You are on page 1of 22

Izod Impact

Izod impact measures energy required to break a specimen by striking a specific size
bar with a pendulum.

From: Extrusion (Second Edition), 2014

Related terms:

Glass Fibre, Lexan, Mols, Monomers, Tensile Strength, Tensiles, Tensile Bar, Poly-
etheretherketone

View all Topics

An overview of mechanical and physical


testing of composite materials
N. Saba, ... M.T.H. Sultan, in Mechanical and Physical Testing of Biocomposites,
Fibre-Reinforced Composites and Hybrid Composites, 2019

1.2.3.2 Izod impact


The Izod impact test was named after English engineer Edwin Gilbert Izod. The
Izod impact test is like the Charpy impact test and is used to test materials at
low temperature: http://www.wmtr.com/en.izod.html. In this test, a specimen is
machined to a square or round section, with either one, two, or three notches that
have a dimension of 70 mm × 15 mm × 3 mm [8]. The Izod impact test consists
of a pendulum with a determined weight at the end of its arm swinging down and
striking the specimen while it is held securely in a vertical position [11]. The V-notch
maker and impact test machine are displayed in Fig. 1.4 [8]. The impact strength
is determined by the loss of energy of the pendulum as determined by precisely
measuring the loss of height in the pendulum's swing [9]. Researchers also defined
impact strength as the tendency of polymer composites to endure high-energy
impact without breaking or fracturing. They also reported that in fiber-reinforced
polymer composites and hybrid composites the impact properties are governed
by the properties of the individual fibers used for hybridization, interlaminar, and
interfacial adhesion between the fiber and the matrix [8].
Figure 1.4. V-notch maker and impact testing machine [8].

> Read full chapter

Testing Properties
John R. WagnerJr., ... Harold F. GilesJr., in Extrusion (Second Edition), 2014

21.10 Izod Impact


Izod impact measures energy required to break a specimen by striking a specific size
bar with a pendulum. Izod normally refers to a notched specimen impact. However,
in some circumstances, unnotched specimens are tested. The data sheet will note
that it is an unnotched bar or unnotched Izod. The notch (needs to be machined
and not molded into the bar) acts as a stress concentrator, forcing the bar to break
at a specific location. A specific size notch (specified in the ASTM D-256 and ISO
180:1993 test methods) is machined into one side of a test bar at a specific distance
from the end. As shown in Figure 21.12, the notch is placed in the clamp and a
pendulum is released that impacts the bar, measuring the energy required to break
the sample. How tightly the specimen is clamped in the holder plus any flash or
defects in the injection-molded specimen can affect the results. Izod samples are
normally 0.125 × 0.50 × 2.5 in. (3.2 × 12.8 × 63.5 mm) in length, with the notch
machined across the 0.125-in. (3.2 mm) face and into the 0.50-in. (12.8-mm) width.
The notch has specific dimensions and radius, as defined in the ASTM and ISO
procedures.
Figure 21.12. Izod test configuration.

If the Izod bar is cut from a 5-in. (128-mm) long molded bar, either the gate end or
dead end are compared and reported as a group. Different ends of the bars are not
mixed in a sample group. If both ends are tested in the same group and compared,
the standard deviation and the measurement precision suffer due to potentially
different injection molding and packing conditions during specimen formation.
Izod is designated as an impact measurement, since the energy to break a bar is
determined. In reality, Izod measures the notch sensitivity rather than being a true
impact measurement. The notch acts as a stress concentrator and simulates what
occurs if a part is scratched or cut in a specific application, making it easier to break
the product. Figure 21.13 shows an Izod test machine, and Figure 21.14 shows the
notcher.

Figure 21.13. Ceast Izod impact testing machine.


Figure 21.14. Atlas notching machine.

> Read full chapter

Recent Advances in Polyethylene-Based


Biocomposites
Muhd R. Mansor, ... Mohd Z. Akop, in Natural Fibre Reinforced Vinyl Ester and Vinyl
Polymer Composites, 2018

3.5.1.3 Impact tests


Charpy and Izod impact testing are techniques used in impact tests to determine
the fracture characteristics in PE composites. In this test, a pendulum will hit the
specimen and the pendulum’s potential energy is determined according to the mass
and the drop height. This test also can identify whether at declining temperature,
a brittle-ductile transition occurs in polymer composites by using ASTM D-256
method. Analysis on 5% and 10% CaCO3/PE composite by Izod impact test at −40°C
to 70°C showed that the impact strength is increased and the plastic deforma-
tion micromechanism is changed into particle-induced cavitation and fibrillation
(Tanniru and Misra, 2005). Analysis on wood fiber/PE composites with MAPE as
compatibilizer shows increment to 60% of impact strength compared to composites
without compatibilizer due to the level of improvement in adhesion (Yuan et al.,
2008; Lai et al., 2003). However, the addition of hemp fiber will decrease the impact
strength while GTR (ground tire rubber) will increase it (Kakroodi et al., 2013).

Composites that are prepared with LDPE have greater impact strength among
unoccupied samples, but the brittleness will increase as loading of filler increases
due to the growing size of the poor interaction between the hydrophilic filler and the
hydrophobic polymer matrix (Yang et al., 2006). For PP/HA/LLDPE ternary biocom-
posites, the impact strength increases as content of LLDPE and impact resistance
increases with temperature (Younesi and Bahrololoom, 2009).
> Read full chapter

Material selection—which plastic to


use?
Mark T. MacLean-Blevins, in Designing Successful Products with Plastics, 2018

2.1.3.1 Notched impact strength


Often called Notched Izod impact strength, this is a relative measure of a materials
ability to resist impact. It is measured under standard test ASTM D256 and is
performed using a fixed specimen which is hit by a swinging pendulum device.
The notched specimen is cantilevered and struck by the pendulum on the notched
side (Fig. 2.9). As with other mechanical properties this test is performed at a
standard temperature and all published values will cite the testing temperature. Care
must be exercised when applying standard test data at other end-use application
temperatures or conditions. For our generic ABS case example (as shown in Fig. 2.4),
the Notched Izod impact values are published as 1.7–7.7 ft. lb./in.2, or 92–410 J/m2
as 73°F (23°C). These values represent the energy value, or magnitude, required to
break the specimen.

Figure 2.9. An illustrative view of the notched izod test set-up, showing the pendu-
lum striker path and the test specimen held with the notch facing the pendulum.The
distance from the notch to the striker impact line is defined within the test specif-
ication.
> Read full chapter

Introduction to Plastics, Polymers, and


Their Properties
Laurence W. McKeen, in The Effect of Temperature and other Factors on Plastics and
Elastomers (Third Edition), 2014

1.3.1.5 Impact Property Testing of Plastics


Table 1.4 listed a number of tests used to measure the impact resistance of plastics.
Impact tests allow designers to compare the relative impact resistance. The tests are
often used for quality control. However, these tests generally do not translate into
explicit design parameters.

While there are many measurements listed, the measurements use test apparatus
that fall into two types: based on a pendulum or a falling object. The main differences
are sample preparation and measurement units.

Izod Impact Strength and Charpy Impact Strength


The standard tests for Izod impact strength are as follows:

• ISO 180:2000 Plastics—Determination of Izod impact strength

• ASTM D256-06a Standard Test Methods for Determining the Izod Pendulum
Impact Resistance of Plastics

The standard test for Charpy impact strength is

• ISO 179-1:2000 Plastics—Determination of Charpy impact properties—Part


1: Non-instrumented impact test

Both Izod and Charpy tests are based upon a swinging pendulum, such as that
shown in Figure 1.27.
Figure 1.27. Pendulum-type impact strength tester.

Basically, the pendulum is raised to a measured point, and it is then released. The
weighted end of the pendulum gains speed as it swings toward a mounted molded
bar of the test plastic. It strikes the bar, breaks it, and the pendulum loses energy
while breaking the plastic bar. Therefore it does not swing as high. The energy lost
by the pendulum is equated with the energy absorbed by the test specimen during
the breaking process.

There are different ways to mount the test specimen, and there are different speci-
men sizes and preparation methods. The different sample mounting configurations
for the Izod and Charpy tests are shown in Figure 1.28. Figure 1.29 shows the details
of the notch. The sharpness of the bottom of the notch affects the test result. Table
1.11 shows the different notch radii possible.
Figure 1.28. Izod and Charpy impact test sample configurations.

Figure 1.29. Izod and Charpy impact test notch details.

Table 1.11. Izod and Charpy Impact Notch Radius Options

Notch Izod Notch Radius (mm) Charpy Notch Radius (mm)


A 0.25 0.25
B 1.00 1.00
C 0.10

The impact resistance is usually reported as energy per unit length or per unit area.

Gardner and Falling Dart Impact Strength


Other impact strength tests that use gravity are the Gardner impact and Falling Dart
tests. These are described in the following standards:

• D5420-04 Standard Test Method for Impact Resistance of Flat, Rigid Plastic
Specimen by Means of a Striker Impacted by a Falling Weight (Gardner Impact)
• D5628-06 Standard Test Method for Impact Resistance of Flat, Rigid Plastic
Specimens by Means of a Falling Dart (Tup or Falling Mass)

D3763-06 Standard Test Method for High Speed Puncture Properties of Plas-
tics Using Load and Displacement Sensors
• ISO 7765-2:1994 Plastics film and sheeting—Determination of impact resis-
tance by the free-falling Dart method—Part 2: Instrumented puncture test

The Gardner test uses a piece of equipment like that shown in Figure 1.30. A weight
is lifted to a given height and it is dropped onto a test plaque. The Falling Dart is
based on the same principle, but the weight is free falling rather than guided through
a tube as in the Gardner equipment.

Figure 1.30. Gardner impact test apparatus.

> Read full chapter

Modelling fracture processes in bones


A. Abdel-Wahab, ... V.V. Silberschmidt, in Computational Modelling of Biomechanics
and Biotribology in the Musculoskeletal System, 2014

9.3.2 Numerical models of Izod impact test


In this part of the study, three finite-element models were developed: Model A,
Model B and Model C. Model A is a 2D X-FEM model used to simulate the fracture
of cortical bone exposed to impact loading in the Izod test set-up. Model B is a 3D
formulation of Model A, whereas Model C is a 3D X-FEM model for quasi-static
fracture analysis. The impact tests were simulated with the finite-element software
Abaqus 6.10/Standard (Abaqus, 2010) using Models A and B to verify the applicability
of the X-FEM to analyse the failure behaviour of the cortical bone tissue under impact
loading conditions. In addition, Model C was developed to elucidate fracture devel-
opment in the cortical bone tissue for a different loading regime. A full description
of Models A and B can be found elsewhere (Abdel-Wahab et al., 2010b; Abdel-Wahab
and Silberschmidt, 2011).

Model geometry and boundary conditions


In Model A, the impact Izod test was simulated with the Abaqus/Standard finite-ele-
ment software using a 2D formulation. The real geometry and masses of the
hammer and specimen with 300 μm notchs were used in simulations (see Fig. 9.4).
The following model assumptions were made: (1) plainstrain conditions of the
specimen; (2) homogeneous and isotropic material properties for both the specimen
and the hammer; and (3) frictionless contact between the hammer and the specimen.
Before employing Model A to analyse the fracture behaviour of the bone specimen
with X-FEM, special emphasis was put on analysing the effect of the type of material
mechanical model – without using X-FEM at this stage – on results of simulations.
The specimen material in simulations was presented using elastic, elastic-plastic or
viscoelastic formulations, while the hammer was modelled using an elastic material
model due to its significantly higher stiffness. A node on the proximal part of the
hammer shown as ‘pivot’ in Fig. 9.4b was constrained in both x and y directions and
set free to rotate around z-axis to simulate its centre of rotation. In simulations,
the initial position of the hammer was close to the specimen; its angular velocity
of 5.33 rad/s corresponded to the initial angle of 58° (initial energy of 0.5 J). The
specimen support was modelled as rigid; all the degrees of freedom of the bottom
part of the specimen were constrained (see Fig. 9.4b and 9.4c). In Model B, 3D
geometry and real masses of the hammer and specimen with a 300 μm notch were
used (see Fig. 9.5).
9.4. (a) Real hammer; (b) Model A; (c) hammer–specimen interaction and mesh
around notch.

9.5. (a) Set-up of Izod test; (b) Model B; (c) hammer–specimen interaction; (d)
meshing of hammer and specimen (specimen is shown larger).
In Abaqus 6.10 (Abaqus, 2010), a kinematic coupling constraint is used to transmit
rotation to a structure while permitting a radial motion. Hence, this feature was
employed to constrain the rotational and translational degrees of freedom of the
hammer except around one axis only; it is z-axis in this case (see Fig. 9.5). To get the
exact movement of the hammer as happens in real tests, all the nodes of the inner
cylinder surface of the upper block of the hammer were kinematically coupled to a
reference point at the middle of the cylinder, then the reference point was restrained
to translate along x, y, and z and to rotate around the x or y axes (Fig. 9.5b). The
reference point was only free to rotate around the z-axis. Using Abaqus/Complete
Abaqus Environment (CAE), two sets were defined: the reference point (set1) and the
inner surface (set2). The following equation was used to define the coupling between
those sets:

[9.1]

where ur1 and ur2 are the rotational degrees of freedom for set1 and set2, respectively.

Additionally, two surfaces were chosen to define a surface-to-surface contact be-


tween the specimen and the hammer. These surfaces are shown as S1 and S2 in
Fig. 9.5d. The master surface was chosen to be S2 with S1 chosen to be the slave
one. The mesh of the master surface was adjusted to be finer than that of the slave
surface. Finite sliding with a frictionless tangential behaviour was chosen between
the two surfaces. The initial condition of the hammer with an angular velocity of
5.33 rad/s and the boundary conditions of the specimen were the same as in Model
A.

To compare the force–time history data of experiments and simulations, the force
due to contact pressure between the piezoelectric force sensor and counterpart of the
specimen elements was recorded in the history output of the finite-element software
Abaqus/Standard. Also, the status of the X-FEM that shows the crack path was used
as an output along with the distributions of stress and strain components and their
principal values. In order to compare the fracture behaviour of the cortical bone
specimen under quasi-static and impact loading, Model C (Fig. 9.6) was developed.
It consists of a cantilever-beam specimen of cortical bone with the same geometry,
mesh and material properties as in Model B. In Model C, the hammer was excluded
from the analysis and a displacement of 2 mm was applied instead at the same
position of the hammer–specimen interaction, see Fig. 9.6b.
9.6. (a) Meshed 3D quasi-static specimen; (b) applied displacement and boundary
conditions of 3D quasi-static model (Model C).

Material properties
The Resil Impactor hammer used in this study was made of carbon steel. Its elastic
material properties are given in Table 9.3. These data were obtained directly from
the manufacturer.

Table 9.3. Elastic material properties of Resil Impactor hammer

Part Material Elastic modulus Poisson ratio Density (kg/m3)


(GPa)
Hammer Carbon steel 210 0.3 7850

For numerical analysis of the cortical bone tissue, two important conclusions
based on our experimental studies were recalled: (1) experimental Izod impact tests
demonstrated that the studied bovine femoral cortical bone tissue had a nearly
uniform character of fracture energy for different cortex positions and was not
sensitive to the notch depth within the physiologically pertinent range (Abdel-Wahab
and Silberschmidt, 2011); (2) quasi-static tests provided a low level of anisotropy
(between 1.2 and 2) compared to structural composites (Abdel-Wahab et al., 2010a).
Taking this into account, it is reasonable to consider only specimens cut from the
anterior longitudinal cortex position when dealing with fracture – for Models A, B
and C.

The elastic, plastic and viscoelastic data of cortical bone tissue were quantified by
the authors experimentally (Abdel-Wahab and Silberschmidt, 2011). The average
elastic modulus and Poisson ratio of anterior longitudinal specimens are given in
Table 9.1. The average true stress-plastic strain behaviour of the latter specimens is
given in Table 9.4. Density of the cortical bone tissue was measured as 1860 kg/m3.
The viscous behaviour of bones was introduced into the FEMs A and B in terms of
the Prony series expansion of the dimensionless relaxation modulus. These material
constants are the shear relaxation modulus ratios ei = 0.1346, bulk relaxation ratios ki
= 0 and relaxation time = 117.85. Only the linear-elastic material model was used
for Model C for quasi-static loading.

Table 9.4. True stress–plastic strain data for anterior longitudinal cortical bone tissue
specimen

Plastic strain (%) Stress (MPa)


0 105.20
0.019 109.95
0.094 124.19
0.165 135.23
0.222 144.00
0.351 165.43
0.439 182.49
0.488 191.89
0.525 200.38
0.565 209.32

In these simulations, the X-FEM-based cohesive segments method was used to


simulate crack initiation and propagation along an arbitrary, solution-dependent
path in the bulk material. In Models A, B and C, the enrichment area was chosen
as the entire bone specimen. For three-dimensional Models A and C, the crack was
introduced as a plane with dimensions of 300 μm × 4 mm to reproduce the notch
depth and specimen thickness of real experiments, respectively. On the other hand,
it was introduced as a 300 μm-long line to represent the notch depth for Model B,
used in 2D simulations.

In general, damage modelling allows simulation of crack initiation and eventual


failure of an enriched area in the solution domain. The initial response is linear,
while the failure mechanism includes a damage initiation criterion and a damage
propagation law. The former was defined based on the maximum principal strain of
0.6% (Bayraktar et al., 2004; Pattin et al, 1996). When the damage initiation criterion
is met, the damage propagation law starts to govern the fracture process. In this
study, damage evolution was defined in terms of fracture energy (per unit area)
and linear softening was chosen. The mixed-mode fracture behaviour was chosen
and respective energies for the fracture modes were introduced into X-FEM. The
fracture-toughness values used were 1374 N/m, 4710 N/m and 4016 N/m for Mode
I, Mode II and Mode III, respectively (Feng et al., 2000). It is worth mentioning
that these simulations were developed before the authors characterized fracture
properties themselves (see Table 9.2).

> Read full chapter

Polymer nanocomposites for laser addi-


tive manufacturing
J.H. Koo, ... H. Wu, in Laser Additive Manufacturing, 2017

8.3.1.4 Izod impact testing


To determine material resistance, notch-A Izod impact strength tests were conduct-
ed in accordance with ASTM D256. Several impact specimens were sintered using
a laser power of 50 W, then subsequently filed and sanded to the correct testing
parameters, if necessary. Once all five samples were of appropriate dimensions, as
prescribed in the D256 standard, a 2.56-mm notch was created in each specimen
using a motorized notch cutter. Each specimen was then mounted in the Izod
impact tester and subsequently struck using a standard cantilever beam-weight
configuration. The resulting energy absorbed by the specimen before breakage was
recorded using the energy difference of the pendulum. Averaging the results of each
sample, it was determined that the Izod impact strength had a mean of 0.754 J/cm2
with a standard deviation of 0.0425 J/cm2.

> Read full chapter

PLASTICIZERS USE AND SELECTION


FOR SPECIFIC POLYMERS
TERESHATOV V. VASILIY, ... MAKAROVA A. MARINA, in Handbook of Plasticizers
(Second Edition), 2012

11.21.4 EFFECT OF PLASTICIZERS ON POLYMER AND OTH-


ER ADDITIVES
Figure 11.17 shows the effect of two plasticizers on Izod impact strength of
polyamide 11. p-methylbenzene sulfonamide has dramatic effect on impact
strength.191 Even several percents of plasticizer rapidly improves impact strength.
When using p-butyl hydroxybenzoate larger additions are required to obtain similar
effects. Figure 11.18 shows that both plasticizers have none or very minor influence
on tensile strength of polyamide 11. Both graphs illustrate that substantial improve-
ments can be made by application of plasticizers as it is being recently exploited in
practical applications.

Figure 11.17. Izod impact strength of polyamide 11 plasticized with variable quanti-
ties of p-methylbenzene sulfonamide, A, and p-butyl hydroxybenzoate, B.[Data from
Li Q F; Tian M; Kim D G; Wu D Z; Jin R G J. Appl. Polym. Sci., 83, No.7, 14th Feb.
2002, p.1600–7.]Copyright © 2002

Figure 11.18. Tensile strength of polyamide 11 plasticized with variable quantities of


p-methylbenzene sulfonamide, A, and p-butyl hydroxybenzoate, B.[Data from Li Q
F; Tian M; Kim D G; Wu D Z; Jin R G, J. Appl. Polym. Sci., 83, No.7, 14th Feb. 2002,
p.1600–7.]Copyright © 2002

Figure 11.19 shows that even small additions of plasticizer increase rapidly fatigue
life of monofilaments obtained from plasticized polyamide-6,6. Considering that the
minimum fatigue life required by commercial product is 50 minutes, at least 3 wt%
plasticizer is needed to make product useful in practical applications.203
Figure 11.19. Effect of N-methylbenzene sulfonamide on fatigue life of
polyamide-6,6.[Data from US Patent 6,249,928.]

> Read full chapter

PLASTICIZERS USE AND SELECTION


FOR SPECIFIC POLYMERS
GEORGE WYPYCH, ... MAKAROVA A. MARINA, in Handbook of Plasticizers (Third
Edition), 2017

11.21.4 EFFECT OF PLASTICIZERS ON POLYMER AND OTH-


ER ADDITIVES
Figure 11.17 shows the effect of two plasticizers on Izod impact strength of
polyamide 11. p-methylbenzene sulfonamide has dramatic effect on impact
strength.2 Even several percents of plasticizer rapidly improves impact strength.
When using p-butyl hydroxybenzoate larger additions are required to obtain similar
effects. Figure 11.18 shows that both plasticizers have none or very minor influence
on tensile strength of polyamide 11. Both graphs illustrate that substantial improve-
ments can be made by application of plasticizers as it is being recently exploited in
practical applications.
Figure 11.17. Izod impact strength of polyamide 11 plasticized with variable quanti-
ties of p-methylbenzene sulfonamide, A, and p-butyl hydroxybenzoate, B.[Data from
Li Q F; Tian M; Kim D G; Wu D Z; Jin R G, J. Appl. Polym. Sci., 83, No.7, 14th Feb.
2002, p.1600-7.]Copyright © 2002

Figure 11.18. Tensile strength of polyamide 11 plasticized with variable quantities of


p-methylbenzene sulfonamide, A, and p-butyl hydroxybenzoate, B.[Data from Li Q
F; Tian M; Kim D G; Wu D Z; Jin R G, J. Appl Polym. Sci., 83, No.7, 14th Feb. 2002,
p.1600-7.]Copyright © 2002

Figure 11.19 shows that even small additions of plasticizer increase rapidly fatigue
life of monofilaments obtained from plasticized polyamide-6,6. Considering that the
minimum fatigue life required by commercial product is 50 minutes, at least 3 wt%
plasticizer is needed to make product useful in practical applications.14
Figure 11.19. Effect of N-methylbenzene sulfonamide on fatigue life of
polyamide-6,6.[Data from US Patent 6,249,928.]

The enhanced chain mobility relative to the concentration of plasticizer (N-butylben-


zenesulfonamide) can impact interfacial compatibilization and phase coalescence.27
The threshold level of 5–10 wt% plasticizer provides an optimum level of chain
mobility.27

> Read full chapter

Integrated Simulation of Injection


Molding Process and Mechanical Be-
havior
Weihong Zhang, Yingjie Xu, in Mechanical Properties of Polycarbonate, 2019

4.4.4 Simulation results and comparisons with experiments


In this section, numerical results for the notched Izod impact of polycarbonate speci-
mens at different mold temperatures are presented and compared with experiments.

The numerical results of the plastic strain distribution within the undamaged spec-
imens at different mold temperatures are presented in Figure 4.13. The greatest
straining occurs at the notch, and there is also a high-strain region opposite to the
notch where the bending strains are concentrated. The plastic strain is always higher
for the specimen with lower mold temperature than with the opposite. This is due
to the processing-affected yield stress of polycarbonate. Since the yield stress of
polycarbonate is reduced as the mold temperature decreases, plastic deformation
will be more easily developed and the plastic strain will be accumulated to a higher
level in the specimen with lower mold temperature.
Figure 4.13. Plastic strain distribution within the undamaged specimens
at different mold temperatures. For a color version of this figure, see
www.iste.co.uk/zhang/polycarbonate.zip

Figure 4.14 presents the variation of internal energy, i.e. the total energy absorbed by
the specimens at different mold temperatures during the impact. Internal energies
can be found to vary in a similar way at different mold temperatures. The energies
gradually increase with time and suddenly decrease at the final stage, which means
that the specimens are fractured. Since the elastic properties of polycarbonate
specimens at different mold temperatures are assumed to be constant, the variations
of the energy with time are exactly the same before the yielding of polycarbonate.
Thus, the energy variation curves of the specimens at different mold temperatures
are overlapping by up to nearly 3 ms. However, distinct separations of the energy
curves are found after yielding, especially in the fracture process.
Figure 4.14. Internal energy variation of impacted specimens at different mold
temperatures. For a color version of this figure, see www.iste.co.uk/zhang/polycar-
bonate.zip

Figure 4.15. shows the detailed fracture process of the specimen at the mold
temperature of 60°C.

Figure 4.15. Fracture process of the specimen at the mold temperature of 60°C. For
a color version of this figure, see www.iste.co.uk/zhang/polycarbonate.zip

Table 4.1 lists the fracture energies obtained by numerical simulations and exper-
imental tests. The fracture energy is measured by dividing the total lost energy by
the cross-sectional area at the notch (ISO 180 2000).

As shown in Table 4.1, both experimental and numerical results clearly indicate that
the fracture energy increases as the mold temperature increases. Since the increase
in the mold temperature leads to the increase in the yield stress of polycarbonate,
more energy of the striker will be consumed for the plastic deformation and fracture
of the polycarbonate specimen. The predicted fracture energies of the polycarbonate
specimens at 60, 80 and 120°C mold temperatures are shown to be in good agree-
ment with the experimental data, indicating the predictive ability of the proposed
integrated simulation framework.

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like