0% found this document useful (0 votes)
10 views73 pages

FRC Unit 4 2025

The document discusses Hooke's Law and the stiffness and compliance matrices for various material types, including general anisotropic, specially orthotropic, transversely isotropic, and isotropic materials. It explains the mathematical relationships between stress and strain, the transformation of elastic constants, and the implications of material symmetry on the number of independent elastic constants. Additionally, it covers failure theories such as Maximum-Stress and Tsai-Wu, along with testing methods for tensile properties of composite laminates.

Uploaded by

lukerosa28
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
10 views73 pages

FRC Unit 4 2025

The document discusses Hooke's Law and the stiffness and compliance matrices for various material types, including general anisotropic, specially orthotropic, transversely isotropic, and isotropic materials. It explains the mathematical relationships between stress and strain, the transformation of elastic constants, and the implications of material symmetry on the number of independent elastic constants. Additionally, it covers failure theories such as Maximum-Stress and Tsai-Wu, along with testing methods for tensile properties of composite laminates.

Uploaded by

lukerosa28
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 73

HOOKE'S LAW AND STIFFNESS AND COMPLIANCE MATRICES

General Anisotropic Material


In general, the state of stress at a point in a body is described by the components
of the stress tensor sij, as shown in Fig. Correspondingly there is a strain tensor,
eij, with nine components.
For tensile/compressive stresses,
the first digit refers to the plane
The plane perpendicular to perpendicular to the direction of
axis 1 is plane 1 and so on. force.

(First digit refers to the For shear stresses, the first digit
plane and second digit to refers to the plane containing the
the axis) force.
Normal strains,
rest are shear.
The most general linear relationship that connects stress to strain is known as the
generalized Hooke's law and can be expressed mathematically as
(ij refer to the stress component and kl
(5.40) refer to the strain component)
I, j, k and l can each take one of three values

Where Eijkl is a fourth order tensor. The elements of Eijkl are known as the elastic
constants. The above equation also can be written in the matrix form as

(9 x 1) (9 x 9) (9 x 1)
(Stiffness matrix)
e23 = e32 (Geometry)
t23 = t32 (Newton’s law of
moment equilibrium)

In general:
ekl = elk
tkl = tlk
The first two subscripts on the elastic constants correspond to those of stress,
whereas the last two subscripts correspond to those of strain. It is seen that each
stress component is related to all nine components of the strain tensor, and there are
81 elastic constants defining the tensor Eijkl·
Fortunately, this tensor exhibits certain symmetry properties that reduce the total
number of independent components to 21 for a material that does not have any axes
of symmetry. Such a material is called aeolotropic or anisotropic.
The first set of reductions in elastic constants is obtained by considering the
symmetry of strain. It can be shown easily that because of the symmetry of the strain
tensor, there is no loss of generality if Eijkl is assumed symmetric with respect to the
last two indices; in other words,
The most general linear relationship that connects stress to strain is known as the
generalized Hooke's law and can be expressed mathematically as

where Eijkl is a fourth order tensor. The elements of Eijkl are known as the elastic
constants. The above equation also can be written in the matrix form as
(Eijkl is symmetric)
Reduces to 21
independent
constants
�11 �1111 �1122 �1133 �1123 �1131 �1112 �11
�22 �2211 �2222 �2233 �2223 �2231 �2212 �22
�33 �3311 �3322 �3333 �3323 �3331 �3312 �33
�23 = �2311 �2322 �2333 �2323 �2331 �2312 �23
�31 �3111 �3122 �3133 �3123 �3131 �3112 �31
�12 �1211 �1222 �1233 �1223 �1231 �1212 �12

�11 �1111 �1122 �1133 �1123 �1131 �1112 �11


�22 �1122 �2222 �2233 �2223 �2231 �2212 �22
�33 �1133 �2233 �3333 �3323 �3331 �3312 �33
�23 = �1123 �2223 �3323 �2323 �2331 �2312 �23
�31 �1131 �2231 �3331 �2331 �3131 �3112 �31
�12 �1112 �2212 �3312 �2312 �3112 �1212 �12
The response of an anisotropic material to impressed forces can be predicted with
the help of these 21 constants, and generally it will be different along each axis. If
one pushes along a given direction, changes in length as well as in angle will occur
along and between all the axes. It should be noted that since the components both
of the stress tensor and of the strain tensor are functions of the orientation of the
axis system, the elastic constants (the elements of the elasticity tensor) also will be
functions of axis orientation.
The elasticity tensor is a fourth-order tensor, and hence its transformation law
can be written as

where E’mnrs is the elasticity tensor in the transformed (x') axis system, Eijkl is the
elasticity tensor in the original (x) axis system, and aim etc. are the direction cosines
of the new axes with respect to the original axes. Once the elastic constants are
known in one reference coordinate system, the transformation law enables us to
calculate the elastic constants in any other reference coordinate system. In general,
the elastic constants will change with the transformation, but under some specific
transformations, the elastic constants may remain unchanged as a result of
additional symmetries existing in the material properties.
Specially Orthotropic Material (coordinate axes coincide with material axes)

Many materials exhibit symmetry in their elastic properties with respect to certain planes; that
is, the elastic constants do not change when the direction of the axis perpendicular to the
plane of symmetry is reversed. The number of elastic constants will reduce when the number
of planes of symmetry increases. The transformation law can be used to derive the number of
independent elastic constants for various symmetry conditions.

Unidirectional fiber composites come under the category of orthotropic materials that exhibit
symmetry of their elastic properties with respect to two orthogonal planes. The number of
independent elastic constants for orthotropic materials is now derived.

First, consider that one of the planes of symmetry of orthotropic materials is the x1x2 plane .
This symmetry requires that the elastic constants do not change under the following coordinate
transformation:
The invariance of elastic properties under the preceding coordinate transformation
imposes certain restrictions on the elasticity tensor. These restrictions are actually
the conditions necessary to satisfy the invariance condition (i.e., E’ijkl = Eijkl) and are
obtained by applying the transformation law. To this end, examine the dependence
of components of E’ijkl on Eijkl.
Since there are only three nonzero direction cosines, the expansion of the
transformation law is simplified. The result given states that the invariance
conditions are satisfied for the first two components examined, but not for the
third one. To satisfy the invariance condition for the third one, it is necessary to
set E1113 equal to zero. In a similar manner, it can be verified easily that the
condition of no change in the elastic constants under the coordinate
transformation would require that 8 of the 21 elastic constants should be zero.
These 8 components that must be set equal to zero are

Now to complete the symmetry requirements for an orthotropic material, consider


that the second plane of symmetry is the x2x3 plane. This means that the elastic
constants do not change under the following coordinate transformation:
The direction cosines corresponding to the preceding transformations are
It can be verified easily that the following additional constants must be zero:

Specially Orthotropic Material


(coordinate axes coincide with material
axes)
By a careful examination of Eq. (5.59), it is quickly realized that it is now
unnecessary to use four subscripts of the original elasticity tensor to describe the
nine nonzero elastic constants of orthotropic materials. It is more convenient to
write Hooke's law for an orthotropic material in the contracted notation as

Specially Orthotropic Material


Equation (5.61) represents three-dimensional stress-strain relations for an
orthotropic material when the reference coordinate axes coincide with the
material axes.

Therefore, these are called stress-strain relations for a specially orthotropic


material. The stiffness matrix contains 12 nonzero elements, with only 9 of those
being independent (C11 , C 22, C33, C12, C13, C23, C44, C55, and C66).

When coordinate axes do not coincide with the material axes, the same
orthotropic material is called a generally orthotopic material. Stress-strain
relations for a generally orthotropic material can be obtained by tensor
transformation, as used in the preceding sections.

The stiffness matrix for a generally orthotropic material is usually fully populated;
that is, it has no zero elements. However, all 36 elements are obtained only from
the 9 independent elements mentioned above.
Transversely Isotropic Material (properties are equal in directions 2 and 3)

It was pointed out earlier that for unidirectional composites, mechanical properties
in all directions perpendicular to the longitudinal direction generally are assumed
to be equal. Thus, for a unidirectional composite, the transverse plane (plane
perpendicular to the longitudinal axis) is a plane of isotropy, and a unidirectional
composite is an example of a transversely isotropic material.
In general, an orthotropic material is called transversely isotropic when one of its
planes of symmetry is isotropic. Since transversely isotropic material has more axes
of symmetry than a specially orthotropic material, its stiffness matrix has a smaller
number of independent elements. If it is assumed that plane 2-3 is the plane of
isotropy, the following relations between the elements of stiffness matrix can be
shown to exist:
Substituting in these equations
Transversely Isotropic Material (properties are equal in directions
2 and 3)
The stiffness matrix therefore can be written as
Isotropic Material

A material is called isotropic when its properties are independent of direction.


As a result, every coordinate axis is an axis of symmetry. It can be shown
that the following relations exist between the elements of the stiffness matrix:

Therefore, the stiffness matrix for an isotropic material becomes

There are only two independent elastic constants for isotropic materials ( C11 and C12).
Specially Orthotropic Material under Plane Stress

In most structural applications, composite laminates are loaded in the plane


of the laminate. Such a loading is called a plane-stress condition in which all
out-of-plane stress components are zero. If axis 3 is the out-of-plane direction,
a plane-stress condition gives
Eqn. 5.61 becomes

0
0 0
0 0
(stress) (stiffness matrix) (strain)
(5.40)

(strain) (compliance matrix) (stress)


Therefore, the following relations between the elements
of the stiffness matrix and compliance matrix may be
obtained by the matrix inversion
For a specially orthotropic 2D
lamina:
Relations between Engineering Constants and Elements of Stiffness and Compliance
Matrices

Relations between the five engineering constants of Eq. (5.10) and four independent
elastic constants of Eq. (5.70) and Eq. (5.72) can be established easily by considering a
specially orthotropic lamina with the longitudinal and transverse directions as the
material axes of symmetry. For such a lamina, stress-strain relations in terms of
stiffness and compliance matrices can be written as follows by changing subscripts of
stresses and strains from 1 and 2 to L and T, respectively:
Now consider that the lamina is subjected to a general state of stress consisting
of sL, sT, and tLT - The resulting strains are given by Eq. (5.10) in terms of engineering constants.
The strains in terms of elements of the compliance matrix are given by Eq. (5.75), which can be
written in the expanded form as

Earlier it was seen that


Problem
Obtain the stiffness and compliance matrices for
a unidirectional orthotropic lamina that has the
following elastic constants:

nTL = 0.035, Q11 = 20.25 GPa, Q22 = 2.025 GPa, Q12 = 0.709 GPa, Q66 = 0.7 GPa,
S11 = 0.05 GPa-1, S22 = 0.5 GPa-1, S12 = -0.0175 GPa-1, S66 = 1.429 GPa-1
Transformation of Stiffness and Compliance Matrices
Maximum-Stress Theory

This theory states that failure will occur if any of the stresses in the principal
material axes exceed the corresponding allowable stress. Thus the following
inequalities must be satisfied to avoid failure:
Problem
A unidirectional glass-epoxy lamina as shown, has the
following allowable stresses

According to the max. stress


theory, determine if the
lamina will fail.

sL = 37.05 MPa, sT = --12.05 MPa, tLT = -57.475 MPa


Ans: The lamina will not fail.
Problem
Determine elements in the stiffness and
compliance matrices for the unidirectional
spcially orthotropic lamina containing 60%
volume of carbon fibers in an epoxy matrix. Ef =
220GPa, nf = 0.2, Em = 3.6GPa, nm = 0.35.

nTL = 0.017, Q11 = 134.03 GPa, Q22 = 8.82 GPa, Q12 = 2.29 GPa, Q66 = 3.254 GPa,
S11 = 0.0075 GPa-1, S22 = 0.1139 GPa-1, S12 = -0.002 GPa-1, S66 = 0.307 GPa-1
Tsai–Wu Failure Theory

The Tsai–Wu failure criterion is a phenomenological material failure theory which is widely
used for anisotropic composite materials which have different strengths in tension and
compression. Under plane stress conditions, the Tsai–Wu failure theory predicts failure in
an orthotropic lamina if and when the following equality is satisfied:

SLt = Longitudinal failure strength in tension


SLc = Longitudinal failure strength in compression
STt = Transverse failure strength in tension
STc = Transverse failure strength in compression
SLT = Shear failure strength in LT plane
Note that F1, F2, F11 , F22 , and F66 can be calculated using the tensile, compressive, and
shear strength properties in the principal material directions. Determination of F12
requires a suitable biaxial test. For a simple example, consider an equal
biaxial tension test in which s11 = s22 = s at failure. Using Equation 6.4, we
can write
Example of use of Tsai-Wu theory: off-axis failure
Estimate the failure strength of a unidirectional lamina in an off-axis tension test using
the Tsai–Wu theory. Assume that all strength coefficients for the lamina are known.
An off-axis tension test on a unidirectional lamina is performed at a fiber orientation angle
q with the loading axis. The stress state in the gage section of the lamina is shown in the
figure. The stress sxx in the loading direction creates the following stresses in the principal
material directions:
sxx

sxx
Test Methods and Analysis
Tensile test:

Tensile properties, such as tensile strength, tensile modulus, and Poisson’s ratio
of flat composite laminates, are determined by static tension tests in accordance with ASTM
D3039. The tensile specimen is straight-sided and has a constant cross section. A compliant
and strain-compatible material is used for the end tabs to reduce stress concentrations in
the gripped area and thereby promote tensile failure in the gage section. Any high-
elongation (tough) adhesive system can be used for mounting the end tabs to the test
specimen.
• The tensile specimen is held in a testing machine by wedge action grips and pulled at a
recommended cross-head speed of 2mm/min (0.08 in./min).

• Longitudinal and transverse strains are measured employing electrical resistance strain
gages that are bonded in the gage section of the specimen.

• Longitudinal tensile modulus E11 and the major Poisson’s ratio n12 are determined from the
tension test data of 0o unidirectional laminates. The transverse modulus E22 and the minor
Poisson’s ratio n21 are determined from the tension test data of 90o unidirectional laminates.

• For an off-axis unidirectional specimen (0o < q < 90o), a tensile load creates both extension
and shear deformations. Since the specimen ends are constrained by the grips, shear forces
and bending couples are induced that create a non-uniform S-shaped deformation in the
specimen (Figur e 4.2). For this reason, the experimentally determined modulus of an off-
axis specimen is corrected to obtain its true modulus.
COMPRESSIVE PROPERTIES
Compressive properties of thin composite laminates are difficult to measure owing to
sidewise buckling of specimens. A number of test methods and specimen designs have been
developed to overcome the buckling problem. Three of these test methods are described as
follows.

Celanese test: This was the first ASTM standard test developed for testing fiber-reinforced
composites in compression; however, because of its several deficiencies, it is no longer a
standard test. It employs a straight-sided specimen with tabs bonded at its ends and 10o
tapered collet-type grips that fit into sleeves with a matching inner taper. An outer cylindrical
shell is used for ease of assembly and alignment. As the compressive load is applied at the
ends of the tapered sleeves, the grip on the specimen tightens and the gage section of the
specimen is compressed by the frictional forces transmitted through the end tabs. Strain
gages are mounted in the gage section to measure longitudinal and transverse strain data
from which compressive modulus and Poisson’s ratio are determined.

The challenge with the Celanese design was the lack of alignment rods meant it was easy for
the wedges to become misaligned and introduce a bending stress on to the specimen,
promoting a premature failure. This issue was overcome through the presence of alignment
rods in the IITRI fixture but eventually these misalignment issues saw the Celanese fixture
removed from the ASTM D3410 standard in recent years.
IITRI test: The IITRI test was first developed at the Illinois Institute of Technology Research
Institute and was later adopted as a standard compression test for fiber-reinforced
composites (ASTM D3410). It is similar to the Celanese test , except it uses flat wedge grips
instead of conical wedge grips. Flat wedge surfaces provide a better contact between the
wedge and the collet than conical wedge surfaces and improve the axial alignment. Flat
wedge grips can also accommodate variation in specimen thickness.

The IITRI test fixture contains two parallel guide pins in its bottom half that slide into two
roller bushings that are located in its top half. The guide pins help maintain good lateral
alignment between the two halves during testing. The standard specimen length is 140mm,
out of which the middle 12.7mm is unsupported and serves as the gage length.

Either untabbed or tabbed specimens can be used; however, tabbing is preferred, since it
prevents surface damage and end crushing of the specimen if the clamping force becomes
too high.
Sandwich edgewise compression test: In this test, two straight-sided specimens
are bonded to an aluminum honeycomb core that provides the necessary support
for lateral stability. Compressive load is applied through the end caps, which are
used for supporting the specimen as well as preventing end crushing. The average
compressive stress in the composite laminate is calculated assuming that the core
does not carry any load. Table 4.5 shows representative compressive properties for
carbon fiber–epoxy and boron fiber–epoxy laminates obtained in a sandwich
edgewise compression test. The data in this table show that the compressive
properties depend strongly on the fiber type as well as the laminate configuration .

Compressive test data on fiber-reinforced composites are limited. From the


available data on 0o laminates, the following general observations can be made.
1. Unlike ductile metals, the compressive modulus of a 0o laminate is not equal to
its tensile modulus.
2. Unlike tensile stress–strain curves, compressive stress–strain curves of 0o
laminates may not be linear.
3. The longitudinal compressive strength of a 0o laminate depends on the fiber type,
fiber volume fraction, matrix yield strength, fiber length–diameter ratio, fiber
straightness, fiber alignment as well as fiber–matrix interfacial shear strength.
4. Among the commercially used fibers, the compressive strength and modulus of Kevlar
49-reinforced composites are much lower than their tensile strength and modulus.
Carbon and glass fiber-reinforced composites exhibit slightly lower compressive
strength and modulus than their respective tensile values, and boron fiber-reinforced
composites exhibit virtually no difference between the tensile and compressive
properties.
With end loading compression, the flatness and parallelism of the ends of the samples,
along with that of the compression platens is of vital importance in order to ensure an
even load transfer across the ends of the samples. This is why standards often state a
requirement that the ends of the samples and the working area of your compression
platens should be flat to within 0.02mm. If the samples aren’t parallel, you will end up
with the load transfer being off axis in comparison to the fiber direction, which will
promote bending and a premature buckling failure.

Likewise, with shear loaded compression samples, the flatness and parallelism of the
tabbed surfaces on the side of the sample are of upmost importance. Again, this is in
order to prevent excessive bending.
FLEXURAL PROPERTIES
Flexural properties, such as flexural strength and modulus, are determined by ASTM test method D790. In
this test, a composite beam specimen of rectangular cross section is loaded in either a three-point
bending mode or a four-point bending mode.
In either mode, a large span–thickness (L/h) ratio is recommended. We will consider only the three point
flexural test for our discussion.
The maximum fiber stress at failure on the tension side of a flexural specimen is considered the flexural
strength of the material. Thus, using a homogeneous beam theory, the flexural strength in a three-point
flexural test is given by
where m is the initial slope of the load–deflection curve.
Three-point flexural tests have received wide acceptance in the composite material
industry because the specimen preparation and fixtures are very simple. However, the
following limitations of three-point flexural tests should be recognized.

1. The maximum fiber stress may not always occur at the outermost layer in a
composite laminate. An example is shown in Figure 4.24. Thus, equation 4.10 gives
only an apparent strength value. For more accurate values, lamination theory should
be employed.

2. In the three-point bending mode, both normal stress sxx and shear stress txz are
present throughout the beam span. If contributions from both stresses are taken into
account, the total deflection at the midspan of the beam is
IMPACT PROPERTIES
The impact properties of a material represent its capacity to absorb and dissipate
energies under impact or shock loading. A variety of standard impact test methods
are available for metals (ASTM E23) and unreinforced polymers (ASTM D256). Some
of these tests have also been adopted for fiber-reinforced composite materials.

CHARPY, IZOD, AND DROP-WEIGHT IMPACT TEST

Charpy and Izod impact tests are performed on commercially available machines in
which a pendulum hammer is released from a standard height to contact a beam
specimen (either notched or unnotched) with a specified kinetic energy. A horizontal
simply supported beam specimen is used in the Charpy test, whereas a vertical
cantilever beam specimen is used in the Izod test. The energy absorbed in breaking
the specimen, usually indicated by the position of a pointer on a calibrated dial
attached to the testing machine, is equal to the difference between the energy of
the pendulum hammer at the instant of impact and the energy remaining in the
pendulum hammer after breaking the specimen.
The drop -weight impact test uses the free fall of a known weight to supply the energy to
break a beam or a plate specimen. The specimen can be either simply supported or fixed .
The kinetic energy of the falling weight is adjusted by varying its drop height . The impact
load on the specimen is measured by instrumenting either the striking head or the
specimen supports. Energy absorbed by the specimen is calculated as
FATIGUE PROPERTIES
The fatigue properties of a material represent its response to cyclic loading, which is a common
occurrence in many applications. Fatigue behavior of a material is usually characterized by an S–N
diagram, which shows the relationship between the stress amplitude or maximum stress and number of
cycles to failure on a semi-logarithmic scale. This diagram is obtained by testing a number of specimens
at various stress levels under sinusoidal loading conditions. For a majority of materials, the number of
cycles to failure increases continually as the stress level is reduced.
The tension–tension fatigue cycling test procedure is described in ASTM D3479. It uses a straight-sided
specimen with the same dimensions and end tabs as in static tension tests.

FATIGUE TEST METHODS

The majority of fatigue tests on fiber-reinforced composite materials have been performed with uniaxial
tension–tension cycling.

Both stress-controlled and strain-controlled tests are performed. In a stress-controlled test, the
specimen is cycled between specified maximum and minimum stresses so that a constant stress
amplitude is maintained. In a strain-controlled test, the specimen is cycled between specified maximum
and minimum strains so that a constant strain amplitude is maintained.

A unique feature of a fiber-reinforced composite material is that it exhibits a gradual softening or loss in
stiffness due to the appearance of microscopic damages long before any visible damage occurs. As a
result, the strain in the specimen increases in stress-controlled tests, but the stress in the specimen
decreases in strain-controlled tests
smean
where
af is the shape parameter in fatigue L0 is the location parameter for the fatigue
life distribution (cycles).
CREEP
Creep is defined as the increase in strain with time at a constant stress level. In polymers,
creep occurs because of a combination of elastic deformation and viscous flow, commonly
known as viscoelastic deformation. The resulting creep strain increases nonlinearly with
increasing time. When the stress is released after a period of time, the elastic deformation
is immediately recovered.

The deformation caused by the viscous flow recovers slowly to an asymptotic value called
recovery strain.

Creep strain in polymers and polymer matrix composites depends on the stress level and
temperature. Many polymers can exhibit large creep strains at room temperature and at
low stress levels. At elevated temperatures or high stress levels, the creep phenomenon
becomes even more critical . In general, highly cross-linked thermoset polymers exhibit
lower creep strains than thermoplastic polymers. With the exception of Kevlar 49 fibers,
commercial reinforcing fibers, such as glass, carbon, and boron, do not creep.
Under uniaxial stress, the creep behavior of a polymer or a polymer matrix composite is
commonly represented by creep compliance, defined as
FRACTURE BEHAVIOR AND DAMAGE TOLERANCE
The fracture behavior of materials is concerned with the initiation and growth of critical cracks that may
cause premature failure in a structure.
In fiber-reinforced composite materials, such cracks may originate at manufacturing defects, such as
microvoids, matrix microcracks, and ply overlaps, or at localized damages caused by in-service loadings,
such as subsurface delaminations due to low-energy impacts and hole-edge delaminations due to static
or fatigue loads. The resistance to the growth of cracks that originate at the localized damage sites is
frequently referred to as the damage tolerance of the material.

CRACK GROWTH RESISTANCE


Many investigators have used the linear elastic fracture mechanics (LEFM) approach for studying the
crack growth resistance of fiber-reinforced composite materials. The LEFM approach, originally
developed for metallic materials, is valid for crack growth with little or no plastic deformation at the crack
tip. It uses the concept of stress intensity factor KI, which is defined as
Equation 4.58 shows that the stress intensity factor increases with both applied stress and crack length.
An existing crack in a material may propagate rapidly in an unstable manner (i.e., with little or no plastic
deformation), when the KI value reaches a critical level. The critical stress intensity factor, KIc, also called
the fracture toughness, indicates the resistance of the material to unstable crack growth.

The critical stress intensity factor of metals is determined by standard test methods, such as ASTM E399.
No such standard test method is currently available for fiber-reinforced composite materials. Most
investigators have used static tensile testing of prenotched straight-sided specimens to experimentally
determine the stress intensity factor of fiber-reinforced composite laminates.

Three types of specimens, namely, center-notched (CN), single-edge notched (SEN), and double-edge
notched (DEN) specimens, are commonly used.

You might also like