Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Effects of Crosshead Speed on

the Quasi-Static Stress–Strain


Yi Zhang1
Department of Mechanical Engineering,
University of Alberta,
Relationship of Polyethylene
10-203 Donadeo Innovation Centre
for Engineering,
Pipes
9211-116 Street NW,
Edmonton, AB T6G 1H9, Canada Quasi-static stress–strain relationship of polyethylene (PE) pressure pipe that plays an
e-mail: yz4@ualberta.ca important role on its long-term performance has been established by removing the vis-
cous stress component from the experimentally measured total stress. Work reported here
P.-Y. Ben Jar is focused on the influence of crosshead speed on the notched pipe ring (NPR) specimens
Department of Mechanical Engineering, that are prepared from PE pressure pipe of 2 in. in diameter. Viscous component of the
University of Alberta, stress–strain relationship was determined using a spring–damper–plastic element model,
10-203 Donadeo Innovation Centre calibrated using results from stress relaxation tests. Crosshead speeds considered for the
for Engineering, initial stretch of the stress relaxation tests are 0.01, 1, and 10 mm/min which due to the
9211-116 Street NW, relatively uniform deformation in the gauge section generate the same order of difference
Edmonton, AB T6G 1H9, Canada in the strain rates. Results from the study suggest that the quasi-static stress–strain rela-
e-mail: ben.jar@ualberta.ca tionship is affected by the crosshead speed used to generate the deformation, and the
trend of change is opposite to the total stress counterpart that includes the viscous
component. [DOI: 10.1115/1.4033777]

Keywords: PE pipe, long-term mechanical properties, relaxation, strain rate

1 Introduction 2 Experimental Details


PE pipe has been widely used to transport natural gas because 2.1 Materials and Specimens. All specimens used for the
of its good mechanical and physical properties and high resistance testing were prepared from commercial PE4710, cell classification
to environmental degradation. Statistics shows that over 90% of 445576C HDPE pipe with inner diameter and nominal wall thick-
the newly installed low-pressure gas pipeline systems are made of ness of 52.5 mm and 5.84 mm, respectively, manufactured by
PE [1]. However, unexpected, catastrophic failures of PE pipeline Endot Industries, Rockaway, NJ. Resin for the HDPE pipe is PE-
were still reported in the last decade [2–4], suggesting limitations 100, which has the minimum required strength of 10 MPa. NPR
in our current understanding of high density polyethylene (HDPE) specimens were machined from the HDPE pipe, of which the
mechanical properties. dimensions follow those recommended in ASTM D2290-12,
As a semicrystalline polymer with glass transition occurring except that the notch profile is flat, instead of round, in order to
well below the room temperature [5], PE is known for its nonlin- have a relatively uniform stress distribution in the ligament
ear, rate-dependent stress response to deformation. This phenom- region. Note that ligament width for the NPR specimens was cho-
enon has been studied extensively [6–16], all suggesting that sen to be 5.84 mm in order to have the aspect ratio of width to
stress required to generate a given level of deformation increases thickness to be close to 1. In this way, contraction should be simi-
with the increase of strain rate. Some of those studies also suggest lar between the width and the thickness directions [20]. A pipe
that yield stress [17] and strain hardening modulus [6] increase section and one of the NPR specimens used in the study are shown
linearly with the increase of the logarithmic value of the strain in Fig. 1(a) and dimensions for the NPR specimens in Fig. 1(b).
rate.
The above rate-dependent stress response to deformation is
generally believed to be due to the viscous component of the me- 2.2 Mechanical Testing. The D-split tensile test, first pro-
chanical properties. Since PE pipes are designed for a service life posed to characterize mechanical properties of composite materi-
of more than 50 yrs, a conservative evaluation of their long-term als [21–23], was adopted for the experimental testing, for which
performance should be based on the quasi-static stress–strain rela- the setup is depicted in Fig. 2(a). All tests were conducted using a
tionship, that is, without the viscous component [18]. In this study, universal test machine (QUASAR 100) at room temperature.
the quasi-static stress–strain relationship is established in the hoop Two loading modes were used. One is monotonic tension and
direction of PE pipe using NPR specimens. In view that necking the other stress relaxation. The former was applied to the NPR
does not occur in this type of PE specimens for a wide range of specimens until one of the ligaments breaks, at a constant cross-
crosshead speeds [19], strain rate used to generate the deformation head speed of 0.01, 1, or 10 mm/min which correspond to the ini-
can be varied by changing the crosshead speed. tial strain rate of 7  105, 7  103, or 7  102 s1,
Work reported here is to determine whether the change of respectively. Although these strain rates only represent the initial,
crosshead speed affects the quasi-static stress–strain relationship lower bound of the range of strain rates that can be introduced at
for PE pipe, and whether the trend of change is similar to that for each crosshead speed, similar to that found previously [1], the
the total stress–strain relationship, i.e., including the viscous range of strain rates generated by each crosshead speed is distinc-
component. tively different from the others.
The loading scheme for the stress relaxation test is depicted in
1
Corresponding author. Fig. 2(b) in which a predefined strain (to be named relaxation
Contributed by the Pressure Vessel and Piping Division of ASME for publication
in the JOURNAL OF PRESSURE VESSEL TECHNOLOGY. Manuscript received March 9, 2016;
strain here) is introduced to the ligament section at one of the
final manuscript received May 29, 2016; published online September 27, 2016. above three crosshead speeds. During the stress relaxation pro-
Assoc. Editor: Kunio Hasegawa. cess, stroke was kept constant for a period of 10,000 s and load

Journal of Pressure Vessel Technology Copyright V


C 2017 by ASME APRIL 2017, Vol. 139 / 021402-1

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jpvtas/935788/ on 02/14/2017 Terms of Use: http://www.asme.org


Fig. 1 Information on specimens used in the study: (a) a pipe section (left) and a NPR
specimen (right), and (b) specimen dimensions

Fig. 2 Information for the stress relaxation tests: (a) test setup (also used for the mono-
tonic tensile tests) and (b) schematic diagram for the stroke and area strain as functions
of time

was recorded as a function of time. Unloading was always intro- proposed by Strobl’s group [24–26]. In this approach, the viscous
duced at the crosshead speed of 0.1 mm/min. stress component at a given strain level is determined by regener-
It should be noted that due to the short gauge length, relaxation ating results from a stress relaxation test using an analytical
strain e introduced to the NPR specimens is represented by area model. The model used in the current study is shown in Fig. 3
strain which is calculated based on the logarithmic ratio of the which consists of two branches. The top, viscous branch is respon-
change of ligament width, as given below, under the assumption sible for the time-dependent stress component rr, which contains
that aspect ratio of the ligament cross section remains close to 1 a spring and a damper connected in series to allow stress to
during the test [20] decrease to zero eventually during the stress relaxation process.
The bottom, elastoplastic branch is responsible for the quasi-static
e ¼ 2  In ðw0 =wÞ (1) stress rep , which contains a spring and a finite plasticity element
connected in series. This model is a simplified version of the
where w0 and w are the initial and deformed width in the ligament model proposed by Hong et al. [25] by combining the crystal and
section. network branches into the elastoplastic branch.
As to be shown later, results from the monotonic tensile tests Based on the model in Fig. 3, value for rep at a given strain e is
suggest that strain for the ligament fracture is around 1.8 at the equivalent to the difference between the total stress and rr at the
crosshead speed of 0.01 mm/min, 1.7 at 1 mm/min, and 1.1 at
10 mm/min (with softening detectable at a strain level around
0.8). Therefore, the relaxation strains (e) for the stress relaxation
tests were calculated using Eq. (1) and were chosen to be in the
range from 0.05 to 1.6 for the crosshead speeds of 0.01 and 1 mm/
min, and from 0.05 to 0.7 for 10 mm/min.

3 Model for Determining the Quasi-Static Stress


The quasi-static stress–strain relationship for the PE pipe is
determined by removing the viscous stress component from the Fig. 3 Schematic diagram of the viscous model, modified from
experimentally measured total stress using an approach originally Ref. [25]

021402-2 / Vol. 139, APRIL 2017 Transactions of the ASME

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jpvtas/935788/ on 02/14/2017 Terms of Use: http://www.asme.org


beginning of the stress relaxation process with the same e. Value Note that the above trend of change is not applicable to the two
for rr at the beginning of the stress relaxation process is denoted curves from the crosshead speeds of 1 and 10 mm/min at the strain
rr ð0Þ and is determined by calibrating the model using the stress level above 0.8, as at such a high strain level, stress for the
relaxation curve obtained experimentally. By determining rep val- curve from 10 mm/min is actually slightly lower than that from
ues at several e levels, the quasi-static stress–strain relationship is 1 mm/min. This, we believe, is because at 10 mm/min, specimen
determined. Details of this procedure are given in Ref. [25] and with deformation above this strain level is about to fracture, and
are described briefly below. thus should have had significant damage to reduce the resistance
Following the suggestion by Hong et al. [25], the Eyring’s to deformation.
viscosity law is assumed to govern the relaxation process in PE, Figure 4 also shows the representative phenomenon of strain at
that is, fracture decreasing with the increase of crosshead speed. This
  phenomenon has been reported in the literature [32–34] and has
d rr 1 rr been attributed to the decrease of fracture energy with the increase
¼  sinh (2)
dt r0 sr r0 of crosshead speed [33,34]. Similarly, work in the literature also
shows that elongation at break decreases with the increase of
with crosshead speed [19,27]. Furthermore, from the microscopic
viewpoint, deformation and fracture process of HDPE has been
e_ 0 Er Er reported to be significantly affected by the change of strain rate.
s1
r ¼ ¼ (3) That is, at a low strain rate, plastic deformation is promoted and
r0 g0
ductile drawing occurs because the rate of chain slip is greater
where r0 ; sr , and Er are reference stress, characteristic relaxation than or comparable to that for the overall tensile deformation,
time, and relaxation modulus, respectively. Solution for Eq. (2) is which leads to the ductile fracture, resulting in a large fracture
strain [35]. By increasing the strain rate, the rate for chain slip can
     no longer match with the overall deformation rate, thus generating
rr ðtÞ rr ð0Þ t
¼ 2 atanh tanh exp  (4) the relatively brittle fracture.
r0 2r0 sr Figure 5 summarizes the experimentally determined stress
decay (Dr )as a function of time for all relaxation strains consid-
Based on the above expression, the stress decay (Dr) defined as ered in the study. Each plot in Fig. 5 contains all curves obtained
at the given crosshead speed, in which the curves with the higher
Dr ¼ rr ð0Þ  rr ðtÞ (5) Dr are from a specimen at a higher relaxation strain level. This
trend of change is consistent with that reported previously [25].
can be expressed as As mentioned earlier, the curves in Fig. 5 were regenerated
    using Eq. (6), with sr being 1.6  104 s and values for rr ð0Þ and r0
rr ð0Þ t adjusted to fit the experimental curves. Figure 6 compares the
Dr ¼ rr ð0Þ  2r0 atanh tanh exp  (6)
2r0 sr curves generated from Eq. (6) (open symbols) with those obtained
experimentally at the relaxation strain of 0.2 (lines), depicting the
The stress decay for the first 10,000 s, determined from the stress excellent agreement. All values for rr ð0Þ and r0 used to regener-
relaxation tests, is used to calibrate the model in Fig. 3. Following ate the curves in Fig. 5 are summarized in Table 1.
the suggestion by Hong et al. [25], sr in Eq. (6) is set to have a The quasi-static stress–strain curves, obtained using the above
constant value of 1.6  104 s, independent of the relaxation strain process to remove the viscous stress component, are presented in
used for the testing. Values for the remaining constants in Eq. (6), Fig. 7. The figure shows that the quasi-static stress component at a
rr ð0Þ and r0 , are then determined by fitting the curve generated given strain level decreases with the increase of the crosshead
by Eq. (6) with that obtained experimentally. The above process is speed, of which the trend is opposite to that shown by the total
applied to several relaxation strain levels to construct the quasi- stress response in Fig. 4 which includes the viscous stress compo-
static stress–strain curve. nent. This opposite trend of change with the increase of crosshead
speed confirms the results reported previously [18] and serves as
another evidence to support the influence of strain rate on the
4 Results and Discussion
True stress–strain curves from three crosshead speeds, obtained
directly from the monotonic tensile tests of NPR specimens, are
summarized in Fig. 4. It should be pointed out that in view of high
reproducibility of results from this type of tests [27], only one
specimen was used for the monotonic tensile test at each cross-
head speed to depict the influence of crosshead speed on the over-
all stress–strain relationship under the tensile loading. Figure 4
shows the general trend of the test results, with increase of stress
with the increase of crosshead speed. This is consistent with that
reported previously [6–16], that is, higher the crosshead speed,
larger the true stress generated at a given strain level.
The trend of change in Fig. 4 is known to be caused by the
change in mechanisms involved in the deformation processes. At
a low crosshead speed, the amorphous phase is relatively soft, and
thus, the overall stress response is dominated by the stress
response to deformation in the amorphous phase. It has also been
suggested that at a sufficiently low strain rate, melting and recrys-
tallization may occur during the plastic deformation, which fur-
ther softens PE and reduces its resistance to deformation [28,29].
With the increase of the crosshead speed, mobility of the amor-
phous phase is reduced to increase its resistance to deformation Fig. 4 Plots of experimentally determined true stress-area
[30,31], thus increasing the involvement of the crystalline phase strain curves under monotonic tension at three crosshead
in the deformation process. speeds

Journal of Pressure Vessel Technology APRIL 2017, Vol. 139 / 021402-3

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jpvtas/935788/ on 02/14/2017 Terms of Use: http://www.asme.org


damage generation in PE, i.e., higher the strain rate, larger the thus delaying the cavitation process [36]. On the other hand, when
damage generated at a given strain level [27]. the mobility of the amorphous phase is insufficient to generate
Damage in semicrystalline polymers, when subjected to tensile enough flow to accommodate the macroscopic deformation,
loading, has been suggested to be initiated from cavitation in the cavitation occurs [38].
amorphous phase, also known as void formation [36]. Cavitation It has been observed that the critical strain for the onset of the
is known to be affected by both temperature and strain rate volume strain increase depends on the strain rate, that is, at a
[30,36,37]. As strain rate decreases or temperature increases, mo- higher strain rate, the growth rate for the volume strain starts
bility of the amorphous phase increases, thus decreasing the increasing at a smaller deformation level [30]. Since volume strain
potential for cavitation. It has also been suggested that at a suffi- is another indicator for the damage generation in semicrystalline
ciently low strain rate, the crystalline phase is able to deform plas- polymers, the critical strain for the increase of the growth rate for
tically, which further reduces the stress on the amorphous phase, the volume strain may also serve as an indicator for the damage
generation. However, due to the short gauge length of the NPR
specimens, volume strain could not be measured accurately, thus
not included in the test program.
Figure 8 presents the variation of reference stress r0 as a func-
tion of relaxation strain. It has been suggested [25,39] that de-
pendence of r0 on the relaxation strain involves several stages
that reflect the difference in mechanisms involved in the deforma-
tion process. Figure 8 shows the four stages for the variation of r0
with the relaxation strain, separated by three critical strains,
marked as A, B, and C in Fig. 8. Values for these critical strains
do not seem to depend on the crosshead speed, which is consistent
with that reported in the literature [25,37,39,40]. The first critical
strain, around 0.04, is an indication of the beginning of individual
inter- and intralamellar slips; the second critical strain, around
0.11, is for the onset of collective lamellar slips; and the third crit-
ical strain, around 1.2, is for the start of fibril formation. A fourth
critical strain has also been suggested in the literature [25,37,39],
to represent the involvement of chain disentanglement in the de-
formation process. However, the fourth critical strain is not shown
in Fig. 8, possibly due to the high strain level required for its
occurrence. These critical strain values may vary with the type of
PE and its crystalline structure used for the study [10,25,41]. In
Fig. 8, the first two critical strains have values similar to those
reported in Ref. [25]. The third critical strain, from the crosshead
speeds of 0.01 and 1 mm/min, is 1.2 which is similar to that for
HDPE reported in Ref. [41], but different from the value of 0.6
reported in Refs. [25] and [39].
It has been well established that three distinctive regions can be
recognized in the bilogarithmic plot of hoop stress versus time-to-
failure from creep tests of PE pipe: (i) ductile failure with visible
plastic deformation occurring at a relatively high stress level, (ii)
brittle failure at a relatively a low stress level which is often
referred to as slow crack growth, and (iii) failure controlled by

Fig. 6 Comparison between curves generated from Eq. (6)


Fig. 5 Stress decay (Dr) as a function of relaxation time at dif- (presented by markers) and curves obtained from stress relaxa-
ferent relaxation strains which were introduced at the cross- tion tests (lines) at the relaxation strain of 20%, with the initial
head speeds of (a) 0.01 mm/min, (b) 1 mm/min, and (c) 10 mm/ stretch introduced at the crosshead speeds of 0.01, 1, and
min (larger the relaxation strain, higher the stress decay) 10 mm/min

021402-4 / Vol. 139, APRIL 2017 Transactions of the ASME

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jpvtas/935788/ on 02/14/2017 Terms of Use: http://www.asme.org


Table 1 Summary of values for rr ð0Þ and r0 in Eq. (6), with sr 51:63104

0.01 mm/min 1 mm/min 10 mm/min

Relaxation strain, e rr ð0Þ ðMPaÞ r0 ðMPaÞ rr ð0Þ ðMPaÞ r0 ðMPaÞ rr ð0Þ ðMPaÞ r0 ðMPaÞ

0.05 4.1 0.82 12.4 1.03 17.4 1.06


0.1 5.12 1 14.7 1.32 18.8 1.10
0.2 5.6 1.05 16.1 1.35 20.7 1.22
0.3 5.97 1.08 16.7 1.43 21.2 1.24
0.4 6.01 1.09 16.9 1.53 21.5 1.27
0.45 6.18 1.11 17.7 1.55 21.8 1.36
0.7 — — — — 24.4 1.56
0.8 6.55 1.16 18.8 1.60 — —
1.2 8 1.25 21.5 1.80 — —
1.6 14.2 2.5 35.3 2.80 — —

in Fig. 8. We believe that the quasi-static stress at the second criti-


cal strain, which is around yield point, may be relevant to the
occurrence of the ductile-to-brittle transition. Therefore, informa-
tion obtained from the approach described in the paper may fur-
ther our knowledge in determining the knee position using short-
term test. This part of study is currently being investigated when
this paper is prepared.
In view that reference stress r0 shows small variation with the
change of crosshead speed used for the testing, especially com-
pared to the variation of rr ð0Þ shown in Table 1, further analysis
was conducted to establish the quasi-static stress–strain curve by
assuming that r0 is independent of the crosshead speed used for
the testing. Figure 9 presents the quasi-static stress–strain curves
generated based on the assumption that values for r0 follow the
variation for the crosshead speed of 0.01 mm/min in Table 1 and
compares the curves with those presented in Fig. 7. Figure 9
shows that only small difference exists between the original
curves in Fig. 7 and those based on r0 being independent of the
crosshead speed. As a result, it is believed that the dominant pa-
rameter in Eq. (6) to cause dependence of the stress decay on the
Fig. 7 Quasi-static stress–strain curves for the three cross- crosshead speed is rr ð0Þ, that is, the viscous stress component at
head speeds used in the study the beginning of the stress relaxation process. The above phenom-
enon will be investigated in detail in the near future, to explore its
validity for a broad range of crosshead speed.
degradation [42–45]. Transition from ductile to brittle failure in
the plot is known as “knee position.” It is not easy to determine
the stress and time for the knee position at room temperature. 5 Conclusions
Therefore, extrapolation is required from data collected at high
temperatures. Work described in this paper has identified critical An experimental study has been conducted to assess the effects
strains for the onset of various deformation mechanisms, as shown of crosshead speed on quasi-static stress–strain relationship for PE

Fig. 9 Variation of the quasi-static stress–strain curve by


assuming r0 being independent of crosshead speed. Curves in
Fig. 8 Variation of reference stresses r0 with the applied relax- dashed line are generated using Eq. (6) with r0 being equivalent
ation strain to those for the crosshead speed of 0.01 mm/min in Table 1.

Journal of Pressure Vessel Technology APRIL 2017, Vol. 139 / 021402-5

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jpvtas/935788/ on 02/14/2017 Terms of Use: http://www.asme.org


pressure pipes. The quasi-static stress–strain relationship is deter- [17] Dasari, A., and Misra, R. D. K., 2003, “On the Strain Rate Sensitivity of High
mined by removing the viscous stress component based on a Density Polyethylene and Polypropylenes,” Mater. Sci. Eng.: A, 358(1–2),
pp. 356–371.
model that consists of spring, damper, and finite plasticity ele- [18] Jar, P. Y. B., 2014, “Transition of Neck Appearance in Polyethylene and Effect
ments, after the model has been calibrated using results from of the Associated Strain Rate on the Damage Generation,” Polym. Eng. Sci.,
stress relaxation tests. Opposite to the trend observed before that 54(8), pp. 1871–1878.
total stress at a given strain level increases with the increase of [19] Zhang, Y., and Jar, P. Y. B., 2015, “Phenomenological Modelling of Tensile
Fracture in PE Pipe by Considering Damage Evolution,” Mater. Des., 77, pp.
crosshead speed, as shown in Fig. 4, the results in Fig. 7 suggest 72–82.
that the corresponding quasi-static stress component decreases [20] Muhammad, S., and Jar, P. Y. B., 2011, “Effect of Aspect Ratio on Large De-
with the increase of crosshead speed, indicating the influence of formation and Necking of Polyethylene,” J. Mater. Sci., 46(4), pp. 1110–1123.
strain rate on the long-term performance of the PE pipe. [21] Rafiee, R., 2013, “Apparent Hoop Tensile Strength Prediction of Glass Fiber-
Reinforced Polyester Pipes,” J. Compos. Mater., 47, pp. 1377–1386.
Based on the change of reference stress r0 with strain, results [22] Shlitsa, R. P., and Novikova, E. A., 1983, “Characteristics of the Use of the
from the study also suggest that the critical strain for the transition Split-Disk Method for Investigating Modern Winding Composites,” Mech.
of mechanisms involved in the deformation process is independ- Compos. Mater., 18(4), pp. 502–508.
ent of the crosshead speed used in the study. Rather, the change of [23] Chen, J. F., Li, S. Q., Bisby, L. A., and Ai, J., 2011, “FRP Rupture Strains in
the Split-Disk Test,” Composites, Part B, 42(4), pp. 962–972.
crosshead speed, thus change of the strain rate, is mainly to affect [24] Hong, K., and Strobl, G., 2008, “Characterizing and Modeling the Tensile De-
the critical strain for the onset of fracture. However, further study formation of Polyethylene: The Temperature and Crystallinity Dependences,”
is needed to collect additional evidence for this phenomenon, Polym. Sci. Ser. A, 50(5), pp. 483–493.
which will be conducted in the near future. [25] Hong, K., Rastogi, A., and Strobl, G., 2004, “A Model Treating Tensile Defor-
mation of Semicrystalline Polymers: Quasi-Static Stress–Strain Relationship
and Viscous Stress Determined for a Sample of Polyethylene,” Macromole-
cules, 37(26), pp. 10165–10173.
[26] Na, B., Zhang, Q., Fu, Q., Men, Y., Hong, K., and Strobl, G., 2006, “Viscous-
Acknowledgment Force-Dominated Tensile Deformation Behavior of Oriented Polyethylene,”
Macromolecules, 39(7), pp. 2584–2591.
The work was supported by Natural Sciences and Engineering [27] Zhang, Y., and Jar, P. Y. B., 2015, “Quantitative Assessment of Deformation-
Research Council of Canada (NSERC) and China Scholarship Induced Damage in Polyethylene Pressure Pipe,” Polym. Test., 47, pp. 42–50.
[28] Yeh, I.-C., Andzelm, J. W., and Rutledge, G. C., 2015, “Mechanical and Struc-
Council (CSC). Sincere appreciation is due to technical staff in tural Characterization of Semicrystalline Polyethylene Under Tensile Deforma-
the Department of Mechanical Engineering at the University of tion by Molecular Dynamics Simulations,” Macromolecules, 48(12),
Alberta, particularly, Campbell, Waege, and Bubenko for speci- pp. 4228–4239.
men preparation and Faulkner for fabrication of the extensometer [29] Lee, S., and Rutledge, G. C., 2011, “Plastic Deformation of Semicrystalline
Polyethylene by Molecular Simulation,” Macromolecules, 44(8),
used in the testing. pp. 3096–3108.
[30] Addiego, F., Dahoun, A., G’Sell, C., and Hiver, J.-M., 2006, “Characterization
of Volume Strain at Large Deformation Under Uniaxial Tension in High-
References Density Polyethylene,” Polymer, 47(12), pp. 4387–4399.
[1] Kiass, N., Khelif, R., Boulanouar, L., and Chaoui, K., 2005, “Experimental [31] Blaise, A., Baravian, C., Andre, S. p., Dillet, J. r. m., Michot, L. J., and Mokso,
Approach to Mechanical Property Variability Through a High-Density Polyeth- R., 2010, “Investigation of the Mesostructure of a Mechanically Deformed
ylene Gas Pipe Wall,” J. Appl. Polym. Sci., 97(1), pp. 272–281. HDPE by Synchrotron Microtomography,” Macromolecules, 43(19),
[2] Azevedo, C. R., 2007, “Failure Analysis of a Crude Oil Pipeline,” Eng. Failure pp. 8143–8152.
Anal., 14(6), pp. 978–994. [32] Xiao, X., 2008, “On the Measurement of True Fracture Strain of Thermoplas-
[3] Shalaby, H. M., Riad, W. T., Alhazza, A. A., and Behbehani, M. H., 2006, tics Materials,” Polym. Test., 27(3), pp. 284–295.
“Failure Analysis of Fuel Supply Pipeline,” Eng. Failure Anal., 13(5), [33] El-Bagory, T. M. A. A., Sallam, H. E. M., and Younan, M. Y. A., 2014, “Effect
pp. 789–796. of Strain Rate, Thickness, Welding on the J–R Curve for Polyethylene Pipe
[4] Majid, Z. A., Mohsin, R., Yaacob, Z., and Hassan, Z., 2010, “Failure Analysis Materials,” Theor. Appl. Fract. Mech., 74, pp. 164–180.
of Natural Gas Pipes,” Eng. Failure Anal., 17(4), pp. 818–837. [34] El-Bagory, T. M. A. A., Alkanhal, T. A. R., and Younan, M. Y. A., 2015,
[5] Peacock, A., 2000, Handbook of Polyethylene: Structures: Properties, and “Effect of Specimen Geometry on the Predicted Mechanical Behavior of Poly-
Applications, CRC Press, Boca Raton, FL. ethylene Pipe Material,” ASME J. Pressure Vessel Technol., 137(6), p. 061202.
[6] Hillmansen, S., Hobeika, S., Haward, R., and Leevers, P., 2000, “The Effect of [35] Dasari, A., and Misra, R. D. K., 2004, “Microscopic Aspects of Surface Defor-
Strain Rate, Temperature, and Molecular Mass on the Tensile Deformation of mation and Fracture of High Density Polyethylene,” Mater. Sci. Eng.: A,
Polyethylene,” Polym. Eng. Sci., 40(2), pp. 481–489. 367(1–2), pp. 248–260.
[7] Dusunceli, N., and Colak, O. U., 2006, “High Density Polyethylene (HDPE): [36] Pawlak, A., 2007, “Cavitation During Tensile Deformation of High-Density
Experiments and Modeling,” Mech. Time-Depend. Mater., 10(4), pp. 331–345. Polyethylene,” Polymer, 48(5), pp. 1397–1409.
[8] Colak, O. U., and Dusunceli, N., 2006, “Modeling Viscoelastic and Viscoplastic [37] Hossain, D., Tschopp, M. A., Ward, D. K., Bouvard, J. L., Wang, P., and Hor-
Behavior of High Density Polyethylene (HDPE),” ASME J. Eng. Mater. Tech- stemeyer, M. F., 2010, “Molecular Dynamics Simulations of Deformation
nol., 128(4), pp. 572–578. Mechanisms of Amorphous Polyethylene,” Polymer, 51(25), pp. 6071–6083.
[9] Ayoub, G., Za€ıri, F., Na€ıt-Abdelaziz, M., and Gloaguen, J. M., 2010, [38] Castagnet, S., Gacougnolle, J.-L., and Dang, P., 2000, “Correlation Between
“Modelling Large Deformation Behaviour Under Loading–Unloading of Semi- Macroscopical Viscoelastic Behaviour and Micromechanisms in Strained a Pol-
crystalline Polymers: Application to a High Density Polyethylene,” Int. J. Plast., yvinylidene Fluoride (PVDF),” Mater. Sci. Eng.: A, 276(1), pp. 152–159.
26(3), pp. 329–347. [39] Hobeika, S., Men, Y., and Strobl, G., 2000, “Temperature and Strain Rate Inde-
[10] Hiss, R., Hobeika, S., Lynn, C., and Strobl, G., 1999, “Network Stretching, Slip pendence of Critical Strains in Polyethylene and Poly(Ethylene-Co-Vinyl
Processes, and Fragmentation of Crystallites During Uniaxial Drawing of Poly- Acetate),” Macromolecules, 33(5), pp. 1827–1833.
ethylene and Related Copolymers. A Comparative Study,” Macromolecules, [40] Patlazhan, S., and Remond, Y., 2012, “Structural Mechanics of Semicrystalline
32(13), pp. 4390–4403. Polymers Prior to the Yield Point: A Review,” J. Mater. Sci., 47(19),
[11] Zhang, C., and Moore, I. D., 1997, “Nonlinear Mechanical Response of High pp. 6749–6767.
Density Polyethylene—Part I: Experimental Investigation and Model Eval- [41] Jiang, Z., Tang, Y., Rieger, J., Enderle, H.-F., Lilge, D., Roth, S. V., Gehrke,
uation,” Polym. Eng. Sci., 37(2), pp. 404–413. R., Heckmann, W., and Men, Y., 2010, “Two Lamellar to Fibrillar Transitions
[12] Zhong, S., Shi, J., and Zheng, J., 2013, “Study on Constitutive Modeling for in the Tensile Deformation of High-Density Polyethylene,” Macromolecules,
Large Deformation Behavior of Polyethylene Considering Strain Rate Effect,” 43(10), pp. 4727–4732.
ASME Paper No. PVP2013-97778. [42] Peres, F. M., and Sch€on, C. G., 2007, “An Alternative Approach to the Evalua-
[13] Drozdov, A. D., and Christiansen, J. d., 2008, “Thermo-Viscoelastic and Visco- tion of the Slow Crack Growth Resistance of Polyethylene Resins Used for
plastic Behavior of High-Density Polyethylene,” Int. J. Solids Struct., Water Pipe Extrusion,” J. Polym. Res., 14(3), pp. 181–189.
45(14–15), pp. 4274–4288. [43] Krishnaswamy, R. K., 2005, “Analysis of Ductile and Brittle Failures From
[14] G’Sell, C., Hiver, J. M., Dahoun, A., and Souahi, A., 1992, “Video-Controlled Creep Rupture Testing of High-Density Polyethylene (HDPE) Pipes,” Polymer,
Tensile Testing of Polymers and Metals Beyond the Necking Point,” J. Mater. 46(25), pp. 11664–11672.
Sci., 27(18), pp. 5031–5039. [44] Hoang, E. M., and Lowe, D., 2008, “Lifetime Prediction of a Blue PE100 Water
[15] Drozdov, A. D., and Yuan, Q., 2003, “The Viscoelastic and Viscoplastic Behav- Pipe,” Polym. Degrad. Stab., 93(8), pp. 1496–1503.
ior of Low-Density Polyethylene,” Int. J. Solids Struct., 40(10), pp. 2321–2342. [45] Frank, A., Pinter, G., and Lang, R. W., 2009, “Prediction of the Remaining
[16] Ritchie, S., 2000, “A Model for the Large-Strain Deformation of Polyethylene,” Lifetime of Polyethylene Pipes After up to 30 Years in Use,” Polym. Test.,
J. Mater. Sci., 35(23), pp. 5829–5837. 28(7), pp. 737–745.

021402-6 / Vol. 139, APRIL 2017 Transactions of the ASME

Downloaded From: http://pressurevesseltech.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jpvtas/935788/ on 02/14/2017 Terms of Use: http://www.asme.org

You might also like