Sie sind auf Seite 1von 6

Proceedings of the ASME 2012 Pressure Vessels & Piping Conference

PVP2012
July 15-19, 2012, Toronto, Ontario, CANADA

PVP2012-78774

ADVANCED ASSESSMENT OF THE INTEGRITY OF DUCTILE COMPONENTS


Michael Daly
Materials Performance Centre
The University of Manchester
The Mill
Sackville Street
Manchester, M60 1QD, UK
Tel: +44 (0)7775712870
Email: michael.daly@postgrad.manchester.ac.uk

Andrew H Sherry
The Dalton Nuclear Institute
The University of Manchester
Pariser Building G Floor
Sackville Street
Manchester, M13 9PL, UK
Tel: +44 (0)161 275 4431
Email: andrew.sherry@manchester.ac.uk

John K Sharples
Serco Assurance
Walton House,
Birchwood Park
Warrington,
Cheshire, WA3 6AT, UK
Tel: +44 (0)192 525 3462
Email: john.sharples@sercoassurance.com

ABSTRACT
Nuclear Reactor Pressure Vessels (RPV) are manufactured
from medium strength low alloy ferritic steel, specifically
selected for its high toughness and good weldability. The ability
of the pressure vessel to resist crack growth is crucial given that
it is one of the fundamental containment safety systems of the
reactor. For most of their lifetime, the pressure vessel operates
at sufficiently elevated temperatures to ensure the material is
ductile. However, the development of ductile damage, in the
form of voids, and the ability to predict the ductile crack
growth in RPV materials requires further work.
The Gurson-Tvergaard-Needleman (GTN) model of void
nucleation, growth and coalescence provides one tool for
predicting ductile damage development. The model is normally
calibrated against fracture toughness test data. However, recent
work [1] has demonstrated the benefit of refining calibrations
against measured void volume fractions generated from
notched and pre-cracked specimen tests.
This paper described the measurement of void distributions
below the fracture surface of a range of notched and precracked specimens. The void distribution below the fracture
surface is shown to be dependent upon the local stress
triaxiality and plastic strain distribution. As a result, precracked specimens show a greater concentration of voids close

to the fracture surface, whilst notched tensile specimens show a


lower volume fraction of voids close to the crack surface. In
both specimen types, voids are observed to extend between 2.5
and 3.5 mm below the fracture surface.
INTRODUCTION
The ductile fracture mechanism is characterised by the
nucleation, growth and coalescence of voids that form at
inclusions or second phase particles in the vicinity of a notch or
crack tip. Voids will often form by either decohesion of the
interface between the matrix and second phase particles,
including inclusion, or by cracking of the particles themselves
[2]. Voids then grow under the influence of increasing plastic
strain and high hydrostatic stress within the material. The crack
will finally propagate once the neighbouring voids coalesce
and/or reach a critical size that leads to a macroscopic flaw.
There exists a range of mechanistically-based models that
have been developed to describe the behaviour of materials
during ductile fracture. One of these is the Gurson, Tvergaard
and Needleman (GTN) model [3] [4] [5] which requires
material specific parameters to be calibrated to enable ductile
crack growth to be simulated. The model assumes that the
material is homogeneous and behaves as a continuum. The
voids are accounted for by influencing the global flow
1

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 10/10/2013 Terms of Use: http://asme.org/terms

Copyright 2012 by ASME

behaviour of the volume and their effects are averaged


throughout the material. Crucially, the model takes into
consideration the dependence of yielding on the hydrostatic
stress exerted on the material by introducing a strain softening
term. This strain softening term accounts for the initiation,
growth and coalescence of voids and is used in conjunction
with the hardening of the matrix material with the yield
dependent term of the hydrostatic stress. The model is defined
by the following yield function, :

>? 2
3G2 >A
>? , >A , >,
B   = C E + 2G1  cosh .
1 1 + G3  2  = 0
B
2>N
(1)

Where:
e = macroscopic Von Mises Stress
m = macroscopic mean stress
= flow stress for the matrix material
 = current void fraction
The values for q1, q2 and q3 were introduced by Tvergaard
and Needleman to better simulate the experimental
observations. These are often taken as q1 = 1.5, q2 = 1.0 and
q3 = q12. The rate of void growth is related to the plastic part of

the strain rate tensor  and the void nucleation rate is related

to the equivalent plastic strain rate  :
=  +  =    !! + #  $%
"

"

(2)

The first term expresses the growth rate of existing voids


assuming the matrix material is incompressible and the second
term defines the quantity of new voids that have nucleated as a
result of the increasing plastic strain.
The scaling coefficient, is characterised by:
# =

'

(' *+

$," - .
*

*
"
$% / '
0'

1 2

(3)

Where:
3 = volume fraction of void nucleating particles
sN = standard deviation
N = mean value
eqP = equivalent plastic strain.
Additional aspects introduced by Tvergaard and Needleman
was to take into consideration the initial void fraction f0, a
critical void volume fraction for coalescence fc, and a critical
void fraction that corresponds to the failure of the matrix, fF.
=4

56 8

9 / 8
: / 8

 8  56 > 8

<

(4)

Where:
fc = critical void volume fraction (typically fc = 0.15 for carbon
steel)

fF = actual void volume fraction at final fracture


fu* = modified void volume fraction (typically fu* = 1/q1)
These parameters are normally calibrated against fracture
test data obtained in controlled experiments. A range of test
specimens have been used as the basis to calibrate the GTN
model parameters, including notched tensile specimens with a
range of notch-root radii and pre-cracked specimens such as
Compact-Tension (CT) specimens or Single Edge-Notched
Bend SEN(B) specimens.
However, previous experiments undertaken on notched and
CT specimens of a high strength and low toughness aluminium
alloy AL2024-T351 have demonstrated that there exists a
difference in the void distribution below the fracture surface in
notched and pre-cracked specimens [1]. In notched specimens
there was a lower critical void volume fraction for fracture than
that measured in cracked specimens. Voids were also observed
to extend further below the fracture surface in notched
specimens when compared with that in cracked specimens.
The aim of the current work is to assess this observation
with respect to a ferritic steel pressure vessel, to enable the
GTN model to be quantified for this material. This paper
describes the experimental procedure used to test notched and
cracked specimens of an A508 Class 3 ferritic steel forging and
the methodology applied to quantify the distribution of voids
below the fracture surface in both specimen types. The results
are discussed in the context of the previous work reported on
AL2024-T351 and conclusions drawn.
EXPERIMENTAL
Material
The material used throughout this experiment was an A508
Class 3 steel. The specimens were extracted from the outer ring
of an upright wedge-shaped block originating from the larger
ring forging. All the specimens were extracted from the same
location and in the same orientation (described below). The
chemical composition of the ferritic steel is as indicated in
Table 1 below.
Sn
0.008

Mo
0.556

Cu
0.042

Ni
0.717

Co
0.000

V
0.005

S
0.0003

P
0.001

Si
0.191

Al
0.011

Mn
1.285

Cr
0.336

Table 1: Composition of A508 Class 3 steel forging (wt%)


Mechanical testing
The tensile properties of the material were determined on
standard round-bar test specimens oriented in the hoop
direction. Three tensile specimens were tested on a Zwick
1464 at room temperature using a strain rate of 0.025% s-1
according to BS EN ISO 6892 procedure [6].
Six notched tensile tests were performed according to the
ESIS P6-98 standard [7]. The tests were performed on

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 10/10/2013 Terms of Use: http://asme.org/terms

Copyright 2012 by ASME

specimens machined in the hoop orientation of gauge length


80mm with an outer radius of 4.5mm, an inner radius of 2.5mm
and a notch root radii, r = 2 or 10mm. The first specimen of
each notch root radii was pulled to fracture while the final two
specimens in each data set were interrupted at different points
during the tests, following maximum load.
Ten fracture toughness tests were performed according to
the ESIS P2-92 standard [8] using CT specimens with standard
dimensions of thickness, B = 25mm, width, W = 50mm and a
crack length to specimen width ratio, a/W = 0.5. Specimens
were 20% side-grooved following fatigue pre-cracking. Tests
were performed using both the unloading compliance and the
multi specimen methods. Out of the ten compact test
specimens, two were left intact in order to preserve the crack
tip.
Metallographic analysis
A range of optical and scanning electron microscopy was
undertaken of the parent material to characterise the general
microstructure of the steel. In addition, both notched and CT
specimens were examined following testing to characterise the
failure mechanism and the area fraction of voids below the
fracture surface.
For the parent material, metallographic sections were taken to
view the material in the axial-radial plane. For tested samples,
sections were machined through the fractured specimen halves
in the region where plain strain fracture was expected to take
place. All samples were mounted into a resin compound and
progressively polished to a mirror finish of 0.25m using a
diamond paste on a soft rotating cloth. For the microstructural
analysis of the parent material, colloidal silica (OPS) polish
was used followed by etching in a 10% Nital solution to
highlight the grain boundaries and particles. For fractured
samples, neither OPS polishing or etchants were used to ensure
the highest contrast between voids and the base material.

Figure 1. Schematic of the region of interest that was


machined, polished and analysed.
An image analysis was performed using the open source
software ImageJ [9] and plugin [10] that creates a grid of
Regions Of Interests (ROI). Within each ROI, ImageJ
quantifies the number of black and white pixels. The grid was
chosen with 100m (228pixels) 100m (228pixels) cells with
a total width and height of the grid corresponding to the length
of the fracture surface being analysed and a distance below the
crack up to where no voids were visible at the chosen
magnification. Figure 2 illustrates an example of the resulting
grid and numbering system used to quantify the void area
fraction as a function of distance below the fracture surface.

Figure 2. Example showing the grid used to quantify the


area fraction of voids as a function of distance below the
fracture surface (left) and numbering system (right).

Quantitative metallographic analysis


A series of optical metallographs were taken of sectioned
and fractured notched and pre-cracked specimens starting from
the fracture surface and imaging sequential areas down to 5mm
below the fracture surface. The metallographs were taken at
20 magnification, each with an area of 550450m2. These
were then assembled to form larger assemblies, or mosaics, of
up to 250 pictures showing in fine detail the area below the
fracture surface in each specimen. To maintain the high
resolution and pixel count, all pictures were saved as TIFF files.
Each picture was taken at both normal (grey) contrast and at
very high contrast to further highlight the black areas identified
mostly as voids and the white areas identified as the metal. In
some cases, some scratch lines and other sample preparation
remnants were also visible as dark areas.

When an ROI overlapped the fracture surface and


metallographic mount, the ROI was discounted unless a
substantial number of voids were identified and could be
quantified using the ImageJ area tool. In all cases, threshold
adjustment was performed to reduce the background noise so
as to eliminate any sample preparation residues and scratch
marks from the analyses.
The void area fraction was calculated for every first cell
below the crack line and similarly for subsequent cells. The
average was calculated for every cell of the same location
below the crack. Finally, the total average was calculated for
each specimen type studied.

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 10/10/2013 Terms of Use: http://asme.org/terms

Copyright 2012 by ASME

RESULTS
Tensile test results from the three tests performed are
summarised in Table 2. The average yield stress was 446 MPa
and the ultimate tensile stress was 594 MPa.
The fracture toughness properties of the A508 Class 3 steel
are illustrated as a J R-curve in Figure 3 which includes data
from both the unloading compliance tests and monotonically
loaded tests together. The data from both test types are in
agreement and the initiation toughness, measured by the
intersection of the blunting line including 0.2 mm tearing and
the power-law curve fit to the data is ~ 475 kJ/m2.
Average
Specimen No.
T1
T2
T3
Modulus (GPa)
211
210
210
208
0.2% proof stress
455
457
446
436
(MPa)
1.0% proof stress
458
467
458
450
(MPa)
Ultimate tensile
595
602
594
586
stress (MPa)
Elongation (%)
29
27
27
28
Reduction in
75
75
74
75
Area (%)
Table 2: Tensile test results (average of three tests)

(b)

Figure 4. Optical (a) and scanning electron (b) micrograph


of A508 Class 3 steel.

Figure 5. Scanning electron micrograph showing voids


initiated at carbides within the bainitic microstructure
Void Area Fraction vs Distance Below Crack
0.04
0.035

Void Area Fraction

Figure 3. J R-curve for A508 Class 3 material tested in the


hoop-radial direction at 23oC.
(a)

Average CT

0.03

Average 2mm
Notch
Average 10mm
Notch

0.025
0.02
0.015
0.01
0.005
0

3650
3450
3250
3050
2850
2650
2450
2250
2050
1850
1650
1450
1250
1050
850
650
450
250
50

Distance from fracture surface (m)

Figure 6. Variation in area fraction of voids below the


fracture surface in notched and CT specimens

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 10/10/2013 Terms of Use: http://asme.org/terms

Copyright 2012 by ASME

Figure 4 illustrates the typical bainitic microstructure of the


A508 Class 3 RPV material using both optical and scanning
electron microscopy.
Microvoids were observed to initiate by the decohesion of
carbides from the matrix as shown in Figure 5. Voids were
observed to be present well below the fracture surface in all
specimens, and a number of microcracks (coalesced voids)
were observed below the fracture surface in some cases.
The variation of void area fraction as a function of distance
below the fracture surface in notched and CT specimens is
illustrated in Figure 6. The data represents an average of all
specimens studied and of all positions below the fracture
surfaces, Figure 2. The following observations can be made:

The void area fraction is highest at the fracture surface


and reduces as a function of distance below the fracture
surface to zero some distance below the fracture
surface.
The void area fraction just below the fracture surface is
approximately 1.5 10-2 for notched specimens and
2.5 10-2 for CT specimens.
The void area fraction reduces to approximately zero at
approximately 2.5 mm below the fracture surface in all
cases. There does appear to be some oscillation in the
void area fraction between 2.5 and 3.5 mm in CT
specimens.
There appears to be a secondary peak in void area
fraction in all specimens between 0.5 and 1.0 mm
below the fracture surface. The origin of this secondary
peak is not fully understood, but may relate to
secondary cracking observed in some specimens where
a second crack propagates parallel to the primary crack
surface.

to subtle differences in the mechanism of ductile damage


accumulation prior to final rupture. Further work is needed to
understand this trend in A508 Class 3 RPV steels.
Figure 7 also indicates that in AL2025-T351 voids are far
more extensive below the fracture surface in notched specimens
than in CT specimens. In the current work, the area fraction of
voids is observed to be similar in notched and cracked
specimens. In both cases, voids are observed to approximately
2.5 mm below the fracture surface. This is similar to that
observed in the notched AL2025-T351 material, though far
greater than that observed in the CT specimens, Figure 7.
The area fraction data is far more scattered in the A508
Class 3 material than the volume fraction data for the AL2025T351. This may relate either to the inherent scatter generated in
area fraction measurements, compared with volume fraction
measurement, i.e. related to the probability of a particular
metallographic section intersecting a void at its maximum (or
minimum dimension), or to differences in the fracture process
deriving from inhomogeneities in the microstructure of RPV
steels compared with AL2025-T351. Further quantitative
analyses using three-dimensional approaches including X-ray
tomography and serial sectioning will provide further
understanding of this aspect.

DISCUSSION
Taylor and Sherry [1] quantified the void volume fraction
below the fracture surface in failed notched tensile and CT
specimens of AL2024-T351 aluminium alloy using optical and
X-ray tomography respectively. As illustrated in Figure 7, they
observed that the critical value fF, i.e. just below the fracture
surface, in CT specimens was almost twice that measured in
notched tensile specimens of root radius 2 mm (NT2) and
10 mm (NT10), in CT specimens fF 1.1 10-2 whilst in both
notched specimens the value was close to 0.6 10-2. The
current results, whilst relating to area fraction and not volume
fraction, also suggest a higher void volume fraction just below
the fracture surface in CT specimens (2.5 10-2) than in
notched specimens (1.5 10-2). This observation, whilst
requiring further quantitative three-dimensional metallographic
analysis to confirm, does infer a similar trend to that observed
previously and, if confirmed, suggests that critical parameters
in the GTN model may not be directly transferable between
notched and cracked geometries. The reason for the difference
observed in AL2024-T351 was deemed to relate to differences
in the strain gradient in the vicinity of a notch and crack leading

Figure 7: Void area fraction below crack surface for


aluminium alloy [1].
CONCLUSION
This paper has described preliminary work undertaken to
characterise the ductile fracture properties and fracture
mechanism in a A508 Class 3 steel. The main conclusions from
the work are as follows:
1.

2.

The mechanical and fracture toughness properties have


been quantified in the hoop direction. The average
yield stress is 446 MPa and the initiation toughness
defined by the 0.2 mm blunting line is ~ 475 kJ/m2.
Ductile fracture has been observed to occur by the
decohesion of carbides from the matrix.

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 10/10/2013 Terms of Use: http://asme.org/terms

Copyright 2012 by ASME

3.

4.

5.

The area fraction of voids as a function of distance


below the fracture surface has been quantified using an
image analysis approach on polished sections.
Consistent with related work on an aluminium alloy,
the critical area fraction of voids for fracture is higher
in CT specimens than in notched tensile samples.
The area fraction of voids below the fracture surface in
CT and notched specimens is observed to be broadly
similar in both specimen types, extending to between
2.5 and 3.5 mm below the fracture surface.
These results may have implications for the calibration
of ductile damage models such as the GTN model
against notched specimen data, when being applied to
predict the behaviour of cracked geometries.

ACKNOWLEDGMENTS
The authors are grateful to Serco Assurance in Risley for
their support in the use of their material testing and microscopy
equipment.

BIBLIOGRAPHY
[1] K. L. Taylor and A. H. Sherry, "The characterization and
interpretation of ductile fracture mechanisms in AL2024T351 using X-ray and focused ion beam tomography,"
Vols. 60 (2012) 1300-1310, 2012.

[2] T. L. Anderson, Fracture Mechanics: Fundamentals and


Applications, Taylor and Francis, 2005.
[3] A. L. Gurson, "Continuum Theory of Ductile Rupture by
Void Nucleation and Growth: Part1- Yield Criteria and
Flow Rules for Porous Ductile Media," vol. 99, no. pp 215, 1977.
[4] V. Tvergaard, "Influence of voids on shear band
instabilities under plane-strain conditions," vol. 17, no. pp.
389407, 1981.
[5] V. Tvergaard and A. Needleman, "Analysis of the cupcone fracture in a round tensile bar," vol. 32, no. 1, pp.
157-169, 1984.
[6] British Standards, ""Metallic materials. Tensile Testing.
Method of test at ambient temperature" BS EN ISO 68921:2009".
[7] European Structural Integrity Society, Procedure to
measure and calculate material parameters for the local
approach to fracture using notched tensile specimens,
ESIS Standard No. P6-98, 1998.
[8] "European Structural Integrity Society, Procedure for
determining the fracture behaviour of materials, ESIS
Standard No. P2-92, 1992".
[9] "Image J," http://rsbweb.nih.gov/ij.
[10] Karl-Franzens-Universitat,
"Patch
Detector Plus,"
http://microscopy.uni-graz.at.

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 10/10/2013 Terms of Use: http://asme.org/terms

Copyright 2012 by ASME

Das könnte Ihnen auch gefallen