Sunteți pe pagina 1din 7

On the valid ligament range of specimens for the essential work

of fracture method: The inconsequence of stress criteria


Ferenc Tuba
a,
, Lszl Olh
b
, Pter Nagy
a
a
Department of Polymer Engineering, Faculty of Mechanical Engineering, Budapest University of Technology and Economics, M} uegyetem rkp. 3, 1111
Budapest, Hungary
b
Audi Hungaria Motor Kft., Kardn utca 1, 9027 Gy} or, Hungary
a r t i c l e i n f o
Article history:
Received 23 May 2012
Received in revised form 28 October 2012
Accepted 20 December 2012
Keywords:
Essential Work of Fracture (EWF)
Stress criterion
Ductile fracture
Plane-stress/plane-strain transition
Thin-walled structures
a b s t r a c t
The Essential Work of Fracture (EWF) method is widely used for the toughness determina-
tion of thin, ductile sheets under quasi plane-stress conditions. To dene the valid ligament
range of tests empirical role of thumbs and theoretical approaches are generally used. The
lower ligament limit is typically ascribed to the plane-strain/plane-stress transition,
which is accompanied by an increase in maximum net-section stress. Nevertheless, during
the measurements neither the assumed plastic-rigid behavior nor true geometric similarity
exists, which unts the theoretical stress criteria based on these approximations. In the
present study several materials were considered to explain the experimental results and
to clarify the possible ambiguities related the observed maximum stresses.
2012 Elsevier Ltd. All rights reserved.
1. Introduction
The importance of aw assessment methods for thin-walled structures has been emphasized in a recent overview by
Zerbst et al. [1]. In materials where the ligament yielding parameter dened as the ratio of net-section stress (r
ns
) and yield
stress (r
Y
) is below 0.5, the stress intensity factor concept is applicable and widely accepted. Up to a ligament yielding
parameter of 0.8 the plastic zone corrected stress intensity factor can be used. However, for materials of high ductility
and low strain hardening exponent the ligament yielding parameter is generally above 0.8, where neither the linear elastic
concepts nor the adjusted methods are appropriate. The elasticplastic concepts like J-integral or crack tip opening displace-
ment are theoretically not limited to small scale yielding, but the geometries of thin-walled structures generally do not meet
the thickness requirements of standard tests (ASTM E1820-11, ISO 12135). Although the resistance to crack initiation is gen-
erally independent of geometry, the specimen thickness inuences the ligament-yielding and the slope of R-curves, thus the
parameters obtained by standard methods (plane-strain conditions) cannot be simply transferred to thin structures (rather
plane-stress) [1].
The Essential Work of Fracture (EWF) method was proposed by Cotterell and Reddel [2] based on Brobergs theory [3] for
the description of plane-stress ductile fracture in thin sheets where the size of the plastic region is comparable with the crack
length. The theory was later extended for polymers by Mai and Cotterell [4], while more recently several reviews [5,6]
addressed its feasibility for polymer based systems.
In this method the ductile fracture is not characterized frominitiation measurements, but fromthe energy consumed dur-
ing the complete fracture of the specimen of thickness B. It supposes that one can split the total fracture work (W
f
) into a
0013-7944/$ - see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.engfracmech.2012.12.011

Corresponding author. Tel.: +36 1 463 1487; fax: +36 1 463 2003.
E-mail address: tuba@pt.bme.hu (F. Tuba).
Engineering Fracture Mechanics 99 (2013) 349355
Contents lists available at SciVerse ScienceDirect
Engineering Fracture Mechanics
j our nal homepage: www. el sevi er. com/ l ocat e/ engf r acmech
dissipative work of outer screening plastic zone (W
p
) and into an essential one (W
e
) of the inner fracture process zone. If the
ligament (L) of a sheet specimen yields before crack initiation and the plastic zone is conned, then the plastic work is pro-
portional to the plastic volume (BL
2
), while the essential one is proportional to fracture area (BL). The work performed in the
fracture process zone is assumed to be proportional to the initial ligament length, thus the testing of geometrically similar
specimens allows the separation of the specic essential (w
e
) and non-essential (w
p
) fracture terms:
W
f
BL
w
e
bw
p
L 1
where b () is a plastic zone shape dependent correction factor. It should be also noted that since in plane-stress the thick-
ness has to be vanishingly small w
e
and w
p
are not true material constants, but functions of sheet thickness (B) [2,7,8].
One of the fundamentals of the tests is the determination of the valid ligament range. Since the concept is based on the
extrapolation of experimental data to zero ligament length, small changes in slope could result in signicant deviations in
the fracture parameters.
One of the problems is related to the full-ligament yielding at larger ligaments and the self-similarity of the measure-
ments. The upper ligament limit (L
max
) can be given from the size of the plastic zone, which can be approximated from
the assumption that at large ligament lengths w
f
is asymptotic to the plane-stress fracture toughness [10]. The self-similarity
is generally checked arbitrarily, i.e. the similarity transformation is neglected, but as an alternative solution a displacement
criterion can be also used [11].
Nomenclature
B specimen thickness
E tensile modulus
e
u
elongation at failure
l specimen length
L ligament length
L
min
lower ligament limit of the tests
L
max
upper ligament limit of the tests
L mean of the studied ligament lengths
m constrain factor
P
max
maximum load during the fracture
s ordinate intercept of the P
max
/B versus L plots
w specimen width
w
e
specic essential work of fracture
w
f
specic total work of fracture
w
p
specic plastic work of fracture
W
e
essential work of fracture, i.e. the work consumed within the fracture process zone
W
f
total work of fracture
W
p
plastic work of fracture, i.e. the work dissipated outside of the fracture process zone
b plastic zone shape dependent correction factor
r slope of the P
max
/B versus L plots
r
Y
yield stress
r
m
mean of the maximum net-section stresses
r
ns
maximum net-section stress during the fracture
Abbreviations
DENT double edge notched tensile specimen
EPBC ethylene-propylene block copolymer
EWF essential work of fracture
HIPS high impact polystyrene
PBT poly(butylene-terephthalate)
PC polycarbonate
PCL poly(e-caprolactone)
PEEK poly(ether-ether-ketone)
PEI poly(ether-imide)
PEN poly(ethylene-naphthalate)
PET poly(ethylene-terephthalate)
PETG poly(ethylene-terephthalate-co-ethylene-glycol)
PVC poly(vinyl-chloride)
350 F. Tuba et al. / Engineering Fracture Mechanics 99 (2013) 349355
The major problem is related, however, to the small ligament lengths. Since the crack initiates under a complex 3D stress
state and propagates under quasi plane-stress conditions only after a given transition distance [8,10], there is a minimal lig-
ament length during the tests (L
min
). This has to be determined as correct as possible, because if L is too small then the stress
triaxiality will rise in the specimen and there is a risk of transition to mixed plane-strain/plane-stress regime where the
linear approach of EWF concept is inappropriate. On the other hand if L
min
is too large then the condence limits of w
e
will
increase. The reason is that the condence band of linear regression is hyperbolic and it is centered at the mean of variables,
i.e. at the mean of ligament lengths L [12]. The precision of the method can be improved by an increased sample size [9],
but the higher is L, the larger will be the condence range (Fig. 1; note: the sample size is the same, only the ligament range
and thus L differs). Accordingly, a large sample size is a necessary but not sufcient condition of high accuracy and good
reproducibility.
The minimal ligament length was originally proposed by Cotterell and Reddel [2] as (35)B, but Marchal et al. [13] found
slightly higher values, (68)B, during statistical analyses. Furthermore, Pardoen et al. [8] found that these conditions are not
severe enough in very thin sheets thus for general use, a combination of the two criteria was suggested in our previous work
[11].
For the L
min
determination a stress criterion based on Hills theory [7] is generally used. Hill [7] has shown for plastic-rigid
materials that in deeply notched sheets under plane-stress conditions no stress can exceed the value 1.15r
Y
, where r
Y
is the
tensile yield stress. Hence, a stress criterion for the maximum net-section stress the maximum engineering stress observed
during the fracture of the specimen; r
ns
= P
max
/BL, where P
max
is the maximum load was proposed by Cotterell et al. [10].
The lower bound of this criterion is the yield stress while the upper one can be approximated by 1.15r
Y
. Nevertheless, by
using Hills theory for L
min
determination several authors [1427] observed that the range of minimum ligament length is
around (3040)B or even higher, which could cause problems during the extrapolation of the experimental data (see Fig. 1).
An other stress criterion suggested by Clutton [28] is based on experimental results and uses the mean of the maximum
stresses r
m
r
ns
instead of theoretical considerations. The lower and upper bounds of this criterion are 0.9r
m
and 1.1r
m
,
respectively. The latter criterion is, however, determined from the test results and does not ensure full ligament yielding
[10], it only facilitates the rejection of measurement errors.
1
The aim of the present study is to examine the adequacy of the above criteria and to investigate the differences between
theoretical plastic-rigid and elasticplastic (viscoelasticviscoplastic) material behavior.
2. Theory and discussion
The EWF experiments are generally performed on double edge notched tensile (DENT) specimens (Fig. 2). This specimen
type is particularly suitable for mode I testing because the transverse stress between the notches is tensile and there are no
buckling problems [10].
By plotting the thickness specic load (P
max
/B) against the ligament length the results lie on a straight line with a positive
intercept (Fig. 3). As it was stated by Cotterell and Reddel [2] true geometric similarity occurs when the line passes through
the origin. They assumed that this positive intercept arose from crack growth before reaching the maximum load.
Ever since several studies and experimental results were released addressing the behavior of metals as well as of poly-
mers. As it is shown in Fig. 3, the load versus ligament length plots are linear implying geometric similarity and having
positive intercept. Since the data were obtained by several research groups [11,1526] from wide range of materials, at var-
ious temperatures the results suggest the potential inuence of other, so far disregarded reasons.
Fig. 1. The effect of minimal ligament length on the error of essential work of fracture parameters.
1
For this purpose there are other approximations as it will be shown later on.
F. Tuba et al. / Engineering Fracture Mechanics 99 (2013) 349355 351
The observed lines can be described with Eq. (2), while the experimentally obtained data are summarized in Table 1.
P
max
B
s mr
Y
L 2
where s (N/mm) is the ordinate intercept of the P
max
/B versus L linear ts supposed to be related to the non-plane-stress
crack initiation and initial elastic deformation of the samples and m () is an experimentally determined constraint factor
similar to the ligament yielding parameter of Zerbst et al. [1], which could be given by the following equation:
m
r
r
Y
3
where r (MPa) is the slope of the P
max
/B versus L linear ts.
Obviously, Eq. (2) is only a phenomenological expression, since at zero ligament length the load will be also zero, but for
the tested valid data range it can be used.
In the reference papers (Table 1) the plane-strain/plane-stress transition was interpreted as the intersection of lines Eq.
(2) and 1.15r
Y
L. However, the by using the 1.15r
Y
expression one assumes perfectly plastic-rigid material behavior, which is
only a rough approximation for real materials. The used minimum ligament length values are, therefore, much higher ((30
70)B) then it was originally suggested by Cotterell and Reddel [2] ((35)B), which can be clearly assigned to the inherent
inuence of the s constant and makes the stress criteria based on Hills theory inappropriate.
Fig. 2. Geometry of DENT specimens.
Fig. 3. Thickness specic load as a function of ligament length (Note: data were obtained from Refs. [11,1526]; results without temperature indicated
were measured at ambient conditions; the lines are least square regression ts).
352 F. Tuba et al. / Engineering Fracture Mechanics 99 (2013) 349355
As shown in Fig. 4 the slope values of investigated materials are closely related to their yield stress, which suggests the
yielding in the entire ligament range. Outliers are the copper [24] and one of the PETGs [15].
In case of copper [24] the signicant strain hardening of the annealed material caused the rise of the engineering stress,
which also led to higher slope values. In this material, rst an extended plastic zone had to be formed owing to the high rate
of hardening. This was followed by a balanced diminution of thickness (apparent increase of calculated engineering stress)
and rate of hardening. The localization of neck could only start after the hardening rate had became low enough [7], but in
this case the thickness decrease of the specimens could not be neglected.
In case of polymers the strain hardening is less signicant, thus the localized neck forms much easily. The difference be-
tween the slope value and yield stress of PETG [15] is treated herein as a measurement error, since similar materials had m in
a range of 0.951.
It was assumed previously that the intercept value s is related to the initial elastic deformations under 3D stress state. To
support this theory in Fig. 5 the ordinate intercept values are plotted as a function of materials elastic moduli. From this
aspect the s parameter can be treated as the stiffness of the test specimen prior yielding, which is therefore geometry
and material dependent, as well. The geometry dependence can be observed in the last three rows of Table 1. As the material
thickness increases the s becomes higher, too. Further investigations have to made, however, to describe the supposed rela-
tionship between s and the tensile modulus.
Nevertheless, it should be also noted that in [27] the specimen slenderness, and therefore the stress triaxiality related to
L/B [29], was not the same for the tested samples. It is known, however, (see e.g. [1,29]) that the stress triaxiality has an effect
on crack resistance, especially during stable crack propagation. Strictly taken, to fulll the geometric similarity prerequi-
site of EWF method the DENT specimens ought to have the same slenderness, but this would be a rather impractical solution.
Since generally specimens of constant thickness are tested, the stress triaxiality will vary with ligament range. Although at
high L/B ratios this constraint effect becomes less signicant, evaluating samples of same slenderness range seems to be a
straightforward approach in order to diminish the effect of stress triaxiality variation. Nevertheless, the variation of slender-
ness through the ligament range could also contribute to the non-zero ordinate intercept, so further experimental work and
numerical analyses are required to explore whether its inuence is signicant or not.
By coming back to experimentally observed lines, Eq. (2) can be rearranged to Eq. (4), since P
max
/B is equal to r
ns
L.
r
ns
mr
Y
1
s
mr
Y
1
L

4
where the s/mr
u
constant is related to the non-ideal material behavior crack initiation under 3D stress state, elastic defor-
mation prior initiation, non-ideal geometric similarity (i.e. differences in stress triaxiality), etc.
As shown in Eq. (4) the s/mr
Y
term has an inherent effect on the net-section stress. It raises r
ns
at small ligament lengths,
which can be misinterpreted based on Cluttons criterion [28] as a measurement error or after Cotterell et al. [10] based on
Hills theory [7] as a plane-stress/plane-strain transition.
As a consequence both stress criteria becomes overconservative at small ligament lengths, which will diminish the accu-
racy and precision of the measurements (see Fig. 1). Since Eq. (4) is non-linear the use of Cluttons criterion 0.9r
m
< -
r
ns
< 1.1r
m
for measuring-error rejection is not straightforward either. For this purpose the statistical analysis of the
linearized form (Eq. (2)) would be more preferable. Similarly as it was done by Williams and Rink [23]; any data point which
lies outside the 95% condence limit (2 standard deviations) should be excluded from the further analysis. Exactly the same
treatment can be made for the elongation at break values as it was outlined in our previous paper [11] to lter invalid
Table 1
Data from the least square regression ts of Fig. 3 as well as experimental yield stresses and minimal ligament lengths from Refs. [11,1427].
Material [Ref.] Data obtained from linear regression Data obtained from reference paper
Ordinate intercept
s (N/mm)
Slope r
(N/mm
2
)
Constraint
factor m ()
Lower ligament
limit used L
min
(mm) (xB)
Yield stress
r
Y
(MPa)
Tensile modulus
E (GPa)
Copper [24] 154.7 116.0 1.45 10 (100B) 80 ?
PC [25] 54.0 48.9 1.04 9 (72B) 47 2.3
HIPS [20] 19.7 17.1 1.07 7 (27B) 16 1.8
PEN at 23 C [16,22] 194.1 117.3 0.86 5 (40B) 137 9.0
PEN at 80 C [16,22] 118.1 85.1 1.06 5 (40B) 80 5.0
PEN at 23 C [14] 103.4 57.4 0.93 5 (10B) 62 1.9
PVC [17] 55.4 39.3 0.89 6 (40B) 44 3.4
PEEK at 23 C [21] 55.8 62.9 1.00 4 (32B) 63 3.5
PEEK at 100 C [21] 23.3 35.4 1.01 4 (32B) 35 2.5
EPBC [23] 22.4 27.7 1.03 6 (60B) 27 1.5
PCL [11] 15.3 19.1 1.12 6 (12B) 17 0.3
PBT [22,26] 50.1 47.7 1.04 8 (64B) 46 2.4
PET [22,25] 133.2 103.4 1.03 6 (48B) 100 5.4
PETG [15]; B = 0.2 mm 85.4 48.6 0.77 6 (28B) 63 3.1
PETG [27]; B = 0.5 mm 26.6 41.6 0.95 5 (25B) 44 2.0
PETG [27]; B = 3.1 mm 39.8 45.6 0.95 5 (1.7B) 48 1.9
PETG [27]; B = 6.1 mm 47.1 47.2 1.00 5 (0.8B) 47 1.7
F. Tuba et al. / Engineering Fracture Mechanics 99 (2013) 349355 353
data resulting from instability problems during crack propagation. As a surplus, the latter treatment facilitates the approx-
imation of the minimum ligament length where the steady-state fracture process zone is already present.
3. Conclusions
In this paper the effect of elastic deformations and the ligament length dependence of maximal net-section stress were
analyzed on DENT specimens for a wide range of materials. Since stress criteria are conventionally used to determine the
lower ligament bound of EWF method, the identication of inuencing factors was necessary.
In the original EWF concept true geometric similarity and plastic-rigid behavior were assumed the maximum load re-
sults should, therefore, lie on a straight line and pass through the origin. However, experimental results have shown that
there is a positive intercept after the linear regression of the maximum load values.
In this work it was shown that after this initial value, which is assumed to be related to the initiation under complex 3D
stress-state, elastic deformation and stress triaxiality variations, the load increment as a function of ligament length remains
constant. The slope of thickness specic load versus ligament length plot is closely related to the yield stress (r
Y
) and its
value exceeds the 1.15r
Y
proposed by Hill only when there is signicant strain hardening in the material. This observa-
tion seemed to be geometry and temperature independent, as well. The ordinate intercept of maximum load versus ligament
length plot seems to be related to elastic modulus of the material, but it is geometry dependent and can be treated in rst
approximation as the stiffness of the specimen.
The increase of net-section stress values is, therefore, inherently related to the positive intercept s/mr
u
term of the Eq.
(4), thus the experimentally observed stress increase at small ligament lengths does not necessarily represent the transition
to mixed plane-stress/plane-strain regime. The stress criterion of Clutton [28] is, therefore, adequate for rejecting errors in
measurement, but if used so, the intrinsic effect of ordinate intercept could result in the misinterpretation of the experimen-
tal data. The criterion presented by Cotterell et al. [10] ensures full ligament yielding but is overconservative, because it
assumes true geometric similarity and plastic-rigid material behavior without strain hardening. In real materials, however,
neither the differences in stress triaxiality (non-ideal geometric similarity) nor the elastic deformations prior to yielding nor
can be neglected as it was shown herein.
Fig. 4. Yield stress versus the slope values of linear regression plots.
Fig. 5. Ordinate intercept values as a function of tensile modulus.
354 F. Tuba et al. / Engineering Fracture Mechanics 99 (2013) 349355
To summarize, based on this study the test protocol of EWF method proposed by Clutton [28] should be modied and
expanded as follows:
For the lower ligament length determination the method outlined in our previous paper [11] should be used.
For the upper ligament length determination the approximation based on linear elastic fracture mechanics, proposed by
Cotterell et al. [10] is appropriate. Additionally, the width restriction for ligament length (L < w/3; see e.g. [5]) seems to be
too conservative, it lowers the accuracy of measurements, thus it can be neglected.
The slenderness (L/B) should be in the same range for all specimens and greater than 4 in order to diminish the variations
of stress triaxiality.
The maximum load (P
max
/B), the ultimate elongation (e
u
) and the specic work of fracture (w
f
) should be plotted against
the ligament length (L). The least square regression lines should be calculated and data lying outside the 95% condence
limit (2 standard deviations) have to be rejected. After this the regression should be repeated and the ordinate intercept,
slope and 95% condence limits of parameters should be given.
Acknowledgements
This work is connected to the scientic program of the Development of quality-oriented and harmonized R+D+I strategy
and functional model at BME project. This project is supported by the New Szchenyi Plan (Project ID: TMOP-4.2.1/B-09/1/
KMR-2010-0002). Ferenc Tuba is also indebted to the project Talent care and cultivation in the scientic workshops of BME
(Project ID: TMOP-4.2.2.B-10/12010-0009) for the nancial support and to Marta Rink (Politecnico di Milano, Italy) and
Jzsef Karger-Kocsis (Budapest University of Technology and Economics, Hungary) for the helpful discussions.
References
[1] Zerbst U, Heinimann M, Donne CD, Steglich D. Fracture and damage mechanics modelling of thin-walled structures : an overview. Engng Fract Mech
2009;76:543.
[2] Cotterell B, Reddel JK. The essential work of plane stress ductile fracture. Int J Fract 1977;13:26777.
[3] Broberg KB. Critical review of some theories in fracture mechanics. Int J Fract 1968;4:119.
[4] Mai Y-W, Cotterell B. On the essential work of ductile fracture in polymers. Int J Fract 1986;32:10525.
[5] Brny T, Czigny T, Karger-Kocsis J. Application of the essential work of fracture (EWF) concept for polymers, related blends and composites: a review.
Prog Polym Sci 2010;35:125787.
[6] Martinez AB, Gamez-Perez J, Sanchez-Soto M, Velasco JI, Santana OO, Ll Maspoch M. The essential work of fracture (EWF) method analyzing the post-
yielding fracture mechanics of polymers. Engng Fail Anal 2009;16:260417.
[7] Hill R. On discontinuous plastic states, with special reference to localized necking in thin sheets. J Mech Phys Solids 1952;1:1930.
[8] Pardoen T, Marchal Y, Delannay F. Thickness dependence of cracking resistance in thin aluminium plates. J Mech Phys Solids 1999;47:2093123.
[9] Pegoretti A, Castellani L, Franchini L, Mariani P, Penati A. On the essential work of fracture of linear low-density-polyethylene. I. Precision of the testing
method. Engng Fract Mech 2009;76:278898.
[10] Cotterell B, Pardoen T, Atkins AG. Measuring toughness and the cohesive stress-displacement relationship by the essential work of fracture concept.
Engng Fract Mech 2005;72:82748.
[11] Tuba F, Olh L, Nagy P. The role of ultimate elongation in the determination of valid ligament range of essential work of fracture tests. J Mater Sci
2012;47:222833.
[12] Himmelblau DM. Process analysis by statistical methods. New York: John Wiley & Sons Inc; 1970.
[13] Marchal Y, Walhin J-F, Delannay F. Statistical procedure for improving the precision of the measurement of the essential work of fracture of thin sheets.
Int J Fract 1997;87:18999.
[14] Karger-Kocsis J, Moskala EJ. Molecular dependence of the essential and non-essential work of fracture of amorphous lms of poly(ethylene-2,6-
naphthalate) (PEN). Polymer 2000;41:630110.
[15] Arkhireyeva A, Hashemi S. Effect of temperature on fracture properties of an amorphous poly(ethylene terephthalate) (PET) lm. J Mater Sci
2002;37:367583.
[16] Arkhireyeva A, Hashemi S. Fracture behaviour of polyetylene naphthalate (PEN). Polymer 2002;43:289300.
[17] Arkhireyeva A, Hashemi S. Combined effect of temperature and thickness on work of fracture parameters of unplasticized PVC lm. Polym Engng Sci
2002;42:50418.
[18] Gamez-Perez J, Velazquez-Infante JC, Franco-Urquiza E, Pages P, Carrasco F, Santana OO, et al. Fracture behavior of quenched poly(lactic acid). Express
Polym Lett 2011;5:8291.
[19] Hashemi S. Determination of the fracture toughness of polybutylene terephthalate (PBT) lm by the essential work method: effect of specimen size
and geometry. Polym Engng Sci 2000;40:798808.
[20] Hashemi S. Work of fracture of high impact polystyrene (HIPS) lm under plane stress conditions. J Mater Sci 2003;38:305562.
[21] Hashemi S. Effect of temperature on fracture toughness of an amorphous poly(ether-ether ketone) lm using essential work of fracture analysis. Polym
Test 2003;22:58999.
[22] Hashemi S, Arkhireyeva A. Inuence of temperature on work of fracture parameters in semi-crystalline polyester lms. J Macromol Sci Part B Phys
2002;B41:86380.
[23] Williams JG, Rink M. The standardisation of the EWF test. Engng Fract Mech 2007;74:100917.
[24] Levita G, Parisi L, McLoughlin S. Essential work of fracture in polymer lms. J Mater Sci 1996;31:154553.
[25] Hashemi S. Fracture toughness evaluation of ductile polymeric lms. J Mater Sci 1997;32:156373.
[26] Hashemi S. Temperature dependence of work of fracture parameters in polybutylene terephthalate (PBT). Polym Engng Sci 2000;40:143546.
[27] Karger-Kocsis J, Czigny T, Moskala EJ. Thickness dependence of work of fracture parameters of an amorphous copolyester. Polymer 1997;38:458793.
[28] Clutton E. Essential work of fracture. In: Moore DR, Pavan A, Williams JG, editors. Fracture mechanics testing methods for polymers, adhesives and
composites. Oxford: Elsevier; 2001. p. 17795.
[29] Schwalbe K-H, Newman Jr JC, Shannon Jr JL. Fracture mechanics testing on specimens with low constraint standardisation activities within ISO and
ASTM. Engng Fract Mech 2005;72:55776.
F. Tuba et al. / Engineering Fracture Mechanics 99 (2013) 349355 355

S-ar putea să vă placă și