underThe
responsibility
of218419991.
theby
Scientific
Committee of ECF21.
* Corresponding
author.
Tel.: +351
2452-3216
2016
Authors.
Published
Elsevier B.V.
E-mail address:
amd@tecnico.ulisboa.pt
Peer-review
under responsibility
of the Scientific Committee of ECF21.

2452-3216 2016 The Authors. Published by Elsevier B.V.

Peer-review under responsibility of the Scientific Committee of PCF 2016.

Copyright 2016 The Authors. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
Peer review under responsibility of the Scientific Committee of ECF21.
10.1016/j.prostr.2016.06.079

Longkui Zhu et al. / Procedia Structural Integrity 2 (2016) 612621


Author name / Structural Integrity Procedia 00 (2016) 000000

613

1. Introduction
The deleterious effect of stress corrosion cracking (SCC) on engineering materials has been well documented for
over a century in civil and military industries such as nuclear, aerospace and so on. Abundant effort was dedicated to
the protection of the damage in order to seek out a reasonable predictive approach. Several SCC mechanisms were
proposed in the last decades on the basis of the two-dimensional experimental phenomena. Among these, there were
two well-known models for the stainless steel/MgCl2 solution system, including the slip dissolution model and the
corrosion-enhanced plasticity model (CEPM). The former by Logan (1952), Vermilyea (1972), Kolman et al. (1999),
Newman and Gutman (2007), and Hall (2009) postulated that either emerging slip planes or simply exposing fresh
metal surfaces by rupturing protective films acted as anodes. This speeded up metal dissolution prior to reappearance of the planes, or re-establishment of protective films, while repetition of this sequence made the cracks
continuously propagate. However, in CEPM by Magnin et al. (1990 and 1996), Flanagan et al. (1991) and Chateau et
al. (2002), the local stress increases due to dislocation pile-ups, while the critical stress intensity factor, KIC,
decreased due to hydrogen, leading to the discontinuous microcrack initiation in front of the main crack tips.
Both of them emphasized the necessity of dislocation motion and slipping. Nevertheless, Sieradzki and Newman
(1985) put forward the film-induced cleavage fracture mechanism in the brass/aqueous ammonia system, that the
crack advanced brittly on account of the restricted dislocation emission in the metal matrixes. That is, SCC of ductile
materials occurred possibly without dislocation slipping. Based on the autocatalysis theory of the occluded cell, the
pH value actually decreased inside the crack tip owing to the hydrolysis of metal ions. The crack tip acted as an
anode to be dissolved, then the film was able to be formed. The comparison of elastic behavior of the aircraft
structural A97075 Al-alloy and bulk Al2O3 oxide by Callister (1985) and Gutman (2007) demonstrates that the
fracture strain of aluminum oxide, f=0.0007, was much less than the elastic limit of the aluminum alloy, YS=0.002.
It is probable that anodic dissolution and the rupture of brittle films promoted SCC advance, whereas the ductile
matrix did not deform nonlinearly.
The argument of the two models was inconsistent whether the cracks propagated continuously or not. From the
view of anodic dissolution, an SCC process was continuous, but the crack could also nucleate discontinuously in
terms of dislocation pile-ups. In previous experiments, however, this kind of dislocation motion was not observed,
meaning the Stroh-Cottrell mechanism might be ineffective. Marrow et al. (2006) and King et al. (2008)
systematically studied intergranular SCC of a sensitized austenitic stainless steel by observing fractography and Xray computed tomographic images. It is detected that the discontinuous surface crack was actually continuous within
the specimen. Additionally, for transgranular SCC, some cracks usually propagated along low-index crystal planes.
Magnin et al. (1990) investigated SCC of 316 alloys in the 153 C boiling MgCl2 solution, and suggested that the
transgranular cracking was related to both microshear on the {1 1 1} planes and microcleavage mainly on the {1 0
0} planes. Li et al. (1989) employed etch-pitting and stereographic observations to analyze the SCC crystallography
of the same SCC system. The results show that the cracking occurred predominantly on the {1 0 0} planes. In 304L
and 310 stainless steels, the SCC cracks occurred primarily on the {1 0 0} planes, while secondary cracks on the {1
1 0} planes were also detected, referred to Meletis et al. (1984 and 1986) and Dickson et al. (1987).
Although lots of experimental and simulative research focused on the micro and macroscopic SCC phenomena
and mechanisms, accurately predicative models have not been created until now. In particular, the experimental
stress applying to SCC specimens was often larger than their yielding strength, while the service loads for
engineering materials were generally far below their yielding strength when SCC took place. On account of the
inconsistency, there were a few ductile fracture processes under highly experimental loads, which might induce the
formation of the ineffective models. In this work, a three-dimensional SCC model at low stress levels was proposed.
Afterward, we discussed detailedly the microscopic SCC mechanisms for the stainless steel/MgCl2 solution system.
2. Outline of experimental results
As already mentioned, the transgranular SCC fracture in ductile materials was a three-dimensional and
crystallographic process. The case of 316L austenitic stainless steel in a 45% MgCl2 solution was one where the
normal stress =20-40 MPa was applied to the specimens. The effects of localized deformation and three-

614

Longkui Zhu et al. / Procedia Structural Integrity 2 (2016) 612621


Author name / Structural Integrity Procedia 00 (2016) 000000

dimensional features on SCC initiation and propagation were studied in detail Zhu et al. (2013, 2014 and 2015) and
Qiao et al. (2011). The findings as follows:
(1)SCC without surface slipping
As shown in Fig. 1, it is found that there was no surface slip band in the initiation zone or around the tip of Crack
1. With crack propagation, SCC mechanisms could transfer from the low-stress model into the high-load one, shown
in Fig. 2. At the initiation site in Fig. 2(b), there was no surface slip band in the area between labals 1 and 2. Then,
the area between labels 2 and 3 was examined, shown in Fig. 2(a), (c), and several slip bands emerged on the sample
surface. Finally, the crack propagated in the area between labels 2 and 3. Crossing slipping took place when the
stress distributed at the crack tip was up to a critical scale, illustrated in Fig. 2(a) and (d).

Fig. 1. The surface morphology of Crack 1 in a specimen subjected to 20 MPa normal stress: (a) the whole crack, (b) the crack tip, where there
was no surface slip band around Crack 1, referred to Zhu et al. (2015).

Fig. 2. (a) The whole surface morphology of a fractured specimen subjected to 20 MPa normal stress; (b) partial surface morphology of the area
between labels 1 and 2 in (a); (c) partial surface morphology of the area between labels 2 and 3 in (a); (d) partial surface morphology of the area
between labels 3 and 4 in (a), some slip bands were virginal above the arrow and the other were coated by corrosion products on the right of the
arrow, referred to Zhu et al. (2015).

(2) Inner continuity of discontinuous surface SCC cracks


Three-dimensional SCC morphology was analyzed through non-destructive X-ray computed tomography. The
surface morphology and fractograph of Crack 2 were characterized by SEM, as shown in Fig. 3. Two discrete crack
sections split by the ligament L2 on the side surface between the dotted lines 1 and 2 in Fig. 3(a1), connected

Longkui Zhu et al. / Procedia Structural Integrity 2 (2016) 612621


Author name / Structural Integrity Procedia 00 (2016) 000000

615

with each other inside the specimen in Fig. 3(b). Likewise, this kind of external discontinuity and internal continuity
phenomenon was explicitly reflected in the fractograph in Fig. 3(a3). A protrusion at the right edge was just the
ligament L2. It initiated at an inner point and gradually propagated to the right edge, directly leading to the
formation of the discontinuous crack tip on the surface. In conclusion, the three-dimensional crack morphology
together with the fractograph indicates that the discontinuous SCC crack continuously propagated inside the
specimen, and the surface microcracks were separated by the ligaments.

Fig. 3. (a1) Optical micrograph of the Crack 2s tip on the side surface of the specimen by wire-electrode cutting; (a2) SEM morphology after the
specimen was pulled apart along the crack; (a3) SEM fractograph of the Crack 2 tip; (b) 3-D rendering of the Crack 2 between the dotted lines
1 and 2 in (a3), where the crack was indicated in yellow, referred to Zhu et al. (2014).

(3) Microcleavage on {1 0 0} planes


As a type of brittle fracture, SCC usually nucleates and propagates along a certain crystal planes. Two-surface
trace analysis was employed to characterize crystallographic feathers of the cracking planes. As seen in Fig. 4, the
first surface containing microcracks was the (0 1 4) crystal plane detected by EBSD. The angle between the [2 4 -1]
crystal orientation and the microcracks was either 78 or 102. According to the crystallography, the relationships
between the [u v w] crystal orientation and the microcracks on the (0 1 4) crystal plane are:

2 u 4 v (1) w
(1)

2 4 (1) u v w
2

0 u 1v 4 w 0

cos 78 or cos 102


(2)

Thus, the orientation of the microcracks was either [1 -0.24 0.06] or [1 -0.76 0.19]. Then, a 7 m thick layer was
removed by mechanical polishing, and the freshly exposed surface was the second one, shown in Fig. 4(b). On the
first and second surfaces, the distances between the labels M1 and M1 and the microcrack were measured, shown

Longkui Zhu et al. / Procedia Structural Integrity 2 (2016) 612621


Author name / Structural Integrity Procedia 00 (2016) 000000

616

schematically in Fig. 4(c). The tilt angle between the cracking plane and the first surface were determined as either
7 or 183. In terms of the crystallography, the expressions for the (h k l) cracking plane are:
(3
uh vk wl 0
)

0h 1k 4l
0 1 4
2

h k l
2

cos 7 or cos 183

(4)

When the microcrack orientation [u v w] was [1 -0.24 0.06], the (h k l) cracking plane was either (0.03 0.38 1)
or (-0.03 0.13 1), while the [1 -0.79 0.19] microcrack orientation corresponded to either (0.08 0.36 1) or (-0.08 0.15
1). However, there was a certain error when measuring the angles, and the angles between the cracking planes and
the low index crystal planes were estimated. For the microcrack, there was a 8 minimum angle between (1 0 0) and
the obtained cracking planes, but for (1 1 1) this minimum angle is 44. As a consequence, the cracking plane should
be (1 0 0).

Fig. 4. Morphology of (a) the original surface and (b) mechanically polished surface after removing a 7 m thick layer, and (c) schematic diagram
of the angle formed by the microcrack and the first surface between the labels M1 and M1 , referred to Zhu et al. (2015).

Fig. 5. (a) Microcracks 1 to 7 on a side surface of 316L stainless steel single crystal and (b) typical river-like fractograph of the same area in
(a), where the area marked by dotted lines was defined as MPD and a number of discontinuous secondary cracks and steps marked by arrows
emanated from MPD and angularly extended to both sides of MPD, leading to the formation of the discontinuous surface microcrack, referred to
Zhu et al (2014).

(4) Macroscopic propagation directions in the river-like fractograph

Longkui Zhu et al. / Procedia Structural Integrity 2 (2016) 612621


Author name / Structural Integrity Procedia 00 (2016) 000000

617

Finite element simulation by Zhu et al. (2013 and 2014) indicates that the normal stress and strain concentrated
inside the crack tip without a pitting defect. When the defect was formed, normal stress and strain concentration sites
simultaneously transferred to the defect edges rather than the defect bottom. The stress and strain distribution was in
favor of preferential SCC initiation inside the crack tip and at the defect edges. Figure 5 shows the typical river-like
SCC fractograph and the discontinuous surface microcracks of the austenitic stainless steel in the 45 % boiling
MgCl2 solution. It is found that the area marked by dotted lines was smooth and continuous, approximately from
several to more than ten microns in width, defined as a macroscopic propagation direction (MPD). A number of
secondary cracks and steps, marked by arrows, emanated from the MPD and were extended to the two sides of the
MPD at an angle. Obviously, some of them reached the edge of the fractograph, resulting in the formation of the
discrete surface microcracks 1 to 7, shown in Fig. 5(a). As a consequence, the continuous MPD directly caused
the inner SCC continuity at the micron scale, while the microscopic surface SCC discontinuity was induced by the
secondary cracks and steps emanating from the MPD.
3. A three-dimensional SCC model under low loads
On the basis of the above experimental and simulative results of the stainless steel/MgCl2 solution system, a
three-dimensional SCC model was created under low loads of 0.1-0.2 times of the yielding strength for engineering
materials. There were two points to be concerned before modeling the SCC process. It is considered that SCC cracks
nucleated when the stress intensity factor, KI, at the crack tip exceeded the threshold stress intensity factor, KISCC:
(5)
K K SCC
When the shear stress, , was equal to or greater than the critical shear stress, C, dislocation emission and slipping
were able to occur:
(6)
C

Fig. 6. A three-dimensional SCC model at low loads, where (d) presented a dislocation emission process at high stress levels, but under very low
loads SCC might nucleate and propagate without dislocation slipping in this process.

In the light of Eqs (5) and (6) as well as the theory of environmentally assisted cracking, the low-load SCC model
was put forward. The included steps as follows:
(1) It is assumed that the SCC crack front was linear and smooth, shown in Fig. 6(a). A microdefect was apt to
initiate inside the crack tip owing to the stress concentration and the smaller pH value.
(2) Once the microdefect was formed, seen in Fig. 6(b), the normal stress concentrated at the defect shoulders
rather than the defect base. The low-potential zone, induced by the different stress distribution, dissolved
preferentially at the low stress levels.

Longkui Zhu et al. / Procedia Structural Integrity 2 (2016) 612621


Author name / Structural Integrity Procedia 00 (2016) 000000

618

(3) The microdefect gradually grew and enlarged, shown in Fig. 6(c). Meanwhile, the normal stress increased
little by little at the defect shoulder.
(4) Slipping and dislocation emission by the shear stress took place on occasion of C at the defect shoulder,
shown in Fig. 6(d). Under very low loads, however, SCC with rather small KISCC could nucleate and propagate
without dislocation slipping in this process.
(5) while KI KISCC under the action of the normal stress, SCC microcracks initiated and grew at the defect
shoulders, shown in Fig. 6(e). The microscopic cracking was able to be caused by localized dissolution,
microcleavage, microshear or synergistic effects of them.
(6) With the defect growth, a macroscopic propagation direction was formed near the mid-thickness of the
specimen. The microcracks emanating discontinuously from the defect shoulders propagated to the two sides of
MPD. Finally, some of the microcracks reached the sample surfaces, resulting in the formation of microsteps,
disconnected surface microcracks such as 1 to 4 as well as the river-like fractograph shown in Fig. 6(f).
4. Discussion
4.1. Microcleavage on low-surface-energy crystal planes
As one of the transgranular SCC mechanisms, microcleavage under normal stress usually originated from a
certain planes in crystal materials. In terms of the Griffiths theory, the brittle fracture depended on the surface
energy of the cracking plane, for instance the critical stress intensity factor of the brittle fracture, KIC:

K C

2E
(1 2 )

(7)

where is the surface energy of the cracking plane, E is the elastic modular, is the poissons ratio. The previous
results by Caglioti et al. (1971) and Meletis et al. (1984) show that the {1 0 0} planes had the lowest surface energy
for austenitic stainless steels. In this circumstance, the cleavage planes were preferentially formed on the {1 0 0}
planes when KI KIC. The SCC microcracks actually propagated along the {100} planes in our experimental results.
Besides, there were also a few SCC cracks along other cleavage planes such as {1 1 0}, referred to Meletis et al.
(1984 and 1986) and Dickson et al. (1987). On these low-surface-energy planes, the normal stress devoted to
microcleavage was composed of the applied normal stress and the microscopic deformation of the crystal lattices.
For example, hydrogen atoms which were generated by cathodic reactions infiltrated into octahedral interstices of
face-centered cubic austenitic stainless steel, inducing the inhomogeneous stress distribution of microscopically
local areas. Therefore, the SCC cracks nucleated preferentially at the stress concentrated sites of the low-surfaceenergy crystal planes.
With SCC propagation in specimens subjected to constant loads, the stress around the crack tip gradually
increased. Orowan and Irwin considered that except from the surface energy, it was necessary for the crack initiation
and propagation to release energy by plastic deformation. For cracked specimens, the stress and strain were highly
localized, and dislocations emitted around the crack tips at the high stress levels. The presence of dislocations
modified the stress intensity factor at the crack tip, referred to Chateau (2002):

K C K a

(8)

where KIa is the applied stress intensity factor and KD is the contribution of each dislocation. From the view of the
linear defects, the more dislocations emitted at the crack tip, the larger the SCC propagation rate was. In our
experiments, it is obvious that the slip bands emerged near the long crack tip, and the crack emanated from the nonsurface-slipping area, then extended into the necking zone with numerous surface slip bands and secondary cracks.
As a consequence, the high stress promoted the microscopic cleavage of transgranular SCC.

Longkui Zhu et al. / Procedia Structural Integrity 2 (2016) 612621


Author name / Structural Integrity Procedia 00 (2016) 000000

619

4.2. Slipping and dislocation emission


The shear deformation including slipping and dislocation motion was considered as a main mechanism of SCC
initiation and propagation. The results in this work show that SCC was capable of advancing with and without
slipping in single-crystal austenitic stainless steels. In theory, the {1 1 1} <1 1 0> slip systems in face-centered cubic
metals started when the shear stress, , applied to the {1 1 1} slip plane along the <1 1 0> direction was equal to or
greater than the critical shear stress, C. Based on the Schmidts law, the shear stress, , is represented by:
(9)
cos cos
where is the applied stress, is the angle between the applied stress axis and the normal line of the {1 1 1} slip
plane, is the angle between the applied stress and the <1 1 0> slip direction, and =coscos is the Schmidts
factor. As a result, was lower than C when was constant and was very low, or when was constant and was
quite small. In this circumstance, the slip system did not start and there was no slip band on the specimen surfaces.
Under the high loads or with the SCC cracks extending in specimens subjected to constant loads, the stress enlarged
step by step and the surface slipping took place again around the crack tip. Analogously, as the microscopic slipping
process, the dislocation emission also depended on the stress distributed in the local areas.
Different from the microcleavage under the normal stress, the microshear might also avail to SCC initiation and
propagation. The gliding along slip planes {1 1 1} intersected on the cracking surfaces, which could form the
dislocation pile-ups, named after the Stroh-Cottrell mechanism. Although the phenomenon that a microcrack
initiated by means of the dislocation pile-ups has not been detected until now, it is not denied that the microshear
played an important role on SCC advance under the high loads. As stated above, for instance, the dislocation motion
to the cracking planes could accelerate the microcleavage process. Meanwhile, the stress concentration areas, caused
by the shear deformation, acted as the preferential SCC nucleation sites or anodes to be dissolved. Hence, it is
considered that the microshear was one of the main SCC mechanisms at the high stress levels, possibly in favor of
microcleavage and anodic dissolution.
4.3. Localized dissolution
SCC was an electrochemical process in essence, in which anodic dissolution of freshly exposed surfaces induced
the crack growth, referred to Logan (1952), Vermilyea (1972), Kolman et al. (1999), Newman and Gutman (2007),
and Hall (2009).In our experiments, the effect of dissolution on SCC was different with the extension of the cracks
from the non-surface-slipping areas to the necking zones under the constant loads. That is, SCC was primarily
controlled by localized dissolution under very low loads. In this case, the stress was not enough to cleavage or shear
deformation. According to the localized dissolution mechanism and the experimental results, the new defects were
able to be formed on the crack fronts and the normal stress for cleavage should firstly play a role on SCC, then the
dislocations slip along {1 1 1} planes when C. Inversely, the freshly formed surfaces by cleavage or yielding
served as anodes to be dissolved. Consequently, the synergistic effects of localized dissolution, microcleavage and
microshear induced SCC initiation and propagation at the high stress levels, while anodic dissolution and
microcleavage were main SCC mechanisms under the low loads. In terms of the Faradays law, the instantaneous
crack growth rate, a, can be expressed as a function of the instantaneous anodic current density, ia,

M
i
zF a

(10)

where M is the molecular weight; z is the charge of the metal cation; is the metal density; F is the Faradays
constant. In the light of Eq. (10), however, the calculated SCC propagation rate did not correspond with the
experimental obtaining, which should be caused by the localized dissolution not being the single SCC mechanism.
Thus, the above three microscopic mechanisms ought to be comprehensively considered in the process of SCC
advance.

620

Longkui Zhu et al. / Procedia Structural Integrity 2 (2016) 612621


Author name / Structural Integrity Procedia 00 (2016) 000000

5. Conclusion
A three-dimensional SCC model under low loads was created, where a microdefect preferentially initiated on the
crack front near the mid-thickness of the specimen due to anodic dissolution; then, the defect gradually enlarged and
secondary microcracks emanated from the defect shoulders, angularly extending to the two sides of the specimen.
Finally, some of the secondary microcracks reached the sample surfaces, resulting in the formation of the
discontinuous surface cracks and the disconnected steps of the typically river-like fractograph. It is concluded that
the synergistic effects of microcleavage and localized dissolution induced the SCC advance on the low surface
energy crystal planes. The Microshear was also a primarily microscopic SCC mechanism at the high stress levels.
Acknowledgements
The authors wish to thank You He, Yanan Fu and Honglan Xie at the Shanghai Synchrotron Radiation Facility
for their experimental help. The authors also acknowledge funding provided by the National Nature Science
Foundation of China under grant 51431004 and the National Key S&T Special Projects under the grant
2011zx06901.
References
R.C. Newman, C. Healey, Stability, validity and sensitivity to input parameters of the slip-dissolution model for stress-corrosion cracking, Corros.
Sci. 49 (2007) 4040-4050.
H.L. Logan, Film-rupture mechanism of stress corrosion, J. Res. Natl. Bur. Stand. 48 (1952) 99-105.
D.A. Vermilyea, A theory for the propagation of stress corrosion cracks in metals, J. Electrochem. Soc. 119 (1972) 405-407.
D.G. Kolman, J.R. Scully, Continuum mechanics characterization of plastic deformation-induced oxide film rupture, Philos. Mag. A 79 (1999)
2313-2338.
E.M. Gutman, An inconsistency in film rupture model of stress corrosion cracking, Corros. Sci. 49 (2007) 2289-2302.
M.M. Hall Jr., Critique of the Ford-Andresen film rupture model for aqueous stress corrosion cracking, Corros. Sci. 51 (2009) 1103-1106.
M.M. Hall Jr., Flim rupture model of aqueous stress corrosion cracking under constant and variable stress intensity factor, Corros. Sci. 51 (2009)
225-233.
T. Magnin, R. Chieragatti, R. Oltra, Mechanism of brittle fracture in a ductile 316 alloy during stress corrosion, Acta Metall. Mater. 38 (1990)
1313-1319.
T. Magnin, A. Chambreuil, B. Bayle, The corrosion-enhanced plasticity model for stress corrosion cracking in ductile fcc alloys, Acta Mater. 44
(1996) 1457-1470.
J.P. Chateau, D. Delafosse, T. Magnin, Numerical simulations of hydrogen- dislocation interactions in fcc stainless steels.: part II: hydrogen
effects on crack tip plasticity at a stress corrosion crack, Acta Mater. 50 (2002) 15231538.
W.F. Flanagan, P. Bastias, B.D. Lichter, A theory of transgranular stress-corrosion cracking, Acta Metall. Mater. 39 (1991) 695-705.
J.P. Chateau, D. Delafosse, T. Magnin, Numerical simulations of hydrogen- dislocation interactions in fcc stainless steels.: part I: hydrogendislocation interactions in bulk crystals, Acta Mater. 50 (2002) 1507-1522.
K. Sieradzki, R. C. Newman. Brittle behaviour of ductile metals during stress-corrosion cracking, Philos. Mag. A 51 (1985) 95-132.
W. D. Callister Jr., Materials Science and Engineering, John Willey & Sons, Inc., 1985.
T.J. Marrow, L. Babout, A.P. Jivkov, P. Wood, D. Engelberg, N. Stevens, P.J. Withers, R.C. Newman, Three dimensional observations and
modeling of intergranular stress corrosion cracking in austenitic stainless steel, J. Nucl. Mater. 352 (2006) 62-74.
A. King, G. Johnson, D. Engelberg, W. Ludwig, J. Marrow, Observations of intergranular stress corrosion cracking in a grain-mapped polycrystal,
Science 321 (2008) 382-385.
S. Li, J.I. Dickson, J.P. Bailon, The influence of the stress intensity factor on the fractography of stress corrosion cracking of 316 stainless steel,
Mater. Sci. Eng. A119 (1989) 59-72.
E.I. Meletis, R.F. Hochman, The crystallography of stress corrosion cracking in face centered cubic single crystals, Corr. Sci. 24(1984) 843-862.
E.I. Meletis, R.F. Hochman, A review of the crystallography of stress corrosion cracking, Corr. Sci. 26 (1986) 63-77, 79-90.
J.I. Dickson, D. Groulx, S. Li, The fractography of stress corrosion cracking of 310 stainless steel: crystallographic aspects and the influence of
stress intensity factor, Mater. Sci. Eng. 94 (1987) 155-173.
L.K. Zhu, Y. Yu, J.X. Li, L.J. Qiao, Z. C. Li, A.A. Volinsky, Stress corrosion cracking under low loads: Surface slip and crystallographic
analysis, Corr. Sci. 100 (2015) 619-626.
L.K. Zhu, Y. Yu, J.X. Li, L.J. Qiao, A.A. Volinsky, Stress corrosion cracking under low stress: Continuous or discontinuous cracks?, Corr. Sci.
80 (2014) 350-358.
L.K. Zhu, Y. Yan, L.J. Qiao, A.A. Volinsky, Stainless steel pitting and stress corrosion cracking under ultra-low elastic load, Corros. Sci. 76
(2013) 360-368.


10

Longkui Zhu et al. / Procedia Structural Integrity 2 (2016) 612621


Author name / Structural Integrity Procedia 00 (2016) 000000

621

L.J. Qiao, K.W. Gao, A.A. Volinsky, X.Y. Li, Discontinuous surface cracks during stress corrosion cracking of stainless steel single crystal,
Corros. Sci. 53 (2011) 3509-3514.
G. Caglioti, G. Rizzi, J.C. Bilello, Surface energy for brittle fracture in metals from phonon frequencies, J. Appl. Phys. 42 (1971) 4271-4276.

Das könnte Ihnen auch gefallen