Sei sulla pagina 1di 15

Engineering Fracture Mechanics 104 (2013) 1–15

Contents lists available at SciVerse ScienceDirect

Engineering Fracture Mechanics


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

Initiation and propagation fracture toughness of solid wood


under the mixed Mode I/II condition examined by mixed-mode
bending test
Hiroshi Yoshihara ⇑
Faculty of Science and Engineering, Shimane University, Nishikawazu-cho 1060, Matsue, Shimane 690-8504, Japan

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

Article history: Mixed-mode bending, double cantilever beam, and end-notched flexure tests were con-
Received 23 July 2012 ducted using specimens of spruce, and the initiation and propagation fracture toughness
Received in revised form 2 December 2012 under various mixed Mode I/II and pure Modes I and II conditions were determined. In
Accepted 22 March 2013
addition to the actual fracture tests, finite element analyses were conducted and the results
Available online 2 April 2013
were compared with those obtained from the actual tests. The elliptical criterion was appli-
cable in representations of the mixed Mode I/II initiation fracture toughness relationship,
Keywords:
whereas the linear criterion was applicable to the propagation fracture toughness
Mixed-mode bending test
Double cantilever beam test
relationship.
End-notched flexure test Ó 2013 Elsevier Ltd. All rights reserved.
Crack length correction
Compliance combination method

1. Introduction

Several test methods have been examined for measurement of the fracture mechanics properties of solid wood under var-
ious mixed Mode I/II conditions, including combined tension and shear loading [1], compact tension-shear loading (Arcan
mixed-mode loading) [2–4], single-edge-notched or centre-notched tension of a specimen with an off-axis crack [5,6], sin-
gle-leg bending (SLB) [7], double-cantilever beam with a crack deviating from the neutral axis [8,9], mixed-mode crack
growth [10], and dual actuator load frame (DALF) tests [11]. Among these methods, the mixed-mode bending (MMB) test,
which was originally developed by Reeder and Crews to examine the fracture mechanics properties of fibre reinforced plas-
tics (FRP) [12,13], has become the primary method for fracture mechanics analysis within this decade. The MMB test has
recently been applied for analyses of the mixed Mode I/II fracture mechanics properties of solid wood and wood-based mate-
rials [14–17].
The MMB test offers three advantages. (1) The mixed-mode ratios can be easily controlled using a simple apparatus. (2)
The test simply combines the Mode I double-cantilever beam (DCB) and Mode II end-notched flexure (ENF) tests, which are
derived using beam theory. Because the load-deformation equation in the MMB test is also derived using beam theory, the
mixed Mode I/II fracture mechanics properties can be obtained because the strain energy release rates and the fracture
toughness are based on energetic considerations that are mathematically well-defined in the DCB and ENF tests [18]. (3)
It is easy to allow the crack to propagate stably when the initial crack length is appropriately determined. This feature allows
for determination of the so-called R-curve. Furthermore, the initiation fracture toughness also can be obtained under various
mixed-mode ratios.

⇑ Tel.: +81 852 32 6508; fax: +81 852 32 6123.


E-mail address: yosihara@riko.shimane-u.ac.jp

0013-7944/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.engfracmech.2013.03.023
2 H. Yoshihara / Engineering Fracture Mechanics 104 (2013) 1–15

Nomenclature

2H depth of the specimen


2L total span length of the MMB and 3ENF specimen
A crack length
a0 initial crack length
aeqI Mode I equivalent crack length
aeqI0 initial Mode I equivalent crack length
aeqII Mode II equivalent crack length
aeqII0 initial Mode II equivalent crack length
B crack/specimen width
CLA load–deflection compliance at point A
CLA0 initial load–deflection compliance at point A
CLB load–deflection compliance at point B
CLB0 initial load–deflection compliance at point B
CSD load-longitudinal strain compliance at point D
CSD0 initial load-longitudinal strain compliance at point D
CSE load-longitudinal strain compliance at point E
CSE0 initial load-longitudinal strain compliance at point E
Ex Young’s modulus in the longitudinal direction
Ey Young’s modulus in the tangential direction
F jx nodal force at the crack tip in the x-direction
F jy nodal force at the crack tip in the y-direction
Gxy shear modulus in the longitudinal-tangential plane
GI Mode I energy release rate
GII Mode II energy release rate
GVCCT
I Mode I energy release rate obtained from the VCCT
GVCCT
II Mode II energy release rate obtained from the VCCT
GIc Mode I initiation fracture toughness
GIIc Mode II initiation fracture toughness
GIR Mode I propagation fracture toughness
GIIR Mode II propagation fracture toughness
P applied load
PNL load at the onset of nonlinearity
Q weight of the loading lever
W distance from the centre of gravity of the loading lever to the loading nose applied at the midspan of the spec-
imen
x length direction of the specimen, which corresponds to the longitudinal direction of wood
y traverse direction of the specimen, which corresponds to the tangential direction of wood
mxy Poisson’s ratio in the longitudinal-tangential plane
dA deflection at point A
dB deflection at point B
dF deflection at point F
dix relative crack face displacement between the nodes adjacent to the crack tip in the x-direction
diy relative crack face displacement between the nodes adjacent to the crack tip in the y-direction
Da length in the x- and y-directions of the element at the delamination front
DaI Mode I propagation crack length
DaII Mode II propagation crack length
DI Mode I additional crack length obtained by Williams’s analysis
DII Mode II additional crack length obtained by Wand and Williams’s analysis
D0I Mode I additional crack length calculated from the compliance combination method
D00II Mode II additional crack length calculated from the compliance combination method
ex longitudinal strain at the mid-point on the top surface of the cracked portion in the specimen
3ENF three-point bend end-notched flexure
DCB double-cantilever beam
MMB mixed-mode bending
VCCT virtual crack closure technique
H. Yoshihara / Engineering Fracture Mechanics 104 (2013) 1–15 3

However, similar to the DCB and ENF tests, the mixed Mode I/II fracture toughness cannot be accurately obtained using
the MMB test if the influences of crack tip rotation and deflection caused by shearing are not considered; therefore, these
influences must be corrected using a data reduction method. In the ASTM D6671/6671M-06 [19], the MMB method for uni-
directional fibre reinforced polymer matrix composites is standardised, and the crack tip rotation effect is considered by
introducing an additional crack length into the equation based on the analyses conducted by Williams [20,21] and Wang
and Williams [22]. However, in the actual fracture test, the additional crack length value may be more pronounced in prox-
imity of the fracture process zone (FPZ), which is the region where the material progressively softens ahead of the crack tip.
In the tests under pure Mode I or Mode II conditions, the influence of the FPZ is significant [23–42]; therefore, concern exists
that the fracture mechanics properties under the mixed Mode I/II conditions cannot be accurately predicted without consid-
ering the influence of the FPZ. In the previous studies conducted by Oliveira et al. [15] and de Moura et al. [16,17], who exam-
ined the MMB test of solid wood, the influence of the FPZ was considered by introducing the concept of an equivalent crack
length, which contains the additional crack length induced by the rotation around the crack tip and the FPZ.
In previous studies on the DCB and ENF tests, it has been suggested that the simultaneously measuring the deflection and
longitudinal strain at a certain point of the specimen along with the applied load is effective for easily determining the frac-
ture mechanics properties of wood by the fracture test alone while considering the influence of crack tip rotation, deforma-
tion by the shearing force, and the FPZ ahead of the crack tip [29,31,35]. This data reduction method, which was originally
proposed by the author, was named the ‘‘compliance combination method,’’ and it may prove promising for analysis of the
MMB test results.
In the present study, the MMB, DCB, and ENF tests were conducted using specimens of sitka spruce. The mixed Mode I/II
initiation fracture toughness was analysed using elementary beam theory, Williams’s modification for the crack tip rotation,
and the compliance combination methods under various mode mixtures. Additionally, the R-curve was determined using the
compliance combination method, and the mixed Mode I/II propagation fracture toughness and the initiation fracture tough-
ness were evaluated. Two-dimensional finite element analyses (2D-FEAs) were performed independently of the fracture
tests, and the mixed Mode I/II strain energy release rates at the delamination front were evaluated using the virtual crack
closure technique (VCCT) and the above mentioned data reduction methods. A comparison of the results demonstrated
the applicability and practicality of the data reduction method examined here in addition to the fracture mechanics prop-
erties of solid wood under the mixed Mode I/II conditions.

2. Mixed-mode bending analyses

Fig. 1 shows a diagram of the mixed-mode bending test for the mixed Mode I/II analysis. A specimen, which has a
width of B, a depth in the cracked portion of H, and a crack length of a, is supported by a span whose length is 2L. A
load of P is eccentrically applied to the loading lever at point F, and it is separated at the ends of the cracked portions
(points A and A0 ) and the midspan (point B), as shown in Fig. 1. The distance between the loaded point of the lever and
midspan is defined as c.
As shown in Fig. 1, the loads Pupper and Plower are applied at the ends of the upper and lower cracked portions, respec-
tively, whereas Pmid and Pend are applied at the midspan and at the end of the crack-free portion, respectively. These
loads are derived from those that contribute to the Modes I and II deformations, PI and PII, respectively, as follows
[12,13,19]:

Fig. 1. A schematic diagram of the mixed-mode bending (MMB) test.


4 H. Yoshihara / Engineering Fracture Mechanics 104 (2013) 1–15

8
>
> Pupper ¼ P I þ P4II
>
>
<
Plower ¼ PI þ P4II
ð1Þ
>
> Pmid ¼ PII
>
>
:
Pend ¼ P2II

The load, P, is partitioned into PI and PII as:


(
PI ¼ 3cL
4L
P
ð2Þ
PII ¼ cþL
L
P

The PI value should be positive; therefore, c P L/3.


When the deflections at the end of the upper cracked portion (point A in Fig. 1) and at the midspan (point B or B0 in Fig. 1)
are defined as dA and dB, respectively, the load–deflection compliances at points A and B, which are CLA and CLB, respectively,
are given as dA/P and dB/P, respectively. CLA and CLB is usually obtained by elementary beam theory. According to Williams
[20,21] and Wang and Williams [22], however, the cracked portion often behaves like a beam with a longer crack because
of the rotation around the crack tip, and the additional crack lengths that correspond to Modes I and II, defined as DI and DII,
respectively, are different [15–17,32–36]. By considering the rotation around the crack tip, CLA and CLB are given as follows:
8
>
> C LA ¼ 2ð3cLÞ ða þ DI Þ3
>
< Ex BH3 L
3
C LB ¼  2E3cL 3 ða þ DI Þ þ
cþL
8Ex BH3 L
½2L3 þ 3ða þ DII Þ3  ð3Þ
>
> x BH L
>
: ¼  14 C LA þ 8EcþL
BH3 L
½2L3 þ 3ða þ DII Þ3 
x

DI and DII are derived as follows [19–21]:


"  (   )#12
1 Ex C 2
D; DII ¼ H 32 ð4Þ
13k Gx;y 1þC

where k = 0.85 and 4.8 for the DCB and 3ENF specimens, respectively, Gxy is the shear modulus in the length-depth plane, and
C is represented as follows:
1
ðEx Ey Þ2
C¼ ð5Þ
kGxy
where Ey is the Young’s modulus in the depth direction. This modification method is called the ‘‘Williams’s crack tip rotation
modification,’’ and an analysis based on the Williams’s modification is adopted in the ASTM D6671/D6671M-06 [19]. Accord-
ing to elementary beam theory, DI = DII = 0.
In the actual DCB and ENF tests, the additional crack length values are more pronounced than that derived by Eq. (4) be-
cause of the deformation caused by the transverse shear force, the fracture process zone (FPZ) induced at the region ahead of
the crack tip, fibre bridgings, and other influences which cannot be predicted by beam theory [16,17,27–42]. Instead of using
DI and DII, Eq. (3) is modified by introducing D0I and D0II as the Modes I and II additional crack lengths, respectively, in which
the above-mentioned influences are contained. The Modes I and II equivalent crack lengths, defined as aIeq and aIIeq, respec-
tively, are derived as a + D0I and a + D0II , respectively.
In the MMB test, the Modes I and II energy release rates, which are defined as GI and GII, respectively, are simply derived
from the DCB and ENF analyses when the influence of rotation around the crack tip and the transverse shear force are con-
sidered as follows [12,13,15–17,19]:
8 2 2
h 2
i
< GI ¼ 3ð3cLÞ P
4E B2 H3 L2
ða þ DI Þ2 þ sE xH
12Gxy
x
ð6Þ
: G ¼ 9ðcþLÞ2 P2 ða þ D Þ2
II 16E B2 H3 L2 II
x

where s is Timoshenko’s shear factor, which has a value of 6/5 for a beam with a rectangular cross section. Based on elemen-
tary beam theory, DI = DII = 0, and sExH2/12Gxy is reduced from Eq. (6). As described above, however, the cracked portion be-
haves like a beam with a longer crack due to the FPZ and fibre bridgings in addition to the rotation around the crack tip and
the deformation induced by the shearing force [15–17,23–41]. As described above, the correction crack lengths D0I and D0II are
therefore introduced to the energy release rates in Eq. (6) instead of DI and DII, as follows:
8
> 2 2
D0I Þ2 3ð3cLÞ2 P 2 a2
< GI ¼ 3ð3cLÞ 2P ðaþ ¼ 4E B2 H3 L2Ieq
4E B H3 L2
x x
ð7Þ
> 2 2 2
: G ¼ 9ðcþLÞ2 P2 ðaþD0II Þ2 ¼ 9ðcþLÞ P aIIeq
II 16E B2 H3 L2 16E B2 H3 L2
x x

When determining GI and GII using Eq. (6), separate tests also should be conducted to determine the elastic constants Ex, Ey,
Gxy, and mxy in light of the DI and DII values. However, when measuring the longitudinal strain at a specific point of a
H. Yoshihara / Engineering Fracture Mechanics 104 (2013) 1–15 5

specimen, GI and GII can be determined by the fracture test alone using Eq. (7). When the longitudinal strain of the specimen,
ex, is measured at a point located at point D or E, as shown in Fig. 1, it is derived using the following equation, which omits
the influence of the crack length and the transverse shear force:
8
< 6PclD
Ex BH2 L
ðPoint DÞ
x ¼ ð8Þ
: 3PðcþLÞl2 E ðPoint EÞ
4Ex BH L

where lD and lE are the distances from the loading point at the cracked portion (point A) to points D and E, respectively
(0 < lD < a and a < lE < 2L). In the actual MMB test, the flexural Young’s modulus, Ex, was evaluated using Eq. (8), and the Ex
value obtained at point E was revealed to coincide well with that obtained from the four-point bending tests using crack-
free specimens, the details of which are described below. Therefore, the load-longitudinal strain relationship obtained at
point E was used in the analysis. The load-longitudinal strain compliance at point E, defined as CSE is represented as:

e 3ðc þ LÞlE
C SE ¼ ¼ ð9Þ
P 4Ex BH2 L
From Eqs. (3) and (9), aIeq and aIIeq are obtained as follows:
8 h i13
>
>
< aIeq ¼ a þ D0I ¼ 3ðcþLÞHl
8ð3cLÞ
E
 CCLA
SE
h   i1 ð10Þ
>
>
: aIIeq ¼ a þ D0II ¼ 2HlE C LB þ C LA  2 L3 3
C SE C SE 3

By substituting Eq. (10) into Eq. (7), Ex, a + D0I , and a + D0II can be eliminated, and GI and GII can be determined from the frac-
ture test alone. In the actual MMB test, however, the influence of the weight is significant; therefore, the equation should be
corrected according to the method determined in ASTM D6671/D6671M-06 [19,43]. As shown in Fig. 1, when the lever
weight and the distance from the centre of gravity to the centre loading point are defined as Q and W, respectively, then
GI and GII can be derived using Eqs. (7) and (10) as follows [19,43]:
8 h i2
>
> 2
2 C SE 3ðcþLÞHlE C LA 3
< GI ¼ ð3cLÞ ½ð3cLÞPþðWLÞQ
ðcþLÞBHLlE
 8ð3cLÞ C SE
h   i23 ð11Þ
>
> 2 2
: GII ¼ ð3cþLÞ ½ ð3cþLÞPþðWþLÞQ
4ðcþLÞBHLlE
 C SE
 2HlE CCLB SE
þ 4CC LA
SE
 23 L3

Using the data reduction method based on Eq. (11), namely, the ‘‘compliance combination method,’’ the GI and GII values can
be evaluated by the fracture test alone without measuring the crack length or any elastic constants, which are implicitly con-
tained in the load–deflection compliances, CLA and CLB, and the load-longitudinal strain compliance, CSE.

3. Finite element calculations

Independently of the actual fracture tests, which are detailed below, two-dimensional finite element analyses (2D-FEAs)
were conducted, and the influence of the data reduction method on the evaluation of the Modes I and II energy release rate
values was examined. The ANSYS 12 program was used for the FE analyses, which is a library program of Shimane University.
Figs. 2a–c shows the FE mesh used in the calculations and the boundary conditions corresponding to the MMB, DCB, and ENF
test simulations. The horizontal length of the model was 390 mm, and the model width, B, was 12 mm. The depth of the
model, 2H, was 18 mm. The model consisted of four-noded plane elements. The mesh was refined to be finer closer to
the crack tip, as shown in Fig. 2d. At the delamination front, the dimensions of the element were 0.5 and 0.5 mm in the
x- and y-directions, respectively. The crack length, a0, was 130 mm. Table 1 presents the elastic properties used in the present
calculations. Hereafter, the x- and y-directions correspond to the longitudinal and tangential directions of the wood. Point E
was determined as the node located at the midspan on the bottom surface, which coincides with point B0 , as shown in Fig. 1.
Therefore, the lE value was 180 mm (=L). The elastic properties were determined using the present four-point bending, asym-
metric four-point bending, and compression tests, which are described in the following.
During the simulated MMB tests, the distance between the supports (points A0 and C), 2L, was 360 mm, and a total load of
100 N was applied to points A and B in Fig. 3a under various ratios. The vertical and horizontal displacements of point A0
were constrained, whereas the vertical displacement of point C was constrained. The loads applied to points A and B varied
based on the following procedure. Using elementary beam theory, DI = DII = 0, and the deflection caused by the transverse
shear force is neglected. Hence, the total energy release rate, G, is derived from Eq. (6) as:
 
3 39c2  18cL þ 7L2 P2 a2
G ¼ GI þ GII ¼ ð12Þ
16Ex B2 H3 L2
6 H. Yoshihara / Engineering Fracture Mechanics 104 (2013) 1–15

Fig. 2. The finite element (FE) meshes used in the simulations. Unit = mm. Pupper and Pmid values are listed in Table 2.

Table 1
The elastic constants used for the finite element analysis and data reduction.

Ex (GPa) Ey (GPa) Gxy (GPa) mxy


15.8 ± 1.0 0.73 ± 0.10 0.61 ± 0.04 0.49

Results are averages ± SD. x- and y-directions correspond to the longitudinal and tangential directions of solid wood.
H. Yoshihara / Engineering Fracture Mechanics 104 (2013) 1–15 7

Fig. 3. Typical examples of the load–deflection relationships obtained in the MMB tests under different mixed-mode ratios.

Therefore, the ratio of each mode can be obtained as:


8 2
< GI ¼ 4ð3cLÞ
G 39c2 18cLþ7L2
ð13Þ
: GII ¼ 3ðcþLÞ2
G 39c2 18cLþ7L2

The mixed-mode ratio was determined by varying the value of c in Eq. (13) as shown in Table 2. Then, the loads applied at
points A and B, which correspond to Pupper and Pmid, respectively, were determined from Eqs. (1) and (2). The Pupper and Pmid
values are also listed in Table 2. During the FE calculation, the load–deflection compliances at points A and B0 , which corre-
spond to CLA and CLB, respectively, were obtained. The load-longitudinal strain compliance at point E (B0 ), which corresponds
to CSE, was obtained from the load-longitudinal strain relationships. Based on elementary beam theory and Williams’s mod-
ification, the GI and GII values were obtained from Eq. (6). In the compliance combination method, the GI and GII values were
obtained from Eq. (11) while the weight of the lever was ignored (Q = 0).
During the simulated DCB test, a vertical load, P, of 100 N was oppositely applied to the loading points (points A and A0 in
Fig. 3b) while the node at the mid-depth of the right end was fixed. The load–deflection compliance at point A and the load-
longitudinal strain compliance at point D, given as CLA and CSD, respectively, were obtained. The GI value was calculated from
all three data reduction methods described above. The equations used for the data reductions are demonstrated in previous
papers [31,35].
In the simulated ENF test, the distance between the supports (points A0 and C), 2L, was 360 mm, and a vertical load, P, of
100 N was applied to the midspan (point B) as shown in Fig. 3c. The load–deflection compliance at the node located behind
the loading point (point B0 ), CLB, and the load-longitudinal strain compliance at point E (B0 ), CSE, was obtained. The GII value
was calculated from the data reduction methods in a similar manner to that adopted in the MMB test analysis. The equations
used for the data reductions have been demonstrated in previous reports [29,35].
The GI and GII values were also calculated using the virtual crack closure technique (VCCT), the details of which are de-
scribed in [35,44]. The GI and GII values obtained from the data reduction methods and VCCT were compared, and the validity
of the three data reduction methods for obtaining the GI and GII values was examined.

4. Experimental

4.1. Materials

Sitka spruce (Picea sitchensis Carr.) lumber, with a density of 440 ± 20 kg/m3 at 12% moisture content (MC) and eight or
nine annual rings contained in a radial length of 10 mm, was used for the tests. The annual rings were sufficiently flat;

Table 2
The target mixed mode ratio and the applied load, corresponding to the c value used in the finite element analysis of the mixed-
mode bending test.

c (mm) GI/G GII/G Pupper (N) Pmid (N)


70 0.019 0.981 38.889 138.889
90 0.129 0.871 50.000 150.000
110 0.263 0.737 61.111 161.111
130 0.380 0.620 72.222 172.222
160 0.509 0.491 88.889 188.889
230 0.674 0.326 127.778 227.778
345 0.780 0.220 191.667 291.667

Pupper and Pmid correspond to the loads applied at points A and B in Fig. 1, respectively. GI/G and GII/G are obtained from Eq. (13).
8 H. Yoshihara / Engineering Fracture Mechanics 104 (2013) 1–15

therefore, their curvature could be ignored. This lumber contained no defects, such as knots or grain distortions; therefore,
the specimens cut from it could be regarded as ‘‘small and clear.’’ The lumber was stored for approximately 1 year in a room
at a constant temperature of 20 °C and a relative humidity of 65% before the test, and it was confirmed to be in an air-dried
condition. These conditions were maintained throughout the tests. The equilibrium MC condition was approximately 12%.
The Young’s modulus in the longitudinal direction, which corresponds to Ex, was 15.8 ± 1.0 GPa. Five specimens were used
for one test condition.

4.2. Measurement of the elastic constants used in the finite element and experimental analyses

For the FE and experimental analyses, the Young’s moduli in the longitudinal and tangential directions, Ex and Ey, respec-
tively, and the shear modulus and Poisson’s ratio in the longitudinal-tangential plane, Gxy and mxy, respectively, are required.
The Ex, Ey, Gxy, and mxy values were measured using the four-point bending, asymmetric four-point bending, and four-point
bending using a strain gauge, respectively; the details of these tests are described below.
The Young’s modulus in the longitudinal direction, Ex, was measured using four-point bending tests. A beam specimen
with dimensions of 390 mm (longitudinal direction)  12 mm (radial direction)  9 mm (tangential direction) was prepared
from the lumber mentioned above. During the four-point bending test, the specimen was supported by the 360 mm spans
such that the top and bottom surfaces coincided with the longitudinal-tangential planes, and a load was applied to the inner
tri-sectional points at a crosshead speed of 2 mm/min. In this span/depth ratio (= 40), the influence of deflection caused by
the transverse shear force can be effectively eliminated [45,46]. Based on elementary beam theory, the Ex value was obtained
from the relationship between the load and deflection at the midspan.
The Young’s modulus in the tangential direction, Ey, was measured using compression tests. A short-column specimen,
whose dimensions were 40 mm (tangential direction)  20 mm (longitudinal direction)  20 mm (radial direction), was pre-
pared. Uniaxial strain gauges (gauge length = 2 mm; FLA-2–11, Tokyo Sokki Kenkyujo Co., Ltd., Tokyo, Japan) were bonded at
the centres of the longitudinal-tangential planes, and a compression load was applied along the long axis of the specimen at
a crosshead speed of 1 mm/min. The Ey value was obtained from the initial inclination of the stress–strain relationship.
For measuring the shear modulus Gxy and Poisson’s ratio mxy, a beam specimen with dimensions of 390 mm (longitudinal
direction)  12 mm (radial direction)  18 mm (tangential direction) was prepared from the lumber mentioned above. Tri-
axial strain gauges (gauge length = 2 mm; FRA-2–11, Tokyo Sokki Kenkyujo Co., Ltd., Tokyo, Japan) were bonded at the cen-
tres of the longitudinal-tangential planes of the specimen. During the four-point bending test, the specimen was supported
by the 360 mm spans such that the top and bottom surfaces coincided with the longitudinal-tangential planes, and a load
was applied to the inner tri-sectional points at a crosshead speed of 2 mm/min. The mxy value was determined from the lon-
gitudinal-transverse strain relationship. To prevent the specimen from experiencing a large deformation, the specimen was
unloaded when the longitudinal strain at the bottom plane attained a value of 0.001. To measure the shear modulus, Gxy, the
unloaded specimen was used for the following asymmetric four-point bending test.
In the asymmetric four-point bending test, the specimen was placed on the supporting blocks such that the side surfaces
of the specimen coincided with the longitudinal-tangential planes, and the specimen was asymmetrically supported and
loaded. The total span length, which corresponds to the distance between the left supporting point and the right loading
point, was 360 mm. The Gxy value was obtained from the load-shear strain relationship based on elementary beam theory.

4.3. Mixed-mode bending tests

All of the specimens were cut from the lumber mentioned above such that they were side-matched according to the
dimensions of 390 mm (longitudinal direction)  12 mm (radial direction)  18 mm (tangential direction). The crack was
produced along the longitudinal direction in the longitudinal-tangential plane, which is the so-called TL-system. The crack
was first cut with a band saw (thickness = 0.3 mm), and it was then extended ahead of the crack tip using a razor blade. After
the crack was cut in the specimen, loading blocks of medium-density fibreboard (MDF) with dimensions of 30 mm in length,
30 mm in height, and 12 mm in thickness were bonded using epoxy resin; the bonded positions were located at the upper
and lower cantilever portions, opposite to each other, as shown in Fig. 1. The initial crack length, a0, which was defined as the
distance from the line of load application in the cracked portion to the crack tip, was 130 mm. The a0 value was determined
as that which exceeds 0.7 times the half span length L, theoretically confirming stable crack propagation in the fracture test.
A load was applied to the specimen using the loading lever, as shown in Fig. 1, at a crosshead speed of 2 mm/min until the
crack had propagated close to the midspan. The total testing time was approximately 10 min.
A displacement gauge was placed above the loading pin of the upper cracked portion (point A) and below the bottom
surface of the midspan (point B0 ). Using these settings, the take-up of play in loading was effectively reduced. The load–
deflection compliances at point A and B0 (E), CLA and CLB, respectively, were obtained. The longitudinal strain, ex, was mea-
sured using a strain gauge (gauge length = 2 mm; FLA-2–11, Tokyo Sokki Kenkyujo Co., Tokyo), which was bonded at the top
surface of the mid-point of the upper cracked portion (point D; lD = a0/2 = 65 mm) and the bottom surface of the midspan
(point E (B0 ); lE = L = 180 mm). The Young’s modulus value obtained from the load-longitudinal strain relationship at points
D was 19.5 ± 3.3, whereas that at point E was 16.2 ± 1.2 GPa, which was close to that obtained from the four-point bending
test of a crack-free specimen (15.8 ± 1.0 GPa). In the bending loading, the longitudinal strain varies in the length direction of
the beam and the variation may affect the accuracy of the flexural Young’s modulus value measured using a strain gauge.
H. Yoshihara / Engineering Fracture Mechanics 104 (2013) 1–15 9

Additionally, it was of concern that the measurement error was induced because the output from the strain gauge was influ-
enced from the pointwise material property variation at the region where the strain gauge was bonded [47]. However, these
concerns can be reduced when the specimen is homogeneous and the strain gauge is short [45,46]. As described above, the
Young’s modulus value could be obtained accurately when measuring the longitudinal strain at the midspan. Therefore, the
value of CSE measured at point E was used in the compliance combination method. Fig. 4 shows the typical load/loading-line
displacement relationships obtained from the conditions under the lever lengths, c, of 345 and 70 mm.
The initial linear portions of the load–deflection compliances, CLA and CLB, were defined as CLA0 and CLB0, respectively. Sim-
ilarly, the linear portion of the initial load-longitudinal strain compliance, CSE, was defined as CSE0. The initiation fracture
toughness values of Modes I and II were determined by substituting the Young’s modulus, Ex, the correction crack lengths,
DI and DII, the initial load–deflection and load-longitudinal strain compliances, CLA0, CLB0, and CSE0, and the load at the onset
of nonlinearity, PNL, into the corresponding equation contained in Eqs. (6) and (11). In this study, the critical load for initial
crack propagation was determined as the load at the onset of non-linearity, which is similar to several previous studies
[35,38,39,41,48], although this determination might be provisional.
Using the compliance combination method, the R-curve that corresponds to each mode can be obtained. The Modes I and
II equivalent crack lengths at the initiation of crack propagation, defined as aIeq0 and aIIeq0, respectively, are derived by substi-
tuting the initial load–deflection compliances, CLA0 and CLB0, and the load-longitudinal strain compliance, CSE0, into Eqs. (10).
The Modes I and II propagation crack lengths, which are defined as DaI and DaII, respectively, are derived as aIeq  aIeq0 and
aIIeq  aIIeq0, respectively.
Fig. 4 shows the typical R-curves obtained under the lever lengths, c, of 345 and 70 mm. As shown in Fig. 3a, when the
lever length is long and the Mode I component is large, the load decreases with increasing deflection. As shown in Fig. 3b, the
load does not significantly decrease when the lever length is short and the Mode II component is large. Because of the de-
crease in load, the R-curve has a plateau when the Mode I component is dominant, as shown in Fig. 4a. In contrast, the energy
release rate tends to increase when the Mode II component is large because the load continuously increases during crack
propagation, as shown in Fig. 4b. When the crack tip is close to the midspan, the stress condition around the crack tip that
is disturbed by the loading nose might have an influence on the increase in the fracture toughness. This phenomenon might
be pronounced when the Mode II component increases. Additionally, the distance between the crack tip and the loading
point was short relative to the length of the FPZ, so confinement of the FPZ, which is more pronounced as the Mode II

Fig. 4. The resistance curves (R-curve) obtained in the MMB tests under the different mixed-mode ratios.
10 H. Yoshihara / Engineering Fracture Mechanics 104 (2013) 1–15

component is increased, might induce the monotonous increase in the R-curve [37]. The propagation fracture toughness is
usually determined when the fracture toughness value can be regarded as a constant value [12,13,15–17]. When the R-curve
had a significant plateau region, as shown in Fig. 4a, the propagation fracture toughness was determined as that at the pla-
teau region. Otherwise, the propagation fracture toughness was provisionally determined as that at the inflection point of
the R-curve.

4.4. Double cantilever beam tests

In a similar approach to the MMB tests, all of the specimens were cut from the lumber described above. The dimensions of
the specimen were similar to those of the specimen used in the MMB test. The crack was cut and the loading blocks were
bonded to the specimen in the same manner as that used in the MMB tests. The initial crack length, a0, was also 130 mm,
which was similar to that in the MMB test. A load was applied to the specimen by pins through universal joints at a cross-
head speed of 2 mm until the crack propagated to the mid-length. The total testing time was approximately 10 min.
The loading-line displacement, dA, was measured using a displacement gauge, whereas the longitudinal strain, ex, was
measured using an uniaxial strain gauge, which was similar to that used in the MMB test. The strain gauge was bonded
at the midpoint between the loading line and the crack tip on the top cantilever portion (lD = a0/2 = 65 mm). From the P  dA
and P  ex relationships, the load–deflection compliance, the load-longitudinal strain compliance, and the load at the onset of
nonlinearity, CLA, CSD, and PNL, respectively, were obtained. The initiation fracture toughness, GIc, and the Mode I R-curve
were determined using the compliance combination method, the details of which are given in [31,35]. Fig. 5a shows the
Mode I R-curve obtained from the DCB test. The Mode I R-curve had a significant plateau; therefore, the Mode I propagation
fracture toughness, GIR, was determined as that at the plateau region.

4.5. End-notched flexure tests

The specimen used in this test had dimensions similar to those of the specimens used in the MMB and DCB tests. The
crack was cut in the same manner as for the MMB and DCB tests. After the crack was cut, two sheets of 0.1-mm-thick Teflon
film were inserted between the crack surfaces to reduce the friction between the upper and lower cantilever beams. The
specimen was supported using 360 mm spans, and a load was applied to the midspan at a crosshead speed of 2 mm/min
until the crack propagated to the mid-span. The total testing time was approximately 10 min.
A displacement gauge was placed below the midspan, and the loading-line displacement, dB, was obtained. The longitu-
dinal strains were measured using an uniaxial strain gauge, which was similar to that used in the MMB and DCB tests,
bonded at the bottom surface of the midspan (lE = L = 180 mm). From the load–deflection and load-longitudinal strain rela-
tionships, the load–deflection compliance CLA, the load-longitudinal strain compliance, CSE, and the load at the onset of non-
linearity, PNL, were obtained. The initiation fracture toughness, GIIc, and the Mode II R-curve were determined using the
compliance combination method, the details of which are demonstrated in [31,35]. Fig. 5b shows the Mode II R-curve ob-
tained from the ENF test. Because of the concentration of stress around the loading point and the confinement of the FPZ,
the Mode II R-curve (GII  DaII relationship) often was observed to continuously increase without demonstrating a plateau
region; therefore, the Mode II propagation fracture toughness, GIIR, was determined as that at the inflection point of the R-
curve.

5. Results and discussion

From the four-point bending, asymmetric four-point bending, and compression tests, the Young’s moduli in the length
and depth directions, Ex and Ey, respectively, the shear modulus, Gxy, and Poisson’s ratio, mxy, were obtained and are listed

Fig. 5. The resistance curves (R-curve) obtained in the DCB and ENF tests.
H. Yoshihara / Engineering Fracture Mechanics 104 (2013) 1–15 11

Fig. 6. The ratios of the Modes I and II energy release rates obtained by the data reduction methods to those obtained by the VCCT method, given as GI/GVCCT
I
and GII/GVCCT
II , respectively. EBT, WM, and CC represent the analyses based on elementary beam theory, Williams’s modification, and the compliance
combination method, respectively.

in Table 1. As described above, these constants were used for the finite element calculations. Additionally, the additional
crack lengths, DI and DII, were determined by substituting these constants into Eq. (9).
Fig. 6 shows the ratios of the Modes I and II energy release rates obtained by each data reduction method to those ob-
tained by the VCCT method, denoted as GI/GVCCTI and GII/GVCCT
II , respectively. The GI value obtained from both the Williams’s
modification and the compliance combination methods are quite close to the GVCCT I value. In contrast, the GII value obtained
from the compliance combination method is larger than the GVCCT II value, and this trend becomes more pronounced as the
Mode II component increases. Nevertheless, this value is closer to the GVCCTII value than that obtained based on elementary
beam theory. This result indicates the effectiveness of the crack length correction. The GVCCTII value is closer to the GII value
obtained from the Williams’s modification method than that obtained from the compliance combination method. As dem-
onstrated below, however, the additional crack length values D0I and D0II obtained in the actual fracture tests are often larger
than the DI and DII values derived by Eq. (4), respectively. Additionally, the orientation of the fibres along the direction of
crack length propagation prevents the rotation of the crack tip. For these reasons, the DI and DII values obtained from Eq.
(4) are not applicable to the analysis of the Modes I and II fracture toughness values of solid wood.

Fig. 7. The relationships between the Modes I and II additional crack lengths obtained by the compliance combination method, D0I and D0II , respectively, and
the mixed-mode ratio, GII/G, and a comparison with those derived from Williams’s modification, DI and DII.
12 H. Yoshihara / Engineering Fracture Mechanics 104 (2013) 1–15

By substituting the initial load–deflection and load-longitudinal strain compliances and the initial crack length
(a0 = 130 mm) into Eq. (11), the Modes I and II additional crack lengths, D0I and D0II , respectively, can be obtained. Fig. 7 shows
the D0I  GII/G and D0II  GII/G relationships and the comparison with the DI and DII values obtained from Eq. (4). The D0I and D0II
values are significantly larger than the DI and DII values, respectively. Based on the DCB and ENF test analyses, Silva et al.
[30], de Moura et al. [32,33,36], Dourado et al. [34], Yoshihara and Satoh [35], and Morel et al. [37] suggested that Eq. (4)
is no longer valid for solid wood when a fracture process zone (FPZ), which is the region where the material progressively
softens, develops at the crack tip, and when the influence of the FPZ is more pronounced for the Mode II component than for
the Mode I component. Additionally, the influences of deflection caused by the transverse shear force and fibre bridgings, and
other influences which cannot be predicted by beam theory also enhances the increase of D0I and D0II values. This result is also
applicable for the MMB test; therefore, the large values of D0I and D0II may be because of the above-mentioned reasons. Con-
sidering the influence of the FPZ, Oliveira et al. [15] and de Moura et al. [16,17] determined the fracture toughness from the
profile of the R-curve under several mixed-mode conditions obtained from the MMB tests while the concept of the FPZ at the
crack tip is considered. Although their method is more complex than the data reduction methods examined here, its consid-
eration is valuable.
Fig. 8 shows the relationship between aIeq and aIIeq, obtained from Eq. (10) during the crack propagation. As described
above, the D0II value is significantly larger than the D0I value, so the aIIeq value is always larger than aIeq. From Eq. (10), how-
ever, the relationship between aIeq and aIIeq can be represented as aIIeq = aIeq + (D0II – D0I ) by eliminating a, and the experimen-
tally obtained relationship coincides well with this equation.
Fig. 9 shows the initiation and propagation fracture toughness envelopes in the GI  GII plane obtained from the different
data reduction methods. As described in the introduction section, there have been several studies on the fracture properties
of solid wood under the mixed Mode I/II conditions, and several empirical equations are used for describing the fracture cri-
terion. Based on several previous studies, the fracture criterion is often derived in a general form as follows:

8  m  n
>
< GGI þ GGIIcII ¼ 1 ðInitiation fracture toughnessÞ
Ic
 p  q ð14Þ
>
: GI þ GII ¼ 1 ðPropagation fracture toughnessÞ
GIR GIIR

Table 3 shows the Modes I and II initiation fracture toughness values, GIc and GIIc, respectively, and the propagation fracture
toughness values GIR and GIIR, respectively, that correspond to each data reduction method. The Modes I and II fracture
toughness values were determined by the DCB and ENF tests, respectively. Many examples have derived the empirical values
to the parameters m and n for solid wood and fibre reinforced plastics by conducting various mixed Mode I/II tests [1–17,49].
For the MMB test results, Kossakowski derived as m = 1 and n = 2.28 for the TL-system and m = 1 and n = 2.25 for the RL-sys-
tem of Scots pine (Pinus sylvestris L.) [14]. Based on Figs. 9a–c, it is sufficient to use the elliptical criterion (m = n = 2) for rep-
resenting the GI  GII envelope of the initiation fracture toughness independently from the data reduction methods. To
represent the GI  GII envelope for the propagation fracture toughness, de Moura suggested that the linear criterion
(p = q = 1) is applicable for maritime pine (Pinus pinaster Ait.) [16]. The results obtained in this research are in agreement
with their suggestion as shown in Fig. 9d. These envelopes may be influenced by the wood species, the fracture system,
the specimen configuration, etc.
Fig. 10 shows typical examples of the variation in the mixed-mode ratio during crack propagation. According to Eq. (13),
which is based on elementary beam theory, the crack length does not influence the mixed-mode ratio. When considering the
additional crack length and the lever weight as in Eq. (11), the mixed-mode ratio deviates from that predicted by Eq. (11) and

Fig. 8. The relationships between the Modes I and II equivalent crack lengths aIeq and aIIeq, respectively, obtained from Eq. (11).
H. Yoshihara / Engineering Fracture Mechanics 104 (2013) 1–15 13

Fig. 9. The relationship between GI and GII for the initiation and propagation of fracture and the envelopes of the linear and elliptical fracture criteria.

Table 3
The Modes I and II initiation fracture toughnesses GIc and GIIc, respectively, and the Modes I and II propagation fracture
toughnesses GIR and GIIR, respectively, obtained from the DCB and ENF tests.

Data reduction method GIc (J/m2) GIIc (J/m2)


Initiation fracture toughness
Elementary beam theory 75.3 ± 14.5 174.6 ± 19.1
Williams’s modification 95.0 ± 18.5 192.4 ± 21.1
Compliance combination 116.4 ± 9.5 294.0 ± 49.0
Data reduction method GIR (J/m2) GIIR (J/m2)
Propagation fracture toughness
Elementary beam theory – –
Williams’s modification – –
Compliance combination 212.5 ± 19.5 1061 ± 156

The results are average ± SD.

alternates during crack propagation, as shown in Fig. 10. Therefore, the fracture criterion does not retain its elliptical enve-
lope, which is shown in Fig. 9a–c, but instead is transformed into a linear relationship, as shown in Fig. 9d.
To measure the fracture mechanics properties under mixed Mode I/II conditions accurately, it is necessary to follow fur-
ther works. (1) To derive the critical load for crack propagation, several definitions can be followed [48], so it is necessary to
determine the critical loads for determining the initiation and propagation fracture toughness values. (2) In the MMB test,
the influence of FPZ ahead of the crack tip is significant; this is all true for both the DCB and ENF tests. The FEAs considering
the FPZ may aid in the accurate estimation of fracture mechanics properties under various mixed Mode I/II conditions. Nev-
ertheless, the additional crack lengths obtained in this study were longer than the length of the FPZ obtained in several pre-
vious studies [15–17,30,32–34,36]. The source of the large values of additional crack length can be revealed by microscopic
observations [25,26], full field observations such as digital image correlation (DIC) [40], and acoustic emission (AE) methods
[14], which are used to detect the initiation of crack propagation. In this manner, the precise determination method can be
14 H. Yoshihara / Engineering Fracture Mechanics 104 (2013) 1–15

Fig. 10. The variation in the mixed-mode ratio during the crack propagation.

enhanced. (3) The R-curve tends to increase monotonously under the Mode II dominant condition because of FPZ confine-
ment [37]. To avoid confinement of the FPZ, the distance between the crack tip and the loading point should be long, so
the specimen dimensions must be sufficiently large. The fixture should be designed in consideration of the specimen dimen-
sions. (4) When a cracked specimen is subjected to a flexural loading condition, the load–deflection behaviours often deviate
from those predicted by beam theory [48]. Compliance calibration method, in which the relationship between the loading-
line compliance and crack length is calibrated by conducting separate tests under varying the crack length independently of
the fracture tests, is effective in measuring the fracture mechanics properties while considering the deviation. In the DCB and
ENF tests, the fracture mechanics properties obtained from the compliance combination method coincided well with those
obtained from the compliance calibration method [29,31,35]. The accuracy of the fracture mechanics properties obtained
from the compliance combination here should be proved by comparing against those obtained from the compliance calibra-
tion method, in which numerous specimens with various crack lengths should be used.

6. Conclusions

MMB tests were conducted using specimens of sitka spruce to determine the mixed Mode I/II initiation and propagation
fracture toughness. The initiation fracture toughness was determined using three data reduction methods, including elemen-
tary beam theory, Williams’s crack tip rotation modification, and compliance combination methods. Using the compliance
combination method, the resistance curve (R-curve) that corresponds to each mixed-mode ratio can be obtained; therefore,
the propagation fracture toughness was determined using this data reduction method. In addition to the fracture tests, finite
element analyses (FEAs) were conducted and the validity of the MMB test methods was examined.
The FEA results demonstrated that the crack length correction is effective in the MMB, DCB, and ENF tests, and the Mode I
additional crack length was longer than the Mode II additional crack length. In the actual fracture tests, however, the Mode II
additional crack length was significantly larger than the Mode I additional crack length over the entire range of mixed-mode
ratios examined here.
The initiation fracture toughness envelope in the GI  GII plane tended to be elliptical, whereas the propagation fracture
toughness envelope was almost linear. The transformation of the fracture toughness envelope resulted because of alterna-
tion in the mixed-mode ratio due to crack tip rotation, deformation caused by the shearing force, and the fracture process
zone (FPZ) ahead of the crack tip.
H. Yoshihara / Engineering Fracture Mechanics 104 (2013) 1–15 15

References

[1] Wu EM. Application of fracture mechanics to anisotropic plates. Trans ASME J Appl Mech 1967;34:967–74.
[2] Valentin G, Caumes P. Crack propagation in mixed mode in wood: a new specimen. Wood Sci Technol 1989;23:43–53.
[3] Sato K, Yamamoto H, Takagi Y. Influence of PEG treatment on fracture of wood under mixed mode loading. J Soc Mater Sci Jpn 2000;49:373–8.
[4] Moutou Pitti R, Dubois F, Pop O, Absi J. A finite element analysis for the mixed mode crack growth in a viscoelastic and orthotropic medium. Int J Solid
Struct 2009;46:3548–55.
[5] Williams JG, Birch MW. Mixed mode fracture in anisotripic media. ASTM STP 1976. p. 125–37.
[6] Mall S, Murphy JF, Shottafer JE. Criterion for mixed mode fracture in wood. J Engng Mech ASCE 1983;109:680–90.
[7] Oliveira JMQ, de Moura MFSF, Morais JJL. Application of the end loaded split and single-leg bending tests to the mixed-mode fracture characterization
of wood. Holzforschung 2009:63 597–602.
[8] Jernkvist LO. Fracture of wood under mixed mode loading I. Derivation of fracture criteria. Engng Fract Mech 2001;68:549–63.
[9] Jernkvist LO. Fracture of wood under mixed mode loading II. Experimental investigation of Picea abies. Engng Fract Mech 2001;68:565–76.
[10] Moutou Pitti R, Dubois F, Pop O. A proposed mixed-mode fracture specimen for wood under creep loadings. Int J Fract 2011;167:195–209.
[11] Nicoli E, Dillard DA, Frazier CE, Zink-Sharp A. Characterization of mixed-mode I/II fracture properties of adhesively bonded yellow-poplar by a dual
actuator test frame instrument. Holzforschung 2012;66:623–31.
[12] Reeder JR, Crews Jr JH. Mixed mode bending method for delamination testing. AIAA J 1990;28:1270–6.
[13] Reeder JR, Crews Jr JH. Redesign of the mixed mode bending delamination test to reduce nonlinear effects. J Compos Technol Res 1992;14:12–9.
[14] Kossakowski PG. Mixed mode I/II fracture toughness of pine wood. Arch Civil Engng 2009;2:199–227.
[15] Oliveira JMQ, de Moura MFSF, Silva MAL, Morais JJL. Numerical analysis of the MMB test for mixed-mode I/II wood fracture. Compos Sci Technol
2007;67:1764–71.
[16] de Moura MFSF, Oliveira JMQ, Morais JJL, Xavier J. Mixed-mode I/II wood fracture characterization using the mixed-mode bending test. Engng Fract
Mech 2010;77:144–52.
[17] de Moura MFSF, Oliveira JMQ, Morais JJL, Dourado N. Mixed-mode (I + II) fracture characterization of wood bonded joints. Construct Build Mater
2011;25:1956–62.
[18] Adams DF, Carlsson LF, Pipes RB. Experimental characterization of advanced composite materials. 3rd ed. Boca Raton: CRC Press; 2003.
[19] ASTM D6671/D6671M-06. Standard test method for mixed mode I–mode II interlaminar fracture toughness of unidirectional fiber reinforced polymer
matrix composites. ASTM Int; 2006.
[20] Williams JG. On the calculation of energy release rates for cracked laminates. Int J Fract 1988;36:101–19.
[21] Williams JG. The fracture mechanics of delamination tests. J Strain Anal 1989;24:207–14.
[22] Wang Y, Williams JG. Corrections for Mode II fracture toughness specimens of composites materials. Compos Sci Technol 1992;43:251–6.
[23] Boström L. The stress-displacement relation of wood perpendicular to the grain Part 1. Experimental determination of the stress-displacement
relation. Wood Sci Technol 1994;28:309–17.
[24] Boström L. The stress-displacement relation of wood perpendicular to the grain Part 2. Application of the fictitious crack model to the compact tension
specimen. Wood Sci Technol 1994;28:319–27.
[25] Vasic S, Smith I. Bridging crack model for fracture of spruce. Engng Fract Mech 2002;69:745–60.
[26] Vasic S, Smith I, Landis E. Fracture zone characterization. Micro-mechanical study. Wood Fiber Sci 2003;34:42–56.
[27] Morel S, Bouchaud E, Schmittbuhl J. Influence of the specimen geometry on R-curve behavior and roughening of fracture surfaces. Int J Fract
2003;121:23–42.
[28] Morel S, Dourado N, Valentin G, Morais J. Wood: a quasibrittle material R-curve behavior and peak load evaluation. Int J Fract 2005;131:385–400.
[29] Yoshihara H, Mode II. Initiation fracture toughness analysis for wood obtained by 3-ENF test. Compos Sci Technol 2005;65:2198–207.
[30] Silva MAL, de Moura MFSF, Morais JJL. Numerical analysis of the ENF test for mode II wood fracture. Composites Part A 2006;37:1334–44.
[31] Yoshihara H, Kawamura T. Mode I fracture toughness estimation of wood by DCB test. Composites Part A 2006;37:2105–13.
[32] de Moura MFSF, Silva MAL, de Morais AB, Morais JJL. Equivalent crack based mode II fracture characterization of wood. Engng Fract Mech
2006;73:978–93.
[33] de Moura MFSF, Morais JJL, Dourado N. A new data reduction scheme for mode I wood fracture characterization using the double cantilever beam test.
Engng Fract Mech 2008;75:3852–65.
[34] Dourado N, Morel S, de Moura MFSF, Valentin G, Morais J. Comparison of fracture properties of two wood species through cohesive crack simulations.
Composites Part A 2008;39:415–27.
[35] Yoshihara H, Satoh A. Shear and crack tip deformation correction for the double cantilever beam and three-point end-notched flexure specimens for
mode I and mode II fracture toughness measurement of wood. Engng Fract Mech 2009;76:335–46.
[36] de Moura MFSF, Dourado N, Morais J. Crack equivalent based method applied to wood fracture characterization using the single edge notched-three
point bending test. Engng Fract Mech 2010;77:510–20.
[37] Morel S, Lespine C, Coureau JL, Planas J, Dourado N. Bilinear softening parameters and equivalent LEFM R-curve in quasibrittle failure. Int J Solid Struct
2010;47:837–50.
[38] Yoshihara H. Examination of the mode I critical stress intensity factor of wood obtained by single-edge-notched bending test. Holzforschung
2010;64:501–9.
[39] Yoshihara H. Influence of loading conditions on the measurement of mode I critical stress intensity factor for wood and medium-density fiberboard by
the single-edge-notched tension test. Holzforschung 2010;64:735–45.
[40] Murata K, Nagai H, Nakano T. Estimation of width of fracture process zone in spruce wood by radial tensile test. Mech Mater 2011;43:389–96.
[41] Yoshihara H, Usuki A. Mode I critical stress intensity factor of wood and medium-density fibreboard measured by compact tension test. Holzforschung
2011;65:729–35.
[42] Yoshihara H. Mode II critical stress intensity factor of wood measured by the asymmetric four-point bending test of single-edge-notched specimen
while considering an additional crack length. Holzforschung 2012;66:989–92.
[43] Chen JH, Sernow E, Schulz E, Hinrichsen G. A modification of the mixed-mode bending test apparatus. Composites Part A 1999;30:871–7.
[44] Rikards R, Buchholz FG, Wang H, Bledzki AK, Korjakin A, Richard HA. Investigation of mixed mode I/II interlaminar fracture toughness of laminated
composites by using a CTS type specimen. Engng Fract Mech 1998;61:325–42.
[45] Yoshihara H, Kubojima Y, Ishimoto T. Several examinations on the static bending test methods of wood using todomatsu (Japanese fir). Forest Prod J
2003;52(2):39–44.
[46] Yoshihara H, Tsunematsu S. Feasibility of estimation methods for measuring Young’s modulus of wood by three-point bending test. Mater Struct
2006;39:29–36.
[47] Yadama V, Davalos JF, Loferski JR, Holzer SM. Selecting a gauge length to measure parallel-to-grain strain in southern pine. Forest Prod J
1991;41(10):65–8.
[48] Davies P, Blackman BRK, Brunner AJ. Mode II delamination. In: Moore DR, Pavan A, Williams JG, editors. Fracture mechanics testing methods for
polymers adhesives and composites. ESIS Publication 28. Amsterdam: Elsevier; 2001.
[49] Szekrényes A. Prestressed composite specimen for mixed-mode I/II cracking in laminated materials. J Reinf Plast Compos 2010;29:3309–21.

Potrebbero piacerti anche