You are on page 1of 28

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/263557657

Basic Experimental Characterization of Polymer Matrix Composite Materials

Article in Polymer Reviews · May 2013


DOI: 10.1080/15583724.2013.776588

CITATIONS READS

60 29,378

3 authors, including:

Byron Pipes
Purdue University
279 PUBLICATIONS 10,602 CITATIONS

SEE PROFILE

All content following this page was uploaded by Byron Pipes on 22 October 2015.

The user has requested enhancement of the downloaded file.


Polymer Reviews

ISSN: 1558-3724 (Print) 1558-3716 (Online) Journal homepage: http://www.tandfonline.com/loi/lmsc20

Basic Experimental Characterization of Polymer


Matrix Composite Materials

L. A. Carlsson , D. F. Adams & R. B. Pipes

To cite this article: L. A. Carlsson , D. F. Adams & R. B. Pipes (2013) Basic Experimental
Characterization of Polymer Matrix Composite Materials, Polymer Reviews, 53:2, 277-302, DOI:
10.1080/15583724.2013.776588

To link to this article: http://dx.doi.org/10.1080/15583724.2013.776588

Published online: 18 Apr 2013.

Submit your article to this journal

Article views: 362

View related articles

Citing articles: 1 View citing articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=lmsc20

Download by: [Purdue University] Date: 22 October 2015, At: 06:54


Polymer Reviews, 53:277–302, 2013
Copyright © Taylor & Francis Group, LLC
ISSN: 1558-3724 print / 1558-3716 online
DOI: 10.1080/15583724.2013.776588

Basic Experimental Characterization of Polymer


Matrix Composite Materials

L. A. CARLSSON,1 D. F. ADAMS,2 AND R. B. PIPES3


1
Department of Mechanical and Ocean Engineering, Florida Atlantic University,
Boca Raton, FL, USA
2
Wyoming Test Fixtures, Inc., Salt Lake City, UT, USA
3
School of Aeronautics and Astronautics, Purdue University, West Lafayette,
Downloaded by [Purdue University] at 06:54 22 October 2015

IN, USA

Composite materials are used in high-performance applications where achieving low


weight is of concern. Since their introduction in the early 1960s, composite materials
consisting of strong fibers in a polymer matrix have received considerable success.
Because of the great variety of fiber and matrix combinations, there is no general data
base of material properties available for the designer of composite structures. Design of
composite materials demands accurate experimental determination of the mechanical
properties of the composite materials considered. The emphasis of this article is on
the experimental characterization of composite materials in the most basic loading
configurations, viz. tension, compression, and shear. The description of the test methods
is limited to in-plane mechanical response, with the exception of short-beam shear
testing.

Keywords composites, testing, data reduction

1. Introduction
Composite materials used for advanced applications such as airframes, space structures,
race cars, and sporting goods employ continuous fiber reinforcements in the form of a
ply. Figure 1 shows two types of such plies where unidirectional fibers and woven fabric
bundles are impregnated by a polymer resin to form the ply. For most composites in use
today, the ply is the basic unit or building block, whether it is in the design, the analysis,
or the fabrication process stage. This lamina may be a unidirectional prepreg, a fabric, a
chopped strand mat, or another fiber form. However, some composites are made from dry
pre-made fiber forms subsequently impregnated with liquid resin, for example, vacuum-
assisted resin transfer molding (VARTM).1 Still, such composites utilize well-recognizable
plies. Hence, whatever the material form of the fabrication process, the properties of the
individual ply (regions, layers, or whatever form the composite takes) must be known for
design and analysis purposes. Determination of the stiffness and strength properties of the
individual ply will be the topic of this article.
Typical fiber and resin properties are listed in Table 1.2 It is noted that glass fibers are
relatively dense, and as a result are less attractive for extreme light-weight applications such

Received November 7, 2012; accepted January 9, 2013.


Address correspondence to L. A. Carlsson, Department of Mechanical and Ocean Engineering,
Florida Atlantic University, 777 Glades Road, Boca Raton, FL 33431, USA. E-mail: carlsson@fau.edu

277
278 L. A. Carlsson et al.
Downloaded by [Purdue University] at 06:54 22 October 2015

Figure 1. Orthotropic composite plies: a) unidirectional, b) woven fabric (plain weave).

as air frames and space frames. The fibers are much stiffer and stronger than the matrix
resins and typically break at smaller strains. The ply unit is highly orthotropic, especially
the unidirectional ply (Fig. 1a), where the stiffness and strength along the fiber direction (x1 )
are much higher than in directions transverse to the fiber (x2 and x3 ). The woven fabric ply
(Fig. 1b), is more balanced and has similar properties in the x1 and x2 directions, but has low
out-of-plane stiffnesses and strengths due to the planar 0◦ and 90◦ fiber pattern. To achieve
a material more resistant to out-of-plane loads, there have been several developments to
orient fibers in the thickness direction (x3 ) or to stitch a laminate preform with planar fiber

Table 1
Typical fiber and matrix properties2

Density, Modulus∗ , Tensile strength∗ , Strain at


Material g/cm3 E, GPa σ ult, MPa break∗ ,%
Carbon fiber (AS4) 1.81 235 3.700 1.57
Glass fiber (E) 2.60 73 3,450 4.73
Kevlar fiber (49) 1.44 131 3,800 2.90
Epoxy resin (3501–6) 1.27 4.43 69 1.70
Vinylester resin (510A) 1.23 3.40 86 4–5

Modulus and strength data for the fibers refer to the axial (“strong”) direction of the fiber.
Polymer Matrix Composite Materials 279

orientation, but such composites are used only for special applications demanding high
strength in the thickness direction.
Test methods used for composite materials were initially very similar to those used
for testing metals and plastics, but it was soon realized that the highly orthotropic behavior
and peculiar failure modes of composite materials demanded specialized test methods.
Aerospace companies developed their own procedures and test methods such as the “Boeing
compression test”.3 Several test methods were developed for measuring the same property.
Some tests were easy to use but produced unreliable test results, while other more complex
test methods produced reliable results given the operator possessed appropriate skills.
In the US, the government sponsored much of the early development work to generate a
database and standardized test methods. Such attempts, however, were largely unsuccessful
because of the continuous emergence of new materials that did not respond to loads in the
same manner as the earlier developed ones.
Despite the many obstacles discussed here, consensus organizations, such as ASTM and
Downloaded by [Purdue University] at 06:54 22 October 2015

ISO, have made tremendous progress in developing test procedures suitable for composite
materials. Most of the test methods reviewed and described in this article and numerous
other composite tests are described in Vol. 15 of The ASTM Annual Books of Standards.4
It should be pointed out that the tests reviewed here are intended for basic materials
characterization, that is, tension, compression, and shear, and such tests are considered well
established. Still, new basic tests continue to appear while other tests are fading from use.
Many other tests exist.4 (See also Tarnopolski and Kincis,5 and Adams et al.6)

2. Basic Loading Configurations of Composite Materials


The basic unit (ply) considered, Fig. 1, may be loaded by any of the six stress components
illustrated in Fig. 2, i.e., σ 1 , σ 2 , σ 3 , τ 23 , τ 13 , and τ 12 . σ 1 , σ 2, and σ 3 are the normal stresses
while τ 23 , τ 13 , and τ 12 are the shear stresses. Associated with these stresses are the strains
ε1 , ε2 , ε3 , γ 23 , γ 13 , and γ 12 , defined in a standard small strain manner.7
It must be pointed out that stresses and strains vary very much throughout the volume
of the composite ply. The fibers are much stiffer than the matrix, (see Table 1), and fiber-
to-matrix modulus ratios, Ef / Em , of the order of 100 are not uncommon. As a result, the
matrix may deform more than the fibers when the composite is under load, especially under
loadings transverse to the fiber direction. As a result, concepts such as stress and strain
refer to suitable defined volume averages, (see Hashin8 for example),


1
σij = σij (x, y, z) dV (1a)
V
V

1
εij = εij (x, y, z) dV (1b)
V
V

where σ ij refers to the six stress components indicated in Fig. 2, and εij refers to the
associated strain components at a material point (x, y, z). V refers to the volume of the
element considered, and the overbar indicates average.
Hence, it is important to realize these definitions of stresses and strains in heterogeneous
materials such as composites. As a result of the volume averages, mechanical properties
such as modulus (E), Poisson’s ratio (ν), and strength (σ ult ) are defined in terms of volume
280 L. A. Carlsson et al.
Downloaded by [Purdue University] at 06:54 22 October 2015

Figure 2. Definition of stress components and ply coordinate system for an orthotropic ply.

averaged stresses and strains, that is,

σa
E= (2a)
εa
εt
ν=− (2b)
εa
σult = Pult /A (2c)

where subscripts “a” and “t” refer to axial and transverse directions of a test specimen (see
Fig. 3), and Pult is the failure load and A is the cross-sectional area of the specimen.

Figure 3. Schematic of composite tensile test specimen and definition of axial (a) and transverse
(t) directions.
Polymer Matrix Composite Materials 281

Figure 4. Electrical resistance strain gage.

The heterogeneity of composite materials is also important to recognize in experimental


testing. For strain measurements, electrical strain gages (Fig. 4) are typically glued to the
surface of a test specimen. Electrical strain gages have been found to be very versatile,
Downloaded by [Purdue University] at 06:54 22 October 2015

reliable, and accurate.9,10 Strain gages are typically small in order to capture the strain at
the desired point in a structure or test specimen, but if the composite is very inhomogeneous,
such as a coarse-weave fabric (Fig. 1b), the active strain measuring area of the gage must
be large enough to cover a representative area of the weave structure.11
Extensometers can also be used to measure strain. Most extensometers are single-
axis, although biaxial units capable of measuring both axial and transverse strains (for
Poisson’s ratio) are available.12–14 The strain measured from an extensometer is based on
the displacement over a distance, typically about 25 mm, which should be adequate even for
coarse-weave fabric composites. Extensometers are less common in composites testing than
strain gages. An extensometer is relatively expensive, and may be damaged or destroyed
upon failure of the test specimen which often tends to be violent because of the release
of elastic strain energy upon failure. Linear variable differential transformers (LVDTs) are
commonly used to measure displacement.15
There are also optical (non-contacting) strain measurement devices such as laser exten-
someters12,13 which do not influence the response of the specimen. The resolution of such
instruments, however, is generally not as high as for strain gages. Recently much attention
has been given to another type of optical technique that is able to provide quite accurate
displacement and strain measurements over a finite area of the specimen. A very powerful
method is the digital image correlation (DIC) method where an area of the specimen is illu-
minated by a laser.16 The displacements of points on the surface of the body are expressed
in pixels when the pictures are analyzed. The scale can vary from μm to m depending on
the accuracy required. A reference image is created for the undeformed object which is
compared to an image of the deformed object. The locations of points on the surface of
the deformed object are obtained by digital image correlation and by identifying the same
points on the surface of the undeformed object it is possible to map the displacements
and calculate the distribution of strains for the entire image. Although this method is most
often applied to structural analysis, it has also found its place in materials testing, (see for
example, references17,18).

2.1 Elastic Constitutive Relations For an Orthotropic Ply


Consider the orthotropic composite ply (Fig. 1), under in-plane stresses σ 1 , σ 2 , and τ 12 in
Fig. 5. Inherent in this presentation is an assumption of plane stress, where the out-of-plane
stresses σ 3, τ 23 , and τ 13 shown in Fig. 1 are assumed to be small and neglected. Such
282 L. A. Carlsson et al.
Downloaded by [Purdue University] at 06:54 22 October 2015

Figure 5. Composite ply under in-plane loading.

an assumption is common both in structural analysis of composites and in basic testing,


although the design of most structural configurations demand 3D analysis and a full set of
orthotropic elastic constants.
Some methods to determine out-of-plane mechanical properties are described in Refer-
ence4 and a recent text.19 Analysis to predict 3D elastic properties of laminates based on the
in-plane elastic properties of the orthortropic plies and the lay-up sequence are described
by Sun and Li20 and Hyer and Knott.21
For a 2D plane stress state, the in-plane elastic constitutive relation between stresses
and strain is,22
⎡ ⎤ ⎡ ⎤⎡ ⎤
ε1 S11 S12 0 σ1
⎣ ε2 ⎦ = ⎣ S12 S22 0 ⎦ ⎣ σ2 ⎦ (3)
γ12 0 0 S66 τ12

Figure 6 illustrates the deformation of an element of the composite under each of the
in-plane stresses σ 1, σ 2 , and τ 12. For loading by σ 1, the element elongates along the x1
direction and contracts along the x2 direction, Fig. 6a, similarily for the element loaded by
σ 2 (Fig. 6b). For in-plane shear loading, the originally straight angle changes to π2 - γ12 .

Associated with the loading by σ 1 only shown in Fig. 6a are the strains ε1 and ε2 ,
σ1
ε1 = (4a)
E1
−ν12
ε2 = (4b)
ε1
where E1 and ν 12 are the effective modulus and Poisson’s ratio for loading along the x1
direction.
Equations (4) allow identification of the following compliance elements in Eqs. (3),
1
S11 = (5a)
E1
−ν12
S12 = (5b)
E1
Polymer Matrix Composite Materials 283
Downloaded by [Purdue University] at 06:54 22 October 2015

Figure 6. Loading of a composite material element by single stresses. (a) σ 1 , (b) σ 2 , (c) τ 12.

For loading by σ 2 only, Fig. 6b,


σ2
ε2 = (6a)
E2
−ν21
ε1 = (6b)
ε2

where E2 and ν 21 are the effective modulus and Poisson’s ratio for loading along the x2
direction.
This in Eqs. (3) leads to

1
S22 = (7a)
E2
ν21
S12 =− (7b)
E2

Finally, for pure shear loading by τ 12 , Fig. 6c,


τ12
γ12 = (8a)
G12

where G12 is the effective shear modulus of the composite in the x1 x2 plane.
In Eqs. (3) this leads to

1
S66 = (8b)
G12
284 L. A. Carlsson et al.

Table 2
Typical elastic constants for continuous fiber composites,2 and minimum aspect ratio.

Carbon/ E-Glass/ Woven Woven


Property Epoxy Epoxy Carbon/ Epoxy Glass/ Epoxy
Fiber volume fraction, Vf .63 .55 .62 .45
E1 , GPa 147 41 77.0 29.7
E2 , GPa 10.3 10.4 75.0 29.7
G12 , GPa 7.0 4.3 6.5 5.3
ν 12 .27 .28 .06 .17
Lg /w 6.74 4.54 5.06 3.48

In summary, the compliance elements Sjj may be related to the engineering constants (E1 ,
Downloaded by [Purdue University] at 06:54 22 October 2015

E2 , G12 , ν 12 , ν 21 )
1
S11 = (9a)
E1
ν12 ν21
S12 =− =− (9b)
E1 E2
1
S22 = (9c)
E2
1
S66 = (9d)
G12
The Poisson’s ratios ν 12 and ν 21 are not independent, see Eq. (9b),
E2
ν21 = ν12 (10)
E1
Table 2 lists typical elastic constants for carbon and glass fiber composites.2 For a unidi-
rectional composite (Fig. 1a) E2 is much less than E1 . Hence ν 21 is much less than ν 12 . For
a woven fabric composite (Fig. 1b) E1 ≈ E2 and ν 12 ≈ ν 21 .

2.2 Failure of a Composite Ply


Failure of a composite ply due to the in-plane stresses σ 1 , σ 2 , and τ 12 shown in Fig. 5 occurs
in very different modes according to the stress acting and its sign (tension or compression)
for the normal stresses σ 1 and σ 2 . The failure mechanism and strength depend strongly on
whether the ply consists of unidirectional fibers (Fig. 1a) or woven fabric (Fig. 1b). The
unidirectional ply is very strong when loaded in the fiber direction, but weak when loaded
transverse to the fibers. Figure 7 illustrates the failure modes for the unidirectional ply due
to stresses σ 1 , σ 2 , and τ 12 . When σ 1 is tensile (σ 1 > 0) the ply fails by failure of both the
fibers and matrix, Fig. 7a. The ultimate stress is σ 1 = XT1 . For compression loading by σ 1
(σ 1 < 0), the ply fails due to a fiber instability mode which often leads to broken fibers in
a narrow zone (kink band) (see Fig. 7b). The magnitude of the compression failure stress
is |σ 1 | = XC1 .
For any other applied stress, Figs. 7c–e show that the failure is governed by matrix
and fiber/matrix interface failure and no broken fibers. Woven fabric composites loaded by
Polymer Matrix Composite Materials 285
Downloaded by [Purdue University] at 06:54 22 October 2015

Figure 7. Failure modes of unidirectional composite under plane stress.

the stresses σ 1 and σ 2 have similar strengths (X1T ∼


= X2T and X1C ∼= X2C ) while the shear
strength (S6 ) is quite small as it is governed by the matrix. Table 3 defines symbols for
the ply strengths, while Table 4 provides some typical strength data for unidirectional and
woven fabric composites.2

Table 3
Symbols for basic in-plane strengths of an orthotropic ply

Direction/ Plane Stress Strength


x1 σ1 XT1 , XC1
x2 σ2 XT2 , XC2
x1 x2 τ 12 S6
Note: all strengths are defined in terms of their magnitude.
286 L. A. Carlsson et al.

Table 4
Typical in-plane strengths (MPa) for continuous fiber composites (same in Table 2)2

Composite XT1 XC1 XT2 XC2 S6


Carbon/ epoxy 2,280 1,725 57 228 76
E-glass/ epoxy 1,140 620 39 128 89
Woven carbon/ epoxy 963 900 856 900 71
Woven glass/ epoxy 367 549 367 549 97.1

2.3 St. Venant’s Principle and End Effects in Composites


In the testing and evaluation of material properties, it is generally assumed that load
introduction effects are confined to a region close to the grips or loading points, and
a uniform state of stress and strain exists within the test section. The justification for
Downloaded by [Purdue University] at 06:54 22 October 2015

such a simplification is usually based on the St. Venant principle, which states that the
difference between the stresses caused by statically equivalent load systems is insignificant
at distances greater than the largest dimension of the area over which the loads are acting.7
This estimate, however, is based on isotropic material properties. For anisotropic composite
materials, Horgan et al.23–26 showed that the application of St. Venant’s principle for plane
elasticity problems involving orthotropic materials is not justified in general.
For the particular problem of a rectangular strip made of highly orthotropic material
and loaded at the ends, it was demonstrated that the stress approached the uniform St. Venant
solution much more slowly than the corresponding solution for an isotropic material.25 The
size of the region where end effects influence the stresses in the strip may be estimated by,

bln100 E1
λ≈ (11)
2π G12

where b is the maximum dimension of the cross section (typically the specimen width) and
E1 and G12 are the longitudinal elastic and shear moduli.
In this equation λ is defined as the distance over which the self-equilibrated stress
decays to 1/e of its value at the end (e = 2.718). When the ratio E1 /G12 is large the
decay length is large and end effects are transferred a considerable distance along the gage
section. Testing of highly orthotropic materials thus requires special consideration of load
introduction effects. Arridge et al.27 for example, found that a very long specimen with an
aspect ratio ranging from 80 to 100 was needed to avoid the influence of clamping effects
in tension testing of highly orthotropic, drawn polyethylene film.
For a conservative estimate of the gage length required to achieve a uniform stress in
a test specimen, Eq. (11) would indicate a minimum aspect ratio (gage length/specimen
width) of:

Lg ln100 E1
> (12)
w π G12

where Lg is the gage length and w is the specimen width. Table 2 lists typical stiffness
constants and aspect ratios (Lg /w) required for testing composites. A unidirectional car-
bon/epoxy composite would require a gage length about 7 times longer than the width
of the specimen. ASTM standard D303928 for tensile testing of composites (to be further
discussed) suggests a gage length of 12.5 cm and a width of 1.25 cm for unidirectional
Polymer Matrix Composite Materials 287

composites. This would provide sufficient decay of end effects for all composites listed in
Table 2 according to the above analysis.
For characterization of compression strength, compression tests (to be discussed)
utilize short specimens, or a specimen with a short unsupported gage section. The short
gage section is necessitated to avoid buckling failure, which may become an issue regarding
uniformity of stress for highly orthotropic test specimens, see Bogetti et al.29

3. Basic Mechanical Characterization of the Orthotropic Ply


In this section, we will describe the most basic composites tests, viz., tension, compression,
and shear, in some detail. It should be pointed out that numerous test procedures exist for
compression and shear testing. In this paper, attention will be placed only on the most
accepted test methods that have become ASTM standards. Many other test methods exist
some of which are ASTM standards.4 An important source for variability of test results and
Downloaded by [Purdue University] at 06:54 22 October 2015

strengths less than anticipated is attributed to the machining of test specimens from a larger
panel. Care must be taken when cutting composite materials for use as test specimens.30–32
The material can be damaged in the process, resulting in reduced strength properties. One
way this damage can be induced is by excessive heat buildup in the cutting zone. Many
methods require the use of tabs bonded to the test specimens. Tabs are used to reduce stress
concentrations and protect the specimen from the aggressive action of the loading device,
for example, wedge grips.33–35 However, the tab bonding procedure can introduce its own
sources of error. The type of tabbing material can be inappropriate for the application or
be of poor quality. The adhesive used can be inadequate or improperly applied. A detailed
description of the test specimen preparation is provided in Reference 6.
Emphasis is here placed on a mechanical test designed to determine the mechanical
response of the basic building blocks of a multi-ply composite laminate, that is, the uni-
directional ply. With known elastic and failure properties of the plies in a laminate, it is,
in principle, possible to predict the stiffness and strength properties of the laminate. For
prediction of stiffness, the classical laminated plate theory22 is widely accepted. Analysis of
ply failures in a multi-ply laminate, however, is not straightforward. One issue with testing
of laminates is the well-known free-edge stress problem.36 This problem is manifested by a
three-dimensional stress state in regions near free edges, and such stresses may separate the
plies. Another problem is that multi-ply laminates that are cured at elevated temperatures
and tested typically at room temperature will have residual thermal stresses that can be a
substantial fraction of the ply failure stress. Such stresses will influence the strength of the
laminate.
Perhaps more serious is that the ply failure stress depends on both the ply thickness
and the orientation of neighboring plies, (see e.g., References37,38.) Still, a basic mechanical
characterization of composites starts with testing of ply properties, but the analyst must be
aware of the uncertainties associated with scale-up, and the issues associated with testing
of laminates.

3.1 Tensile Testing


Tensile properties are often the first to be thought of when a composite or any other type of
material is considered for a design application. Although tensile properties do not always
ultimately determine the design, they are nevertheless among those of primary importance.
The difficulty of performing an acceptable tensile test typically increases as the or-
thotropy of the material increases, that is, as the ratio of the axial stiffness (or strength) to the
288 L. A. Carlsson et al.

transverse stiffness (or strength) and the longitudinal shear stiffness (or strength) increases.
Thus, a unidirectional composite lamina is often the most difficult material form to test.
For this reason, particular emphasis will be on suitable methods for testing unidirectional
composites.
A significant source of potential problems with test results is in the specimen. The
elastic stiffness and strength properties of orthotropic composites depend strongly on the
fiber orientation. Hence, it is extremely important that the test specimen be checked for
proper fiber alignment, particularly for an axial test of a unidirectional 0◦ ply, because
modulus and strength decrease rapidly with off-axis angle.6 Fiber alignment within ± 1◦
is attainable if sufficient attention is given to the fiber orientation of the individual plies
of prepreg and the establishment of a visible reference axis during laminate fabrication.35
A few fiber threads of a contrasting color, for example, glass fibers in a carbon–fiber
composite, can be inserted during prepreg manufacture to assist in attaining alignment. As
a final check, after cure of the unidirectional composite, the panel may be fractured along
Downloaded by [Purdue University] at 06:54 22 October 2015

the fiber direction. A clean break indicates that the individual plies are well aligned. If a
clean break is not obtained, the panel should be discarded.

3.2 Load Introduction


Most commonly, clamping grips are used to introduce tensile load in the specimen. By this
procedure the load is induced via shear at the clamp-specimen interface. This shear force
is equal to the clamping force times the effective coefficient of friction at the interface.
There is a limit to how much clamping force can be applied without crushing or otherwise
degrading the specimen material. It is recognized that the through-thickness compressive
strength of most composites is low relative to the axial tensile strength. However, the
faces of the clamp can be altered to increase friction. Machined patterns of increasing
aggressiveness (e.g., from swirls to crosshatches to straight grooves) can be used, with
deeper and sharper profiles more effectively penetrating the surface of the specimen and
thus producing a higher effective coefficient of friction.
Wedge grips are commonly used. Two types of wedge grips are in common use:
mechanical and hydraulic. The wedging action of mechanical wedge grips is in direct
proportion to the magnitude of the tensile force being applied, and inversely proportional
to the angle to which the wedges are machined, which is typically about 10◦ . If the wedge
angle is too low, the clamping force may crush the specimen being gripped. The wedges
of hydraulic grips are loaded by hydraulic pressure prior to the start of the tensile test. The
magnitude of the hydraulic pressure must be predetermined. Suppliers of commercially
available hydraulic wedge grips provide guidelines for the pressure required as a function
of the anticipated total axial force to be applied by the grips.
For strong materials (large loads), the specimen must be gripped firmly to prevent
slipping, or more aggressive grip faces must be used. In either case, stress concentrations
are induced in the unprotected specimen ends, promoting premature (grip) failures. Thus, it
becomes necessary to use tabs. A sketch of a typical tabbed, straight-sided tensile specimen
is shown in Fig. 8. The length of the tabbed end regions is Lt and the gage length is Lg . The
tabs are commonly tapered where θ is the taper angle. The tabs, however, induce new stress
concentrations because of the relatively abrupt change in specimen thickness at the tab
ends. Thus, the tabs are typically tapered to reduce the abruptness of the thickness change.
ASTM D 303928 does not suggest a specific taper angle, and indicates only that the angle
should be between 5 and 90◦ . The results of a round robin experimental study reported by
Hojo et al.39 suggest that the tensile strength of unidirectional carbon/epoxy composites is
Polymer Matrix Composite Materials 289

Figure 8. Composite tensile test specimen configuration.


Downloaded by [Purdue University] at 06:54 22 October 2015

virtually constant over a range of tab taper angles between 10 and 90◦ . Notice also that
ASTM D 303928 recognizes the possibility of successfully testing untabbed specimens of
lower strength materials such as fabric-reinforced composites.

3.3 Specimen Configurations and Test Procedures


Recommended specimen configurations and dimensions are presented in ASTM D 3039.28
Although considerable detail is presented, these standards recognize and accept that certain
composite materials and test conditions may require modifications. For unidirectional
composites of 0◦ fiber orientation, a specimen width of 12.7 mm and a specimen thickness
of 6 to 8 plies are common, assuming typical ply thickness on the order of 0.1 mm.
Unidirectional 90◦ specimens are typically 25 mm wide and 16 to 24 plies thick. Loading
eccentricity may arise because of variations in tab and specimen thicknesses. Tolerances for
tab and specimen thicknesses are ±1 and ±4%, respectively. The tab length (Fig. 8) should
be at least 38 mm, and the tab material should be 1.6 to 3.2 mm thick. The gage length
(distance between tabs) is commonly 125 to 155 mm. Variations of the specimen width
should not exceed 1%. If Poisson’s ratio is desired, a 0/90 strain gage rosette should be
bonded in the center gage section region of the specimen (Fig. 8) or a biaxial extensometer
should be used. If only axial stiffness and strength are desired, a longitudinal strain gage or
a uniaxial extensometer is sufficient.
Accurate measurement of the specimen cross-sectional area in the gage section is par-
ticularly important, along with careful alignment of the specimen in the grips of the testing
machine. Observers should wear adequate eye protection during the test procedure. Com-
posite materials, and particularly axially-loaded high strength unidirectional composites,
can splinter and fragment violently upon failure.

3.4 Compression Testing


The buckling stability of compression test specimens is an important issue. If the unsup-
ported length of the test specimen is too long, the specimen may fail due to buckling in-
stability, resulting in strength values far below the material strength. When fiber-reinforced
composites containing unidirectional plies oriented in, or at small angles to, the loading
direction are loaded in compression, the fibers may buckle in a small region of the test
290 L. A. Carlsson et al.

section.40 This is followed by the formation of kink zones,41 and failure of the fibers at the
boundaries of the kink zones because of locally large bending stresses (Fig. 7b).
Compression loading of a composite perpendicular to the fibers involves failure of
the matrix and fiber-matrix interface, often exhibited as a shearing type (inclined failure
planes), such as illustrated in Fig. 7d. A detailed discussion of the various compressive
failure modes observed in fiber-reinforced composites has been presented by Fleck42 and
Odom and Adams.43 However, since instability at any scale is dependent upon the support
of load-bearing fibers, and that the fibers in actual composite structures are not fully
constrained, the compression strength will be test configuration dependent and likely not a
material property.
Many compression tests are available. Each method should be judged by its ability to
produce compression failure without introducing loading eccentricity and severe stress con-
centrations at the loaded ends, while avoiding global buckling instability (Euler buckling)
of the specimen. The fulfillment of the criteria makes the determination of the compressive
Downloaded by [Purdue University] at 06:54 22 October 2015

strength difficult.

3.5 Shear-Loading Test Method


The Illinois Institute of Technology Research Institute (IITRI) test fixture is a well-
established standard test procedure for compression testing of various forms of composites44
and is designated as ASTM D3410.45 A schematic of a typical IITRI fixture is shown in
Fig. 9. Its principal features are the use of flat wedge grips, and a pair of alignment rods
and linear bearings. Either untabbed or tabbed specimens are permitted. However, tabbed
specimens are most commonly used (see Fig. 10) to protect the test material from the high
clamping forces of the wedge grips, and surface damage if aggressive grip faces are used.
The flat wedge grips permit variation of specimen thickness within the available range of
movement of the wedges. Typically, the range is between 4 and 5 mm providing consider-
able flexibility in terms of specimen preparation. With the use of interchangeable inserts of
different thicknesses in the upper and lower blocks of the fixture, a broad range of specimen
thicknesses can be accommodated. For example, some commercially available fixtures can
accommodate specimens ranging in thickness from 4 to 15 mm, and obviously a fixture to
test a specimen in any practical thickness range could be designed. The standard specimen
length is 140 mm, with a 12.7 mm gage length (unsupported length). (See Fig. 10.) A
specimen width of 12.7 mm is common, although some fixtures can accommodate a width
up to 38 mm. The alignment rods and linear ball bearings of the IITRI fixture will not bind,
and provide minimal frictional resistance.

3.6 Combined Loading Compression (CLC) Test Methods


The concept of combined load introduction evolved from the desire to eliminate end
crushing of end-loaded compression specimens. A combined loading compression (CLC)
test method was developed at the University of Wyoming.46, 47 This method combines shear
loading and end loading, and is designated as ASTM Standard D 6641.48
The CLC test fixture is outlined in Fig. 11. The shear-loading component is achieved
by clamping pairs of lateral support blocks, which have high-friction contact surfaces, to
each end of the specimen. The end-loading component is induced directly because each
end of the specimen is flush with the outer surfaces of the support blocks. The ratio of shear
loading to end loading can be controlled by varying the torque in the clamping screws.
The goal is to achieve just enough shear loading to avoid end crushing of the specimen.
Polymer Matrix Composite Materials 291
Downloaded by [Purdue University] at 06:54 22 October 2015

Figure 9. IITRI compression test fixture.

Because the fixture grip surfaces contain relatively small tungsten carbide particles, they do
not significantly penetrate and damage the specimen surfaces. Combined with the fact that
the through-the-thickness clamping forces on the specimen do not have to be very high, it

Figure 10. IITRI compression test specimen.


292 L. A. Carlsson et al.
Downloaded by [Purdue University] at 06:54 22 October 2015

Figure 11. Schematic of combined loading compression (CLC) test fixture.

is typically possible to test an untabbed, straight-sided specimen. This simplifies specimen


preparation considerably.
Cross-ply, angle-ply, quasi-isotropic lay-up, and similar laminate forms can all be
readily tested using an untabbed specimen. However, testing untabbed, straight-sided spec-
imens of highly orthotropic, high compressive strength, unidirectional composites using
the CLC fixture does present some problems. Relatively high clamping forces are required
to prevent end crushing, which induces undesirable stress concentrations. Tabs can be
used to increase the end-loading area and thus reduce the clamping force (shear-loading
component) required.
A new combined loading compression test fixture was recently introduced by Bech
et al.49 This fixture employs a unique load introduction mechanism, purely mechanical
including levers and four support columns with linear bearings. The specimen employs
end tabs and steel end cups bonded to the specimen to prevent crushing of the ends. The
load from a test machine is transferred to two levers. In the current design, 34 of the load is
transferred in shear, and 14 by end loading, and this ratio remains independent of the applied
load. This fixture is ideal also for fatigue testing of composites.

3.7 Indirect Determination of Unidirectional Compressive Strength


As discussed above, compression testing of unidirectional composites presents several
challenges due to the high strength in the fiber direction. In order to reduce the strength,
Polymer Matrix Composite Materials 293

Hart-Smith50 proposed compression testing a general laminate containing 0◦ plies, and then
indirectly determining the strength of these 0◦ plies using a suitable analysis. That is, the
compressive strength of a unidirectional ply, X1c , is estimated by multiplying the measured
compressive strength, XC , of the chosen laminate by a factor, the back-out factor (BF).
BF is calculated from the stiffness properties of the unidirectional material and classical
laminated plate theory. The procedure utilizes the following equation,

X1c = BF Xc (13)

Using a simple [0/90]ns cross-ply laminate has significant advantages. The 90◦ plies on the
surfaces protect the immediately adjacent 0◦ plies from fixture- and tabbing-induced stress
concentrations. Of course, as n becomes larger, the importance of protecting these 0◦ plies
next to the surface becomes less important because they become a smaller percentage of
the total number of primary load-bearing plies (i.e., 0◦ plies) present.
The BF for a [0/90]ns cross-ply laminate is given by51
Downloaded by [Purdue University] at 06:54 22 October 2015

1 E1 (E1 + E2 ) − (ν12 E2 )2
BF = 2 (14)
1 (E1 + E2 ) − (ν12 E2 )2
4
where E1 and E2 are the axial and transverse moduli of the unidirectional lamina, re-
spectively, and v12 is the major Poisson’s ratio of the unidirectional lamina. These three
stiffness constants must be determined from axial (E1 , v12 ) and transverse (E2 ) tests of a
unidirectional composite. This is not a difficulty, however, because the composite does not
have to be loaded to high stress levels to determine these stiffness properties.

3.8 Compression Test Procedures


Proper alignment of the testing machine, and of the associated test fixture relative to
the machine is particularly important for compression loading, considering the stability
(buckling) issues involved. Any misalignment of the test specimen can induce bending
and promote failure due to buckling. The use of a spherical seat platen or other alignment
device in the load train is sometimes encouraged to accommodate misalignments. However,
a more positive approach is to achieve proper alignment by careful checking and adjusting
of the test setup prior to beginning testing. Specimen bending and buckling cannot usually
be detected visually during the test, or by microscopic examination of the failed specimen.
As discussed in the various standards, the use of two axially oriented strain gages, mounted
on opposite faces in the gage section, is the only reliable method of detecting bending and
buckling. Although some bending can be tolerated, buckling cannot. The governing ASTM
standards define the amount of bending in a compression specimen as,
ε1 − ε2
By = % bending = ∗ 100 (15)
ε1 + ε2
where ε1 and ε2 are the strain readings from each side of the specimen. By , may be monitored
from the beginning of the test until the specimen fails. By should not exceed 10% at any
load up to failure. The sign of By will provide the direction of bending of the specimen.
The ends of end-loaded and combined-loading specimens must be flat, mutually paral-
lel, and perpendicular to the specimen axis. Specified parallel and perpendicular tolerances
are typically on the order of 0.03 mm. Specimen end flatness is particularly important, so
that the loading is introduced along the central axis of the specimen and local end crushing
294 L. A. Carlsson et al.

is avoided. Uniform specimen thickness from end to end is also important. It is not uncom-
mon for composite panels to have a taper from side to side or end to end, induced by poor
alignment of the curing platens or caul plates during panel fabrication. This non-uniformity
can result in loading misalignments when specimens are subsequently tested. Uniformity
of specimen thickness from end to end within 0.06 mm is commonly specified.
Whenever specimen tabs are used, many additional sources of error exist. Just as for
specimen thickness, variations in thickness of any of the tabbing strips can occur, along
with variations in thickness of the adhesive bond lines. Not just end-to-end variations,
but also differences in thickness of the tab and adhesive components on one side of the
specimen relative to those on the other, are important. Symmetry must be maintained to
achieve uniform axial loading.
In all cases the specimen should fail in the gage section. Unfortunately, failure often
occurs at the very end of the gage section. This is due to the unavoidable stress concentrations
induced in these regions by the tabs (if used), grips, or lateral supports. Thus, although
Downloaded by [Purdue University] at 06:54 22 October 2015

undesirable, such failures are commonly accepted.

3.9 Shear Testing


A shear test of a composite material is performed to determine the shear modulus or shear
strength, or both. Ideally, both properties can be determined from the same test, but this is
not always the case. In addition, the shear response of a composite material is commonly
nonlinear, and full characterization thus requires generating the entire shear stress–strain
curve to failure. However, many shear test methods are not capable of generating the entire
curve, and sometimes not even a portion of it. Figure 6c defines the in-plane shear stress,
τ12 and shear strain, γ12 . The in-plane shear modulus is denoted by G12 , and the shear
strength byS6 .
The major deficiency of all existing shear test methods for composite materials is the
lack of a pure and uniform state of shear stress in the test section. Thus, compromises have
to be made.

3.10 Iosipescu Shear Test Method


The Iosipescu shear test method and specimen configuration shown in Fig. 12 are based
on the original work with metals by Iosipescu,52 from which the test method derives its
name. Adams and Walrath53, 54 extended the test method to composite materials, which is
now ASTM standard D 5379.55 Analysis of the specimen under load reveals that a state
of uniform shear stress exists in the center of the notched specimen on the cross section
through the notches, although not in the immediate vicinity of the notch roots.56 In addition,
the normal stresses (the nonshear stresses) are low everywhere on this cross section. By
orienting the longitudinal axis of the specimen along any one of the three axes of material
orthotropy, any one of the six shear stress components, representing the three independent
shear stress components (Fig. 2) can be developed. For in-plane shear testing, a specimen
with a 0◦ orientation (fibers parallel to the long axis of the specimen) forms a much more
robust specimen and is strongly preferred over a 90◦ orientation. A [0/90]ns (cross-ply)
specimen is even more robust. Because there is no shear coupling between the plies of a
[0/90]ns laminate, this orientation will theoretically produce the same shear properties as
those of a unidirectional composite. In practice, it is likely to produce shear strengths closer
to the ideal shear strength of the composite material because premature failures are less
Polymer Matrix Composite Materials 295
Downloaded by [Purdue University] at 06:54 22 October 2015

Figure 12. Iosipescu test fixture (a) and test specimen (b). The fiber direction should be aligned with
longitudinal axis of the specimen for inplane shear testing.

likely to occur. That is, the cross-ply laminate is likely to produce more accurate (and in
this case higher) shear strengths.
The Iosipescu specimen is shown in Fig. 12b. The top and bottom edges must be
carefully machined to be flat, parallel to each other, and perpendicular to the faces of the
specimen, to avoid out-of-plane bending and twisting when the load is applied. The standard
fixture, shown in Fig. 12a can accommodate a specimen thickness up to 12.7 mm, although
a thickness on the order of 2.5 mm is commonly used.
If shear strain is to be measured, a two-element strain gage rosette with the elements
oriented ±45◦ relative to the specimen longitudinal axis can be attached to the central test
section region, such as shown in Fig. 12b. A single-element gage oriented at either 45◦ or
−45◦ can also be used although this is not common practice. If out-of-plane bending and
twisting of the specimen are a concern, back-to-back strain gages can be used to monitor
these undesired effects. However, this is normally not necessary.
The (average) shear stress across the notched section of the specimen is calculated
using the simple formula

τ = P /A (15)
296 L. A. Carlsson et al.

where P is the applied force, and A is the cross-sectional area of the specimen between
the notches. For a unidirectional composite specimen tested in the 0◦ orientation, detailed
stress analyses56, 57 indicate that an initially nonuniform elastic stress state exists. However,
if any inelastic material response occurs, and particularly if there is initiation and arrested
propagation of a crack parallel to the reinforcing fibers at each notch tip (which will occur
well before the ultimate loading is attained), the local stress concentrations are significantly
relieved. The stress distribution then becomes even more uniform across the entire cross
section of the specimen.
Shear strain, γ12 , is obtained from the ±45◦ strain gage readings,

γ12 = |ε(45◦ )| + |ε(−45◦ )| (16)

If only a single-element gage mounted at plus or minus 45◦ is used,


Downloaded by [Purdue University] at 06:54 22 October 2015

γ12 = 2|ε(45◦ )| (17)

The shear modulus, G12 , is obtained as the initial slope of the shear stress–shear strain
curve.
Premature damage in the form of longitudinal matrix cracks initiating from the notch
roots is a common occurrence in 0◦ unidirectional specimens. Load drops are observed at
about two thirds of the eventual ultimate load when these cracks initiate and propagate,
but they quickly arrest and the specimen then carries additional load until the shear failure
occurs, which involves multiple matrix cracks parallel to the fibers concentrated in the
region of the specimen between the two notches. The ultimate stress carried by the test
section is the shear strength, S6 .
The upper half of the test fixture is loaded in compression through a suitable adapter,
attaching it to the crosshead of the testing machine. The applied load and strain signals are
monitored until the specimen fails.

3.11 Two-Rail Shear Test Method


Both the two- and three-rail shear tests are included in the same ASTM Standard D 4255.58
The commonly used tensile-loading version of the two-rail shear test fixture is shown
schematically in Fig. 13a. Note that the specimen is loaded at a slight angle relative to its
axis. An essentially pure shear loading is applied to the gage section of the specimen. The
shear stress along the length of the specimen (parallel to the rails) is relatively constant,
except near the ends. The two-rail shear test specimen is shown in Fig. 13b. The specimen
is quite large, 76.2 × 152.4 mm. For this and other reasons, the rail shear test methods are
used somewhat less frequently than other shear test methods.
For testing unidirectional composites, the fibers can be oriented either parallel (90◦
orientation) or perpendicular (0◦ orientation) to the rails. However, a fiber orientation
perpendicular to the rails is much preferred to minimize bending and edge effects. Testing
a [0/90]ns (cross-ply) laminate may be even better; the specimen is even more robust than
a unidirectional lamina with fibers oriented perpendicular to the rails. Cross-ply laminates
tend to produce higher shear strengths of the composite material because premature failures
are less likely to occur and have less influence on the response even if they occur.
For measuring the shear strain, a single-element or a dual-element strain gage is bonded
to the specimen at the center of the gage section, see Fig. 13b. The calculations of shear
Polymer Matrix Composite Materials 297
Downloaded by [Purdue University] at 06:54 22 October 2015

Figure 13. Rail shear (two-rail) test fixture (a) and test specimen (b).

stress and shear strain are the same as for the Iosipescu specimen (Eqs. 15–17). The cross-
sectional area in the present case is the length of the specimen parallel to the rails times the
specimen thickness.
The specimen to be tested is clamped between the pairs of rails, as indicated in
Fig. 13a. It is important that the rails do not slip during the test. When a two-rail shear test
is conducted, a large clamping force is typically needed. If the rails slip, the clamping bolts
can bear against the clearance holes in the specimen, inducing local stress concentrations
leading to premature failure and unacceptable test results. Most currently commercially
available fixtures have thermal-sprayed tungsten carbide particle gripping surfaces.59

3.12 [±45]ns Tensile Shear Test Method


The [±45]ns tensile coupon shown in Fig. 14 can be employed to determine the shear
properties of the unidirectional ply in the principal material coordinate system.60 The test
specimen is relatively simple to prepare and requires no special test fixture other than
standard tensile grips. The test method is ASTM Standard D 3518.61 The state of stress
in each lamina of the [±45]ns laminate is not pure shear. Each lamina contains tensile
normal stresses, σ1 and σ2 , in addition to the desired shear stress, τ12 , see Fig. 14. In
addition, an interlaminar shear stress, τxz , is present near the laminate free edge. Normally,
these considerations would lead to the rejection of this test geometry. However, there are
mitigating circumstances that reduce these concerns. First, the shear stress–strain responses
of many types of composite laminate are nonlinear, and may exhibit strain softening
characteristics. Thus, although the biaxial state of stress present in the specimen likely
reduces the measured value of shear strength. The reduction may be small because of
the nonlinear softening response. Second, the interlaminar stresses are small for 45◦ ply
angles. Therefore, the [±45]ns tensile shear test method is an appropriate and a simple test
for determining the shear modulus and strength for the ply.
298 L. A. Carlsson et al.
Downloaded by [Purdue University] at 06:54 22 October 2015

Figure 14. [±45]ns tensile test specimen for measuring in-plane shear response.

The test specimen geometry is shown in Fig. 14. The width of the specimen typically
is on the order of 25 mm. End tabs may not be required because the tensile strength of a
[±45]ns laminate is small due to the absence of 0◦ plies. The specimen is instrumented with
a 0◦ /90◦ biaxial strain gage rosette as shown in Fig. 14. The specimen is tested in tension
to ultimate failure following the procedures outlined for the tension test.
Determination of the shear stress and strain in the principal planes (x1 x2 ) of the ± 45◦
plies, Fig.14, is based on a stress analysis of the [±45]ns specimen.60 The in-plane shear
stress, τ12 is simply

τ12 = σx /2 (18)

where σx is the axial stress (P/A). The shear strain is

γ12 = εx − εy (19)

where εx and εy are the axial and transverse strains, respectively (εy < 0). Hence, the in-
plane shear modulus,G12 , is readily determined by plotting σ2x vs. (εx − εy )and establishing
the slope of the initial portion of the curve. The shear strength, S6 , is defined as the maximum
value of σ2x .

3.13 Short-Beam Shear Test Method


The short-beam shear test, ASTM D2344,62 is an interlaminar shear test method. The
specimen utilizes a [0]n specimen where n is the number of plies which should be sufficient
to introduce enough of the shear load into the specimen. Figure 15 shows the short-beam
shear fixture and test specimen. The short beam shear test is used extensively as a materials
screening and quality control test because the specimen is small and easy to prepare, and
the test is both quick and economical since no strain or displacement measurements are
made.
Polymer Matrix Composite Materials 299

Figure 15. Short-beam shear test. The fiber direction of the specimen should be aligned with the
longitudinal axis of the specimen.

A specimen with a low support span length-to-specimen thickness ratio (typically a


Downloaded by [Purdue University] at 06:54 22 October 2015

ratio of 4) is subjected to three-point loading. The specimen can be very small, consuming a
minimal amount of material. For example, when a span length-to-specimen thickness ratio
of 4 and a 2.5-mm-thick composite are used, the span length is only 10 mm. If one specimen
thickness of overhang is allowed on each end, the total specimen length is still only 15 mm.
In addition, minimum specimen preparation time is required because the length and width
dimensions, and the quality of the cut edges of the specimen, are not critical.
Both bending (flexural) and interlaminar shear stresses are induced. The axial bending
stresses are compressive on the top surface of the beam where the load is applied, and
tensile on the opposite surface. By definition, the neutral axis (neutral plane) is where the
bending stress passes through zero. It is on this neutral plane that the interlaminar shear
stress is theoretically at maximum, varying parabolically from zero on each surface of the
beam. Thus, although a combined stress state exists in general, the stress state should be
pure shear on the neutral plane. By keeping the span length-to-specimen thickness ratio
low, the bending stresses can be kept low, promoting shear failures on the neutral plane.
Unfortunately, the concentrated loadings on the beam at the loading and support points
create stress concentrations throughout much of the short beam63 and a complicated stress
state. As a result, the assumption of a parabolic stress distribution with a maximum at the
neutral plane becomes valid only in regions between the loading and support locations.
Nevertheless, the ASTM standard advocates the following beam theory formula for the
maximum shear stress

τ13 = 0.75P /A (20)

where P is the applied load on the beam, and A is the cross-sectional area of the beam.
This assumption is the reason for the use of “apparent” in the previous title of ASTM D
2344. Whitney and Browning63 found that standard (16 ply) carbon/epoxy specimens did
not fail in an interlaminar mode. The specimens failed in compression near the central load
introduction. Thicker (50 ply) carbon/epoxy specimens, however, failed in interlaminar
shear, not at the center, but rather in the upper quarter of the specimen. Figliolini and
Carlsson64 observed shear failures in carbon/vinylester SBS specimens, although some
specimens failed in compression near the load introduction, evidenced by a kink band
formation. If, however, the failure occurs in an interlaminar failure mode at the mid-plane,
the short beam shear test method usually produces reasonable values of shear strength.65
Modifications of the short-beam shear test have been proposed in order to promote shear
failure at the mid-plane. Rodriguez and Aviles66 developed a sandwich beam consisting of a
300 L. A. Carlsson et al.

thin polymer layer sandwiched between two thick aluminum adherends. Other researchers
suggest the use of four-point bending to reduce the high bending stresses in the specimen.67

Acknowledgements
Thanks are due to Melissa Mancao and Mark S. Hoerber Jr. for preparing the text and
Shawn Pennell for the graphics works.

References
1. Advani, S.G. Flow and Rheology in Polymer Composites Manufacturing, in the series, Composite
Materials, R.B. Pipes, Ed., Elsevier, Amsterdam, 1994; pp 465–511.
2. Daniel, I.M.; Ishai, O. Engineering Mechanics of Composite Materials, 2nd Ed.; Oxford Univer-
sity Press: New York, 2006.
Downloaded by [Purdue University] at 06:54 22 October 2015

3. Boeing Specification Support Standard BSS 7260, Advanced Composite Compression Test. The
Boeing Company, Seattle (originally issued February 1982, revised December 1985).
4. ASTM Annual Book of Standards, Vol.15.03, Space Stimulation; Aerospace and Aircrafts Com-
posite Materials, American Society for Testing and Materials, West Conshohocken, PA, 2010.
5. Tarnopolskii, Y.M.; Kincis, T.Y. Static Test Methods for Composites; Van Nostrand Reinhold:
New York, 1985.
6. Adams, D.F.: Carlsson, L.A.: Pipes, R.B. Experimental Characterization of Advanced Composite
Materials, 3rd Ed.; CRC Press: Boca Raton, FL, 2003.
7. Timoshenko, S.P.; Goodier, J.N. Theory of Elasticity, 3rd Ed.: McGraw-Hill: New York, 1970.
8. Hashin, Z. Analysis of composite materials-a survey. J. Appl. Mech. 1983, 50, 481–505.
9. Dally, J.W.; Riley, W.F. Experimental Stress Analysis, 3rd Ed.; McGraw-Hill: New York, 1991.
10. Tuttle, M.E.; Brinson, H.F. Resistance-foil strain gage technology as applied to composite mate-
rials. Exp. Mech. 1984, 23, 54–65.
11. Masters, J.E.; Ifju, P.G. Strain gage selection criteria for textile composite materials. J. Compos.
Technol. Res. 1997, 19, 152–167.
12. Guide to Advance Materials Testing, Product Catalog, Instron Corporation: Canton, MA, 1997.
13. Material Testing Products, 7th Ed., MTS Systems Corporation: Eden Prairie, MN, 1999.
14. Version 104 Extensometer Catalog, Epsilon Technology Corporation, Jackson, WY, 1997.
15. Linear & Angular Displacement Transducers, Catalog No. 101, Lucas Schaevitz, Inc.,
Pennsauken, NJ, 1990.
16. Sutton, M.A.; Orteu, J.J.; Schreier, H.W. Image Correlation for Shape, Motion and Deformation
Measurements; Springer: New York, 2009.
17. Kamaya, M.; Kawakabo, M.A. A procedure for determining the true stress-strain curve over a
large range of strains using digital image correlation and finite element analysis. Mech. Mater
2011, 43, 243–253.
18. Taher, S.T.; Thomsen, O.T.; Dulieu-Barton, J.M.; Zhang, S. Determination of mechanical prop-
erties of PVC foam using a modified Arcan fixture, Compos. A (in press).
19. Carlsson, L.A.; Adams, D.F.; Pipes, R.B. Experimental Characterization of Advanced Compos-
ites Materials, 4th Ed.; CRC Press: Boca Raton (in press).
20. Sun, C.T.; Li, S. Three-dimensional effective elastic constants for thick laminates. J. Compos.
Mater. 1988, 22, 629–639.
21. Hyer, M.W.; Knott, T.W. Analysis of End-Fitting Induced Strains and Axial Loaded Glassy-Epoxy
Cylinders, Composites Structs for SMES Plants (NISTIR5024), (R.P. Reed; J.D. McColskey,
Eds.), Oct., 1994.
22. Hyer, M.W. Stress Analysis of Fiber-Reinforced Composite Materials; Updated Edition, DES
Tech Publications: Lancaster, PA, 2009.
23. Horgan, C.O. Some remarks on Saint-Venant’s principle for transversely isotropic composites.
J. Elasticity 1972, 2, 335–339.
Polymer Matrix Composite Materials 301

24. Horgan, C.O. Saint-Venant end effects in composites. J. Compos. Mater. 1982, 16, 411–422.
25. Choi, I.; Horgan, C.O. Saint-Venant’s principle and end effects in anisotropic elasticity. J. Appl.
Mech. 1997, 44, 424–430.
26. Horgan, C.O.; Carlsson, L.A. Saint-Venant end effects for anisotropic materials. In Compre-
hensive Composite Materials, A. Kelly; C. Zweben Eds., Vol. 5, 2000, Elsevier: Oxford; pp
5–21.
27. Arridge, R.G.C.; Barham, P.I.; Farell, C.J.; Keller, A. The importance of end effects in
the measurement of moduli of highly anisotropic materials. J. Mater. Sci. 2001, 11, 788–
790.
28. ASTM Standard D 3039, Standard Test Method for Tensile Properties of Polymer Matrix Com-
posite Materials, American Society for Testing and Materials, West Conshohocken, PA, 2008.
29. Bogetti, T.A.; Gillespie, J.W., Jr.; Pipes, R.B. Evaluation of the IITRI compression test method
for stiffness and strength determination. Compos. Sci. Tech. 1988, 32, 57–76.
30. Abrate, S.; Walton, D.A. Machining of Composite Materials. Part I: Traditional, Methods, Com-
pos. Manuf. 1992, 3, 75–83.
31. Srivatsan, T.S.; Lane, C.T.; Bowden, D.M.; Eds., Machining of Composite Materials II, ASM
Downloaded by [Purdue University] at 06:54 22 October 2015

International, Materials Park, OH, 1994.


32. Komanduri, R. Machining fiber-reinforced composites, Mech. Eng. 1993, 4, 58–63.
33. Cunningham, M.E.; Schoultz, S.V.; Toth Jr., J.M. Effect of End-Tab Design on Tension Speci-
men Stress Concentrations, ASTM STP 864, American Society for Testing and Materials, West
Conshohocken, PA, 1985; pp 253–262.
34. Adams, D.F.; Odom, E.M. Influence of specimen tabs on the compressive strength of a unidirec-
tional composite material. J. Compos. Mater. 1991, 25, 774–786.
35. Hart-Smith, L.J. Generation of Higher Composite Materials Allowables Using Improved Test
Coupons, Proceedings of the 36th International SAMPE Symposium, Anaheim, CA, Society
for the Advancement of Material and Process Engineering, Covina, CA, 1991; pp 1029–
1044.
36. Pipes, R.B.; Pagano, N. J. Interlaminar stresses in composite laminates – An approximate elas-
ticity solution. J Appl. Mech. 1974, 41, 668–672.
37. Flaggs, D.L.; Kural, M.H. Experimental determination of the in-situ transverse lamina strength
in graphite/epoxy laminates. J. Compos. Mater. 1982, 16, 103–116.
38. Crossman, F.W.; Wang, A.S.D. The Dependence of Transverse Cracking and Delamination on
Ply Thickness in Graphite/Epoxy Laminates, ASTM STP 775, 1982; pp 118–139.
39. Hojo, M.; Sawada, Y.; Miyairi, H. Influence of Clamping Method on Tensile Properties of Unidi-
rectional CFRP in 0◦ and 90◦ Directions-Round Robin Activity for International Standardization
in Japan. Composites 1994, 25, 278–283.
40. Rosen, B.W. Mechanics of composite strengthening. In Fiber Composite Materials, ASTM
International, Metals Park, OH, 1965.
41. Evans, A.G.; Adler, W.F. Kinking as a mode of structural degradation in carbon fiber composites.
Acta Metall. 1978, 26, 725–738.
42. Fleck, N.A. Compressive failure of fiber composites Adv. Appl. Mech. 1977, 33, 43–117.
43. Odom, E.M.; Adams, D.F. Failure modes of unidirectional carbon/epoxy composite compression
specimens Composites 1990, 21, 289–296.
44. Hofer, Jr., K.E.; Rao, P.N. A new static compression fixture for advanced composite materials.
J. Test. Eval. 1997, 5, 278–283.
45. ASTM Standard D 3410, Standard Test Method for Compressive Properties of Polymer Matrix
Composite Materials with Unsupported Gage Section by Shear Loading, American Society for
Testing and Materials, West Conshohocken, PA, 2008.
46. Adams, D.F.; Welsh, J.S. The Wyoming Combined Loading Compression (CLC) Test Method.
J. Compos. Tech. Res. 1997, 19, 123–133.
47. Wegner, P.M.; Adams, D.F. Verification of the combined load compression (CLC) test method,
Report No. DOT/FAA/AR-00/26, Federal Aviation Administration Technical Center, Atlantic
City, NJ, 2000.
302 L. A. Carlsson et al.

48. ASTM Standard D 6641, Standard Test Method for the Compressive Properties of Polymer Matrix
Composite Laminates Using a Combined Loading Compression (CLC) Test Fixture, American
Society for Testing and Materials, West Conshohocken, PA, 2009.
49. Bech, J.I.; Goutianos, S.; Anderson, T.L.; Torekov, R.K.; Brondsted, P. A new static and fatigue
compression test method for composites. Strain 2001, 47, 21–28.
50. Hart-Smith, L.J. Backing-out composite lamina strengths from cross-ply testing. In Compre-
hensive Composite Materials; A. Kelly; C. Zweben, Eds., Vol. 5; Elsevier: Oxford, 2000; pp
149–161.
51. Welsh, J.S.; Adams, D.F. Unidirectional composite compression strengths obtained by testing
cross-ply laminates. J. Compos. Technol. Res. 1996, 18, 241–248.
52. Iosipescu, N. New accurate procedure for single shear testing of metals. J. Mater. 1967, 2,
537–566.
53. Adams, D.F.; Walrath, D.E. Iosipescu Shear Properties of SMC Composite Materials, ASTM STP
787, American Society for Testing and Materials; West Conshohocken, PA, 1982; pp 19–33.
54. Walrath, D.E.; Adams, D.F. Verification and Application of the Iosipescu Shear Test Method,
Report No. UWME-DR-401–103–1, Department of Mechanical Engineering, University of
Downloaded by [Purdue University] at 06:54 22 October 2015

Wyoming Laramie, WY, 1984.


55. ASTM Standard D 5379, Standard Test Method for Shear Properties of Composite Materials by
the V-Notched Beam Method, American Society for Testing and Materials, West Conshohocken,
PA, 2005.
56. Ho, H.; Morton, J.; Farley, G.I. Non-linear numerical analysis of the Iosipescu specimen for
composite materials. Compos. Sci. Tech. 1994, 50, 355–365.
57. Morton, J.; Ho, H.; Tsai, M.Y.; Farley, G.I. An evaluation for the Iosipescu specimen for com-
posite materials shear property measurement. J. Compos. Mater. 1992, 26, 708–750.
58. ASTM Standard D 4255, Standard Test Method for In-Plane Shear Properties of Polymer Matrix
Composite Materials by the Rail Shear Method, American Society for Testing and Materials,
West Conshohocken, PA, 2007.
59. High Performance Test Fixtures, Product Catalog No. 106, Wyoming Test Fixtures, Inc., Laramie,
WY. (also www.wyomingtestfixtures.com), 2000.
60. Rosen, B.W. A simple procedure for experimental determination of the longitudinal shear mod-
ulus of unidirectional composites. J. Compos. Mater. 1972, 6, 552–554.
61. ASTM Standard D 3518, Standard Test Method for In-Plane Shear Response of Polymer Matrix
Composite Materials by Tensile Test of ±45◦ Laminate, American Society for Testing and
Materials, West Conshohocken, PA, 2001.
62. ASTM Standard D 2344, Standard Test Method for Short Beam Strength of Polymer Matrix
Composite Materials and their Laminates, American Society for Testing and Materials, West
Conshohocken, PA, 2006.
63. Whitney, J.M.; Browning, C.E. On short-beam shear tests for composite materials. Exper. Mech.
1985, 25, 294–300.
64. Figliolini, A.M.; Carlsson, L.A. Mechanical properties of carbon fiber/vinylester composites
exposed to marine environments, Polym. Compos. 2013, (in press)
65. Adams, D.F.; Lewis, E.Q. Experimental assessment of four composite material shear test methods.
J. Test. Eval., 1997, 25,174–181.
66. Rodriguez, J.A.; Aviles, F. A modified short-beam shear specimen for characterization of inter-
facial strength in nanocomposites. Polym. Test, 2012, 31, 792–799.
67. Browning, C.E.; Abrams, F.L.; Whitney, J.M. A Four-Point Shear Test for Graphite/Epoxy
Composites, ASTM STP 797, American Society for Testing and Materials: West Consohocken,
PA, 1983; pp 54–74.

View publication stats

You might also like