Sie sind auf Seite 1von 34

7.

2 Test Methods for Mechanical Properties


Donald F Adams, University of Wyoming, Laramie, WY, United States
Thomas J Whitney, University of Dayton, Dayton, OH, United States
r 2018 Elsevier Ltd. All rights reserved.

7.2.1 Introduction 5
7.2.1.1 Goals of Mechanical Testing 5
7.2.1.2 Types of Tests Typically Required 5
7.2.1.3 Unique Aspects of Testing Composite Materials 6
7.2.1.4 Evolution of Mechanical Testing Methodologies 7
7.2.1.5 Test Method Selection Criteria 8
7.2.2 General Test Equipment Requirements 9
7.2.2.1 Types of Testing Machines Available 9
7.2.2.2 Electromechanical Versus Servo-Hydraulic Machines 9
7.2.2.3 Hydraulic Versus Mechanical Wedge Grips 10
7.2.2.4 Other Types of Grips 10
7.2.2.5 Types of Grip Faces 10
7.2.3 Test Specimens 10
7.2.3.1 Specimen Preparation 11
7.2.3.1.1 Cutting procedures 11
7.2.3.1.2 Dimensional measurements 11
7.2.3.1.3 Specimen quality issues 11
7.2.3.2 Specimen Conditioning and Storage Prior to Testing 11
7.2.4 Measurement of Stresses and Strains 12
7.2.4.1 Stresses 12
7.2.4.2 Strains 12
7.2.4.2.1 Strain/displacement measurement 12
7.2.4.2.2 Extensometers versus strain gages 13
7.2.4.2.3 Digital image correlation 13
7.2.5 Tension Test Methods 14
7.2.6 Compression Test Methods 15
7.2.6.1 Problems Unique to Compression Testing 15
7.2.6.2 ASTM and Other Standard and Nonstandard Test Procedures 16
7.2.6.3 Specimen Loading Configurations 16
7.2.6.3.1 End loading versus shear-loading test methods 17
7.2.6.3.2 Combined loading 17
7.2.6.3.3 Tabbed versus untabbed specimens 18
7.2.6.4 Specific Test Fixtures 18
7.2.6.4.1 Shear-loading fixtures 18
7.2.6.4.2 End-loading fixtures 19
7.2.6.4.3 Sandwich specimens 20
7.2.6.4.4 Combined loading fixtures 21
7.2.6.5 Strain Measurement Instrumentation 23
7.2.7 Shear Test Methods 23
7.2.7.1 Problems Unique to Shear Testing 23
7.2.7.2 ASTM and Other Standards 23
7.2.7.3 Specimen Loading Configurations 23
7.2.7.3.1 Torsional loading 23
7.2.7.3.1.1 Thin-walled tubes 23
7.2.7.3.1.2 Solid rods 24
7.2.7.3.2 Direct shear 24
7.2.7.3.2.1 Iosipescu shear 24
7.2.7.3.2.2 Rail shear 24
7.2.7.3.2.3 Double-notched shear 25
7.2.7.3.3 Induced shear 25
7.2.7.3.3.1 þ 45 Degree laminate tensile shear 25
7.2.7.3.3.2 Short beam shear 27
7.2.7.4 Strain Measurement Instrumentation 28

4 Comprehensive Composite Materials II, Volume 7 doi:10.1016/B978-0-12-803581-8.10032-3


Test Methods for Mechanical Properties 5

7.2.7.4.1 Iosipescu shear 28


7.2.7.4.2 Thin-walled tube and solid rod torsion 28
7.2.7.4.3 745 Degree tensile shear 28
7.2.7.4.4 Two- and three-rail shear 28
7.2.7.4.5 Short beam shear and notched shear 28
7.2.8 Flexure Test Methods 28
7.2.8.1 Problems Unique to Flexure Testing 29
7.2.8.1.1 Solid laminates 29
7.2.8.1.2 Sandwich beams 29
7.2.8.2 ASTM and Other Standards 29
7.2.8.3 Strain/Displacement Measurement Instrumentation 29
7.2.9 Multiaxial Loading Test Methods 30
7.2.9.1 Biaxial Testing 30
7.2.9.2 Triaxial Testing 30
7.2.9.3 Strain Measurement Instrumentation 31
7.2.10 Fracture Mechanics Test Methods 31
7.2.10.1 Historical Introduction 31
7.2.10.2 Fracture Mechanics Principles 31
7.2.11 Nonambient Testing Conditions 32
7.2.11.1 Common Problems/Precautions 32
7.2.11.2 Elevated Temperature Testing 34
7.2.11.3 Subambient Temperature Testing 34
7.2.11.4 Testing Moisture-Conditioned Specimens 34
7.2.12 Conclusions 34
References 35
Further Reading 37
Relevant Websites 37

7.2.1 Introduction

Mechanical testing can be very time-consuming and also very expensive. Thus only as much testing as is necessary to achieve
carefully defined objectives should be conducted, and it is important that it be conducted properly, the first time, so that costs are
not compounded. The goal of this section is to offer guidelines to aid in achieving carefully planned and properly executed test
programs.
The formulas required to calculate stresses from measured loads for the various tests are typically very simple and straightforward,
and thus are not included here. However, complete references to sources of detailed information are included wherever appropriate.

7.2.1.1 Goals of Mechanical Testing


There are at least two primary goals of mechanical testing of composite materials, viz., to obtain preliminary properties data for
comparison purposes, and to generate detailed design data.
New classes of composite materials are still emerging, and improvements of existing classes are still being made on a regular
basis as well. Mechanical testing is required to preliminarily characterize new materials, to permit initial comparisons of them with
existing competitors. Likewise, it is necessary to determine the changes made to existing systems, to verify that in fact that the
desired improvements have been made. Since these types of basic properties measurements are for materials comparison purposes,
and may have to be repeated many times as new developments emerge, the tests utilized should be as straightforward and
economical as practical, and require relatively simple test specimen configurations.
In contrast, when an adequately developed composite material system is ready for use in the actual design of structural
components, it must be thoroughly characterized. Sometimes this even involves a formal “material qualification” procedure, as
defined by the appropriate governing body, for example, the military, Federal Aviation Administration (FAA), Automotive
Composites Consortium, or some similar agency. Whether it is a formal material qualification or not, a major amount of testing
typically is required to characterize all of the properties required for design purposes.

7.2.1.2 Types of Tests Typically Required


The basic mechanical properties of materials in general can be determined by applying tensile, compressive, and shear loadings.
Flexure testing is also often included, although it is actually a combination of these three basic loading modes. The FAA has
recently published a three-volume report that provides an excellent summary of historical research work that has been performed
6 Test Methods for Mechanical Properties

worldwide related to each of these three basic loading modes.1 Another excellent publication is that edited by Jenkins.2 A text by
Traceski3 provides cross-indexing between many of the governing international standards and specifications. The appendices in the
backs of the annual books of ASTM standards provide even more up-to-date cross-indexing since they are issued each year.
Static material properties are usually of primary interest. These represent the response of the material to loadings applied
monotonically to failure over time intervals typically measured in tens of seconds or at most a few minutes.
Impact loadings may be applied in the same manner, i.e., as tensile, compressive, shear, or flexural loadings, but over much
shorter time intervals, typically in durations of a few milliseconds. Thus, inertial (material mass) effects can play a role, unlike for
static loadings.
Fatigue loadings are essentially static loadings, but to only some specified fraction of ultimate, and repetitively applied. For
metals, a cyclic rate of 30 cycles per second (Hz) is common. For composite materials, relatively low cyclic loading rates of 5–10
Hz are typical, not just to simulate actual in-service loading rates, but at least equally important, to minimize hysteretic (internal
friction or self-induced) heating effects within the composite material. Hysteretic heating is a particular concern when fatigue
testing composite materials as significant internal friction (damping) is a typical characteristic, and heat transfer rates are low. In
contrast, most structural metals exhibit low damping and relatively high heat dissipation. By keeping the cyclic rate low when
fatigue testing composite materials, the amount of heat generated per unit time is less and the time per cycle available for heat
dissipation is increased.
Static loadings to some fraction of ultimate, but held for long periods of time, viz., creep loadings, are also of interest. Time
durations under load of weeks and months are typical, shorter durations of tens of minutes to several days are of frequent interest,
and longer durations of many months and even years are the ultimate indication of creep response, but may not be practical.
Superimposed on these basic loading modes and rates are variations in test temperature, specimen preconditioning (moisture
or solvent exposure, salt spray, sunlight or other irradiation, aging, etc.), combined (multiaxial) loadings, and geometric effects
(holes, cutouts, local impact damage, and other material discontinuities).
The above discussion relates to mechanical properties in general, i.e., the response of the composite material to the application
of a mechanical loading. Fatigue, creep, impact, environmental exposure, and other “durability” tests are discussed in detail in
Section 7.2.8, the present section focusing on static mechanical properties. Certain physical properties are also needed for design
and analysis purposes. Perhaps the most important of these to the mechanical designer and stress analyst are the thermal and
moisture expansion coefficients, i.e., the dimensional change of the material when subjected to changes in temperature and
moisture. These are discussed in Section 7.2.9.

7.2.1.3 Unique Aspects of Testing Composite Materials


Composite materials are typically orthotropic, i.e., they have different stiffnesses and strengths in different directions. However,
orthotropy (as opposed to anisotropy) also implies that some order exists. In fact, most man-made composites are specially
orthotropic; their complete array of properties can be determined most simply by making measurements in three specific
orthogonal (mutually perpendicular) directions, the principal directions. However, these directions may not coincide with the
(local and/or global) coordinate directions defined by the designer of the structural component (the geometric coordinates),
which are commonly defined as x, y, and z directions. Thus, to distinguish the material coordinate directions, they are con-
ventionally defined as the 1, 2, 3 directions. If property measurements are not made in the principal (1, 2, 3) directions, the data
reduction is significantly complicated by the presence of additional coupling (property interaction) terms. Thus, when char-
acterizing a composite material, it is both conventional and very advantageous to apply loadings in the principal material
directions. Further details relating to material orthotropy can be found in excellent text books by Gibson,4 Daniel and Ishai,5
Hyer,6 Herakovich,7 and Jones.8
Since most structural components are relatively thin, i.e., essentially two-dimensional (2D) on a local scale, the so-called in-
plane material properties are of primary interest. Of the two in-plane orthogonal directions that exhibiting the higher axial stiffness
is conventionally defined as the 1-direction (the axial direction), the 2-direction (the transverse direction) being in the perpen-
dicular in-plane direction. By definition of mutual orthogonality, the 3-direction is then perpendicular to the plane of the material,
the positive direction being as required to define a right-hand coordinate system.
The measured (principal) material properties can be transformed, using well-known transformation relations, to any other
coordinate system, the geometric coordinates x, y, and z in particular.4–8 If the material 3-direction (the thru-thickness direction)
coincides with the geometric z-direction, the coordinate transformation is particularly straightforward, being simply a rotation
about the 3-(z-) axis.
As evidenced by the above discussion, the properties of a composite material are usually measured in the 1, 2, and 3 directions.
The axial normal (as opposed to shear) properties in tension and compression will not in general be equal. Thus six axial normal
loadings are required. These loadings are typically termed axial (1-direction), transverse (2-direction), and thru-thickness or
interlaminar (3-direction) tension and compression.
To fully define a stress, which is a second-rank tensor, a double subscript notation can be used.4,7,8 The first subscript indicates
the plane on which the stress is acting. (A plane is defined geometrically by the direction of its normal. Thus a plane normal to the
1-direction is the 1-plane.) The second subscript indicates the direction of loading. Thus a loading normal to the 1-plane is in the
1-direction. The corresponding stress is a normal stress (i.e., a tensile or compressive stress as opposed to a shear stress), and is
designated as s11. The corresponding normal stresses in the other two principal directions are s22 and s33, respectively, i.e., the two
Test Methods for Mechanical Properties 7

subscripts for a normal loading are always identical. This has led to the common use of a so-called contracted notation, whereby
only one subscript is used for normal stresses.8
The above discussion applies to normal strains (also second-rank tensors) as well, e1 being used more commonly than e11, etc.8
Likewise, the common notations for the three elastic stiffnesses, which, following conventional terminology, are taken to mean
the normal stiffnesses in each of the three principal directions, are E1, E2, and E3. Sometimes these are alternately expressed as E11,
E22, and E33 in contracted notation since stiffnesses are fourth-rank tensors and are defined by four subscripts in tensor notation.8
Poisson’s ratios must also be determined if the material is to be fully characterized. Poisson’s ratio is defined as the negative of the
ratio of the normal strain induced in one of the two principal lateral directions to the normal strain in the direction of normal
loading. The first subscript indicates the direction of loading, the second the direction of induced strain. The normal loading can be
either tensile or compressive. Thus six Poisson’s ratios exist for each of these two loading modes. Those quantities defined as the
negative of the ratio between the strain in the direction of lower axial stiffness to the strain in the direction of higher axial stiffness (of
the two strains involved) are termed the major Poisson’s ratios. The inverse are the minor Poisson’s ratios. Typically the values of
Poisson’s ratio do not differ significantly between tension and compression loadings, and therefore often no distinction is made.
Whether in tension or compression, the corresponding six Poisson’s ratios and three axial stiffnesses E1, E2, and E3 are
interrelated via the three so-called reciprocal relations, viz., (n)21/E2 ¼ (n)12/E1, (n)3l/E3 ¼ (n)13/E1, and (n)32/E3 ¼ (n)23/E2.4,7,8 Thus
only six of the nine quantities in these relations are independent. Normally, the three minor Poisson’s ratios are taken to be the
dependent quantities, since they are typically the most difficult of the nine properties to measure experimentally, because of their
relatively small magnitudes (Loading in the less stiff direction of the composite induces relatively little deformation in the stiffer
direction).
In addition to (tensile or compressive) normal loadings, to fully characterize an orthotropic composite material, shear loadings
must also be applied in the three principal material directions. Following the same definitions as for normal loadings, an induced
shear stress t12 implies a shear loading applied on (parallel to) the 1-plane in the 2-direction, while t21 indicates a shear loading
applied on the 2-plane in the 1-direction. To maintain moment equilibrium of a local (differential) element of material, t21 ¼ t12,
t31 ¼t13, and t32 ¼ t23.4–8 Corresponding relations exist for the six shear strains, gij. Thus, although it is possible to apply six shear
loadings to a material, only three independent shear stresses and shear strains, for example, t12, t13, t23 and g12, g13, g23 exist. It
will be noted that, although the order of the subscripts is immaterial since tij ¼ tji and gij ¼ gji two subscripts are still required,
unlike for normal stresses and strains, since both the plane and the direction must be defined.
The remaining question is whether the values of these shear stresses and strains at fracture (and/or yield) are independent of the
sense of the shear loading (e.g., whether a shear loading in the positive direction results in failure at the same level as a shear
loading in the negative direction) on the same plane. Typically, shear strength (and shear strain at failure) is in fact independent of
the direction of the applied shear loading. Then only three shear strengths (and/or shear strains at failure) need be measured.
Corresponding to the three applied shear stresses, three independent shear moduli, for example, G12, G13, G23 in contracted
notation, exist.
In summary, nine independent elastic stiffness properties, commonly taken to be G12, G13, G23, (n)12, (n)13, (n)23, G12, G13, G23,
must be experimentally evaluated, recognizing that E1, E2, E3, (n)12, (n)13, (n)23 may differ for tensile and compressive loadings.
Correspondingly, three axial tensile strengths and three axial compressive strengths (and the corresponding axial strains at
failure, if of interest) must be measured, along with three shear strengths (and three shear strains at failure, if of interest), if it can
be assumed that these shear values are independent of the sense of the shear loading, the common assumption.

7.2.1.4 Evolution of Mechanical Testing Methodologies


It is useful to understand the evolution of composite material mechanical test methods that has occurred since the 1950s, in order
to better understand their current status.
By the late 1950s, many serious structural applications of composite materials had begun to emerge. Much of the early interest
was related to aircraft and missile structures, rocket propellant cases, and pressure vessels. In the practical sense, the only fiber-
reinforcement available was E-glass, a high strength but relatively low modulus (E¼72 MPa) fiber.9,10 Obviously, all composite
material test methods at that time were developed to characterize this specific fiber-reinforced composite. Although the technology
was in its infancy, much excellent test methods development work was conducted during this time.
In the very early 1960s, a slightly higher modulus S-glass fiber (E¼ 86 MPa) was introduced.10 This 20% increase in modulus
had only minor implications with respect to test methods. However, both boron and carbon fibers were introduced commercially
shortly thereafter, in approximately the 1964–65 time period.9,10 Boron has an inherent modulus of 414 GPa, 5.7 times or 470%
higher than E-glass, and is almost equally strong. Boron fibers were produced by vapor deposition of boron from a boron
trichloride gas onto a resistively heated tungsten substrate, itself the diameter of a glass fiber. The resulting boron fibers with a
tungsten core were an order of magnitude larger in diameter than glass fibers. Thus boron fiber-reinforced composites better
resisted compressive buckling, and exhibited very much higher compressive strengths than glass fiber-reinforced composites.
The resulting two factors, significantly increased composite stiffness and extremely high compressive strengths, had a major
effect on composite material test methods. In fact, they made most of the carefully developed methods for glass fiber composites
unsuitable.
The first carbon fibers were introduced commercially at about the same time as boron fibers. But these first fibers had axial
moduli only about twice as high as glass fibers, fiber diameters comparable to glass, and typically lower tensile strengths.9,10 One
8 Test Methods for Mechanical Properties

definite virtue was their relatively low density, only about 70% that of glass (and boron). However, the resulting carbon fiber-
reinforced composites exhibited extremely low shear strengths, due to the poor bond between the fibers and the surrounding
polymer matrix. Therefore, it was particularly difficult to transfer loads into a test specimen via shear, as, for example, by using
tabbed tensile specimens, and even to redistribute stresses within the composite (also a shear transfer mechanism).
Thus, these early carbon fiber composites presented their own testing problems, uniquely different from both glass- and boron-
fiber composites. However, this proved to be a moving target, because for the next 15–20 years carbon fibers of progressively
higher modulus (eventually significantly surpassing even boron fibers) and higher tensile strength (equaling or surpassing all
others) became common. Correspondingly, the composite shear strengths attainable became among the highest of all composites,
due to improvements in carbon fiber surface chemistry. These improvements eased the load-transfer problem, but obviously made
the experimental determination of shear strengths more difficult, high strengths almost always being more difficult to measure
than low strengths.
During the 1970s, the use of boron fibers diminished (because of the inherently high cost of the vapor deposition production
process), and applications of carbon fibers focused on the use of what are now termed low modulus carbon fibers (on the order of
E¼220 MPa). This once again encouraged the focus of test methods development on one specific composite material system, now low
modulus carbon fiber composites rather than glass fiber composites. This focus continued throughout the entire decade of the 1970s.
This stability in test methods development was perturbed only slightly by the introduction of aramid (viz., DuPont’s Kevlar 49)
fibers in the early 1970s.9,10 Although this aramid fiber, with a modulus of only 135 GPa and a very low compressive strength, was
not being used in many primary load-bearing structures, it did present its own unique testing problems. The fiber transverse tensile
strength was very low, leading to fiber fibrilation (breaking up of the fiber into individual fibrils). Composite specimens were also
much more difficult to prepare, the organic fiber material being difficult to cut.
During the early 1980s, polyethylene fibers (e.g., Allied’s Spectra) were introduced.11 These fibers exhibited the same general
testing problems as aramid, plus an extremely low composite shear strength.12 This was because polymer matrices did not bond
well to these polyethylene fibers. Once again, a whole new set of test methods development problems was introduced.
However, the major influencing factor of the 1980s, and felt into the new millennium, was the introduction of whole new
classes of polymer matrices. During the 1970s, the composites industry had settled into the primary use of what was essentially
only one type of polymer-matrix material (based upon the Ciba Geigy MY720 epoxy monomer, which in the late 1990s is termed
a brittle structural epoxy). This “stability” was promoted to a large extent by the aircraft industry’s desire to stabilize the material
form, and focus on its introduction into airframe structures. This was achieved by enforcing the requirement that composite
materials be fully qualified for use, which as previously noted involves a costly and time-consuming mechanical testing program.
By the end of the 1970s, the properties of low modulus carbon fiber composites incorporating this matrix material were being fully
utilized, particularly the limited toughness properties.
Recognizing this, in the early 1980s, the US government funding agencies in particular began to encourage the release of new
polymer matrices by the material suppliers. Of course, polymer chemists had been formulating new matrix materials for com-
posites throughout the 1970s, even though there was little commercial market for them at the time. Thus, as if the flood gates had
been opened, almost immediately entire new families of polymers were being made available. These included toughened epoxies,
bismaleimides, polyimides, and the high-temperature thermoplastics, such as poly-etheretherketone (PEEK) and polyphenyl-
sulfone (PPS). This flood of new matrix materials continued throughout the 1980s,13–16 although many of them, including the
high-temperature thermoplastics, have not replaced epoxies to the extent initially anticipated, because of increased processing
complexities and other problems.
During the early part of the 1980s, the composites industry was beginning to realize that the “structural epoxy” then in
widespread use, being very brittle, resulted in very low composite toughness. One important aspect of many of the new polymers
being introduced was their increased toughness. Thus, characterization of composite toughness, impact damage resistance, and the
effects of geometric defects became of particular interest. But almost no corresponding test methods existed. Once again, there was
a major redefinition of required test methods, and a flurry of new developments. Open-hole tension and compression, com-
pression after impact, and fracture toughness (particularly double cantilever beam (DCB) and end-notched flexure (ENF), i.e.,
Modes I and II, respectively) test methods became prominent.
At the end of the 1990s, composite materials, and hence test methods, are once again stabilized, with some degree of
standardization emerging. The following sections reflect this emergence.

7.2.1.5 Test Method Selection Criteria


Within each general type of testing to be discussed in the following sections, multiple test methods exist. Some of these test
methods are not commonly used, for reasons to be discussed, and thus can be readily dismissed. However, at least several others of
each type are in common use, and thus guidelines must be provided for the selection of those that are the most suitable for each
specific testing application.
Whether the test method is an ASTM or other standard is not in itself a dominant selection criterion. Some test methods have
been standardized even though they never have been widely used. Others, once popular, have been superseded by better pro-
cedures, standardized or otherwise, but have not yet been canceled. Standardization is typically slow, often requiring years for a
new test method to pass through the process. Thus an attractive new method may be in widespread use long before it finally
becomes a standard. Correspondingly, removal of obsolete standards can be equally slow.
Test Methods for Mechanical Properties 9

Sometimes, the specific test method to be used is dictated by the customer and cannot be easily altered. Fortunately, this is not
usually the case. Taking into account the expected use of the data to be generated, the cost of generating them may be paramount.
In other cases, maximum accuracy may be the driver. Often the laboratory equipment available, and the skill of that laboratory’s
personnel, have to be taken into consideration when selecting a procedure. Thus, the “best” test method can be strongly dependent
on many factors, not all of which are completely technical.

7.2.2 General Test Equipment Requirements

7.2.2.1 Types of Testing Machines Available


To perform mechanical testing, some type of force application device is required, along with grips or other fixtures to transfer the
applied loading to the test specimen. While special purpose load frames can be and are used, general purpose load frames, termed
universal testing machines, are the most common. These are either electromechanical or servo-hydraulic devices.
In contrast, early day machines used balance beams or load rings with dial indicators to monitor load, and were often powered
by a hand crank and gear arrangement.
Current generation electromechanical testing machines typically have a moving crosshead which moves up and down on two
rotating screws driven by an electric motor. Load is monitored by an electronic load cell connected to a data logger or other
computer-based data acquisition module. The more versatile of these machines can operate in load, displacement, or strain
control. In load control, a specified rate of loading can be applied to the test specimen via electronic feedback from the load cell to
the driving motor. In a similar manner a specified rate of strain can be applied to the test specimen by electronic feedback to the
crosshead driving motor from an extensometer, a strain gage, or other type of strain transducer monitoring the test specimen
response to the applied loading. These strain measuring devices will be discussed in Section 7.2.4.2.1.
There are several dozen universal testing machine manufacturers worldwide, and both quality and purchase price vary widely
among them. An important characteristic of a universal testing machine is loading frame stiffness. It is desired to have the test
specimen and not the testing machine deform under applied load.
Early hydraulic testing machines did not have electronic feedback via servovalves. A manual valve was opened by hand to allow
fluid to pressurize a hydraulic ram which loaded the specimen. A calibrated pressure gage indicated the magnitude of the load
being applied to the specimen at any time. The applied force was controlled by manually increasing and bleeding off the hydraulic
pressure to the ram. The accuracy at which the rate of loading or displacement could be controlled was, thus, totally dependent on
the skill and attentiveness of the human operator. In spite of these limitations, these were very useful testing machines, many of
which are still in use today.
Current generation servo-hydraulic testing machines have servovalves with electronic feedback so that, like the current elec-
tromechanical machines, loading rate, displacement rate, or strain rate can be controlled accurately. It is usually convenient to
locate the hydraulic ram in the base of the machine. Accordingly, the crosshead is fixed during a test, being clamped to the
columns of the load frame. However, the crosshead can be raised or lowered on these columns to accommodate specimen test
configurations of varying heights. This is mandatory since the hydraulic ram has limited travel, typically only 100–200 mm. The
crosshead can be very massive for large capacity machines. Thus, provision is often made for raising and lowering it hydraulically
as well, using long-stroke hydraulic cylinders.
Special purpose testing machines are also used, although not as commonly as universal machines. When a very specific type of
test is to be repetitively conducted, for example, in a quality control laboratory, the versatility of a universal machine is not
required. The machine can then be tailored to the specific application, deleting the features that are not needed, thus reducing
its cost.
Fatigue and creep test apparatuses are other examples. Such tests can take a long time to conduct, tying up an expensive piece of
universal testing equipment. It may be much more economical to design and fabricate a specialized apparatus for the specific
application.

7.2.2.2 Electromechanical Versus Servo-Hydraulic Machines


Both electromechanical and servo-hydraulic universal testing machines are being used extensively, each type having its advantages
and disadvantages. Very simple, small, low cost (but typically low load capacity, i.e., 4 kN or less) electromechanical machines of
various levels of quality are readily available. Most servo-hydraulic machines tend to be larger (20 kN load capacity and much
larger), and correspondingly more expensive. However, for an equal load capacity, and equivalent quality, the two types of
machines tend to be comparable in cost.
Older generation hydraulic pumps on servo-hydraulic machines tended to be noisy, usually justifying a soundproof enclosure
or location of the pump in a separate room. Recent advances in hydraulic pump technology have resulted in quieter units with
smaller footprints, with some units located adjacent to the load frame.17
Electromechanical machines are safer to operate, the screw drives limiting the rate of “runaway” of the crosshead if used
improperly. The ram of a servo-hydraulic machine is capable of moving very rapidly, and may do so very unexpectedly if not
operated properly. This can cause damaged test specimens, destroyed fixturing, and serious human injury. Thus novice laboratory
personnel are best introduced to mechanical testing on an electromechanical testing machine.
10 Test Methods for Mechanical Properties

The above discussion alludes to the fact, however, that electromechanical machines are limited in rate of load application.
Thus, for dynamic (high loading rate testing), servo-hydraulic machines are the necessary choice. Likewise, for cyclic (fatigue)
loading, servo-hydraulic machines are almost mandatory. Even though composite materials typically are tested at relatively low
cyclic rates (5–10 Hz, as previously discussed in Section 7.2.1.2), this is still too fast for a typical electromechanical machine to
respond to.

7.2.2.3 Hydraulic Versus Mechanical Wedge Grips


Mechanical wedge grips have been available almost as long as mechanical testing machines themselves. However, during the
1980s and 1990s, hydraulic wedge grips became very popular, also. In general, hydraulically actuated grips tend to be very
expensive, at least two to three times more expensive than mechanical wedge grips. They also tend to be very heavy and bulky.
Each half of a pair of typical 90 kN capacity grips weigh 32 kg, each half of a pair of 225 kN capacity grips weigh 80 kg.17 Thus it
becomes a nontrivial operation to install and remove the larger ones. Of course, they require a hydraulic power supply, which may
be a self-contained unit supplied with the grips, or pressurized hydraulic fluid supplied by the testing machine pump if a servo-
hydraulic machine is being used.
Mechanical grips have the advantage of being self-regulating, i.e., the specimen is gripped in direct proportion to the load being
applied to it. For hydraulic grips, the gripping (clamping) force is applied by the human operator before the test is started, by
selecting the hydraulic pressure to be used. If too much pressure is applied, the gripped ends of the specimen can be damaged
(crushed). If insufficient pressure is used, the specimen will slip in the grips during the loading, quite possibly also ruining the test,
and test specimen.
On the other hand, for fatigue testing, hydraulic wedge grips are definitely better. With each successive tensile loading and
corresponding Poisson contraction in the gripped region, mechanical wedges move slightly, but then may not release during the
unloading half of the cycle. If so, the grips become tighter and tighter with each cycle, eventually crushing the gripped ends of the
specimen. Hydraulic grips apply a constant force to the ends of the wedges, this force inducing a clamping force which tends to be
somewhat higher than that exerted by the mechanical grips. Thus the specimen ends are always compressed throughout the entire
cycle and the superimposed effect of Poisson’s ratio contraction and release is less significant.

7.2.2.4 Other Types of Grips


In addition to mechanical and hydraulic wedge grips, an almost unlimited number of other types of grips for special applications
are available. One of the more common of these types is a simple bolt clamping grip. This consists of two flat steel plates with the
test specimen sandwiched in between. Bolts passing through the plates are tightened to grip the specimen. Such grips are much less
expensive than wedge grips, an advantage in itself.17,18 However, from a technical standpoint there can also be an advantage when
lightweight grips are desirable, such as for dynamic loading and high rate fatigue cycling. The lower mass reduces undesirable
inertial loadings.
When conducting nonambient temperature testing, the “thermal inertia” of massive grips is also detrimental. Times to achieve
thermal equilibrium, and the thermal energy requirements (particularly for cooling), can be major disadvantages. Also, the rapid
temperature changes required for dynamic thermal loadings may not be attainable.
Other types of special grips are available for testing bundles of fibers, single fibers, thin films, composite tubes, etc.17,18

7.2.2.5 Types of Grip Faces


Traditionally, straight-cut serrated or diamond-cut serrated, hardened steel grip faces have been used. Thermal-sprayed grip
surfaces were introduced in the 1990s are in common use today.19–21 These typically consist of tungsten carbide particles in a
cobalt or nickel matrix. These thermal-sprayed surfaces are relatively smooth, usually having the roughness of 100–150-grit emery
cloth. Thus, they are very gentle to the specimen surface, often permitting the elimination of tabs, which can be a significant cost
saving. Even though the gripping surfaces are much smoother than serrated surfaces, they have equivalent holding power.22 This
has also been verified by the author in other (unpublished) work.
An equivalent coating is the SurfAlloy process.23 This is an electrospark deposition by the pulsed-arc microwelding process.
Again, tungsten carbide is the most common material deposited.
Thermal-sprayed or SurfAlloy grip faces do not perform as well as on irregular surfaces as do heavily serrated faces. The
relatively smooth grip faces tend to ride on the high points, reducing the overall effectiveness of the grips. Serrated grip faces dig
into the surface, overcoming surface irregularities.

7.2.3 Test Specimens

Since there are many types of mechanical tests to be considered, there are equally large number of specimen configurations. Many
of these will be discussed in relation to the individual test methods in subsequent parts of this section. However, there are
common considerations and problems that can be discussed as a group.
Test Methods for Mechanical Properties 11

7.2.3.1 Specimen Preparation


If accurate test data with minimum scatter are to be obtained, careful specimen preparation is mandatory. This includes cutting of
the specimens from a panel or other basic form, application of tabs, if used, application of strain gages or other instrumentation, if
used, and accurate measurement of specimen dimensions.

7.2.3.1.1 Cutting procedures


Flat panels are usually cut into individual specimens by first making rough cuts with an abrasive cutoff wheel. Although, diamond-
particle-imbedded cutoff blades can be used, conventional aluminum oxide particle or similar blades are more common, and
more economical. A band saw with an appropriate blade tooth design is sometimes used, but this may tend to damage the
composite material too severely in the region of the cut. In particular, delamination damage may occur and not be apparent.
After the individual specimens are rough cut, these cut edges can be machined to final dimensions by grinding, again using an
abrasive grinding wheel. This operation is usually performed using a water-base cutting fluid, both to control dust and to cool the
surfaces being cut. Polymer-matrix composites absorb moisture. Thus, this moisture exposure time should be kept to a minimum,
recognizing that half of the moisture that enters the surface continues to diffuse inward even after the specimen surface is placed in
a dry environment, i.e., a brief dry-out cycle, even in a vacuum chamber or desiccator, does not remove the moisture that is
diffusing inward. It can take weeks and months to completely dry out a composite component only 2 or 3 mm thick.
On the other hand, most composites are used in an ambient air environment containing at least some moisture (some relative
humidity (RH)). Thus the small amount of moisture induced during the brief specimen cutting operation is usually not significant.
Alternatives to machining have emerged, the principal being abrasive water jet machining.24 Others include ultra-sonic
vibration methods, electrical discharge machining (EDM), and laser machining.25 In any method, the issue of clean cut planes is
paramount, so as not to induce damage that could change test results or lead to premature failure.

7.2.3.1.2 Dimensional measurements


Stress, one important quantity to be determined in a mechanical test, is defined as being proportional to the applied force, and
inversely proportional to the area resisting the force. Strain, another important quantity, is defined as the change in a given
dimension divided by the magnitude of that dimension initially. Thus the dimensions of each specimen must be measured
relatively accurately. Typically, dimensional measurements are made to the nearest 20–30 mm (0.02–0.03 mm), using a con-
ventional micrometer or dial callipers. Of course, much more accurate measurements can be made if the application justifies the
additional effort.

7.2.3.1.3 Specimen quality issues


In addition to the accurate measurement of specimen dimensions, other parameters must also be controlled. Rough surfaces,
induced during composite fabrication or machining, can create local stress concentrations leading to premature failures. Voids in
the composite (e.g., trapped volatiles), matrix rich or starved regions, misaligned fibers, wavy fibers, and variations in panel
thickness can likewise result in nonrepresentative experimental data being generated.
Tabs must be oriented carefully on the ends of tensile specimens so that forces are introduced axially during a test. Suggestions
for proper tabbing are given by Hercules Inc.26 and by the CMRG.27 If a compression specimen is to be end-loaded, the ends must
be parallel to each other, and perpendicular to the axis of the specimen, again so that the compressive force is introduced
axially.28–30

7.2.3.2 Specimen Conditioning and Storage Prior to Testing


Specimens to be tested in the “dry” condition often are given little consideration with respect to their storage environment prior to
testing. Polymer matrices absorb moisture in proportion to the RH of their environment. Since different geographic regions have
significantly different average humidity levels, the “dry” condition of the specimen can vary accordingly. When the composite is
tested at room temperature, these variations in moisture have only a relatively minor influence. However, at elevated temperatures
the effect can be measurable.13,16
A not uncommon practice is to “dry out” specimens just prior to testing by placing them overnight, or for a few days, in a
heated oven or a desiccator. As noted previously, it can take many weeks to months under such conditions to actually remove the
moisture from such a specimen. The short drying time only removes the moisture near the surface.31
In addition to testing in the dry condition, it is often of interest to test moisture-conditioned specimens, both at room
temperature and at multiple elevated temperature levels.13,16 Specimens are moisture-conditioned either by immersing them in
water maintained at an elevated temperature, or placing them in a heated humidity chamber. Since moisture absorption equi-
librium level is directly proportional to RH, and liquid water is equivalent to 100% RH (saturated) air, there is no difference
between water immersion and saturated air exposure. As a minor consideration, liquid water can more readily leech out con-
stituents from the polymer during exposure. However, most current polymers are sufficiently stable that this is not a problem,
although in past generations of polymers it sometimes was.32
An elevated temperature exposure is desired as the rate of moisture absorption, which is an exponential function of tem-
perature.32 Thus, conditioning times required to reach equilibrium can be significantly reduced, perhaps from many months or
12 Test Methods for Mechanical Properties

even years, to a few months. But the conditioning temperature must not be too high or the composite material will be damaged by
microcracking. At high temperatures, the surfaces of the specimen absorb moisture rapidly, and swell accordingly, while the bulk
of the material does not. It should be noted that the coefficient of moisture expansions of most polymers are as much as 100–1000
times their coefficient of thermal expansions.33 This steep gradient of moisture content induces residual stresses, compressive at the
surface but tensile just below. These tensile stresses can become large enough to cause local failures, i.e., microcracking. The reverse
process occurs, should there be a rapid dry out of the surface, either deliberate or accidental.31 Then the surface stresses become
tensile, with the same resultant microcracking there. This whole problem is amplified in a composite material since the fibers
typically do not absorb moisture, and thus do not moisture-swell, creating additional constraints.
The “safe” conditioning temperature depends on the properties of the polymer matrix, and of the composite itself. Polymers
with higher moisture diffusion coefficients, higher coefficients of moisture expansion, and lower ductility will be more suscep-
tible.31 However, as a general “rule of thumb,” a conditioning temperature of 65–701C is usually “safe.” It should be noted here
that sometimes boiling water immersion, usually for relatively short-time periods, for example, 1 day, 2 days, 5 days, is called for.
Obviously boiling water, at 1001C, can create microcracking problems in many polymer-matrix composites. In addition, as
previously noted, these short-time exposures, even at 1001C, and even if the composite survives, are not anywhere near long
enough to allow a composite of typical thickness to reach moisture equilibrium. The composite will probably be tested shortly
after the conclusion of the boiling water exposure. Thus, not only will a very steep moisture gradient, and hence a gradient of
material stiffness, be present, but significant moisture-induced residual stresses will exist also. Obviously these effects obscure the
meaning of the resulting experimental data.

7.2.4 Measurement of Stresses and Strains

Almost always the stress in the test coupon must be determined, at ultimate and possibly at yield. Very often the modulus (the
initial stiffness of the composite) is also required. If so, then the strain in the specimen must also be monitored. In fact, more and
more commonly a complete stress–strain curve to failure is also desired, often to be used subsequently as input to a nonlinear
stress analysis of the component being designed.

7.2.4.1 Stresses
As previously discussed, stress is defined as the applied force (often termed “load”) divided by the cross-sectional area carrying the
load. Measurement of the dimensions of the cross section was also previously discussed.
In earlier times, load was often an actual “dead load,” i.e., weights supported by the specimen in some manner. Since many
thousands of kilograms of load may be required to fail a specimen, systems of levers were often used to amplify the applied load,
thus reducing the amount of mass (dead load) required.
Later, screw-driven crossheads, initially hand-powered and later motor-driven, were used. These required some type of “load
cell” to measure the force being applied by the mechanical screws. Compound lever beam balances were used, being continually
balanced manually. Spring scales (sometimes termed “fish scales”) were also used, particularly if the applied forces were not too
large. Later, “proofing rings,” circular steel rings with a calibrated dial gage mounted across the inside diameter in the direction of
the applied load, were developed.
Load measurement was somewhat simpler with hydraulic testing machines. The hydraulic pressure being applied to the ram
could be monitored with a fluid pressure gage and the corresponding force being applied by the ram calculated based upon the
piston area.
Electronic load cells are almost universally used today. These are actually not too different in concept from proofing rings. The
load cell usually consists of calibrated foil resistance strain gages mounted on flexure arms (“flexures”) within the cell, i.e., a
displacement is still being measured and calibrated as a force. By varying the stiffness of the flexure arms, cells of any load range
can be fabricated.17,18

7.2.4.2 Strains
7.2.4.2.1 Strain/displacement measurement
The movement of the testing machine crosshead or hydraulic ram is sometimes adequate as a specimen displacement mea-
surement from which to compute strains. This can be true when only limited accuracy is required, or when the displacements are
large, for example, in rubbery materials. Usually, however, greater accuracy is required. The problem is that the specimen may
deform nonuniformly and even slip in the gripped regions. Also, the testing machine itself deforms under load, although as
previously discussed, this is kept to a minimum in a good quality (stiff load frame) machine.
To eliminate these extraneous sources of displacement, it is desirable to measure the displacement in a region of the specimen
away from these “end effects.” This region is termed the gage section of the specimen.
Many types of displacement measurement devices have been used over the years. Before the advent of electronics, motion
magnifying mechanical lever systems incorporating vernier displacement scales, and later dial gages also, were used. For even
greater resolution, optical (mirror) systems, attached directly to the specimen across a known gage length, were developed. The
mirror systems in particular could achieve a high degree of accuracy, by incorporating very long optical path lengths. However,
Test Methods for Mechanical Properties 13

they did require considerable effort to use. These mechanical and optical systems were called extensometers (or compressometers
when they measured compressive rather than tensile strains).
In the 1940s, electrical resistance wire strain gages were introduced. These gages could be bonded adhesively directly to the
surface of the test specimen in its gage section. As the fine diameter wire was stretched or compressed and thus the cross-sectional
area changed, the electrical resistance changed proportionally. This change in electrical resistance was then calibrated to strain. In
the late 1990s, foil strain gages had replaced wire gages, but the principle of operation is exactly the same.
The development of strain gages led to the development of “strain gage extensometers.” The arms of the mechanical extens-
ometer that index on the specimen surface in its gage section serve as flexures, and are strain gaged. The change in electrical
resistance from the strain gage extensometer is calibrated by applying a precisely known displacement. It will be noted that the
principle of operation is essentially the same as that of a load cell.17,18
In recent years, the optical extensometer systems have been revisited. A laser is used to produce two powerful beams which
track two points (targets) a known initial distance apart on the surface of the specimen.17

7.2.4.2.2 Extensometers versus strain gages


When strain gages and strain gage extensometers are both used properly, they can produce a similar degree of resolution, and
similar accuracies. This answers the frequently asked question as to whether strain gages or extensometers are “better,” at least from
a technical viewpoint.
It is true that extensometers, particularly optical extensometers, can be designed to operate over much greater strain ranges than
strain gages. For example, an optical extensometer may be designed to have a maximum travel of 500 mm or more. Thus, using a
25 mm gage length, a maximum strain measurement of the order of 2000% is possible.17 Mechanical extensometers, incorpor-
ating a linear variable differential transformer (LVDT) or similar device and having comparable maximum travel, are also
available.17,18
Correspondingly, strain gage extensometers designed to accommodate 100% strain are available.17,18 Of course, as the
operating strain range of all of these devices is reduced, the measurement accuracy typically increases. Thus, for applications
requiring high accuracy, a strain gage extensometer may be designed for as little as 5 or 10% full-scale strain.
The full-scale strain range of typical strain gages is much more limited. Commonly used strain gages typically have a maximum
strain capability of the order of 3–5%, or even less. The higher strain capabilities are achieved for longer gage lengths. “High
elongation” gages, capable of 10–20% strain, are available but much less commonly used.
Another significant consideration is economics. Strain gages cannot be reused. Extensometers, in theory at least, can be reused,
many times. On the other hand, single-element strain gages cost on the order of $5 each. A strain gage extensometer may cost
$3000.17,18 Even a simple laser extensometer costs of the order of $10,000.17 Thus a strain gage extensometer must be reused 600
times, and a laser extensometer 2000 times, to match the cost of strain gages. These numbers of reuses are probably unreasonable,
even if the extensometer is not accidentally destroyed first. A strain gage extensometer is particularly vulnerable. It is small, and
somewhat fragile, although not excessively so. It can be dropped on the floor – and then accidentally stepped on, usually a fatal
action. Also, to generate a complete stress–strain curve to failure, the strain gage extensometer must be left on the specimen until
the specimen fails. Composite materials often fail in a very explosive manner, potentially damaging the extensometer. Optical
extensometers do not contact the specimen directly and hence are not subjected to this potential damage.
Thirty years ago extensometers were used more commonly than strain gages. At that time strain gages cost on the order of $5
each, in 1980 dollars. Strain gage extensometers cost on the order of $700 each. That is, during the past 30 years strain gages have
changed very little in price, while extensometers have increased in price by over 400%. Strain gages were of excellent quality then,
and remain so today. Extensometers were of adequate quality then, and certainly were of no better quality in the late 1990s. In fact,
some would argue that the quality today is distinctly less. For these reasons there has been a gradual shift in usage from
extensometers to strain gages over the years.
One justified criticism of strain gages is the time required, hence the labor cost, to install one on a specimen. Strain gages are
very small. A gage length of 3 mm is common; less than 2 mm is not uncommon. Some training, and considerable patience, is
required to bond these small strain gages properly. However, once this skill is acquired by an individual with a steady hand, it is
not uncommon to be able to install and wire 4 or 5 strain gages per hour.
Calibration of strain gages is simpler, less time consuming, and is required less often. Thus, on balance, the total labor cost
required to use strain gages versus extensometers is about the same.34

7.2.4.2.3 Digital image correlation


A strain measurement technique known as digital image correlation (DIC) has become popular over the past decade. DIC is an
optical full-field strain mapping technique, in which a camera tracks the movement of points on a test sample’s surface. The
tracking creates in-plane displacement patterns, from which strains are calculated. Multiple cameras triangulating on the points can
map out-of-plane displacements as well. The points are created by a random distribution array of “spots” on the sample surface.
The spots are typically created by priming the surface with a uniform color (typically white or black), followed by randomly
applied droplets of a contrasting color. The contrasting color can be applied by splashing, light spraying, or patting the surface with
a suitable sponge. During the test, the movement of the spots are recorded digitally, and appropriate derivatives of displacement
are taken to arrive at strains.
14 Test Methods for Mechanical Properties

The term “correlation” is used in this technique due to the algorithm that tracks the spots. As each frame in the motion is
analyzed, statistical methods are employed to “decide” which spots have moved from a given location in one frame to a given
location in the next. To make the correlation possible, the spots themselves should not be identical nor regularly repeating. Such a
pattern is “textureless” and no correspondence from one frame to the next can be established without further assumptions on the
deformation.35
DIC technology can be procured as turnkey systems which include cameras and software to conduct complete strain-field
mapping and analysis.36–38 Open source software has also become available for use with separately provided digital image capture
hardware.39
DIC is just now beginning to make its way into standardized test methods. For example, ASTM B 831-1440 includes an
advanced method for shear stress–strain behavior in thin aluminum using DIC, but requires a modified sample. ASTM E 64741 on
measurement of fatigue crack-growth rates discusses DIC in its guidelines for measuring growth rates of small fatigue cracks. The
lone composite test method to apparently include DIC, ASTM D 8067-17,42 allows for the use of DIC to measure strain or out-of-
plane buckling when testing for in-plane shear properties of sandwich panels using a picture frame fixture. The reasons for lack of
inclusion in current methods could include high initial acquisition costs of DIC equipment compared to strain gages, as well as
increase setup time and sample preparation of DIC specimens. However, for circumstances in which strain gages are not practical,
or in development of test standards in which full strain fields must be verified (and not just measured under the grid of a single
strain gage), the use of DIC in composites testing is anticipated to grow.

7.2.5 Tension Test Methods

Tension testing of composite materials is much less controversial at the present time than either compression or shear
testing. There is one ASTM standard43 for testing flat laminates that is widely followed internationally. Many individual nations,
and the international organization for standardization (ISO 527-5), have a similar standard. For testing composite pipes and
tubes, the lesser known standard ASTM D 2105-0144 is available. It principally addresses methods of gripping pipes and
tubes.
Many composite materials are very strong in axial tension, and tensile test specimens must, by their nature, be gripped in some
manner for testing. Unreinforced plastics and metals are typically tested using a dogbone specimen, i.e., the specimen has a
constant-width central (gage) region, but is widened at the ends to reduce the stresses there, by increasing the cross-sectional
area.45 Thus, if local damage is induced it hopefully will not be enough to cause failure in these low stress regions. Unfortunately, a
similar approach cannot be taken with composite materials when there is any significant degree of material orthotropy.46 The low
shear strength of the composite relative to its high axial tensile strength results in the tabs simply being prematurely sheared off
parallel to the gage section width. This reduces the specimen to a straight-sided specimen.
For composite materials that have a low orthotropy ratio, for example, quasi-isotropic laminates, chopped fiber composites,
and most fabric-reinforced composites, it may be possible to use a straight-sided specimen. Of course, for such material config-
urations a dog-boned specimen could be used also, but the increased machining time (cost) required may not be necessary.
Likewise, unidirectional composites tested in transverse tension usually have low strengths and can be tested as straight-sided
specimens.
While it is possible to grip the surfaces of a straight-sided axial tensile specimen directly, limits on the effective coefficient of
friction that can be obtained between the specimen surface and the grips imply that large gripping (clamping) forces must be
utilized. In most cases, serrated grip faces cannot even be considered as they would induce too much surface damage to the
specimen. Thermal-sprayed grip faces are a possibility, but for very strong composites the high gripping forces required will induce
stress concentrations leading to premature failures.
Thus, for high-strength composites tested in axial tension, it is almost always necessary to use tabs. These tabs are adhesively
bonded directly to the straight-sided strip of test material. The tab material, often a glass-fabric-reinforced epoxy composite itself,
absorbs the local damage induced by the grips, thus protecting the test material.47 The limiting factor now becomes the shear
strength of the adhesive. Thus a strong adhesive must be used. The tabs can also be increased in length to increase the shear transfer
area. The practical limit is the length of the grip faces available. Grip lengths of 40–50 mm are common. Even if longer grip faces
are available, maintaining the uniformity of the clamping pressure then becomes problematic.
If gripping problems usually the final solution persist is to reduce the thickness of the material being tested. The area of bonded
adhesive is maintained and the total cross-sectional area in the gage section is reduced. In fact, axially loaded unidirectional
composites are typically tested in thin composite form. For example, ASTM D 3039-1443 recommends a thickness of only 1 mm.
Fig. 1(a) shows the tensile specimen overall configuration recommended in ASTM Standard D 3039-14. An untabbed, dog-boned
specimen, such as recommended in ASTM D 638-14,45 is shown in Fig. 1(b).
When performing an axial tension test, it is often desired to determine both the axial modulus and the major Poisson’s ratio.
Thus strains must be measured in both the axial and transverse directions. A 0/90 two-element biaxial strain gage is very
convenient for this purpose. A biaxial strain-gaged extensometer can also be used. However, such extensometers are even more
expensive than the single-axis extensometers previously discussed.17,18 Alternatively, a standard extensometer can be used in
conjunction with a transverse extensometer specifically designed for this purpose. In either case, this type of instrumentation can
be somewhat heavy to hang onto a thin axial tensile specimen of unidirectional composite material.
Test Methods for Mechanical Properties 15

Fig. 1 Tensile test specimen configurations. (a) Tabbed, straight-sided flat specimen (ASTM D 3039). (b) Untabbed, dod-boned flat specimen
(ASTM D 638).

When performing a transverse tensile test, it is common to not determine the minor Poisson’s ratio, for the reason discussed in
Section 7.2.1.3, viz., the transverse strain is likely to be small, and hence difficult to measure accurately. Since the reciprocal
relations (see Section 7.2.1.3) must be satisfied anyway, it is convenient, and perhaps more reliable, to simply calculate the minor
Poisson’s ratio from the measured transverse modulus, the axial modulus, and the major Poisson’s ratio. This assumes, of course,
that an axial tensile test has been performed.

7.2.6 Compression Test Methods

At the end of the 20th century, static compression testing of composite materials is perhaps the most controversial type of testing.
There is considerable debate as to how to even define the compressive strength of a fiber-reinforced composite,48,49 and there exist
a number of directly competing methods for performing a compression test.49–54

7.2.6.1 Problems Unique to Compression Testing


The most unique problem associated with compression testing is being able to obtain a compression failure before buckling
(elastic instability) occurs.55 Buckling on a gross scale, i.e., buckling of the entire specimen, is common to compression testing of
all materials, and is usually readily eliminated by reducing the specimen length relative to its cross-sectional dimensions, or by
ensuring that the test fixture provides sufficient lateral support.
However, with composites, buckling can occur on a more localized scale also and thus as more restraint is provided, the
buckling is forced to more and more localized sites.56,57 For example, individual plies can buckle locally, or even individual fibers
within a ply can buckle. These modes can be suppressed by properly oriented adjacent plies, but with greater and greater difficulty
as the scale of buckling becomes more localized. Only when these various buckling modes are all suppressed can the true
compressive strength of the reinforcing fibers, and hence of the composite, be attained. During the 1990s, several successful
attempts have been made to achieve, or at least closely approach, the true axial compressive strength of various composite
materials.49,56–58 For example, for a typical low modulus carbon fiber/epoxy matrix composite, compressive strengths exceeding
2.8 GPa have been achieved.49 The measured compressive strength for the same material using conventional, ASTM and other
standardized test methods is about 1.5 GPa, i.e., only a little more than half as high.48
However, it has been demonstrated recently that this “true” compressive strength is not the strength attained in a typical
laminated composite in an actual design application.48,49 A series of tests of commonly used laminate configurations was
conducted, the axial ply strength being “backed out” using classical laminated plate theory. Additional data were collected from
the published literature. For both carbon/epoxy and glass/epoxy laminates, the backed-out axial compressive strength was shown
to be very nearly a constant value for all general laminates tested. It was also shown that a compression test of a simple [90/0]
16 Test Methods for Mechanical Properties

cross-ply laminate produced essentially the same backed-out axial compressive strength as any of the other general laminates. For
example, for the carbon/epoxy composite, a value of about 1.9 GPa was obtained. Not only is this value representative of actual
composite laminate behavior, it is about 25% higher than that obtained using current standard tests of unidirectional composites.
This suggests that the current standard compressive test methods produce strengths that are much too conservative.
On the basis of this recent work, a new test method is currently being proposed to ASTM, MIL-HDBK-17, and SACMA, to test
[90/0] cross-ply laminates and back out the unidirectional composite compressive strength,59 as will be discussed subsequently.

7.2.6.2 ASTM and Other Standard and Nonstandard Test Procedures


As noted above, there are many test methods in use for determining the static compressive properties of composite materials.
Some have been standardized by ASTM and other international or national organizations. Some are individual company
standards which have also been adopted by other groups. Some simply exist, their origin having been lost to the general user
public over the years. In all cases, the specimen is loaded by either direct end loading or by shear loading, or a combination of
the two.

7.2.6.3 Specimen Loading Configurations


Traditionally, specimens for compression testing of conventional homogeneous isotropic materials have been short, solid
cylindrical configurations, either circular or square in cross-section, loaded directly on their ends. Tubes are also sometimes tested.
The specimen length is kept short enough that gross buckling does not occur. But composite materials are usually available for
testing in the form of thin flat plates. Thus lateral support against gross buckling must be provided by the test fixture.
Also, following the example of tensile test methods, shear loading by wedge grips has also become common, perhaps being
made popular by the introduction of the Celanese compression test fixture60 shown in Fig. 2, which became ASTM Standard D
3410-16 in 1975. This has led to a whole family of shear-loading test fixtures during the ensuing years, as will be discussed in
subsequent subsections.
Another form of shear loading is a sandwich beam loaded in four-point flexure. The face sheet on the compression surface is
the actual test specimen, a dummy material being used on the tension surface that is sufficiently strong and/or thick that it does
not fail first. Compressive load is transferred into the compressive face sheet by shear from the core material, which can be any
low axial stiffness material of sufficient shear strength such that it does not fail first. Some type of honeycomb core is a common
choice.
Fixtures incorporating a combination of end- and shear-loading have also been developed, as will be discussed later.

Fig. 2 Celanese compression test fixture. Reproduced from ASTM D 3410-16, 2016. Test Method for Compressive Properties of Polymer Matrix
Composite Materials With Unsupported Gage Section by Shear Loading. West Conshohocken, PA: ASTM International.
Test Methods for Mechanical Properties 17

7.2.6.3.1 End loading versus shear-loading test methods


These two types of compressive loading for composite materials were developed somewhat independently over the years, with very
few direct comparisons being made. However, summaries of the experimental literature have been provided by Berg and Adams,52
Adams and Lewis,61 Welsh and Adams,49 and Adams,51 and of the analytical literature by Xie and Adams.62,63 These summaries
suggest that end-loading produces slightly higher (perhaps 10% higher) axial compressive strengths for unidirectional composites
than shear loading. Theoretical analyses62–65 suggest that this is due to the reduced stress concentration in the gage section of the
test specimen at the junctions with the tabs. Because of the significant scatter in the data not typically obtained by past investi-
gators during compression testing composite materials, this difference between loading methods often escaped detection.61
However, a survey of published experimental data over a number of years does indicate a slightly higher average value for end-
loaded specimens.49,51

7.2.6.3.2 Combined loading


While end loading has been shown both experimentally and analytically to produce lower stress concentrations at the ends of the
gage section, a very high contact stress can be induced at the specimen ends if the loading is not introduced uniformly. This can
lead to end crushing of the specimen, which can negate the test. Thus various attempts have been made over the years to apply a
combination of end and shear loading.66–69 While the concept was sound and these methods were a technical success, the test
fixtures and/or procedures used were not sufficiently attractive to become used extensively.
More recently, Adams and Welsh48 developed what they term the Wyoming combined loading compression (CLC) test
method, and associated fixture shown in Fig. 3. The test fixture is relatively simple and easy to use. A straight-sided [90/0] cross-ply
laminate can be tested without tabs since flame-sprayed gripping surfaces are used. The clamping forces are attained by tightening
four bolts in each fixture half. By controlling the torque applied to the bolts, the ratio of end loading to shear loading can be
controlled. Since only relatively low bolt torques are required (2.2–3.4 Nm), the stress concentrations induced by the clamping
forces are low, as opposed to shear-loading methods. Yet, since only a fraction of the total applied force is being transferred directly
into the specimen ends, end crushing can be prevented even though the specimen is untabbed.

Fig. 3 Dimensioned sketch of an assembled Wyoming combined loading compression (CLC) test fixture. After Adams, D.F., Welsh, J.S., 1997.
The Wyoming combined loading compression (CLC) test method. Journal of Composites Technology and Research 19 (3), 123–133.
18 Test Methods for Mechanical Properties

7.2.6.3.3 Tabbed versus untabbed specimens


There have been a number of studies of the effect of tabs on material strength. Many of these studies have been for tensile
loading47,70,71 but the principles are the same as for compressive loading.62,63,72 These various studies suggest that there are at least
three distinct disadvantages of using tabs on test specimens:

1. Significant additional time, and hence labor cost, is required to add tabs to specimens.
2. If the tabs are not installed properly, they can introduce their own complications. For example, the tabs may not be of uniform
thickness along their length or across their width, the adhesive bond lines may not be of uniform thickness along or across each
tab, the adhesive bond line may not be of the same thickness on each tab surface, the tabs may be poorly bonded and come off
during a test, or they may not be bonded on straight. Each of these irregularities can disrupt the axiality of the loading and thus
negatively influence the test results.
3. The tabs themselves introduce stress concentrations, due to the geometric discontinuity at the gage end of each tab, which is
unavoidable, and due to discontinuities in material properties of the specimen and tabbing material. This latter source can be
minimized by making the tabs of the same material as the specimen. However, usually the specimen material being tested does
not make good tabs. For example, unidirectional carbon/epoxy tabs on a unidirectional carbon/epoxy specimen will tend to
fail in shear parallel to the fiber reinforcement, negating the test.

7.2.6.4 Specific Test Fixtures


As previously noted, there are many different test fixtures in use at the present time. The more common of these will be described
in the following subsections.

7.2.6.4.1 Shear-loading fixtures


Perhaps the oldest of the shear-loading fixtures is the Celanese fixture previously mentioned, and shown in Fig. 2. It was
introduced in 1971 by Park of the Celanese Corporation.60 In 1975, it was standardized as ASTM D 3410-16, the first ASTM
standard specifically for compression testing of composite materials. ASTM D 695-15, “Compressive Properties of Rigid Plastics,”
an end-loaded specimen configuration, had been issued in 1942, 33 years earlier, but as the title implies it was specifically
intended for testing (unreinforced) plastics. The latest version, issued in 2015, still carries the same title.73
Although it became an ASTM standard, the Celanese test fixture suffered from some serious deficiencies. The cone-in-cone
arrangement of the wedge grips (see Fig. 2) requires the tabbed ends of the specimen to be of a precise thickness, viz., 470.05 mm.
Otherwise the male split-cone grips with the tabbed specimen sandwiched in between do not form circular cross-sections, and thus
do not nest properly in the cone-shaped hole in each mating holder. It is difficult to hold the thicknesses of the specimen, two tabs,
and two adhesive bond lines to a cumulative total tolerance of only 0.05 mm. Finish grinding after tab bonding is almost a
necessity. This increases fabrication time and adds one more expense.
The other major deficiency of the Celanese test fixture is the use of a close-fitting outer sleeve to maintain axial alignment. This
sleeve cannot actually maintain alignment; it really just serves as an indicator of misalignment. If the two ends of the fixture do
become misaligned, the close-fitting sleeve binds, creating a redundant load path to carry an unknown portion of the applied load.
If undetected, it results in the indicated compressive strength of the specimen being higher than it actual is. Thus, to obtain a
proper test, the sleeve must be constantly moved around manually during the test, to insure that it is not being loaded.
In spite of these deficiencies, it has been demonstrated (see, e.g., Ref. [50]) that the Celanese fixture can produce compressive
properties equal to the other shear-loading fixtures subsequently developed. But it requires more care, and a higher skill level,
which is a definite disadvantage.
Following the almost immediate criticisms of the Celanese fixture as soon as it was standardized by ASTM, Hofer and Rao74 of
the Illinois Institute of Technology Research Institute (IITRI) introduced what became known as the IITRI compression test fixture,
shown in Fig. 4. The two major deficiencies of the Celanese fixture were eliminated. Flat wedge grips, like tensile wedge grips but
inverted, were used. These could accommodate a range (about 2.5 mm) of specimen thickness, while always maintaining full
contact with the holders. Also, the “alignment” sleeve was replaced by posts and linear ball bushings, which truly do maintain
alignment. It is essentially impossible to bind up linear bushings.
The deficiency of the IITRI fixture is its extreme weight, typically about 45 kg.75 In comparison, the Celanese fixture weighs just
about one-tenth as much, about 4.5 kg. This makes the IITRI fixture very cumbersome to handle. Being so large, and thus requiring
considerable machining, it is also very expensive to fabricate. A typical selling price of the least expensive version of the IITRI
fixture is about $8000.75 For example, this is more than twice as costly as the Celanese fixture, which itself is not an inexpensive
fixture.
One reason for the large mass of the IITRI fixture is that the rectangular holder blocks do not carry the forces induced by the
wedge grips as efficiently as the circular holders of the Celanese fixture. Another reason is that most IITRI fixtures are capable of
testing a specimen up to 38 mm wide, and 15 mm thick in the tabbed regions. It will be noted that the standard Celanese fixture is
designed to accommodate a specimen only 6.3 mm wide and 4 mm thick. The much higher forces required to fail the larger
specimen require a sturdier fixture.
The IITRI test fixture was added to ASTM D 3410-16 in 1987, as Method B. It will be noted that this was a full 10 years after it
was first introduced into the open literature. This is not untypical for new test methods and new test fixtures.
Test Methods for Mechanical Properties 19

Fig. 4 Illinois Institute of Technology Research Institute (IITRI) compression test fixture. Reproduced from ASTM D 3410-16, 2016. Test Method
for Compressive Properties of Polymer Matrix Composite Materials With Unsupported Gage Section by Shear Loading. West Conshohocken, PA:
ASTM International.

To reduce both weight and cost, a smaller version of the IITRI fixture is available.72,75 Shown in Fig. 5, this fixture can
accommodate a 12.7 mm wide specimen up to 7 mm thick. It weighs only 10.5 kg, about one-quarter that of a standard IITRI
fixture. Its cost is similar to that of the Celanese fixture.
There are at least two so-called modifications of the Celanese fixture also. The Wyoming-modified Celanese fixture, shown in
Figs. 5 and 6, retains the efficient circular shape of the holders, but the alignment sleeve has been replaced by posts and linear
bushings, like the IITRI fixture. Also, the wedge grips are tapered circular cylinders rather than cones. Thus, they make full contact
with the holders independent of the specimen thickness, just like flat wedges. However, the circular cylinder wedges distribute the
clamping force reactions more uniformly. The standard design utilizes a tabbed specimen only 114 mm long, i.e., 25 mm shorter
than the IITRI and Celanese fixtures, demonstrating that the 64 mm long tabs commonly used with those fixtures are really longer
than necessary.52,72,76 Specimens up to 12.7 mm wide and 6 mm thick can be accommodated. The fixture weighs 4.5 kg, just about
the same as the standard Celanese fixture, and is only about 70% as expensive. Thus it has become a very popular alternative to the
standard Celanese and IITRI compression fixtures.75
Another modification is the German-modified Celanese fixture,77 shown in Fig. 7. This fixture uses flat wedges like the IITRI,
but circular holders like the Celanese. Unfortunately, it incorporates an alignment sleeve like the Celanese also. This fixture has
experienced somewhat limited use to date, primarily in western Europe.

7.2.6.4.2 End-loading fixtures


By far the most popular end-loading compression test fixture at the present time is the so-called modified ASTM D 695-15
compression test fixture. It actually does not conform to that ASTM standard, and is not an ASTM standard in itself. It was
developed by the Boeing Company in conjunction with Hercules Inc. in 1979,78 and then included in Boeing Specification
Support Standard BSS 7260, first issued in 1982.29 Thus it is often called the Boeing Modified ASTM D 695-15 compression test
fixture. It was later adopted by the Suppliers of Advanced Composite Materials Association in 1989 as SACMA Recommended
Method SRM 1-88.28
20 Test Methods for Mechanical Properties

Fig. 5 Wyoming-modified Illinois Institute of Technology Research Institute (IITRI) compression test fixture: size comparison with standard IITRI
and Wyoming-modified Celanese compression test fixtures.

Fig. 6 Wyoming-modified celanese compression test fixture.

The fixture is shown in Fig. 8. It incorporates I-shaped lateral supports like the ASTM D 695-15 fixture,73 but that is where the
similarity ends. The specimen, rather than being an untabbed dogbone specimen, is straight-sided and tabbed. Actually an
untabbed straight-sided specimen is used to measure modulus, and the tabbed specimen for determining compressive strength.
The specified 4.8 mm gage length between tabs is too short to accommodate a strain gage. The gage length is very short to prevent
gross buckling since a specimen only 1 mm thick is specified. The untabbed specimen cannot be loaded to failure because it will
end-crush prematurely. Having to test two specimens rather than one is inefficient. Also, a complete stress–strain curve to failure
cannot be obtained. Nevertheless, this test method is currently very popular, the fixture being very small and relatively inexpensive,
and the strength and modulus results obtained being very comparable to those measured using the shear-loading fixtures.51,61,79
Also, there is no inherent reason why a thicker specimen with a longer gage length cannot be tested in this fixture, as demonstrated
by Adams and Lewis61 and Westberg and Abdallah.79

7.2.6.4.3 Sandwich specimens


The sandwich beam loaded in four-point bending, the face sheet on the compressive side being the test coupon, has already been
mentioned. This configuration has been used for many years, at least since the late 1950s, but by just a few groups. General
Test Methods for Mechanical Properties 21

Fig. 7 German-modified celanese compression test fixture (all dimensions in millimeter).

Dynamics Corporation, Fort Worth, TX, has long been a leading proponent of its use.80 Its popularity has always been limited,
however, because of the very large size of the specimen, typically 560 mm long and 25 mm wide, which consumes considerable
test material. Also, fabricating a honeycomb core sandwich beam requires special knowledge and skill, which is lacking in most
testing laboratories and the resulting specimen is relatively expensive.
In spite of these criticisms, this test method was finally standardized by ASTM, by including it as Method C in ASTM D 3410-16
at the same time that the IITRI compression test was added as Method B.55 However, just a few years later it was removed from this
standard and established as a separate standard, viz., ASTM D 5467-97.81 It has not gained much more popularity since then.
As an aside, Hofer and Rao74 used almost exactly the same sandwich beam compression specimen as a comparison method
when introducing their IITRI compression test fixture in 1977, 10 years before the sandwich beam and the IITRI compression test
methods became part of ASTM D 3410-16.
A short sandwich panel loaded in edgewise compression has also been standardized.82 While it does primarily test the
compressive strength of the composite face sheets, which are laterally supported against buckling by the honeycomb core, this test
has mostly been used to evaluate the sandwich panel rather than specifically the face sheets.

7.2.6.4.4 Combined loading fixtures


As previously noted, there are relative advantages of both shear loading and end loading of compression specimens. Combined
loading is an attempt to combine the best features of both.
22 Test Methods for Mechanical Properties

Fig. 8 Modified ASTM D 695-15 compression test fixture and specimen (all dimensions in millimeter).

One of the most significant early works was that performed at the Royal Aircraft Establishment (RAE) in England.66–68 The
concept was to adhesively bond the ends of a thickness-tapered unidirectional composite into slots machined in aluminum
blocks. An unknown but significant amount (perhaps 20%) of the applied force was shear-transferred through the adhesive bonds
into the specimen, while the remaining force was transmitted by direct bearing of the aluminum blocks on the ends of the
specimen. While this test method never became very popular, and was never standardized, it did offer useful concepts for future
developments.
Hsiao et al.69 attempted to use a modified version of the IITRI compression test fixture in a combined loading mode. A specified
amount offered was applied in the conventional shear-loading manner through the wedge grips, and then spacers were inserted at
the ends of the wedges so that additional loading would be primarily end-loading. They were able to generate higher compressive
strengths than when using the conventional IITRI fixture.
Actually the original IITRI fixture incorporated grooves and bars at the outer ends of the grips. Their purpose is now somewhat
obscure. They may have been intended to constrain the pairs of grips to move as a unit when gripping the specimen. However, a
more popular concept is that they were to be used to end-load the specimen in combination with the shear-loading wedge grips.
Even with the limited number of current IITRI fixtures that do have these end-loading bars, they are not used regularly. The author
is not aware of any systematic studies of their use.
An end-loading fixture had been developed at the University of Wyoming in the early 1980s76 for use in testing relatively low
strength composites (of the order of 800 MPa or less), such as sheet molding compounds (SMC). It was termed the Wyoming end-
load, side-support (ELSS) compression test fixture and was used extensively for the next 15 years, although it was never stan-
dardized. It utilized a straight-sided, untabbed specimen. The specimen was lightly clamped between the smooth steel blocks of
the fixture to prevent gross buckling. The gage section (unsupported length) was typically 12.7 mm. This fixture performs very well
when used with low strength materials. It was not intended for use with high strength materials, but could have been if specimen
tabs were added. Without tabs the specimen would have end-crushed.
In 1996, some 15 years later, recognizing the value both of using an untabbed specimen and combined loading, researchers at
the University of Wyoming48 added flame-sprayed gripping surfaces and a few other minor refinements to the ELSS fixture, as
shown in Fig. 3. Increased bolt torques were used so that a combined loading rather than pure end-loading was achieved. The
straight-sided untabbed specimen was retained. The intention was to test [90/0] laminates in compression successfully, and then
to “back out” the unidirectional ply axial compressive strength, as discussed previously in Section 7.2.6.1. The fixture has been
renamed the Wyoming CLC test fixture and is receiving considerable attention in the late 1990s.48,59
Test Methods for Mechanical Properties 23

7.2.6.5 Strain Measurement Instrumentation


As for tensile testing, either strain gages or extensometers can be used to measure axial compressive strains. Compressive strain
measurements are a little more challenging, however, since the specimen gage lengths are typically considerably shorter. It is very
difficult to apply a strain gage to a gage length less than about 10 mm. It is even more difficult to attach an extensometer. Thus
strain gages are commonly used. Likewise, because of these difficulties, it is not very common to measure both axial and transverse
strains (to permit the calculation of the compressive Poisson’s ratio). Typically it is simply assumed that it will be relatively close to
the tensile Poisson’s ratio.

7.2.7 Shear Test Methods

There may be a greater number of test methods for determining the shear properties of a composite material than for any other
property. Yet, at the present time, there is relatively little general concern about shear-testing anomalies. Most investigators have
their own preferred method, and relatively few comparisons of experimental results have been made. However, the few com-
parison studies that have been performed indicate that there can be a considerable variation in the results obtained, particularly for
shear strength, as will be discussed.

7.2.7.1 Problems Unique to Shear Testing


As discussed in Section 7.2.1.3, there are three independent shear moduli (stiffnesses) and three independent shear strengths for an
orthotropic composite material. The components in the plane of the laminate, i.e., the 1–2 planes as previously defined, are
termed the in-plane shear modulus (G12) and the in-plane shear strength (S12). The two other components of shear modulus and
of strength are termed the interlaminar shear moduli and strengths.
Some shear-test methods are only capable of measuring in-plane properties, while others only measure interlaminar properties.
Few test methods are capable of measuring both: The Iosipescu shear-test method is one with this dual capability, as will be
discussed. Likewise, a few shear-test methods measure strength but not modulus, for example, the short beam shear-test method,
and a few measure modulus but not strength, for example, the plate-twist test. Some methods, although capable of measuring
both modulus and strength, are not recommended for measuring one or the other of these properties, because of expected
inaccuracies.
One vexing problem when shear-testing many composite materials is deciding when a shear failure occurs. Shear failures
commonly initiate as a network of small-shear cracks in the matrix. Since the fibers usually do not fail, the composite is still
capable of carrying significant load after these shear failures occur. Also, even relatively brittle matrix materials typically exhibit a
distinctly nonlinear shear stress–shear strain response, i.e., a decreasing slope of the stress–strain curve with increasing stress. Thus,
there may not be an abrupt, or even detectable, drop in load at shear failure, the observed composite stress–strain curve
nonlinearity being due to a combination of local shear failures and the matrix material nonlinearity. If the loading is continued to
larger deformations, the intact fibers rotate such that they can begin to carry axial tensile and compressive loads, causing the slope
of the stress–strain curve to begin to increase again. This occurs well beyond the point of shear failure.

7.2.7.2 ASTM and Other Standards


There are a number of ASTM standards for determining the shear properties of composite materials. In-plane test methods include
þ 45 degree laminate tensile shear,83 Iosipescu shear,84 two-rail and three-rail shear,85 and torsion of a hoop-wound cylinder86 or
a pultruded rod.87 Interlaminar test methods include short beam shear,88 Iosipescu shear,84 double-notched shear,89 short beam
shear of a pultruded rod,90 and punch tool shear.91
Other standards organizations in various nations around the world have standardized these same basic test methods as well.3
Obviously, no one or two shear-test methods have yet been generally accepted as the “best” for measuring shear properties, even
though individuals have offered recommendations.92,93

7.2.7.3 Specimen Loading Configurations


There are a number of ways a shear stress can be induced in a composite material. Perhaps one of the most readily recognized ways
is torsional loading. Direct shear, for example, the application of two equal and opposite forces to the test material, is perhaps
equally easy to recognize. In addition, indirect shear-loading methods exist, i.e., the shear stress is an indirect consequence of how
the material is oriented and loaded.

7.2.7.3.1 Torsional loading


7.2.7.3.1.1 Thin-walled tubes
This is an in-plane shear-test method. Almost every publication attempting to compare shear test methods begins by stating that
“torsional loading of a thin-walled composite tube would be the preferred test method, but…” Then it is stated that composite
tubes are usually more difficult and expensive to fabricate than flat panels, may not be representative of the actual component
24 Test Methods for Mechanical Properties

processing method anticipated, may not be a form that can be fabricated using the material of interest, and are difficult to grip
properly for testing. Since basic unidirectional composite shear properties are to be generated, the tube should be wound with
fibers oriented axially or circumferentially (hoop-wound). It is difficult to wind fibers axially, and relatively simple to hoop-wind
them. But a thin-walled hoop-wound tube is very weak in bending (the composite being subjected to a local transverse normal
tensile stress). The tube may be damaged, or even broken, during attachment of end-fittings or strain instrumentation. If the
torsional loading is not well-aligned axially, bending and hence transverse tensile stresses can be induced. The composite is then
being subjected to a combined in-plane shear and transverse tensile loading, which will lead to premature failure. Unfortunately, it
is difficult to detect this induced bending without extensive strain instrumentation. The observed fracture surfaces look very similar
between pure shear and transverse tensile loading also.
As a result of these difficulties, torsion testing of thin-walled composite tubes is not a popular shear test method.

7.2.7.3.1.2 Solid rods


Solid rods of circular cross-section, with the reinforcement oriented axially, are a viable alternative to thin-walled tubes, and also
determine the in-plane shear modulus and strength. A 6-mm-diameter rod, perhaps 100 mm long, is a very viable specimen size,
although rods as small as 3 mm in diameter can be tested.94 Their principal attributes are that they are relatively simple to
fabricate, and very rugged, both in terms of handling and gripping. Rods can be molded to finished dimensions, or machined from
thick fiat laminates.
Unlike for a thin-walled tube, the shear stress across the radius varies from zero at the center (axis of twist) to a maximum at the
outer surface. The variation is linear when the material is in the linear range, but the slope of the shear stress versus radius curve
decreases with radius in the inelastic range. Thus calculation of the maximum stress at shear failure is more difficult.95
One distinct limitation in testing solid circular rods is that it is only practical to fabricate and test unidirectional composites. An
alternative is to test specimens of rectangular cross-section.95,96 However, the reliability of the even more difficult shear stress
calculation becomes questionable, in the inelastic range in particular. A concise discussion is presented in Volume III of the FAA
report previously cited.1

7.2.7.3.2 Direct shear


Inducing pure shear stresses of uniform magnitude directly into the composite specimen is the ideal situation. Unfortunately, none
of the many shear test methods presently available achieve this, although some come much closer than others. Pure shear is taken
to mean the absence of any other stress in the gage section.

7.2.7.3.2.1 Iosipescu shear


The most popular of the direct shear-stress-test methods is the Iosipescu shear test method.75,84,97–99 The test specimen and fixture
are shown in Fig. 9. By appropriately orienting the material in the test fixture, any one of the six shear stress states defined in
Section 7.2.1.3 can be achieved. This is the only test method with such a general purpose use.
For isotropic materials, such as those Iosipescu developed the test for originally, the stress state in the gage section (the region
between the notches) is uniform pure shear. However, for orthotropic materials, such as composites, although the stress is still
essentially pure shear, shear stress concentrations develop at the root of each notch. These stress concentrations are reduced only
slightly by rounding the bottom of each notch and by increasing the notch angle.97 As the material shear response transitions from
linear elastic to inelastic, these stress concentrations do decrease significantly, the actual decrease being dependent on the ductility
of the matrix material.
For a unidirectional composite incorporating a brittle matrix and with the fibers oriented along the specimen axis, the local
stress concentrations induce very visible large shear cracks parallel to the fibers at the notch roots. This premature cracking almost
totally relieves the local stress concentrations. This notch root cracking typically occurs at about two-thirds of the applied load level
at which a network of fine cracks across the gage section of uniform shear stress signifying general shear failure occurs. This
combination of local matrix yielding when possible, and local notch root cracking when it is not, provide a relatively uniform pure
shear stress at failure, the desired stress state. For this reason, the Iosipescu shear-test method is the most common and most
trusted composite shear-test method.

7.2.7.3.2.2 Rail shear


Both two and three-rail shear tests are included in ASTM D 4255-15a.85 The corresponding fixtures are shown in Fig. 10. While
neither fixture is nearly as popular as the Iosipescu shear fixture, the two-rail shear method is very slightly the more popular of the
two.75 The very interesting evolution of each of these test methods, and the two-rail shear fixture in particular, has been detailed by
Hussain and Adams.22,100
The specimens are relatively large (the two-rail shear specimen being 152 mm  76 mm, and the three-rail shear specimen even
larger at 152 mm  137 mm). In addition, six holes must be machined in the two-rail shear specimen, and nine in the three-rail
shear specimen, to clear the bolts used to attach the rails.
The adequate determination of shear strength when using these fixtures has always been questionable because of the stress
concentrations induced by the rails.22,85 An equally significant limitation in the past has been slipping of the rails, which can
negate the test. However, this problem has been significantly reduced with the introduction of tungsten carbide particle surfaces as
described earlier.75
Test Methods for Mechanical Properties 25

Fig. 9 Iosipescu shear test fixture and test specimen (all dimensions in mm). (a) Iosipescu shear specimen in the Wyoming shear test fixture.
(b) Iosipescu shear specimen. Reproduced from ASTM D 5379-12, 2012. Test Method for Shear Properties of Composite Materials by the
V-Notched Beam Method. West Conshohocken, PA: ASTM International.

A very promising modification of the two-rail shear fixture was introduced in the late 1990s.22,100 Roughened rails are clamped
onto the specimen, but the bolts do not pass through the specimen, thus eliminating the need for clearance holes, and the
associated preparation cost. The fixture is shown in Fig. 11. It essentially eliminates the slipping problem. While less research has
been performed on the three-rail shear fixture, the problems are very similar.

7.2.7.3.2.3 Double-notched shear


The ASTM D 3846-0889 specimen configuration is shown in Fig. 12. The modified ASTM D 695-15 compression fixture (see Fig. 8)
can be used to apply the shear loading. The specimen can also be loaded in tension.30 Although an ASTM standard, this test
specimen is justifiably criticized because of the severe normal as well as shear stress concentrations induced at the bottoms of the
notches. This leads to premature local failures which can then immediately propagate across the entire section between
the notches. Although this test method is not used very extensively at present, it was relatively popular for a period of time after the
modified ASTM D 695-15 compression test fixture first became popular in the mid-1980s. This was primarily because the
specimen is relatively simple to fabricate and the same test fixture can be used. Shear strains typically are not measured when using
this test method, and thus the shear modulus and a shear stress–shear strain curve are not obtained.

7.2.7.3.3 Induced shear


By far the most popular of the induced shear test methods are þ 45 degree laminate tensile shear and short beam shear. The
former is restricted to in-plane shear and the latter to interlaminar shear.

7.2.7.3.3.1 þ 45 Degree laminate tensile shear


This in-plane shear test method, as defined by ASTM Standard D 3518-13,83 utilizes a straight-sided [ þ 45]ns laminate tensile
specimen, typically about 25 mm wide. The specimen can be tabbed, but because the axial tensile strength of this laminate is
relatively low, tabs usually are not necessary. The applied tensile loading induces shear stresses parallel to the ply orientations.
Using classical lamination theory, the shear modulus and shear strength can be calculated.
26 Test Methods for Mechanical Properties

Fig. 10 Two-rail and three-rail shear test fixture configurations. Reproduced from ASTM D 4255-15a, 2015. Guide for Testing In-Plane Shear
Properties of Composite Laminates. West Conshohocken, PA: ASTM International.

Fig. 11 Wyoming-modified two-rail shear test fixture.

Unfortunately, axial and transverse stresses are also induced in each ply, leading to the potential for premature failure due to
these combined stresses. As discussed previously relative to the torsion of a thin-walled tube test, failure modes induced by such
combined stress effects are difficult to detect.
Equally detrimental to this test method is the fact that interlaminar shear stresses are also induced between plies, which are
amplified at the edges of the specimen. These stresses can be visualized as being induced by the “scissoring” deformation as each
ply attempts to reorient in the direction of applied loading, adjacent plies rotating in opposite directions. Thus the laminate may
fail in this interlaminar shear mode rather than the intended in-plane mode.
In general, although this is an ASTM standard test method, it is of questionable validity because of the complex stress state in
the specimen. Nevertheless, perhaps because it is sanctioned by ASTM, it remains as a relatively popular shear-test method
undoubtedly due to the simple specimen configuration and no need for special fixturing.
Test Methods for Mechanical Properties 27

Fig. 12 Double notch shear test fixture and specimen. Reproduced from ASTM D 3846-08, 2015. Test Method for In-Plane Shear Strength of
Reinforced Plastics. West Conshohocken, PA: ASTM International.

7.2.7.3.3.2 Short beam shear


This test method has many of the same attractions and disadvantages of the þ 45 degree tensile shear test method, plus additional
ones. A composite beam is loaded in three-point bending. The shear stress induced in a beam subjected to a bending load is
directly proportional to the magnitude of the applied load, and independent of the span length. The bending stresses are directly
proportional to both the applied load and the span length. Thus the support span of the short beam shear specimen is kept short
so that, hopefully, an interlaminar shear failure occurs before a bending failure. This test method is defined by ASTM D 2344-16,88
which specifies a span length-to-specimen thickness ratio of five for low stiffness composites, and four for higher stiffness
composites.
It should be noted that shear modulus, or shear strains in general, cannot be measured adequately in this short beam. Thus,
each test specimen produces just one bit of information, shear strength.
The principal problem in determining this shear strength is that a very complex stress state is induced in the short beam.101 The
shear stress is not adequately defined by the simple beam theory the ASTM standard uses to calculate it. The support and loading
cylinders induce these complex local contact stresses in the beam specimen. For this reason the standard is titled “apparent
interlaminar shear strength…” and is continually criticized as producing invalid shear strength results, which at best should only
be used for material performance comparison purposes.
Interestingly, however, this shear-test method is probably performed more often than any other. Like the 745 degree tensile
shear test, the reason is its simplicity. As will be discussed subsequently, a bending (flexural) test is among the easiest to perform.
Also, the test specimen is very small, thus requiring very little material, for example, assume a specimen 3 mm thick. The span
length for a low modulus composite should then be 15 mm (five times the thickness). Allowing one specimen thickness overhang
at each end of the beam, the total specimen length is 21 mm. The specimen width is typically about half of this. Thus many test
specimens can be fabricated from a very small amount of composite material.
The interlaminar shear strength of any laminate configuration can be determined. However, most commonly a simple uni-
directional composite is tested, to obtain basic ply properties. It should be noted that in this case, if the axially aligned reinfor-
cement fibers can be assumed to be randomly distributed throughout the composite cross-section (i.e., the composite is
transversely isotropic), the interlaminar and in-plane shear strengths S12 and S13 (see Section 7.2.1.3) are identical, i.e., the in-
plane shear strength of a unidirectional composite can be determined using a short beam shear test. This, of course, adds further to
the popularity of this test method.
Recognizing the deficiencies of the short beam shear test as discussed above, Busse and Adams102 performed a detailed study of
this test method. This was a follow-up and extension of a prior study.103,104 Many significant findings were obtained in both
studies. However, the most significant finding relative to the present discussion was that, by significantly increasing the diameters
28 Test Methods for Mechanical Properties

of the support and loading cylinders, and the loading cylinder in particular, and increasing the span length-to-specimen thickness
ratio to about eight, the parabolic distribution of shear stress across the thickness of the specimen predicted by simple beam theory
could be approximated very well in the regions between the loading and support cylinders.102 Correspondingly, the specimens did
fail in a shear mode. Thus, with these modifications, ASTM D 2344-16 has become a technically acceptable, as well as popular,
shear-test method.

7.2.7.4 Strain Measurement Instrumentation


Each type of shear specimen tends to have its own strain measurement requirements. A key point to remember in all cases,
however, is that the foil resistance strain gages and extensometers previously discussed in Section 7.2.4.2.1 only measure normal
strains. The shear strain must always be inferred based upon fundamental mechanics principles. In directions þ 45 and  45
degrees from the directions of shear strain, normal strains are always present. It is these normal strains that are measured by the
strain instrumentation. In fact, if it is known that the shear strain in a given direction is pure (not combined with normal strains),
only one normal strain (at 745 degree) need be measured in order to determine the shear strain.

7.2.7.4.1 Iosipescu shear


As discussed in Section 7.2.7.3.2.1, the shear stress in the gage section between the notches of an Iosipescu shear specimen can be
assumed to be approximately uniform. Thus only one strain gage (or extensometer) oriented at either 745 degree is actually
required. The gage will record a tensile strain if oriented in one 45 degree direction, and a compressive strain in the other 45 degree
direction. The magnitude of the pure shear strain is then twice this value. Typically, however, two gages are used, one at þ 45
degree and one at  45 degree, wired in a Wheatstone half-bridge circuit, to double the sensitivity of the measurement. Then the
shear strain is simply the sum of the magnitudes of the two normal strain measurements.
Since, as discussed in Section 7.2.7.3.2.1, very localized strain concentrations do exist near the notches, a special biaxial strain
gage has been developed that covers the entire width of the Iosipescu shear specimen between the notches. In this way, the average
strain in the specimen is obtained, not influenced by local concentrations. Incidentally this gage has sensing elements oriented at
both þ 45 and  45 degrees. The current cost of these special gages is about the same as a standard biaxial gage.

7.2.7.4.2 Thin-walled tube and solid rod torsion


Although not a commonly available measurement device in most testing laboratories, a rotometer can be used, if available. As
implied, this device measures the relative rotation of two sections of the tube or rod a known distance apart. From this mea-
surement, the (average) shear strain at the surface of the specimen between the sections can be calculated.
It is more common to use strain gages. Since, barring the induced bending problems discussed in Section 7.2.7.3.1.1, the
specimen is in pure shear, most of the comments made relative to the Iosipescu shear specimen apply here also. The shear strain is
in the plane defined by the axial and circumferential directions of the tube or rod. Thus a gage is oriented in either the þ 45 or
 45 degrees direction, or both. Since no strain concentrations are anticipated anywhere on the surface of the specimen, the special
“strain averaging” gage developed for the Iosipescu specimen is not necessary here (although it could be used, if desired for some
reason).

7.2.7.4.3 745 Degree tensile shear


This specimen requires two normal strain measurements, one along the loading axis (x-axis) of the specimen and the other in the
transverse in-plane (y-axis) direction. Either simple strength of materials analysis or classical lamination theory permits the
determination of the corresponding in-plane shear stress and shear strain (in planes in the þ 45 degree directions, i.e., in the
directions of the unidirectional ply orientations). The governing equations for shear stress and shear strain are, respectively83,105:
t12 ¼ sx =2 and g12 ¼ jex j þ jey j
7.2.7.4.4 Two- and three-rail shear
Only a single gage oriented at þ 45 or  45 degrees to the rails is required. For the three-rail shear test, two gages are often used,
one in each shear section of the specimen, although this is not necessary.

7.2.7.4.5 Short beam shear and notched shear


As noted in Sections 7.2.7.3.3.2 and 7.2.7.3.2.1, strains are not measured when using these test methods, and thus no strain
instrumentation is required.

7.2.8 Flexure Test Methods

Flexural testing (bend testing) of composite materials, like tensile testing as discussed in Section 7.2.5, is relatively noncontroversial,
at least at the present time. Flexural testing, like the short beam shear testing discussed previously in Section 7.2.7.3.3.2, is popular
because it is relatively simple to conduct, using simple specimens. It is questionable as a basic composite material properties test since
the specimen is subjected to a combination of nonuniform stresses, viz., maximum tension at the lower surface, maximum
Test Methods for Mechanical Properties 29

compression at the upper surface, and maximum interlaminar shear at the midplane. Elsewhere the beam is in a combined stress
state of normal and shear stresses. Thus, failure can be in tension, compression, shear, or a combination of these stresses.
Either solid laminates or sandwich beams are commonly tested, and there are ASTM standards for both, as will be discussed in
Section 7.2.8.2. Both three- and four-point loading are used, as defined in the standards. This nomenclature is a bit misleading,
although universally used. A “three-point loading” beam specimen is actually loaded in the center by a single load, and supported
near each end. Thus technically it should be termed “single-point loading.” But it never is. Correspondingly, for four-point
loading, two loads are applied symmetrically with respect to the center of the beam length, and the beam is supported near each
end. With four-point loading it is conventional to use either “quarter-point loading” or “third-point loading,” i.e., the two loads
are applied either at one-quarter of the support span length from each support, or at one-third of the span length.
There is some controversy as to whether three- or four-point loading is “better.” To achieve the same bending stresses at the
center of the beam, the same force must be applied at the single loading point as for the double loading points, if quarter-point
loading is used. For third-point loading, the required double point loading is less, by one-third, than either of the two previous
cases. This is beneficial in terms of local damage problems.
For three-point loading, the maximum bending stresses occur only at the center section of the beam, increasing linearly inward
from the support points. For four-point loading, the bending stresses are constant between the two loading points. Thus, for
quarter-point loading, the center one-half of the beam is in constant bending, while for third-point loading only one-third of the
beam is in constant bending.
In summary, perhaps the four-point configuration with three-point loading is the most desirable, because of the reduction in
contact stresses on the beam. Nevertheless, quarter-point loading and also single-point loading are being used at least equally as
often at the present time.75

7.2.8.1 Problems Unique to Flexure Testing


7.2.8.1.1 Solid laminates
Solid laminate flexure specimens are usually loaded and supported by means of circular cylinders. This does induce local stress
concentrations, just as discussed relative to the short beam shear test method in Section 7.2.7.3.3.2. However, in the present case,
since the span length-to-specimen thickness ratios are much larger, the forces required to fail the specimen are lower. Thus the
problem of crushing and other local damage under the loading and support points is not as severe.

7.2.8.1.2 Sandwich beams


Flexural testing of sandwich beams, in particular beams consisting of a honeycomb core and thin composite laminate face sheets,
has been performed for many years, in fact about as long as composites have been seriously considered as a construction
material.106 However, during the past few years, the amount of this type of testing has increased significantly.75
Unlike the sandwich beam test discussed in Section 7.2.6.4.3 as a compression test of the face sheet material, a true flexural test
of a sandwich beam is intended to test all components of the beam, viz., the axial tensile and compressive strengths of the face
sheets, the shear strength of the core material, and the strength of the adhesive which bonds the face sheets to the core.
Correspondingly, the sandwich beam is typically fabricated by personnel skilled in such fabrication techniques, as opposed to the
sandwich compression test coupon which may have to be fabricated by mechanical test laboratory personnel or others who are
much less knowledgable.
Perhaps the most common problem encountered when testing sandwich beams with thin face sheets as opposed to solid
laminates is the greater potential for local damage at the loading and support points. Thus it is common to load the beam via flat
plates (perhaps 25 mm wide or more) rather than cylinders. These loading plates should be free to pivot so that they continue to
apply a distributed load as the beam deflects.

7.2.8.2 ASTM and Other Standards


By far the most popular standard for flexural testing of solid laminates is ASTM D 790-15e2.107 Methods 1 and 2 define three-point
and four-point loading, respectively. Ceramic composites typically utilize very small specimens and are covered by ASTM C 1341-
13.108 Flexural testing of pultruded rod is defined in ASTM D 4476-14,109 and bonded laminates in ASTM D 1184-98.110
Sandwich beam flexural testing is covered by ASTM C 393-16.111 Many aspects of this standard are similar to ASTM D 790-
15e2, as might be expected.

7.2.8.3 Strain/Displacement Measurement Instrumentation


For either solid laminate or sandwich beam flexure, it is very common to measure the beam deflection under load, rather than to
use strain gages. This is perhaps a carry-over from earlier times when electronic strain instrumentation was not available. Since the
beam deflections often tend to be large relative to the deflection of a tensile or compression specimen, it is practical to measure
them. Dial gages and manual readings are still sometimes used, but for automated data acquisition, linear variable differential
transducers (LVDTs) are usually used. The amplified electronic signal can then be fed directly into the computer for data reduction.
30 Test Methods for Mechanical Properties

Alternatively, one or more strain gages can be bonded to one or both surfaces of the beam, at points of known bending
moment. Single-element gages are typically used, Poisson’s ratio in flexure not usually being of interest.

7.2.9 Multiaxial Loading Test Methods

A multiaxial stress state is often an undesirable test condition when determining basic composite material properties, as previously
discussed. However, actual structural components are often subjected to combined stresses, and many types of failure criteria
(defining either yielding or fracture) have been developed, in an attempt to predict when the component will fail. These criteria
utilize the basic tensile, compressive, and shear properties defined in previous sections.
Unfortunately, no universal, or even generally accepted, failure criterion has yet been developed. This problem is not unique to
composites, however; there are no universal criteria for metals and other materials either.
While the theoretical development of failure criteria has proliferated since the 19th century, very few multiaxial stress
experimental results are yet available for verification purposes. Such testing is both time-consuming and costly, and limited test
equipment has been developed.
Multiaxial loading can be either 2D (biaxial) or three-dimensional (3D) (triaxial). Either tubes or flat plates can be tested in
either case.

7.2.9.1 Biaxial Testing


Biaxial testing is much simpler to perform than triaxial testing. Thus the few experimental combined stress data for composite
materials that do exist in the literature are predominately biaxial. It is practical to test either tubes or flat plates. The form in which
the material of interest can be supplied may be the deciding factor, although there are important technical differences as well.
A composite tube can be subjected to axial tension or compression, and torsion, combined with internal pressure. However, the
use of torsion is less common than axial loading, both because of equipment limitations and the decreased generality of the stress
state induced. External pressurization is also desired, so that a compression–compression biaxial stress state can be achieved. But
external pressurization is considerably more difficult to achieve than internal pressurization, with few experimental facilities being
available.
A flat plate specimen is usually tested in a cruciform (cross-shaped) configuration. Any combination of tension and com-
pression loading can be applied to the pairs of arms of the cruciform. The test region (gage section) is the central portion at the
junction of the four arms. But stress concentrations are induced at these junctions (internal corners). Thus, considerable analytical
effort has been expended to define optimum geometries that will minimize these local stress concentrations. Usually, circular
cutouts of some shape such as shown in Fig. 13 are used.
The usual data presentation format in any case is a 2D plot of axial normal (s1) stress versus transverse normal (s2) stress. If
data for all four quadrants of the plot can be generated (i.e., if external pressurization can be attained), the plot is an enclosed area,
the perimeter of which at any combination of cr, and a2 defines failure.

7.2.9.2 Triaxial Testing


As stated in Section 7.2.9.1, very few experimental results for triaxial loading have been generated to date. As for biaxial loading,
either a tube or a cruciform specimen can be used.
For a tube, the internal and/or external pressurization provides a compressive normal stress in the third (interlaminar)
direction. If these pressures are not equal, the interlaminar normal stress will not be uniform through the thickness. Also, these

Fig. 13 Typical flat cruciform biaxial specimen.


Test Methods for Mechanical Properties 31

pressures must be relatively high if they are to induce an interlaminar normal compressive stress of any significance. It is not
practical to achieve significant tensile stresses (e.g., applying a vacuum would induce insignificant interlaminar normal stress).
The internal and external pressures in turn induce proportional circumferential stresses, just as for biaxial testing. By varying the
ratio of internal to external pressure, the desired level of circumferential stress can be achieved. However, the interlaminar normal
stress will vary from a stress equal to the internal pressure on the inside surface of the tube to a stress equal to the external pressure
on the outside surface. If the internal and external pressures are equal, the induced circumferential stress will be zero. Thus the
versatility of the tube test for triaxial loading is somewhat limited.
As an alternative, a 3D cruciform specimen can be used, i.e., the specimen has pairs of arms protruding in three perpendicular
directions. Any combination of tension and compression loading can be applied to the three pairs of arms, thus permitting the
generation of a totally general stress state.
The usual data presentation format is a 3D plot of axial normal (s1) stress versus transverse normal (s2) stress, versus
interlaminar normal stress s3. If data for all eight octants of the plot are generated, the resulting failure envelop is a 3D surface,
every point on this surface representing a combination of s1, s2, and s3 that will cause failure.

7.2.9.3 Strain Measurement Instrumentation


It may not be necessary to measure strains since only a stress-failure envelope is to be plotted and the stresses can be calculated
from the known magnitudes of the applied forces and/or pressures. However, if strains are to be measured, for biaxial testing a
two-element biaxial strain gage rosette can be used, on either the tube or the cruciform specimen surface.
Strain instrumentation becomes more difficult for triaxial testing. When testing a pressurized tube it is difficult to measure the
interlaminar (through the wall thickness) strain. It is the change in tube wall thickness that must be determined. Capacitance and
proximity gages are two possibilities.
The triaxial cruciform specimen has all six sides of its cubical gage region covered by loading arms. Thus there are no exposed
surfaces for the attachment of strain gages. Thus, capacitance or proximity gages, or similar devices, must be used in all three
directions. There is much development work remaining in this area.

7.2.10 Fracture Mechanics Test Methods

The combined stress-failure criteria discussed in Section 7.2.9 predict failure based upon the local 2D or 3D stress state at a point.
In contrast, fracture mechanics predicts failure based upon the stress intensity or strain energy density at the tip of a preexisting
crack in the material being sufficient to cause that crack to grow.

7.2.10.1 Historical Introduction


Fracture mechanics was initially developed for metals and similar isotropic materials in the 1940s, assuming linear material
stress–strain response. Thus this early work was termed “linearly elastic fracture mechanics,” or more commonly, simply LEFM.
Later these works were extended to incorporate inelastic material response.
When composite materials became prominent two to three decades later, initial attempts were made to use the same LEFM concepts
to predict their failure as well. Major efforts in this regard were made during the early 1970s. Perhaps what is the obvious was eventually
realized, viz., composite materials do not exhibit the same failure modes as homogeneous materials such as metals. For example, a
preexisting crack in an isotropic material will often propagate in a “self-consistent” mode, i.e., it will continue along the plane of the
initial crack, in the same direction. Cracks in a composite material, because of the nonhomogeneous (fibers in a matrix), orthotropic,
and often layered nature of the composite, propagate in much more complex and varied modes. As a result, fracture mechanics as
applied to composites gained a very poor reputation and tended to be ignored by the composites community for some time.
With the surge in the introduction of new types of composite materials in the early 1980s, and the corresponding desire to identify
the tougher of these composites, as discussed in Section 7.2.1.4 there was renewed interest in applying the basic principles of fracture
mechanics.112 Fortunately, the prior mistakes were not repeated, and current fracture mechanics efforts have a much sounder basis.

7.2.10.2 Fracture Mechanics Principles


Materials in general are assumed to fracture in one of three modes, viz., Mode I – the opening mode, Mode II – the shearing mode,
or Mode III – the tearing mode, or some combination of two or all three of these modes. The basic fracture modes are indicated in
Fig. 14. They are always designated by Roman numerals, as indicated above.
The fracture mechanics specimen configurations in most common use for composite materials at the present time include the
DCB, ENF, and mixed-mode bending (MMB) delamination test. These specimen configurations and the corresponding methods of
loading are indicated in Figs. 15 and 16. The DCB is a pure Mode I test, ENF a Mode II test, and MMB a mixed Mode I and Mode II
test. In all cases the dominant crack propagation is parallel to the reinforcement, or the plies if it is a laminated composite.
The MMB apparatus113 is capable of inducing any prescribed ratio of Modes I and II loading, including pure Mode I or Mode II.
The various fracture mechanics test methods each require their own specific data collection and reduction procedures, which
need not be presented here. The interested reader is referred to an excellent and concise presentation by Carlsson et al.105
32 Test Methods for Mechanical Properties

Fig. 14 Fracture mechanics failure Modes I, II, and III.

Of primary interest for composite materials is the Mode I fracture toughness.114 Fatigue delamination growth is addressed in
ASTM D 6115-97,115 which uses the double-cantilevered beam sample to determine the fatigue cycles required for propagation of
an existing delamination as a function of cyclic strain energy release rate. The mixed-mode fixture of Crews and Reeder113 has been
codified as ASTM D6671.116 The Mode II end-notch flexure (ENF) sample was developed into ASTM D 7905,117 incorporating an
edge crack in a 3-point bend configuration. The DCB test features stable crack growth (crack growth will cease after the available
elastic energy drops below the threshold required for the crack to propagate). However, the configuration of the ENF sample
renders the growth unstable, i.e., the crack will fully propagate (to the zone underneath the loading nose), once the available
energy exceeds the threshold.105 Very little work has been accomplished with Mode III. An edge crack torsion (ECT) specimen was
under examination by ASTM in the late 1990s.118 However, there are currently no active nor proposed new methods in ASTM for
Mode III testing.

7.2.11 Nonambient Testing Conditions

A brief caution about the low rate of moisture absorption and desorption of moisture in polymers and polymer-matrix com-
posites, and the problems this can cause, was given in Section 7.2.3.1.1. This is one major problem, however, there are also other
potential problems, and corresponding precautions to be taken when conditioning and testing composite materials.

7.2.11.1 Common Problems/Precautions


As previously noted in Section 7.2.3.1.1, once moisture enters the surface of a “dry” polymer (or a polymer-matrix composite),
then even if the environment is soon returned to a “dry” state, for example, 0% RH, half of the moisture in the surface layer will
Test Methods for Mechanical Properties 33

Fig. 15 Interlaminar fracture toughness test methods for composite materials.

Fig. 16 Mixed-mode interlaminar fracture toughness specimen and test fixture.(a) Test specimen and loading. (b) Schematic diagram of
apparatus. After Crews Jr., J.H., Reeder, J.R., 1988. A mixed-mode bending apparatus for delamination testing. NASA Technical Memorandum
100662. Hampton, VA: NASA-Langley Research Center.

continue to propagate inward. Assuming that moisture is entering the material from all surfaces of the composite, then only when
two moisture propagation fronts meet will the moisture reverse direction and propagate back out of the material. In even a
relatively thin composite (e.g., 2 mm thick) at room temperature, the total dry-out process can thus take weeks or months,
depending upon the type of polymer, i.e., its moisture diffusivity.32
On the other hand, since moisture absorption in most polymers is slow, brief exposures to high humidity air or liquid water
result in relatively little, perhaps a negligible amount of, moisture weight gain. Thus, for example, there is usually no cause for
concern with the brief exposures to water during specimen cutting operations.
Moisture diffusivity is a strong function of temperature, and therefore moisture absorption can be minimized by keeping the
exposure temperature low. Correspondingly, absorption and dry out can be hastened by elevating the exposure temperature. The
same general principles apply to other absorbing fluids also, for example, gasoline, hydraulic fluid, and solvents in general,
assuming of course that the fluid does not chemically attack the polymer.
34 Test Methods for Mechanical Properties

Both moisture and temperature reduce the stiffness of the polymer, and usually the strength as well. The possible exception is
when the polymer is very brittle in the room temperature and dry condition. Then moisture and/or an increase in temperature can
soften the polymer, making it less sensitive to local stress concentrations, and thus making it appear to be stronger. Correspondingly,
the matrix-dominated properties of the composite, for example, transverse tensile strength and shear strength, can also increase.

7.2.11.2 Elevated Temperature Testing


The two principal problems encountered in elevated temperature testing are gripping of the specimen if necessary and strain
instrumentation.
Adhesives tend to soften with increasing temperature. Thus a tab adhesive that performs well at room temperature may not
hold at elevated test temperatures. It is then necessary to switch to a higher use temperature polymer. Most high performance
epoxy adhesives are serviceable up to temperatures of 120–1401C.
The commonly used strain gages are limited to use temperatures in the same range as the adhesives discussed in the previous
paragraph, for the same reason. These strain gages incorporate polymer backing materials, which are effected by the elevated
temperature, and they are adhesively bonded to the test specimen just as tabs are.
Strain gages for use at very high temperatures, well above 5001C, are available. However, these gages are expensive, and not
always readily available. Again, they must be firmly attached to the test specimen. Most high-temperature polymer adhesives are
not adequate above about 3001C. Most of the ultrahigh temperature gages were originally developed for metals, intended to be
welded to the specimen surface. For nonmetallic-matrix composites it may be possible to bond these gages with ceramic adhesives,
some of which are usable well above 10001C.119,120
The orthotropic thermal expansion properties of composite materials must be taken into account when using elevated tem-
perature strain gages, so that the differing expansion rates between the specimen and the gage are properly compensated for.
Most mechanical (strain-gaged) extensometers are designed for an operating temperature upper limit of 2001C. The primary
limitation is the same as for standard strain gages themselves, i.e., the temperature capability of the strain gage backing material
and the adhesive. Because of this limiting temperature, the molded plastic case and the wiring are also not designed for higher
temperatures.
Higher temperature extensometers, both water-cooled (5001C) and for even higher temperatures are available, but at a somewhat
higher cost.17,18 A typical alternate solution is to use a conventional extensometer but with long ceramic-rod reach arms into the hot
zone, through a heat shield if necessary, the body of the extensometer remaining at a much lower temperature. These extensometers
are advertised for use up to 6501C, and to 12001C if forced air cooled; their cost is about twice that of a standard extensometer.17,18
Another alternative for high-temperature applications is a laser extensometer. The less expensive of these units are about twice
that of the shielded mechanical extensometers, i.e., about four times the cost of a standard extensometer.17 They are not as accurate
as mechanical extensometers or strain gages, but have the advantage of being noncontacting. Thus the laser beam can be focused
through the window of a high-temperature furnace onto markers on the specimen surface. Thus these extensometers can be used at
very high temperatures.

7.2.11.3 Subambient Temperature Testing


Subambient temperature testing is much less of a problem from a strain instrumentation viewpoint than elevated temperature
testing. Strain gage adhesives and polymer backing materials are much less affected by low temperatures. An advertised lower
temperature limit of  2501C is not uncommon.

7.2.11.4 Testing Moisture-Conditioned Specimens


Here, clearly, polymer-matrix composites are being addressed rather than metal- or ceramic-matrix composites, the latter two types
typically not absorbing moisture.
The often expressed main concern is that the specimen will “dry out” during the time between when it is removed from the
conditioning environment until the mechanical testing is complete. As previously noted in Section 7.2.11.1, dry out as such is not
the problem. Unless the testing temperature is very high and/or the testing time very long, the amount of moisture lost will
normally be small.
What should be of concern is that a severe moisture gradient can be induced since the extreme outer surface does become dry
and attempts to contract accordingly. However, the bulk of the material inward from the surface remains moisture-conditioned,
and hence in a swelled state, thus restricting the contraction of the outer surface and inducing significant residual stresses. In fact,
these stresses can become high enough to induce fracture in the composite. The entire problem can be amplified when a moisture-
conditioned specimen is tested at an elevated temperature since the moisture gradient becomes even steeper.

7.2.12 Conclusions

The mechanical testing of composite materials has matured significantly during the 1990s. This is because the increasingly
sophisticated uses of composite materials have dictated the need for a much greater variety of measured properties, and a need for
Test Methods for Mechanical Properties 35

greater accuracy. Hence, many new test methods have emerged since the mid-1980s. Sufficient time has now elapsed for the
composites industry to assess most of these, and to establish preferred procedures. As a result, the state-of-the-art of mechanical test
methods is now stabilizing. Thus the guidelines presented in this chapter should remain generally applicable for some time to come.

See also: 7.1 Introduction to Mechanical Testing. 7.3 Saint-Venant End Effects for Anisotropic Materials. 7.4 Size Effects in Composites. 7.5
Fiber Test Methods. 7.6 Computed Tomography of Composites. 7.7 Backing-Out Composite Lamina Strengths from Cross-Ply Testing. 7.8
Environmental Durability of Polymer Matrix Composites. 7.10 Radiographic Inspection of Composite Materials. 7.12 Thermal Nondestructive
Evaluation of Composite Materials and Structures. 7.14 Acoustic Emission Analysis. 7.15 Adhesive Bonding of Composite Structures. 7.18
Smart Composite Materials Systems

References

1. FAA Technical Center, 1993. Test methods for composites: A status report; Volume I. Tension Test Methods, Volume II. Compression Test Methods, Volume III. Shear
Test Methods. Report DOT/FAA/CT-93/17, Atlantic City, NJ.
2. Jenkins, C.H., 1997. Manual on Experimental Methods of Mechanical Testing of Composites, second ed. Bethel, CT: Society for Experimental Mechanics.
3. Traceski, F.T., 1990. Specifications & Standards for Plastics & Composites. Materials Park, OH: ASM International.
4. Gibson, R.F., 2012. Principles of Composite Material Mechanics, third ed. Boca Raton, FL: CRC Press, Taylor & Francis Group.
5. Daniel, I.M., Ishai, O., 2005. Engineering Mechanics of Composite Materials, second ed. Oxford: Oxford University Press.
6. Hyer, M.W., 2009. Stress Analysis of Fiber-Reinforced Composite Materials. Lancaster, PA: DEStech Publications Inc.
7. Herakovich, C.T., 1998. Mechanics of Fibrous Composites. New York, NY: Wiley.
8. Jones, R.M., 1999. Mechanics of Composite Materials, second ed. Philadelphia, PA: Taylor & Francis.
9. Schwartz, 1992. Composite Materials Handbook, second ed. New York, NY: McGraw-Hill.
10. Lubin, G., 1982. Handbook of Composites. New York, NY: Van Nostrand Reinhold.
11. Chang, H.W., Lin, L.C., Bhatnagar, A., 1986. Properties and Applications of Composites Made of Polyethylene Fibers. Petersburg, VA: Allied Fibers Technical Center.
12. Adams, D.F., Zimmerman, R.S., Chang, H.W., 1985. Properties of a polymer-matrix composite incorporating Allied A-900 polyethylene fiber. SAMPE Journal 21 (5),
44–48.
13. Zimmerman, R.S., Adams, D.F., 1986. Mechanical properties of neat polymer matrix materials and their unidirectional carbon fiber-reinforced composites. Report UWME-
DR-601-103-1. Laramie, WY: University of Wyoming, Department of Mechanical Engineering.
14. Carlsson, L.A., 1991. Thermoplastic Composite Materials, Composite Materials Series. Oxford: Elsevier Science.
15. Cogswell, F.N., 1992. Thermoplastic Aromatic Polymer Composites. Oxford: Butterworth-Heinemann.
16. Coguill, S.L., Adams, D.F., 1989. Mechanical properties of several neat polymer matrix materials and unidirectional carbon fiber-reinforced composites. Report UW-
CMRG-R-89-114 (NASA Contractor Report 181805), Composite Materials Research Group. Laramie, WY: University of Wyoming.
17. MTS Systems Corporation, 2017. MTS Landmarks Testing Solutions, Eden Prairie, MN. Available at: http://www.mts.com/cs/groups/public/documents/library/
dev_004324.pdf
18. Instron Corporation, 2006. 7 Tips for Materials Testing, Norwood, MA. Available at: www.instron.com
19. Products Finishing, 1998. Products Finishing Directory and Technology Guide. Cincinnati, OH: Gardner Publications.
20. ASM, 1992. Thermal-Spray Buyer’s Guide. Pittsfield, MA: ASM International.
21. Busboy, P.O., Nikitich, J., 1991. HVOF thermal spraying of nitrided parts. Advanced Materials & Processes 141, 35–36.
22. Hussain, A.K., Adams, D.F., 1998. An analytical and experimental evaluation of the two-rail shear test for composite materials. Report No. UW-CMRG-R-98-105.
23. Alloying Surfaces Inc., 1998. The Surfalloys Process, Troy, MI.
24. Davim, J.P. (Ed.), 2013. Machining Composite Materials. New York, NY: Wiley-ISTE.
25. Hocheng, H. (Ed.), 2012. Machining Technology for Composite Materials. Oxford: Woodhead Publishing.
26. Hercules Inc., 1990. Tensile and compression specimen fabrication and testing. Technical Service Bulletin TS-002, Magna, UT.
27. CMRG, 1996. Tabbing guide for composite test specimens. FAA Grant No. 94-G-009, Composite Materials Research Group. Laramie, WY: University of Wyoming.
28. SACMA SRM 1-88, 1989. Compressive Properties of Oriented Fiber-Resin Composites. Arlington, VA: Suppliers of Advanced Composite Materials Association.
29. Boeing BSS 7260, 1988. Advanced Composite Compression Tests. Seattle, WA: The Boeing Company.
30. Hercules Inc., 1990. Mechanical test methods for fiber reinforced resin-matrix composites. Document E23004DT06005, Revision F, Magna, UT.
31. Mahishi, J.M., Adams, D.F., 1984. Analysis of neat resin cracking induced by rapid moisture loss. Composites Technology Review 6 (4), 159–8163.
32. Springer, G.S., 1981. Environmental Effects on Composite Materials. Westport, CT: Technomic Publishing Company.
33. Cairns, D.S., Adams, D.F., 1983. Moisture and thermal expansion properties of unidirectional composite materials and the epoxy matrix. Journal of Reinforced Plastics
and Composites 2 (4), 239–255.
34. Coguill, R.J., 1998. Laboratory Manager, Composite Materials Research Group. Laramie, WY: University of Wyoming.
35. Sutton, M.A., Orteu, J.-J., Schreir, H.W., 2009. Image Correlation for Shape, Motion, and Deformation Measurements. New York, NY: Springer.
36. Trilion Quality Systems, 2017. Available at: www.trilion.com
37. Correlated Solutions, 2017. Non-Contacting Measurement Solutions. Available at: www.correlatedsolutions.com
38. Dantec Dynamics, 2017. Available at: www.dantecdynamics.com/digital-image-correlation
39. Blaber, J., Adair, B., Antoniou, A., 2015. Ncorr: Open-source 2D digital image correlation Matlab software. Experimental Mechanics 55 (6),
40. ASTM B 831-14, 2014. Standard Test Method for Shear Testing of Thin Aluminium Alloy Products. West Conshohocken, PA: ASTM International.
41. ASTM E 647-e1, 2015. Standard Test Method for Measurement of Fatigue Crack Growth Rates. West Conshohoken, PA: ASTM International.
42. ASTM D 8067-17, 2017. Standard Test Method for In-Plane Shear Properties of Sandwich Panels Using a Picture Frame Fixture. West Conshohoken, PA: ASTM
International.
43. ASTM D 3039-14, 2014. Test Method for Tensile Properties of Polymer Matrix Composite Materials. West Conshohocken, PA: ASTM International.
44. ASTM D 2105-01, 2014. Test Method for Longitudinal Tensile Properties of “Fiberglass” (Glass-Fiber-Reinforced Thermosetting-Resin) Pipe and Tube. West
Conshohocken, PA: ASTM International.
45. ASTM D 638-14, 2014. Test Method for Tensile Properties of Plastics. West Conshohocken, PA: ASTM International.
46. Oplinger, D.W., Gandhi, K.R., Parker, B.S., 1982. Studies of tension test specimens for composite material testing. Report AMMRC TR 82-27. Watertown, MA: Army
Mechanics and Materials Research Center.
36 Test Methods for Mechanical Properties

47. Abdallah, M.G., Westberg, R.L., 1987. Effect of tab design of the ASTM D 3039 tension specimen on delivered strength for HMS 1/3501-6 graphite/epoxy material.
Proceedings of the Spring Conference on Experimental Mechanics, Society for Experimental Mechanics, Bethel, CT, pp. 362–366.
48. Adams, D.F., Welsh, J.S., 1997. The Wyoming combined loading compression (CLC) test method. Journal of Composites Technology and Research 19 (3), 123–133.
49. Welsh, J.S., Adams, D.F., 1997. Current status of compression test methods for composite materials. SAMPE Journal 33 (1), 35–43.
50. Adsit, N.R., Chait, R., Papirno, R. (Eds.), 1983. Compression Testing of Homogeneous Materials and Composites, ASTM STP 808. West Conshohocken, PA: American
Society for Testing and Materials.
51. Adams, D.F., 1995. In: Harmston, D., Carson, R., Bailey, G.D., Riel, F.J. (Eds.), Current status of compression testing of composite materials. Proceedings of the 40th
International SAMPE Symposium, Anaheim, CA, vol. 40, pp. 1831–1843. Covina, CA: SAMPE (Chapter 192).
52. Berg, J.S., Adams, D.F., 1989. An evaluation of composite material compression test methods. Journal of Composite Technology and Research 11 (2), 41–46.
53. Haberle, J.G., Matthews, F.L., 1993. The influence of test method on the compressive strength of several fiber-reinforced plastics. Journal of Advanced Materials 25 (1),
35–45.
54. Lamothe, R.M., Nunes, J., Chait, R., Papirno, R. (Eds.), 1983. Compression Testing of Homogeneous Materials and Composites, ASTM STP 808. West Conshohocken,
PA: American Society for Testing and Materials.
55. ASTM D 3410-16, 2016. Test Method for Compressive Properties of Polymer Matrix Composite Materials With Unsupported Gage Section by Shear Loading. West
Conshohocken, PA: ASTM International.
56. Crasto, A.S., Kim, R.Y., 1991. Compression strengths of advanced composites from a novel mini-sandwich beam. SAMPE Quarterly 22 (3), 29–39.
57. Kim, R.Y., Crasto, A.S., 1992. A longitudinal compression test for composites using a sandwich specimen. Journal of Composite Materials 26 (13), 1915–1929.
58. Anquez, L., 1994. In: Newtonin, C. (Ed.), Proceedings of the 30th Polymer Matrix Composites Coordination Group Meeting, (MIL-HDBK-17), New Orleans, LO,
pp. 139–146.
59. Wegner, P.M., Adams, D.F., 1998. Verification of the Wyoming combined loading compression test method. Report UW-CMRG-R-98-116, Composite Materials Research
Group. Laramie, WY: University of Wyoming.
60. Park, I.K., 1971. Tensile and compressive test methods for high-modulus graphite fiber-reinforced composites. Proceedings of the International Conference on Carbon
Fibres, Their Composites and Applications, Paper No. 23. London: The Plastics Institute.
61. Adams, D.F., Lewis, E.Q., 1991. Influence of specimen gage length and loading method on the axial compression strength of a unidirectional composite materials.
Experimental Mechanics 31 (1), 14–20.
62. Xie, M., Adams, D.F., 1995. Effect of specimen tab configuration on compression testing of composite materials. Journal of Composites Technology and Research 17 (2),
77–83.
63. Xie, M., Adams, D.F., 1995. Study of three- and four-point shear testing of unidirectional composite materials. Composites 26 (9), 653–659.
64. Tan, S.C., 1992. Stress analysis and the testing of Celanese and IITRI compression specimens. Composites Science and Technology 44, 57–70.
65. Bogetti, T.A., Glimpse Jr., J.W., Pipes, R.B., 1988. Evaluation of the IITRI compression test method for stiffness and strength determination. Composites Science and
Technology 32, 57–76.
66. Port, K.F., 1982. The compressive strength of carbon fiber reinforced plastics. RAE Technical Report 82083. Farnborough: Royal Aircraft Establishment.
67. Purslow, D., Ceilings, T.A., 1972. A test specimen for the compressive strength and modulus of unidirectional carbon fibre reinforced plastic laminates. RAE Technical
Report 72096. Farnborough: Royal Aircraft Establishment.
68. Ewins, P.D., 1971. Tensile and compressive test specimens for unidirectional carbon fiber reinforced plastics. RAE Technical Report 71217. Farnborough: Royal Aircraft
Establishment.
69. Hsiao, H.M., Daniel, I.M., Wooh, S.C., 1995. Journal of Composite Materials 29 (13), 1789–1806.
70. Cunningham, M.E., Schoultz, S.V., Toth Jr., J.M., 1985. Effect of End-Tab Design on Tension Specimen Stress Concentrations, ASTM STP 864. West Conshohocken, PA:
American Society for Testing and Materials, pp. 253–262.
71. Hart-Smith, L.J., 1991. Making Aerospace Composites & Materials 3 (4), 13–17.
72. Adams, D.F., Odom, E.M., 1991. Influence of specimen tabs on the compressive strength of a unidirectional composite material. Journal of Composite Materials 25 (6),
774–786.
73. ASTM D 695-15, 2015. Compressive Properties of Rigid Plastics. West Conshohocken, PA: ASTM International.
74. Hofer Jr., K.E., Rao, P.N., 1977. A new static compression fixture for advanced composite materials. Journal of Testing and Evaluation 5 (4), 278–283.
75. WTF, 2000. Product Catalog No. 106, Wyoming Test Fixtures, Laramie, WY.
76. Irion, M.N., Adams, D.F., 1981. Compression creep testing of unidirectional composite materials. Composites 12 (2), 117–123.
77. DIN Standard 65 380, 1991. Compression Test of Fiber Reinforced Aerospace Plastics: Testing of Unidirectional Laminates and Woven-Fabric Laminates. Koln: Deutsches
Institut fur Normung.
78. Berg, J.S., Adams, D.F., 1988. An evaluation of composite material compression test methods. Report UW-CMRG-R-88-106. Laramie, WY: University of Wyoming.
79. Westberg, R.L., Abdallah, M.G., 1987. An experimental and analytical evaluation of three compressive test methods for unidirectional graphite/epoxy composites. Report
RI-86B12100. Magna, UT: Hercules Inc.
80. Shockey, P.D., Waddaups, M.E., 1966. Strength and modulus determination of composite materials with sandwich beams. Report No. F2M4691. Fort Worth, TX: General
Dynamics Corp.
81. ASTM D 5467-97, 2015. Test Method for Compressive Properties of Unidirectional Polymer Matrix Composites Using a Sandwich Beam. West Conshohocken, PA: ASTM
International.
82. ASTM C 364-16, 2016. Test Method for Edgewise Compressive Strength of Flat Sandwich Constructions. West Conshohocken, PA: ASTM International.
83. ASTM D 3518-13, 1994. Practice for In-Plane Shear Stress–Strain Response of Unidirectional Polymer Matrix Composite Materials by Tensile Test of 745 Laminate.
West Conshohocken, PA: ASTM International.
84. ASTM D 5379-12, 2012. Test Method for Shear Properties of Composite Materials by the V-Notched Beam Method. West Conshohocken, PA: ASTM International.
85. ASTM D 4255-15a, 2015. Guide for Testing In-Plane Shear Properties of Composite Laminates. West Conshohocken, PA: ASTM International.
86. ASTM D 5448-16, 1993. Test Method for In-Plane Shear Properties of Hoop Wound Polymer Matrix Composite Cylinders. West Conshohocken, PA: ASTM International.
87. ASTM D 3914-02, 2016. Test Method for In-Plane Shear Strength of Pultruded Glass-Reinforced Plastic Rod. West Conshohocken, PA: ASTM International.
88. ASTM D 2344-16, 2016. Test Method for Apparent Interlaminar Shear Strength of Parallel Fiber Composites by Short-Beam Method. West Conshohocken, PA: ASTM
International.
89. ASTM D 3846-08, 2015. Test Method for In-Plane Shear Strength of Reinforced Plastics. West Conshohocken, PA: ASTM International.
90. ASTM D 4475-02, 2016. Test Method for Apparent Horizontal Shear Strength of Pultruded Reinforced Plastic Rods by the Short-Beam Method. West Conshohocken, PA:
ASTM International.
91. ASTM D 732-10, 2010. Test Method for Shear Strength of Plastics by Punch Tool. West Conshohocken, PA: ASTM International.
92. Adams, D.F., Lewis, E.Q., 1995. Current status of composite material shear test methods. SAMPE Journal 31 (1), 32–41.
93. Lee, S., Munroe, M., 1986. Evaluation of in-plane shear test methods for advanced composite materials by the decision analysis technique. Composites 17, 13–20.
94. Adams, D.F., Thomas, R.L., 1969. The solid rod torsion test for the determination of unidirectional composite shear properties. Textile Research Journal 39 (4), 339–345.
95. Chatterjee, S.N., Wung, E.C.J., Yen, C.F., et al., 1991. Composite specimen design analysis I: Analytical studies, volume II: Experimental effort. Report MTL TR 91-5.
Watertown, MA: Army Materials and Mechanics Research Center.
Test Methods for Mechanical Properties 37

96. Sumsion, H.T., Rajapakse, Y.D.S., 1979. Simple torsion test for determining the shear moduli of orthotropic composites. Composite Technology Review 1, 8.
97. Walrath, D.E., Adams, D.F., 1983. Analysis of the stress state in an iosipescu shear test specimen. Report UWME-DR-301-102-1. Laramie, WY: University of Wyoming.
98. Adams, D.F., Lewis, E.Q., 1997. Experimental assessment of four composite material shear test methods. Journal of Testing and Evaluation 25 (2), 174–181.
99. Iosipescu, N., 1967. New accurate procedures for single shear testing of metals. Journal of Materials 2, 537–566.
100. Hussain, A.K., Adams, D.F., 1999. The Wyoming-modified two-rail shear test fixture for composite materials. Journal of Composites Technology and Research 21 (4),
215–223.
101. Whitney, J.M., Browning, C.E., 1985. On short beam shear tests for composite materials. Experimental Mechanics 25 (3), 294–300.
102. Busse, J.M., Adams, D.F., 1998. Investigation of modifications to the short beam shear test method for composite materials. Report UW-CMRG-R-98-111. Laramie, WY:
University of Wyoming.
103. Adams, D.F., Lewis, E.Q., 1995. Experimental study of three- and four-point shear test specimens. Journal of Composites Technology and Research 17 (4), 341–349.
104. Lewis, E.Q., Adams, D.F., 1991. An evaluation of composite material shear test methods. Report UW-CMRG-R-103. Laramie, WY: University of Wyoming.
105. Carlsson, L.A., Adams, D.F., Pipes, R.B., 2014. Experimental Characterization of Advanced Composite Materials, Issued by CRC Press, fourth ed. Boca Raton, FL: Taylor
& Francis Group.
106. Aero Research Ltd., 1945. A fighter fuselage in synthetic material. Technical Notes, Bulletin No. 34, Duxford, Cambridge.
107. ASTM D 790-15e2, 2015. Test Methods for Flexural Properties of Unreinforced and Reinforced Plastics and Electrical Insulating Materials. West Conshohocken, PA:
ASTM International.
108. ASTM C 1341-13, 2013. Test Method for Flexural Properties of Continuous Fiber-Reinforced Advanced Ceramic Composites. West Conshohocken, PA: ASTM
International.
109. ASTM D 4476-14, 2014. Test Method for Flexural Properties of Fiber Reinforced Pultruded Plastic Rods. West Conshohocken, PA: ASTM International.
110. ASTM D 1184-98, 2012. Test Methods for Flexural Strength of Adhesive Bonded Laminated Assemblies. West Conshohocken, PA: ASTM International.
111. ASTM C 393-16, 2016. Test Method for Flexural Properties of Flat Sandwich Constructions. West Conshohocken, PA: ASTM International.
112. O’Brien, T.K., Johnston, N.J., Raju, I.S., Morris, D.H., Simonds, R.A., 1987. In: Johnston, N.J. (Ed.), Comparisons of various configurations of the edge delamination test
for interlaminar fracture toughness. Toughened Composites, ASTM STP 937. West Conshohocken, PA: American Society for Testing and Materials, pp. 199–221.
113. Crews Jr., J.H., Reeder, J.R., 1988. A mixed-mode bending apparatus for delamination testing. NASA Technical Memorandum 100662. Hampton, VA: NASA-Langley
Research Center.
114. ASTM D 5528-13, 2013. Test Method for Mode I Interlaminar Fracture Toughness of Unidirectional Fiber-Reinforced Polymer Matrix Composites. West Conshohocken,
PA: ASTM International.
115. ASTM D 6115-97, 2011. Standard Test Method for Mode I Fatigue Delamination Growth Onset of Unidirectional Fiber-Reinforced Polymer Matrix Composites. West
Conshohoken, PA: ASTM International.
116. ASTM D 6671-13e1, 2015. Standard Test Method for Mixed Mode I-Mode II Interlaminar Fracture Toughness of Unidirectional Fiber Reinforced Polymer Matrix
Composites. West Conshohoken, PA: ASTM International.
117. ASTM D 7905-14, 2014. Standard Test Method for Determination of the Mode II Interlaminar Fracture Toughness of Unidirectional Fiber-Reinforced Polymer Matrix
Composites. West Conshohoken, PA: ASTM International.
118. Li, J., Lee, S.M., Lee, E.W., O’Brien, T.K., 1997. Evaluation of the edge crack torsion (ECT) test for mode III interlaminar fracture toughness of laminated composites.
Journal of Composites Technology and Research 19 (3), 174–183.
119. Sauereisen, M.M., 1998. Inorganic Hi-Temp Cements. Sauereisen, Inc., Pittsburgh, PA.
120. Cotronics Corporation, 1995. High Temperature Epoxy Electrical and Structural Adhesives Handbook. Cotronics Corporation, Inc., Brooklyn, NY.

Further Reading
Adams, D.F., Finley, G.A., 1996. Experimental study of thickness-tapered unidirectional composite compression specimens, Experimental Mechanics 36 (4), 345–352.
Komanduri, R., 1993. Machining fiber-reinforced composites, Mechanical Engineering 115 (4), 58–64.
Lau, W.S., Lee, W.B., 1991. A comparison between EDIVI wire-cut and laser cutting of carbon fibre composite materials, Materials & Manufacturing Processes 6 (2), 331–342.
Vishay Measurements Group. Stress analysis strain gages. Available at: http://www.vishaypg.com/micro-measurements/stress-analysis-strain-gages/

Relevant Websites
https://www.astm.org
ASTM International.
https://www.cmh17.org
Composite Materials Handbook (CMH-17).
https://www.instron.us
Instron, Inc.
https://www.vishay.com/micro-measurements
Micro-Measurements.
https://www.mts.com
MTS Systems Corporation.
https://www.sem.org
Society for Experimental Mechanics.
https://www.wyomingtestfixtures.com
Wyoming Test Fixture.

Das könnte Ihnen auch gefallen