Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Engineering Fracture Mechanics 293 (2023) 109682

Contents lists available at ScienceDirect

Engineering Fracture Mechanics


journal homepage: www.elsevier.com/locate/engfracmech

Constraint loss correction: Enhancing transferability of fracture


toughness of Eurofer97 compact tension specimens between
specimen sizes
B. Aydin Baykal *, Philippe Spätig
Laboratory for Nuclear Materials, Paul Scherrer Institut, Forschungsstrasse 111 5232, Villigen PSI, Aargau, Switzerland

A R T I C L E I N F O A B S T R A C T

Keywords: This paper proposes and implements a constraint loss correction as a supplement to the ASTM
Finite Element Analysis correction factor for the master curve method (ASTM-E1921) for Eurofer97 reduced activation
Nuclear Reactor Safety steel in order to enhance the transferability of results between compact tension (CT) specimen
Fracture Mechanics
sizes. To facilitate the acquisition and scalability of reliable fracture toughness data from one
Brittle Fracture
Constraint Loss
specimen size to another and to apply the Master-Curve method with confidence when using
Miniaturization small specimens, the difference in the structure of the crack tip stress field between large and
small specimens was modeled. The stress intensity factors and critical volumes/stresses of two
specimen sizes (1 T and 0.18 T) were calculated with 3D finite element simulations run over a
large displacement range and at a temperature range of − 120 to − 80 ◦ C and the results were
compared with real data from a database of toughness tests of Eurofer97 steel. The K values from
a miniaturized specimen without or with ASTM correction and with the proposed combined
correction with constraint loss factor (which is variable over the entire loading range) were
compared with simulated CT-1 T stress intensity values. It was found that this new correction
factor was accurate up to a target toughness of 200 MPa.m1/2, which represents a major
improvement on the current state of the art.

1. Introduction

Eurofer97 is a high-chromium reduced activation tempered martensitic stainless steel developed for structural applications in
fusion reactors, typically the first wall and the blanket [1]. These parts will be subjected to high temperature and intense neutron
irradiation during reactor operation, as well as to significant stress due to localized effects resulting from temperature transients and
gradients, possible electro-magnetic loads, weld residual stresses, etc. The integrity of the first wall and blanket has to be ensured
during both normal and emergency operation conditions. Therefore, understanding the irradiation-induced degradation of the me­
chanical properties and in particular the evolution of the fracture toughness during long-term operation is a necessity. One of the main
concerns is loss of toughness due to irradiation, which increases the ductile–brittle transition temperature (DBTT) of ferritic steels
[2,3]. The magnitude of this increase can be several hundreds of degrees ◦ C, potentially even higher than the operating temperature.
Unlike the thick pressure vessel of fission reactor, the blanket is a thin-walled structure. Since fracture toughness of a given component
made of ferritic steel in the ductile–brittle transition is strongly dependent on its size[4], it is crucial to develop reliable models and

* Corresponding author.
E-mail address: bedi.baykal@psi.ch (B.A. Baykal).

https://doi.org/10.1016/j.engfracmech.2023.109682
Received 30 June 2023; Received in revised form 27 September 2023; Accepted 17 October 2023
Available online 19 October 2023
0013-7944/© 2023 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).
B.A. Baykal and P. Spätig Engineering Fracture Mechanics 293 (2023) 109682

Nomenclature

A Area
A* Critical area (for brittle fracture)
B Specimen thickness
b0 Ligament length
CT Compact tension specimen
d Displacement
DBTT Ductile-brittle transition temperature
E Young’s modulus
FEM Finite element method
K Stress intensity factor
KJ Equivalent stress intensity factor
KJc Mode I fracture toughness
KJc.limit Maximum allowable calculated fracture toughness
Kmin Lowest admissible stress intensity factor for fatigue considerations
kASTM ASTM crack front correction factor
kCL Constraint loss correction factor
kM Combined correction factor
Mlim Load limit factor
r Miniaturization ratio
SE(B) Bend bar specimen
T Temperature
T0 Reference temperature for ductile–brittle transition
V Volume
V* Critical volume (for brittle fracture)
ε Engineering strain
εpl Plastic strain
ν Poisson’s ratio
W Specimen width
σ Engineering stress
σ* Critical stress (for brittle fracture)
σYS Yield strength

calibration procedures that account for the effect of specimen/component size on measured fracture toughness.
Fracture toughness evaluation based on the Master-Curve method (ASTM-E1921) has become a widely-accepted way to determine
a reference temperature T0 of the ductile to brittle transition, which characterizes the brittleness of ferritic steels [5]. The reference
temperature T0 indexes the median toughness-temperature curve at a toughness value of 100 MPa.m1/2 obtained with 1 T-size
specimens (B = 1 in = 25.4 mm). However, fracture toughness depends on specimen size and geometry, and is prone to scatter due to
inhomogeneity of the material. Typically, miniaturized specimens with a crack front ranging between 4 and 22 mm are used to
characterize the fracture toughness of irradiated specimens.
Assessing the long-term structural integrity of large-scale steel structures requires fracture toughness testing of a sufficient number
of specimens with equivalent stress state and environment conditions to what the material may be subjected to during operation. Due
to practical issues with testing life-size specimens (high weight, high load capacity requirement, logistical problems, high material
costs) as well as limited availability of irradiated materials, it is unavoidable that such specimens are miniaturized. The most con­
ventional and widely accepted geometries for fracture toughness tests are the compact tension (CT) specimen and bend bars (SE(B)).
One advantage of the CT specimens is that they maintain a higher level of constraint than bend bars at a given applied stress intensity
factor. Nonetheless, in order to maintain accuracy, the effects of miniaturization and geometry on the fracture toughness of CT (or SE
(B)) specimens must be quantified and the model appropriately adjusted to correspond to the fracture toughness of specimens with
different sizes. There have indeed been concerns about overestimation of the fracture toughness of larger-sized structures using current
state of the art methods [6,7].
There are two distinct effects of miniaturization which may affect transferability of fracture toughness values from one size of
specimen to another: First, a different crack front length will have a different probability of a crack initiation event occurring, e.g. [8].
This is known as the statistical effect; the longer the crack front available for initiation, the higher is the probability of crack initiation.
This effect lowers the measured fracture toughness by lowering the critical stress intensity value for geometries/specimen sizes with
longer crack fronts. Second, there is a potential effect of constraint loss, which is divided into two categories: In-plane and out-of-plane
constraint loss. In-plane constraint loss occurs when the plastically deformed area at the crack tip is too large compared to the intact
ligament. Out-of-plane constraint loss occurs when the plane strain condition cannot be provided over a significant part of the crack
front length due to improper thickness/width ratio (most often due to insufficient specimen thickness). Both of these effects can alter

2
B.A. Baykal and P. Spätig Engineering Fracture Mechanics 293 (2023) 109682

the stress field at the crack tip so that fracture toughness values obtained with standard specimens of different sizes can differ
significantly. It is important to note that the statistical effect is dependent only on the ratio of crack front lengths while the constraint
loss is simultaneously dependent on loading level, ligament length and other complex variables such as mode of loading. There is
abundant literature that has been published on constraint loss effect on brittle fracture, dealing either with in-plane conditions only
[9,10], or by considering in- and out-of-plane effects simultaneously [11].

2. Theory

The Master-Curve method introduced by Wallin [12] is the state of the art method for assessment of the fracture toughness-
temperature relationship for 1 T-sized CT specimens. This method accounts for experimental scatter using a failure probability cri­
terion based on experimental data, where a different K vs. T curve is generated for each failure probability Pf. However, the Master-
Curve is only defined for quasi-static loading and mode I fracture of 1 T-sized specimen (a 1 T specimen is a specimen with a width B =
25.4 mm or 1 in). Therefore, using the Master-Curve with different specimen sizes requires the conversion of fracture toughness values
of the miniaturized specimens to values corresponding to 1 T-CT specimen equivalents. In the ASTM-E1921 standard, the specimen
measuring capacity is defined as:
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
E′.b0 .σYS
KJc,limit = (1)
Mlim

where E′ = (1− Eν)2 , b0 is the ligament length, σYS is the yield strength and the recommended value for Mlim is 30Mlim. Every specimen
failing at a KJc value lower than KJc,limit is considered as having a high level of constraint at fracture, or in others words considered as
having negligible constraint loss effects on the measured toughness. Consequently, the specimen size adjustment recommended in the
standard accounts only for the statistical effects. This adjustment is performed using a coefficient k that links the two fracture
toughness values:
K2 = k.K1 (2)
where 1 refers to the miniaturized specimen and 2 refers to the reference specimen (in this case 1 T-CT). The correction factor k
based on the ratio of crack front lengths to account for this effect is defined as:
( )1/4
B1
(3)
1
kASTM = = r4
B2
In addition, this specimen size adjustment is modified by introducing a threshold Kmin, which is often accepted to be 20 MPa.m1/2
and which represents a minimum loading level that generates a stress gradient at the crack tip low enough for an initiated crack to
propagate.
( )1/4
B1
K2 = Kmin + (K1 − Kmin ) (4)
B2
Again, the previous equation is considered valid as long as the measured fracture is lower than KJc,limit. However, care must be
taken that Mlim in Eq. (1) is sufficiently large to avoid constraint loss. The guideline is that varying Mlim should not affect the calculated
reference temperature T0. To this effect, Joyce and Tregoning [6], Odette et al. [13], and Rathbun et al. [14] showed that the minimum
value to satisfy this condition was 50–80 for CT specimens and 100–300 for bend bar specimens, i.e. far above the consensus value of
30. These findings were later confirmed by Mueller et al. [15]. While this approach is accurate and practical, it also features a sig­
nificant disadvantage: Increasing the Mlim factor also lowers the fracture toughness limit defined in Eq. (4) by a factor of (30/Mlim)1/2.
This can push the applicability range of the kASTM factor below the fracture toughness values of interest, especially in miniaturized
specimens where scatter of fracture toughness tends to be higher [13].
To find a more thorough solution to the problem of inaccuracies in fracture toughness measurements using miniaturized specimens,
it is worth remembering the origin of the ASTM correction factor kASTM. This factor is derived from a combination of a local criterion
for brittle fracture based on the attainment of a critical stressed volume (i.e. volume encompassing a critical stress) and small-scale
yielding theory (SSY). In SSY conditions, any given stressed area A (area encompassing a given stress) scales with K4, yielding the
following equations:
Stressed area in SSY:

A ≡ α.K 4 (5)
Critical volume:

V * = Bi .Ai = α.B1 .K14 = α.B2 .K24 (6)

Scaling:

3
B.A. Baykal and P. Spätig Engineering Fracture Mechanics 293 (2023) 109682

( )1/4 ( )1/4
B1 A2
K2 = .K1 = .K1 (7)
B2 A1
which is equivalent to Eq. (3) without a threshold minimum stress intensity factor. It is important to note that the ASTM correction
factor depends only on the ratio of crack front lengths between the tested specimen and the reference specimen. Therefore, this
correction factor does not take the effect of constraint loss into consideration. Equally important is the fact that constraint loss in­
creases with increasing loading, which may explain the necessity of further restricting the range of applicability of the ASTM correction
factor (since constraint loss can be neglected only at low K values).
This paper aims to devise and implement a second correction factor based on constraint loss in order to supplement the ASTM
correction factor, allowing the correction to be valid on a broader range. The approach is based on the local approach to brittle
fracture, where cleavage fracture is triggered when a critical stress σ* is reached in a critical volume of V* around the crack tip. Since
V* is dependent only on material properties, it can be assumed to be constant for the same material (Eurofer97) even with different
specimen sizes. This model is essentially deterministic in the sense that the σ*-V* are calibrated in the transition for a given failure
probability only. This is different from the Beremin’s model based on the Weibull stress, which is obtained by integration the principal
stress over an arbitrary volume usually taken as the plastically deformed volume at the crack tip [16]. Since the volume in a CT
specimen is the side area multiplied by the crack front length (i.e. width of the specimen):
B1 .A1 = B2 .A2 (8)
The idea is to compare K values between two specimen sizes at the same area A affected by the critical stress. Without constraint
loss, the K values would depend only on the ratio of crack front lengths as shown in Eq. (7). Constraint loss causes deviation from this
relationship, since the ratio of K values of different specimen sizes becomes dependent on the loading itself. The constraint loss effect
can be accounted for using the following correction factor, which quantifies the deviation of K values between two specimen sizes with
the same stress state (i.e. critically stressed area):

K1 ⃒
kCL = ⃒⃒ (9)
K2 A1 =A2

Combining this constraint loss correction factor with the ASTM correction factor, which accounts for the statistical effect, yields a
modified combined correction factor as follows:
kM = kASTM .kCL (10)
Plugging this modified factor into Eq. (1), we obtain our proposed new equation for converting stress intensity values between
different specimen sizes:
( )1/4
B1
K2 = kM .K1 = kASTM .kCL .K1 = .kCL (K1 ).K1 (11)
B2
It is important to note that kCL is a function that is dependent on the loading of the miniaturized specimen itself, through the
parameter K1. Therefore, the ratio of K-values for different sized specimens in Eq.11 is dependent on both the ratio of crack front
lengths and the loading state, which accounts for constraint loss. This leads to the main objective of this study, which is to construct and
implement the function kCL(K) for Eurofer97 steel in order to facilitate accurate prediction of stress intensity equivalence between
miniaturized and full-sized specimens.

3. Material and methods

The subject material of this study is Eurofer97, a high chromium, reduced activation tempered martensitic steel developed for
fusion reactor applications under the European Fusion Development Agreement [1]. The standard composition is 8.90 wt% Cr, 0.12 wt
% C, 0.46 wt% Mn, 1.07 wt% W, 0.2 wt% V, 0.15 wt% Ta, and balance Fe. W, V and Ta have been used to replace common alloying
elements like Ni, Nb and Mo in order to achieve the desired reduced activation property under irradiation. This is due to the shorter
half-life of radionuclides produced when exposed to neutron bombardment. The constitutive law used for the material in this study is
based upon specimens that were annealed at 980 ◦ C for 0.5 h and then tempered at 760 ◦ C for 1.5 h before being quenched, resulting in
a fully martensitic structure. The constitutive laws for − 120, − 100 and − 80 ◦ C are plotted in Fig. 1, experimentally obtained and
originally used in Mueller et al. [17]. In the temperature range of − 120 to − 80 ◦ C it has a Young’s modulus of 214 GPa and a Poisson’s
ratio of 0.33.
In order to simulate the loading of a CT specimen, an FEM simulation was created with each constitutive law for − 120 ◦ C, − 100 ◦ C
and − 80 ◦ C, assuming elastic–plastic behavior. The FEM calculations were performed using the commercial code ABAQUS/Standard
that solves nonlinear equilibrium equations. ABAQUS models the effect of the multi-axial stress state using the von Mises stress po­
tential and associated J2 plasticity. Since large deformation levels were expected in all the simulations that were run, a finite strain
(“large displacements theory”) approach was used. The von Mises isotropic hardening rule was used in the simulations. Since CT
specimens’ geometry is directly scalable around the stressed region (within the bounds of ASTM E1921, which defines the specimen
thickness–width relationship as B = 0.5 W, leading to self-similarity due to all dimensions depending on a single parameter) it is
possible to calculate equivalent loads, displacements and volumes directly from the simulation of a CT-1 T specimen for all sizes. This

4
B.A. Baykal and P. Spätig Engineering Fracture Mechanics 293 (2023) 109682

Fig. 1. Constitutive laws for Eurofer97 in the − 120 to − 80 ◦ C temperature range.

Fig. 2. FE model of a CT-1 T quarter specimen and pin, with mesh displayed.

5
B.A. Baykal and P. Spätig Engineering Fracture Mechanics 293 (2023) 109682

cuts down the number of simulations necessary to implement our proposed constraint loss correction. As a general rule, distances scale
linearly, load and area scales with the square and volumes scale with the cube of the miniaturization ratio.
The simulation consisted of one quarter of a CT-1 T specimen with two planes of symmetry along the crack plane (y = 0) and the
middle plane of the specimen (z = B/2), as shown in Fig. 2. The specimen has a hole in which a rigid pin was modeled as a shell. Around
the pin hole two rows of purely elastic elements were included in order to avoid plastic deformation in this region during the
movement of the pin. Boundary conditions include the plane representing the middle of the specimen (z = B/2) being fixed along the z-
axis and the crack middle surface (y = 0) being fixed along the y-axis. The ligament surface was constrained from moving vertically and
the pin was restricted to vertical movement only. The crack was modeled as a semicircular shape with a radius of 5 µm. The mesh is
shown on Fig. 2. It was refined and oriented to capture the large gradients of load and displacement near the crack tip. Along the width
of the specimen, the mesh was refined (biased) toward the free surface. These refinements were included due to the expectation of
larger strain/stress gradient in these regions. Displacement was applied to the pin in steps of 25 µm (50 µm for a full specimen) until the
deformation was too high for the next step to converge, using Abaqus/CAE 2021. The simulations at lower temperatures were thus able
to go to higher displacement values due to experiencing less plastic deformation as shown in the constitutive laws. At such high
loading, the initial root radius of the crack was found to have no effect since the crack was fully open. J-integrals with 30 contours had
been calculated for the same geometry in a previous work [7] on each position of the z-axis and used to confirm that the number of
elements on the z-axis were appropriate for the gradients encountered.

4. Results and discussion

The load–displacement curves for standard and miniaturized CT specimens obtained in the FEM simulations is given in Fig. 3. Due
to lower stiffness and higher deformation with equivalent displacement, the higher temperature simulations converge for a smaller
displacement range, but the behavior observed is not unexpected. These load–displacement curves were utilized to calculate stress
intensity factors throughout the loading range for both specimen sizes. The resulting relationship between displacement and stress
intensity factor in both specimen sizes is given in Fig. 4.
In order to calculate the constraint loss correction factor kCL, the necessary inputs are K and A. K can be calculated from the
load–displacement curve of the given specimen size obtained form the FEM simulation as an equivalent stress intensity factor obtained
from J integral values (KJ=(J.E’)1/2, where E’=E/(1-ν2)). This equivalent stress intensity factor can be used to compare different
specimen sizes’ behavior, but is only equal to the actual stress intensity factor under small scale yielding conditions. The critically
stressed area A can be obtained without a small scale yielding assumption using Eq. (6). For this purpose, a postprocessing subroutine
for Abaqus/CAE was written to calculate the critically stressed volume V. This program works by checking the maximum principal
stress on each node belonging to a particular element. If at least half of the nodes show a maximum principal stress at or above the
critical stress value, the element is flagged as part of the critical volume. After all elements have been checked, the volumes of all
elements found to be critically stressed are added together to obtain the total critically stressed volume V, which can be divided by the
specimen width/crack front length in order to obtain A.
Obviously, in order to be able to calculate the aforementioned parameters that are necessary to obtain the proposed constraint
correction factor, the critical stress for Eurofer97 at the given conditions must be known. Critical stress is usually calculated by plotting
stressed volume vs. stress plots for a given failure probability at different temperatures and searching for a point or range where the
resulting curves cross. In this work, the selected temperatures were − 120 ◦ C, − 100 ◦ C and − 80 ◦ C. The corresponding toughness values
of the median curve (50 % failure probability) and upper bound (99 % failure probability) have been calculated with a modified
master-curve indexed at − 80 ◦ C [18]. A detailed description and example of the calibration procedure have been provided by Lucas
et al.[19]. Fig. 5 clearly shows that the critical stress is in the 2000–2050 MPa range for both 50 % (median) and 99 % (upper bound)
failure probabilities. It is thus concluded that the critical stress does not change over this range of failure probability, but the critical
volume increases. Other materials may have a variation in critical stress or a combination of both the stress and volume as failure
probability changes. For simplicity, a critical stress of 2000 MPa was used in all calculations.
After performing these calculations, we check to see if the small scale yielding condition is satisfied. This condition is represented in
Eq. (5), which shows that a plot of K4 vs. A should be linear with a slope of α if the small scale yielding condition is satisfied. Fig. 6
shows that this condition is indeed satisfied for the temperature range in consideration for the 1 T specimen. However, deviation from
linearity and thus constraint loss occurs at very low K values for the 0.18 T specimen, as low as 40 MPa.m1/2. The deviation of
equivalent stress intensity between the two specimen sizes at the same A value is due to constraint loss and forms the basis of the kCL
factor accounting for constraint loss effects (Eq.9). Then, several A values are selected and the corresponding K values for both
specimen sizes are extracted from the simulation data. After applying Eq. (9) to obtain a set of data points for kCL, a curve can be fitted
(as in Fig. 7) to obtain a continuous kCL(K) function for each temperature. These can be used in Eq. (11) in combination with the ASTM
correction factor to calculate a combined correction factor kM. This factor can then be used to calculate equivalent K values between a
specific pair of different specimen sizes. Figs. 6 and 7 show that the data has good correlation and thus these operations can be
performed with confidence.
It should be noted that the combined correction factor, which contains the constraint loss correction factor, is not constant
throughout the loading range. This is physically logical, since constraint loss is not expected to have a strong effect without sufficient
loading. Indeed, the plots of the kCL(K) functions show that higher loading leads to higher projected constraint loss and thus a stronger
effect on the equivalent stress intensity factor on a different specimen size. Furthermore, it should be noted that the constraint loss
function does not appear to be strongly dependent on temperature in the range considered in this work, i.e. − 120 to − 80 ◦ C. It depends
only on the stress intensity factor K, or in other words, the loading conditions the specimen is subjected to.

6
B.A. Baykal and P. Spätig Engineering Fracture Mechanics 293 (2023) 109682

Fig. 3. Load-displacement curves obtained from FE model for 1 T and 0.18 T specimens.

The ASTM correction factor (defined in Eq. (2) and the proposed combined correction factor (defined in Eq. (11) were then applied
to the K-values obtained from miniaturized specimens to determine how accurate the prediction of equivalent stress intensity factor for
a different specimen size was. The miniaturized specimens were used as the raw data and the 1 T-specimens as target data due to the
practical application of calculating equivalent stress intensity on a 1 T-specimen for use of the master curve method, but it is also
theoretically possible to predict the behavior of a smaller specimen using this method. The results are shown as a function of the front
face displacement of the miniaturized specimens in Figs. 9-11. It should be noted that the x-axis, the displacement of the 0.18 T
specimen, is scaled; the K(1 T) curve is based on the equivalent displacement on a 1 T specimen, which e.g. for a front face
displacement of 0.18 mm on a 0.18 T specimen would correspond to a displacement of 1 mm on a 1 T specimen. This is why the K(1 T)
curves show higher K values than the K(0.18 T) curves unlike in Fig. 4.
In order to check that the simulations actually conform to real data, the K-values obtained were compared with fracture toughness
database obtained by Mueller et al.[20]. As seen in Fig. 8, the 50 % and 99 % failure probability curves fit well with the real data.
The combined correction factor clearly displays superior accuracy compared to the recommended ASTM correction, remaining
reasonably accurate until about 0.25 mm displacement on the CT-0.18 T specimen, which roughly corresponds to K = 85 MPa.m1/2 (or
equivalent to 300 MPa.m1/2 on a full-sized CT-1 T specimen) based on Figs. 8-11. This is a major improvement on the toughness limit
imposed by the M factor when using the ASTM factor only, as shown on Eq. (1).

7
B.A. Baykal and P. Spätig Engineering Fracture Mechanics 293 (2023) 109682

Fig. 4. KJ vs. d curve for 0.18 T and 1 T specimens.

Fig. 5. V vs σ plot showing critical stress and volume for − 120 to − 80 ◦ C temperature range.

At higher loading, the calculated K values diverge sharply from the simulation results, but this is as expected, since the
load–displacement curves (Fig. 3) show that beyond this level of deformation, the material starts behaving more plastically as the rate
of deformation rapidly increases. This is a natural result of the constitutive law of the subject material, and the limit of applicability of
the combined correction factor may vary if applied to different materials.
Fig. 12 shows the same concept applied at − 80 ◦ C to a different sized specimen, in this case CT-0.36 T (a proportionally scaled
specimen, as per ASTM-E1921). Notably, the constraint loss correction is significantly smaller since the two specimen sizes are closer to
each other and the effect of constraint loss is therefore reduced. In spite of this difference, the fit and relationship with the real 1 T
specimen’s stress intensity factors is the same. It can thus be concluded that the proposed correction factor should work for different
miniaturization ratios without issues.
Overall, we can conclude that accounting for constraint loss in miniaturized specimens yields a significantly better accuracy in

8
B.A. Baykal and P. Spätig Engineering Fracture Mechanics 293 (2023) 109682

Fig. 6. A vs. K4 curves for the confirmation of the small scale yielding condition for − 120 ◦ C, − 100 ◦ C and − 80 ◦ C.

9
B.A. Baykal and P. Spätig Engineering Fracture Mechanics 293 (2023) 109682

Fig. 7. KCL curves over ranges of loading for − 120 ◦ C, − 100 ◦ C and − 80 ◦ C.

10
B.A. Baykal and P. Spätig Engineering Fracture Mechanics 293 (2023) 109682

Fig. 8. Fracture toughness-temperature relationship predicted with the combined correction factor kM compared to conventional master curves and
experimental data from Mueller et al. [16].

Fig. 9. Comparison of raw K data, ASTM correction and combined correction factor for a CT-0.18 T specimen at − 120 ◦ C.

predicting the behavior of full-sized specimens and related structures at the cost of being a more complex procedure than simply
considering the statistical effect of the crack front length alone. Improvements to this technique are certainly possible, for example by
introducing additional crack front adjustment based on the stress state along the crack front. Looking at the data obtained by Mueller
et al., the apparent limit toughness of 300 MPa.m1/2 is enough to accurately correct toughness results for most fracture[17] tests
performed in the temperature range in question. It has thus been shown that the proposed kCL and kM correction factors provide a
significant improvement in transferability of toughness results between specimen sizes and geometries and usefulness in practical
applications such as the master curve method/DBTT reference temperature determination for Eurofer97 as a safety critical fusion
reactor material.

11
B.A. Baykal and P. Spätig Engineering Fracture Mechanics 293 (2023) 109682

Fig. 10. Comparison of raw K data, ASTM correction and combined correction factor for a CT-0.18 T specimen at − 100 ◦ C.

Fig. 11. Comparison of raw K data, ASTM correction and combined correction factor for a CT-0.18 T specimen at − 80 ◦ C.

Fig. 12. Comparison of raw KJ data, ASTM correction, and combined correction for a CT-0.36 T specimen at − 80 ◦ C.

12
B.A. Baykal and P. Spätig Engineering Fracture Mechanics 293 (2023) 109682

5. Conclusion

In this work, we have defined and successfully implemented a second correction factor in addition to the generally accepted ASTM
correction (for the statistical effect of miniaturization) which represents the constraint loss effect observed in miniaturized specimens
and avoids overestimation of the toughness of full sized specimens, potentially avoiding critical failures. It was shown that the
constraint loss correction combined with the statistical (ASTM) correction applied to K data from a miniaturized specimen fits better
with the real behavior of a full sized specimen over a larger range than the currently widely accepted correction method. The
transferability of the results to different sized specimens and the assumptions behind the model were checked and confirmed suc­
cessfully, with a new toughness limit of around 300 MPa.m1/2 being observed. On the load–displacement curves this coincided with
strongly plastic behavior.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data availability

Data will be made available on request.

Acknowledgements

The authors are grateful for the financial support provided by the Swiss Federal Nuclear Safety Inspectorate (ENSI) through the
PROACTIV project (Contract No. CTR-00489).

References

[1] Möslang A, Diegele E, Klimiankou M, Lässer R, Lindau R, Lucon E, et al. Towards reduced activation structural materials data for fusion DEMO reactors. Nuclear
Fusion 2005;45:649.
[2] Spätig P, Chen J-C, Odette GR. Chapter 11 - Ferritic and Tempered Martensitic Steels. In: Odette GR, Zinkle SJ, editors. Structural Alloys for Nuclear Energy
Applications. Boston: Elsevier; 2019. p. 485–527.
[3] Williams T, Nanstad R. Chapter 10 - Low-Alloy Steels. In: Odette GR, Zinkle SJ, editors. Structural Alloys for Nuclear Energy Applications. Boston: Elsevier;
2019. p. 411–83.
[4] Mastilovic S, Djordjevic B, Sedmak A. A scaling approach to size effect modeling of J(c) CDF for 20MnMoNi55 reactor steel in transition temperature region.
Engineering Failure Analysis 2022;131.
[5] ASTM E1921-22a Standard Test Method for Determinationof Reference Temperature, To, for Ferritic Steels in the Transition Range. 2022.
[6] Joyce JA, Tregoning RL. Determination of constraint limits for cleavage initiated toughness data. Engng Fract Mech 2005;72:1559–79.
[7] Baykal BA, Spätig P. Analysis of crack front loading to improve reliability of fracture toughness calculations based on miniaturized CT specimens. Procedia
Structural Integrity 2022;42:1350–60.
[8] Wallin K. The size effect in KIc results. Engineering Fracture Mechanics. 1985;22 No.1:149-63.
[9] Cravero S, Ruggierl C. A Two-Parameter Framework to Describe Effects of Constraint Loss on Cleavage Fracture and Implications for Failure Assessments of
Cracked Components. J Braz Soc Mech Sci Engng 2003;25:403–12.
[10] Dodds RH, Fong Shih C, Anderson TL. Continuum and micromechanics treatment of constraint in fracture. International Journal of Fracture 1993;64:101–33.
[11] Nevalainen M, Dodds RH. Numerical investigation of 3-D constraint effects on brittle fracture in SE(B) and C(T) specimens. International Journal of Fracture
1996;74:131–61.
[12] Wallin K. Irradiation damage effects on the fracture toughness transition curve shape for reactor pressure vessel steels. International Journal of Pressure Vessels
and Piping 1993;55:61–79.
[13] Odette GR, Yamamoto T, Kishimoto H, Sokolov M, Spätig P, Yang WJ, et al. A master curve analysis of F82H using statistical and constraint loss size adjustments
of small specimen data. Journal of Nuclear Materials 2004;329–333:1243–7.
[14] Rathbun HJ, Odette GR, He MY, Yamamoto T. Inluence of statistical and constraint loss size effects on cleavage fracture toughness in the transition - A single
variable experiment and database. Engng Fract Mech 2006;134–58.
[15] Mueller PSP. 3D finite element and experimental study of the size requirements for measuring toughness on tempered martensitic steels. Journal of Nuclear
Materials 2009;389:374–84.
[16] Beremin FM, Pineau A, Mudry F, Devaux J-C, D’Escatha Y, Ledermann P. A local criterion for cleavage fracture of a nuclear pressure vessel steel. Metallurgical
Transactions A 1983;14:2277–87.
[17] Mueller PF. Finite element modeling and experimental study of brittle fracture in tempered martensitic steels for thermonuclear fusion applications. Lausanne:
PhD-4518 EPFL; 2009.
[18] Mueller PSP, Bonadé R, Odette GR, Gragg D. Fracture toughness master curve analysis of the tempered martensitic steel Eurofer97. Journal of Nuclear Materials
2009;386–388:323–7.
[19] Lucas GE, Odette GR, Sokolov M, Spätig P, Yamamoto T, Jung P. Recent progress on small specimen test technology. Journal of Nuclear Materials 2002;
307–311:1600–8.
[20] Mueller P, Spätig P, Bonadé R, Odette GR, Gragg D. Fracture toughness master curve analysis of the tempered martensitic steel Eurofer97. Journal of Nuclear
Materials 2009;386–388:323–7.

13

You might also like