You are on page 1of 10

Mechanics of Materials 46 (2012) 113–122

Contents lists available at SciVerse ScienceDirect

Mechanics of Materials
journal homepage: www.elsevier.com/locate/mechmat

Thermodynamic analysis of fatigue failure in a composite laminate


M. Naderi, M.M. Khonsari ⇑
Department of Mechanical Engineering, Louisiana State University, Baton Rouge, LA 70803, USA

a r t i c l e i n f o a b s t r a c t

Article history: We put forward a thermodynamic approach for analyzing fatigue failure in a composite
Received 3 March 2011 laminate. We show that fatigue is an irreversible progression of increasing entropy that
Received in revised form 15 December 2011 accumulates until it reaches a critical value called the fracture fatigue entropy (FFE) at
Available online 24 December 2011
the onset of failure. Extensive series of fatigue tests are carried out that involve load-con-
trolled tension-tension, and displacement controlled fully-reversed bending fatigue with
Keywords: three different stress ratios as well as constant- and variable-loading. The role of hysteresis
Fatigue
energy in the entropy generation is investigated. FFE values are calculated based the exper-
Fracture
Woven Glass/Epoxy laminate
imental data obtained for temperature and hysteresis energy of a woven Glass/Epoxy (G10/
Thermodynamic entropy FR4) laminate. The concept of tallying entropy accumulation and the use of FFE are useful
for determining the fatigue life of composite laminates undergoing cyclic loading.
Ó 2011 Elsevier Ltd. All rights reserved.

1. Introduction gue strength failure criterion and related it to the number


of cycles to failure. Fawaz and Ellyin (1994) reported a
Composite materials—owing to high strength and stiff- semi-log linear relationship between the applied stress
ness, low weight, and high fatigue life—are widely used and the number of cycles to failure. Their analysis was
in the industry and their applications are steadily growing. capable of predicting the fatigue life of a composite lami-
These materials are anisotropic and inhomogeneous and nate for different load ratios and load directions by the
thus exhibit significantly more complicated behavior than use of a reference S–N curve. According to Philippidis and
their metal counterparts, particularly during repetitive cyc- Vassilopoulos (1999), however, this criterion is very sensi-
lic loadings. In general, the fatigue degradation behavior of tive to the choice of the reference curve. Huang (2002) pre-
a composite laminate is characterized by a combination of sented a fatigue life prediction model for plain-woven glass
matrix cracking, delamination, fiber/matrix debonding, and fabric–reinforced polyester composites subjected to biaxial
fiber breakage (Mao and Mahadevan, 2002; Varvani-Farah- repetitive loading. There are a number of other investiga-
ani et al., 2007). Thus, the complexities of multi-mode fati- tors who have also attempted to characterize the fatigue
gue damage mechanisms in composite laminates have failure based on S–N curve method for woven fabric–rein-
hampered the development of a general fatigue failure cri- forced composites. See, for example, Owen and Griffiths
terion (Varvani-Farahani et al., 2007). (1978), Amijima et al. (1991), and Khan et al. (2002).
A survey of literature reveals that there are a number of One of the promising approaches for estimating the fa-
notable research works devoted to the prediction of fatigue tigue life of composite materials is based on energy consid-
life in composite laminates. One of the earliest models of fa- eration. Ellyin and Elkadi (1990) postulated that the strain
tigue life prediction is due to the work of Hashin and Rotem energy can be utilized to predict fatigue failure and
(1973) followed by Sims and Brogdon (1977) who based showed that the fatigue life is related to the cyclic en-
their analyses on the conventional S–N curve method. They ergy—a combination of number of cycles at failure and
replaced the static strength failure criterion with the fati- material constants—through a power-law relation. Using
the critical plane approach (the plane in which cracks ini-
⇑ Corresponding author. Tel.: +1 225 578 9192; fax: +1 225 578 5924. tiate and propagate), Plumtree and Cheng (1999) and Pet-
E-mail address: khonsari@me.lsu.edu (M.M. Khonsari). ermann and Plumtree, (2001) developed a fatigue failure

0167-6636/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.mechmat.2011.12.003
114 M. Naderi, M.M. Khonsari / Mechanics of Materials 46 (2012) 113–122

criterion model. Natarajan et al. (2005) also put forward a


fatigue model based on the strain energy density by per-
forming tension–tension and bending fatigue experiments
on fiber–reinforced polymeric composites. Later, Shokrieh
and Taheri-Behrooz (2006) proposed a fatigue life model
based on the energy method for unidirectional polymer
composite laminates under tension–tension and compres-
sion–compression fatigue loading. To account for the effect
of off- and on-axis strain energy on fatigue life, Varvani-
Farahani et al. (2007) proposed a fatigue life model for
three different damage stages: matrix cracking, fiber ma-
trix interface delamination, and fiber breakage.
Most of the fatigue failure prediction models discussed
above are typically performed under constant amplitude
and simplified loading conditions. To date, a unified fatigue
criterion for a composite laminate subjected to variable
loading remains lacking. Fig. 1. Energy balance, entropy production and entropy flow in an
The premise of this paper is that the strain energy density arbitrary control volume.
associated with permanent degradation during fatigue must
be accompanied by irreversible entropy gain in accordance respectively. By convention, energy added to the control
with the second law of thermodynamics (Bryant et al., volume is assumed to be positive.
2008). Thus, it is hypothesized that entropy—a fundamental In terms of specific quantities, Eq. (1) can be written as
parameter that characterizes disorder—offers a natural follows:
measure for assessment of deterioration (Basaran and Yan, du
q ¼ w  r  Jq ð2Þ
1998; Doelling et al., 2000; Amiri and Khonsari, 2010, dt
2012; Beheshti and Khonsari, 2010; Aghdam et al., 2011; where w is the rate of volumetric mechanical work, Jq is
Naderi and Khonsari, 2011). More recently, Naderi et al. heat flux across the boundary, q is the density, and u de-
(2010). have demonstrated that by tallying the entropy notes the specific internal energy.
production in two metallic specimens (Al 6061-T6 and SS The time rate of internal energy, q du , is composed of two
dt
304) up to the fracture point, one arrives at a unique mate- terms: (i) energy associated with thermal capacity, qc @h ,
@t
rial property—a so-called fracture fatigue entropy (FFE). where c denotes specific heat and h represents tempera-
Within the range of experimental conditions, this parameter ture; (ii) damage energy, ed. The energy associated with
was found to be independent of geometry, load, and thermal capacity is responsible for rising the temperature
frequency. In other words, the necessary and sufficient of the specimen while damage energy is responsible for cre-
condition for a metal to fracture due to cyclic loading is that ation of new surfaces, internal cracks, changes in material
its accumulated production of entropy reaches a certain micro-structures (Taylor and Quinney, 1934; Halford,
level. In this paper, we quantify the mechanical entropy gain 1966; Kaleta et al., 1990). Equation (2) can be written as:
in a displacement-controlled bending fatigue and load-  
@h
controlled tension–tension fatigue in a Glass/Epoxy (G10/ w ¼ r  J q þ qc þ ed ð3Þ
@t
FR4) laminate subjected to either constant- or variable-
amplitude loading and determine the associated FFE. For an open system, which is allowed to exchange heat
The outline of the paper is as follows. In Section 2, we with its environment (Fig. 1), the total exchange of entropy
present the fundamental of the theory associated with en- (dtS) consists of two terms: the entropy flow into or out of
tropy accumulation and introduce the concept of FFE. In the system, deS, and the entropy production within the sys-
Section 3, experimental procedure to assess fatigue life tem, diS (Prigogine, 1967; Basaran and Nie, 2004), i.e.,
prediction using entropy concept is presented. Results of dt S di S de S
entropy calculation are discussed in Section 4, followed ¼ þ ð4Þ
dt dt dt
by the conclusions.
The entropy exchange with the surroundings, deS, can be
either positive or negative. But according to the second
2. Theory and formulation law of thermodynamics (Clausius–Duhem inequality) the
rate of entropy generation within the system must be
The first law of thermodynamics states that the internal non-negative, i.e.,
energy within an arbitrary control volume changes only if
di S P 0 ð5Þ
energy flows into (or out of) the control volume (Fig. 1) (Pok-
rovski, 2005) in accordance to the following expression: In terms of the specific quantities, Eq. (4) can be written as
(Naderi et al., 2010):
dE dW dQ qs_ ¼ r  J s þ c_ ð6Þ
¼ þ ð1Þ
dt dt dt
where s represents the total entropy per unit mass. Js is the
where E is total energy, t denotes time, Q and W are heat entropy flow and c denotes the volumetric entropy
flow and work across the boundary of the control volume, production.
M. Naderi, M.M. Khonsari / Mechanics of Materials 46 (2012) 113–122 115

Using the definition of Helmholtz free energy, 3. Material and experimental procedure
W = u  h  s, the energy balance and the entropy produc-
tion Eqs. (2) and (6) can be rewritten as (Naderi et al., The material studied is Glass/Epoxy (G10/FR4) which is
2010) an unbalanced woven fabric composite with plain weave
and aligned configuration and consists of a continuous fil-
qðW_ þ hs _ ¼ r  J q þ w
_ þ hsÞ ð7Þ ament glass cloth with an epoxy resin binder. The plain
woven glass fabric is stacked in 15 and 24 layers with
_ þ hsÞ
w qðW _ J q  rh two different thicknesses of 3 and 4.85 mm. Each woven
c_ ¼   P0 ð8Þ layer has two unidirectional layers stacked in [0°/90°]. This
h h h2
type of a composite offers high tensile and flexural
By introducing the specific heat capacity, c, and the Fou- strength (see Table 1) and thus finds use in a variety of
rier’s heat conduction law (Jq = k r h) into Eqs. (7) and applications such as electrical equipment, aerospace struc-
(8), one can obtain the general form of the heat conduction tures, and rocket structural components. Specimens are
equation and entropy production inequality as follows prepared with on- and off-axis stacking sequences. In on-
(Naderi et al., 2010 ; Voyiadjis and Faghihi, 2011).): axis stacking, the warp and weft directions are aligned
with the load direction. The former is called lengthwise
kh;ii þ w ¼ qch_ ð9Þ (0°) while the latter is called crosswise (90°). In off-axis
stacking, the angle between the warp and the load direc-
w ed k tion is (a = 15°, 30°, 45°, 60° and 80°, in this study); See
c_ ¼ þ þ 2 h;i h;i P 0 ð10Þ Fig. 2a. Fig. 2a also presents a schematic diagram of the
h h h
experimental setup for tension–tension fatigue test. The
where k is the thermal conductivity. apparatus used is MTS 810 servohydraulic single actuator.
Eq. (10) consists of a combination of three terms: the Sinusoidal fatigue loads are defined in Multipurpose Test-
Ware (MPT) software which controls the fatigue test and
mechanical  dissipation due to permanent deformation
c_ mech ¼ wh , internal variable evolution due to, for example, applies with a frequency in the range of 5–15 Hz and a load
 
damage energy c_ int¼ ehs and the thermal dissipation due ratio of 0 and 0.1. Constant and variable loads are applied
to heat conduction c_ cond ¼ hk2 h;i h;i . in both the on-axis (0° and 90°) and the off-axis (15, 30,
The fracture fatigue entropy (FFE), cf, can be obtained by 45, 60 and 80°) directions. Variable loading tests include
integrating Eq. (10) up to the time tf when failure occurs both high-to-low and low-to-high sequences.
(Naderi et al., 2010): Fig. 2b shows the specimen dimensions used in a fully
reversed bending fatigue test. The apparatus is a compact,
Z tf  
w ed k bench-mounted unit with a variable-speed motor, vari-
cf ¼ þ þ 2 h;i h; i dt ð11Þ
able-throw crank connected to the reciprocating platen,
0 h h h
with a failure cut-off circuit in a control box, and a cycle
Eq. (11) reveals that the cyclic hysteresis energy (w) is a counter. The variable throw crank is infinitely adjustable
crucial parameter for determining FFE. In metals, w—the from 0 to 50.8 mm to provide different levels of stress
so-called cyclic plastic energy—is a function of fatigue amplitude.
parameters such as cyclic strain hardening, fatigue ductil- Test conditions are summarized in Table 2. Tension–
ity coefficient, and fatigue strength coefficient. To estimate tension fatigue tests are carried out with a total of 62 con-
the FFE for a metallic component, one can make use of the stant loads and 16 variable loads (high-to-low and low-to-
cyclic plastic energy formula such as the Morrow’s equa- high) with 3 and 4.85 mm-thick specimens in the range of
tion (Morrow, 1965) for w. Theoretical procedure and 2–10 kN load amplitude. For the bending fatigue tests, 16
experimental verification of thermodynamic fracture en- constant amplitude loads with the 3 mm-thick specimens
tropy was recently reported by the authors (Naderi et al., are performed.
2010; Naderi and Khonsari, 2010). However, in composite The instrumentation includes a high-speed, high-reso-
materials, the hysteresis energy variation is much more lution infrared (IR) thermography used to record the tem-
complicated than their metal counterparts and, to the best perature evolution of the specimen during the entire
of our knowledge, there are no expressions available to experiment. The IR camera is a MIKRON M7500 with tem-
relate cyclic hysteresis energy to composite fatigue proper- perature range between 0 °C and 500 °C, resolution of
ties. It should also be pointed out that in metals, internal 320  240 pixel, accuracy of ±2% of reading, sensitivity/
variables evolution represent only 5–10% of the entropy NETD of 0.08 °C at 30 °C, and image update rate of 7.5 Hz.
generation owing to mechanical dissipation and is often Before fatigue testing, the surface of the specimen is coated
negligible (Clarebrough et al., 1955a,b; Halford, 1966). with black paint to increase the thermal emissivity of the
However, in composite materials, the evolution of internal specimen surface.
variables represents nearly 30–50% of mechanical entropy
generation (Gamstedet et al., 2002). Due to the importance
of internal variable term, the effect of damage energy on 4. Results and discussion
the fracture fatigue entropy must be considered. However,
in this work, only the experimentally-determined hystere- A series of tension–tension and bending fatigue tests
sis energy is used for evaluating the FFE. The detailed of are performed under constant and variable amplitude
this approach is described in Section 4.1.1. loading. All tests are conducted until fracture occurs to
116 M. Naderi, M.M. Khonsari / Mechanics of Materials 46 (2012) 113–122

Table 1
Mechanical properties of G10/FR4.

Tensile strength (MPa) Flexural strength (MPa) Elastic modulus in flexure (GPa)
Lengthwise Crosswise Lengthwise Crosswise Lengthwise Crosswise
275 240 380 310 18 15

Fig. 2. (a) and (b) Specimen geometry and schematic diagram of the experimental setup in tension–tension fatigue test and specimen geometry in bending
fatigue test (all dimensions are in mm).

Table 2
Fatigue test conditions.

Fatigue test Frequency (Hz) Specimen thickness (mm) Number of tests Load amplitude
Constant load Variable load
Tension–tension 5, 10, 15 3, 4.85 62 10 2–10 (kN)
Bending 10 3 16 – 20–50 (mm)

determine the fatigue fracture entropy (FFE) associated 4.1.1. Hysteresis energy
with each specimen. In the following sections the results Fig. 3 shows the area within the hysteresis loop—the so-
of tension–tension and bending fatigue tests are discussed. called strain energy—increases during a series of tension–
tension fatigue tests. The hysteresis area of each cycle is
4.1. Tension–tension fatigue obtained using a MATLAB™ code that utilizes the experi-
mental load and displacement data between loading and
In the following sections, we present the results of unloading paths. Typical experimentally-determined hys-
experimentally-obtained hysteresis energy and tempera- teresis curves at 5% of the total life, 50% of the total life,
ture evolution that are necessary for thermodynamic anal- and just before failure are also presented in Fig. 3. Exami-
ysis followed by experimental entropy generation nation of the strain energy reveals three distinct stages.
evaluation. The first stage is limited to about 5–10% of the total life;
M. Naderi, M.M. Khonsari / Mechanics of Materials 46 (2012) 113–122 117

tion of Fig. 5a and b reveals that similar to the temperature


rise, the entropy production undergoes three separate
stages: an initial rise (Stage I), slow and steady increase
(Stage II), and a drastic increase prior to failure (Stage
III). During Stage I, entropy accumulates due to three
mechanisms: the high energy released at the tips of mi-
cro-cracks in multiple locations in the matrix between lay-
ers, debonding at the cross-over points between warp and
weft fibers and matrix, and the breakage of some fibers
with low strength. Degradation grows and entropy gener-
ation piles up from the beginning of the test and continues
to the onset of Stage II. Due to a comparatively small num-
ber of cycles in this phase, the accumulation of entropy is
not appreciable. As soon as the deterioration in the mate-
Fig. 3. Accumulation of hysteresis energy per cycle during tension– rial reaches the saturation level, the trend of degradation
tension fatigue test at different constant load amplitudes, frequency of changes. At this stage, the process of entropy production
10 Hz, and R = 0. The hysteresis energy evolution has three phases: initial
reduces due to a lower rate of heat generation and a slower
increase, gradual and slow rise and abrupt rise.
damage process. However, entropy accumulation in Stage
II is more pronounced than that of the previous stage due
during the second stage, gradual and slow increase in to a higher span of fatigue life there (around 55% of total
strain energy occurs and lasts about 70–75% of total life; life).
and the third stage the fibers tend to break and high strain
energy is released.

4.1.2. Temperature, entropy evolution, and entropy


production
Fig. 4 shows the surface temperature recorded using an
IR camera plotted versus the number of cycles-to-failure
for a series of tension–tension fatigue test at frequency of
10 Hz and R = 0. The temperature plotted in this figure is
measured at the cross section where failure occurs. It
shows that, under the conditions tested, temperature in-
creases rapidly over the first 20% of the total life, followed
by a more gradual increase. Finally, a sudden increase of
temperature takes place and the specimen fractures. A
similar trend of the temperature evolution for woven and
Polypropylene/Glass fiber composite is reported by Toubal
et al. (2006), Paulo et al. (2010), Naderi et al. (2011).
We now turn our attention to the entropy production
rate and entropy accumulation calculated based on Eqs.
(10) and (11) for tension–tension fatigue test. The results
of entropy generation are plotted in Fig. 5a and b. Examina-

Fig. 5. (a) Entropy production during tension-tension fatigue test of


Fig. 4. Evolution of surface temperature increase during tension-tension 5.25 kN load amplitude at the frequency of 10 Hz, and R = 0, respectively.
fatigue test at different constant load amplitudes, frequency of 10 Hz, and The entropy production undergoes three stages: initial increase, gradual
load ratio R = 0. Temperature increases initially, continues with slow and and slow rise and abrupt rise. (b) Entropy accumulation. Entropy
steady increase, and rises prior to failure. accumulates during fatigue until reaches the critical value at the fracture.
118 M. Naderi, M.M. Khonsari / Mechanics of Materials 46 (2012) 113–122

At the beginning of Stage III, the fatigue damage behav-


ior and the degree of irreversibility vary due to the change
of degradation modes from matrix cracking to fiber break-
age. More energy is released due the occurrence of fiber
breakage during this stage. Both entropy production and
its accumulation also increase with time until failure oc-
curs. It is also to be noted that changing the degradation
mode affects the heat dissipation rate and consequently
the rate of entropy production.
Fig. 6a–c depicts the results of FFE versus number of cy-
cles for tension-tension fatigue tests. Tests are carried out
for two different load ratios of R = 0 and 0.1. The results
of entropy accumulation subjected to variable loading
(high-to-low and low-to-high load) are also shown in
Fig. 6a–c. It is to be noted that according to an order of
magnitude analysis, entropy generation owing to the
mechanical dissipation is dominant and the entropy gener-
ation due to heat conduction is negligible (see Appendix A).
Hence, in the calculation of FFE, heat conduction effect is
neglected. The Results cover different frequencies within
the range of 5 to 15 Hz, constant- and variable-load ampli-
tude, and off-axis angles of 0°, 15°, 30°, 45°, 60°, 80°, and
90°. Within the range of operating conditions tested, FFE
is independent of frequency, load and fiber orientation.
The accumulated fracture energy (total hysteresis en-
ergy) and FFE are plotted in Fig. 7a and b for all the exper-
imental tests performed at different load ratios, load
amplitudes, and load angles. The accumulated fracture en-
ergy (AFE) is calculated by summing up the hysteresis en-
ergy per cycle
  to the total number of cycles at failure
PN f
wt ¼ j¼1 wi . Fig. 7a and b shows that the predictions
of the fatigue life based on AFE and FFE results have similar
trends. Fig. 7 a also reveals that AFE of on-axis and off-axis
angles close to warp and weft directions have different
trends than those of the off-axis angles far from warp
and weft directions. Failure in the weft and the warp direc-
tions and in off-axis angles close to on-axis are mostly
dominated by the normal stress. However, the shear failure
is mostly dominated by shear stress at the angles far from
the warp and the weft directions.

4.2. Bending fatigue

A series of fully reversed bending fatigue tests is carried


out for two different fiber orientations (0° and 90°) at the
frequency of 10 Hz. In these series of tests, the dissipated
strain energy—which is later utilized to determine the en-
tropy accumulation—is estimated using an experimental
procedure described by Meneghetti (2007) and Meneghetti Fig. 6. Fracture fatigue entropy versus the number of cycles to failure for
and Quaresimin (2011). This procedure involves measuring different tension–tension fatigue tests of Glass/Epoxy (G10/FR4) laminate
the cooling rate after a sudden interruption of the fatigue with different specimen thicknesses, frequencies and loads. a) Fracture
fatigue entropy in crosswise (90°) direction. Constant loads are in the
test. Following the work of Meneghetti (2007) and Mene-
range of 4.9–9.5 kN and variable loads are 5.15–6 kN and 6–5.15 KN. (b)
ghetti and Quaresimin (2011), the statement of energy bal- Fracture fatigue entropy in lengthwise (0°) direction. Constant loads are
ance reads: in the range of 5–13.5 kN and variable loads are 5.7–6 KN and 6–5.7 kN.
Fracture fatigue entropy is within the range of 1–3 MJ m3 K1. (c) Off-
@h axis fracture fatigue entropy of 15°, 30°, 45°, 60° and 80°. Loads are in the
w ¼ h þ qc þ ed ð12Þ
@t range of 2.5–5 kN.

where w is the input mechanical energy. The parameter h


denotes the rate of the dissipated thermal energy in a unit the rate of variation of internal energy which consists of
volume due to conduction, convection and radiation. The two terms. The first term depends on the temperature
remaining two terms on the right-hand side account for variation of the material. The second term, ed, is the time
M. Naderi, M.M. Khonsari / Mechanics of Materials 46 (2012) 113–122 119

Fig. 8. Determination of dissipated thermal energy by experimental


measurement of the cooling rate.

rate of variation of the so-called stored energy of cold


work, which is the part of the mechanical input energy
responsible for creation of new surfaces, internal cracks,
and changes in material micro-structures (Bledzki et al.,
1997; Halford, 1966; Kaleta et al., 1991; Taylor and Quin-
ney, 1934).
Referring to Fig. 8, upon suddenly stopping the fatigue
test at time t⁄ when the surface temperature reaches h⁄,
the mechanical input power and the rate of variation of
cold work, ed, in Eq. (12) drop to zero. Hence, just after
t = (t⁄)+ Eq. (12) can be written as:


@h
Fig. 7. Comparison of fatigue life prediction models at various load ratios h ¼ qc ð13Þ
(R = 0, and 0.1), and load/fiber angles (a = 0°, 15°, 30°, 45°, 60°, 80°, 90°). @t t¼ðt Þþ
(a) Accumulated fracture energy concept. (b) Fracture fatigue entropy
criterion.
Measuring the temperature change as a function of time
during the cooling period after stopping the fatigue test

Fig. 9. (a) Comparison of cyclic strain energy based on the MTS results and the results of measuring the cooling curve for tension–tension fatigue test at the
frequency of 10 Hz, R = 0, and 5.75 kN load amplitude. (b) Dissipated energy variation during bending fatigue of G10/FR at the frequency of 10 Hz and
displacement amplitude of 38.1 mm.
120 M. Naderi, M.M. Khonsari / Mechanics of Materials 46 (2012) 113–122

provides the right-hand-side of Eq. (13). The thermal en- 4.2.1. Validation of hysteresis energy estimation
ergy per unit volume per cycle, q, is: In order to examine the validity of Eqs. (12)–(14) for
determining the hysteresis energy, we first apply the pro-
h
q¼ ð14Þ cedure described above to a tension–tension fatigue test
f
and then compare the results to the hysteresis energy ob-
where f is the frequency of the test. tained directly using the MTS instrument.

Fig. 10. SEM images of the surface of the specimen and the fracture surface for bending fatigue test at the frequency of 10 HZ and displacement amplitude
of 38.1 mm. (a) and (b) SEM image of the surface of the specimen around 85% of total life. (c) and (d) SEM image of the surface of the specimen around 95%
of total life. (e) SEM image of the fracture cross section of woven Epoxy/Glass laminate. SEM images show different failure mechanism such as matrix
cracking, debonding between fiber strands and matrix, delamination between warp and weft fiber strand at the cross-over points, and fiber breakage.
M. Naderi, M.M. Khonsari / Mechanics of Materials 46 (2012) 113–122 121

Fig. 9a shows the strain energy evolution of tension–


tension fatigue test at the frequency of 10 Hz, R = 0, and
5.75 kN load amplitude. The curve-fitting toolbox of MAT-
LAB™ is used to generate the cooling curves from which
the rate of drop in temperature, @h @t
, is calculated and used
in Eq. (14) to evaluate the cyclic strain energy. The results
show that dissipated energy obtained by stopping the
fatigue test is in good agreement with the experimental
results of MTS. Since the bending fatigue apparatus is not
capable of measuring hysteresis loop, the method is ap-
plied for bending tests to determine the strain energy per
cycle for different series of tests.

4.2.2. Hysteresis energy in bending fatigue


The results of cyclic dissipated energy obtained by stop-
Fig. 11. Fracture fatigue entropy versus the number of cycles to failure for
ping the fatigue test versus number of cycles to failure is different fully reversed bending fatigue tests of Glass/Epoxy (G10/FR4)
plotted in Fig. 9b for bending fatigue test at the frequency laminate for different loads (30–50 mm displacement amplitudes) and
of 10 Hz and displacement amplitude of 38.1 mm. To dem- frequency of 10 Hz.
onstrate the failure mechanisms based on the distinct
laminates subjected to fatigue load. As demonstrated in
stages in Fig. 9b, SEM images of the surface of the specimen
this work (Figs. 6 and 11), the entropy generation increases
as well as the fracture cross section are shown in Fig. 10a–e
during the fatigue life toward a final value of cf. Thus, frac-
for bending fatigue test of G10/FR4 at the frequency of
ture fatigue entropy, FFE can be utilized as an index of fail-
10 Hz and displacement amplitude of 38.1 mm.
ure. Successful implementation of this concept for
The SEM images of Fig. 10a–e also show the details of
monitoring the fatigue fracture of metals has been recently
matrix cracking, debondings, and delamination of the weft
demonstrated (Naderi and Khonsari, 2010).
and the warp strands. Fig. 10a and b shows the SEM images
of the specimen surface captured after completing approx-
imately 85% of the total life and Fig. 10a and b shows the 5. Conclusions
SEM images of the specimen taken after about 95% of total
life is depleted. The SEM image of the fracture cross section A thermodynamic approach for characterization of wo-
is shown in Fig. 10e. Examination of Fig. 10a and b reveals ven Glass/Epoxy (G10/FR4) laminate degradation is pro-
that some of the debonded regions join each other at inter- posed which utilizes the entropy generated during the
face between the warp and the weft yarns. Debondings con- entire life of the specimens undergoing tension–tension
tinue to propagate and delamination takes place between and fully-reversed bending fatigue tests subjected to con-
warp and weft bundles at the fabric cross-over points as stant and variable loadings. The results reveal that fatigue
well as between the adjacent layers in the Glass/Epoxy lam- failure prediction by tallying entropy criteria is indepen-
inate. As fatigue process progresses to around 95% of the to- dent of loading direction, constant load and variable ampli-
tal life (Figs. 10c–e), breakages at the delaminated areas tude load (high-to-low and low-to-high load), and loading
occur and fiber strands are separated from the surface. Deb- frequency. Both accumulated fracture energy and accumu-
ondings and delaminations are the origin of more cracks in lated fracture entropy have the similar variation and can
the pure matrix region. These cracks also propagate into be considered as a methodology for prediction of fatigue
some of the fiber bundles as shown in the SEM image of life. However, the results of entropy accumulation tend
the fracture cross section (Fig. 10e). to be more unified than that of accumulated fracture en-
The results of dissipated energy for different series of ergy due to the consideration of the temperature in its for-
bending fatigue for two different on-axis fiber directions mulation (Eq. (11)). The implication of these findings is
along with temperature data are then used to calculate that the evolution of entropy generation can be utilized
the FFE presented in Fig. 11. The abscissa of Fig. 11 shows to assess the severity of degradation of the specimen and
the entropy generation at fracture, cf, and the ordinate guard against fatigue failure.
shows the number of cycles to failure. Examination of the
results reveals that by using the FFE parameter, all the re- Appendix A. An order of magnitude analysis
sults of the experimental data are independent of load
amplitude and fiber orientation. An order of magnitude analysis is performed to show
The results presented in Figs. 6 and 11 demonstrate the that entropy generation due to mechanical dissipation is
validity of the hypothesis that the cumulative entropy at dominant and that entropy generation due to heat conduc-
failure for Glass/Epoxy (G10/FR4) laminate for tension– tion is negligible. Referring to Eq. (10)
tension and bending fatigue tests are independent of load-  2
w k Dh
ing direction, and loading type (constant load and variable c_  þ 2  c_ mech þ c_ cond ðA:1Þ
h h Dy
amplitude load).
A possible application of the proposed hypothesis of the where Dh represents the temperature difference between
constant entropy gain at failure is in the development of a two cross sections at a distance, Dy with one of the
methodology for prevention of the failure of composite sections being the failure section (Fig. A.1). The tension–
122 M. Naderi, M.M. Khonsari / Mechanics of Materials 46 (2012) 113–122

Doelling, K.L., Ling, F.F., Bryant, M.D., Heilman, B.P., 2000. An experimental
study of the correlation between wear and entropy flow in machinery
components. J. Appl. Phys. 88, 2999–3003.
Ellyin, F., Elkadi, H., 1990. A fatigue failure criterion for fiber reinforced
composite laminate. Compos. Struct. 15, 61–74.
Fawaz, Z., Ellyin, F., 1994. Fatigue failure model for fiber–reinforced
materials under general loading conditions. J. Compos. Mater. 28,
1432–1451.
Gamstedet, E., Redon, O., Brondsted, P., 2002. Fatigue dissipation and
failure in unidirectional and angle-ply glass/fiber/carbon fiber hybrid
laminates. Key Eng. Mater. 221–222, 35–48.
Halford, G.R., 1966. Energy required for fatigue. J. Mater. 1, 3–18.
Hashin, Z., Rotem, R., 1973. A fatigue failure criterion for fiber reinforced
materials. J. Compos. Mater. 7, 448–464.
Huang, Z.M., 2002. Fatigue life prediction of a woven fabric composite
subjected to biaxial cyclic loads. Compos. Part A – Appl. Sci. Manuf.
Fig. A.1. Schematic of temperature gradient in scale analysis. 33, 253–266.
Kaleta, J., Blotny, R., Harig, H., 1990. Energy stored in a specimen under
fatigue limit loading conditions. J. Test. Eval. 19, 326–333.
Table A.1 Kaleta, J., Blotny, R., Harig, H., 1991. Energy stored in a specimen under
Order of magnitude of tension-tension fatigue test. fatigue limit loading conditions. J. Test. Eval. 19, 326–333.
Khan, R.H., Khan, Z., Al-Sulaiman, F., Merah, N., 2002. Fatigue life
% of total life (Nf) Dh (°C) h1 (°C) w (kJ m3) c_ mech
estimates in woven carbon fabric/epoxy composites at non-ambient
c_ cond
temperatures. J. Compos. Mater. 36, 2517–2535.
5 0.1 22.4 42.4 66,782 Mao, H., Mahadevan, S., 2002. Fatigue damage modelling of composite
60 0.3 47.7 54 10,253 materials. Compos. Struct. 58, 405–410.
90 3 62.5 75 150 Meneghetti, G., 2007. Analysis of the fatigue strength of a stainless steel
based on the energy dissipation. Int. J. Fatigue 29, 81–94.
Meneghetti, G., Quaresimin, M., 2011. Fatigue strength assessment of a
short fiber composite based on the specific heat dissipation. Compos.
tension fatigue test of 5 kN load amplitude, the load ratio Part B – Eng. 42 (2), 217–225.
Morrow, J., 1965. Cyclic plastic strain energy and fatigue of metals. ASTM
of zero, and the frequency of 10 Hz is considered for this
STP 378, pp. 45–87.
analysis. The hysteresis energy data obtained using MTS Naderi, M., Amiri, M., Khonsari, M.M., 2010. On the thermodynamic
instrument, temperature data from IR camera and entropy of fatigue fracture. Proc. R Soc. A 466, 423–438.
Dy = 1.2 mm are substituted in the Eq. (A.1) to calculate Naderi, M., Khonsari, M.M., 2010. Real-time fatigue life monitoring based
on thermodynamic entropy. Struct. Health Monitor., doi:10.1177/
the effect of mechanical dissipation and thermal dissipa- 1475921710373295.
tion on the entropy production. Table A.1 summarizes Naderi, M., Kahirdeh, A., Khonsari, M.M., 2011. Dissipated thermal energy
the scale analysis in details. and damage evolution of Glass/Epoxy using infrared thermography
and acoustic emission. Composite Part B, doi:10.1016/
Examination of the results show that entropy genera- j.compositesb.2011.08.00.
tion due to heat conduction is negligibly small to that of Naderi, M., Khonsari, M.M., 2011. A comprehensive fatigue failure
the mechanical dissipation. Hence, Eq. (A.1) reduces to criterion based on thermodynamic approach. J. Compos. Mater.,
doi:10.1177/0021998311419540.
w Natarajan, V., GangaRao, H.V.S., Shekar, V., 2005. Fatigue response of fabric–
c_   c_ mech ðA:2Þ reinforced polymeric composites. J. Compos. Mater. 39, 1541–1559.
h
Owen, M.J., Griffiths, J.R., 1978. Evaluation of biaxial stress failure surfaces
for a glass fabric reinforced polyester resin under static and fatigue
References loading. J. Mater. Sci. 13, 1521–1537.
Paulo, N.B.R., Ferreira, J.A.M., Richardson, M.O.W., 2010. Fatigue damage
characterization by NDT in polypropylene/glass fibre composites.
Aghdam, A.B., Beheshti, A., Khonsari, M.M., 2011. On the fretting crack Appl. Compos. Mater., doi:10.1007/s10443-10010-19172-10449.
nucleation with provision for size effect. Tribol. Int., doi:10.1016/ Petermann, J., Plumtree, A., 2001. A unified fatigue failure criterion for uni-
j.triboint.2011.10.001. directional laminates. Compos. Part A – Appl. Sci. Manuf. 32, 107–118.
Amijima, S., Fujii, T., Hamaguchi, M., 1991. Static and fatigue tests of a Philippidis, T.P., Vassilopoulos, A.P., 1999. Fatigue strength prediction
woven glass fabric composite under biaxial tension torsion loading. under multiaxial stress. J. Compos. Mater. 33, 1578–1599.
Composites 22, 281–289. Plumtree, A., Cheng, G.X., 1999. A fatigue damage parameter for off-axis
Amiri, M., Khonsari, M.M., 2010. On the thermodynamics of friction and unidirectional fiber reinforced composites. Int. J. Fatigue 21, 849–856.
wear – A review. Entropy 12, 1021–1049. Pokrovski, V.N., 2005. Extended thermodynamics in a discrete-system
Amiri, M., Khonsari, M.M., 2012. On the role of entropy generation in approach. Eur. J. Phys. 26, 769–781.
processes involving fatigue. Entropy 14, 24–31. doi:10.3390/ Prigogine, I., 1967. Etude Thermodynamique des Processus Irreversible,
e14010024. fourth ed. BelgiumLiege, Desoer.
Basaran, C., Yan, C.Y., 1998. A thermodynamic framework for damage Shokrieh, M.M., Taheri-Behrooz, F., 2006. A unified fatigue life model
mechanics of solder Joints. Trans. ASME, J. Electr. Packag. 120, 379–384. based on energy method. Compos. Struct. 75, 444–450.
Basaran, C., Nie, S., 2004. An irreversible thermodynamic theory for Sims, D.F., Brogdon, V.H., 1977. Fatigue behavior of composites under
damage mechanics of solids. Int. J. Damage Mech. 13, 205–223. different loading modes. In: K.L. Reifsnider, K.N. Lauraitis (Eds.), Fatigue
Beheshti, A., Khonsari, M.M., 2010. A thermodynamic approach for of Filamentary Composite Materials, ASTM STP 636, pp. 185–205.
prediction of wear coefficient under unlubricated sliding condition. Taylor, G.I., Quinney, H., 1934. The latent energy remaining in a metal
Tribol. Lett. 38, 347–354. after cold working. Proc. R. Soc. Lon. Ser. – A 143, 0307–0326.
Bledzki, A.K., Gassan, J., Kurek, K., 1997. The accumulated dissipated Toubal, L., Karama, M., Lorrain, B., 2006. Damage evolution and infrared
energy of composites under cyclic-dynamic stress. Exp. Mech. 37, thermography in woven composite laminates under fatigue loading.
324–327. Int. J. Fatigue 28, 1867–1872.
Bryant, M.D., Khonsari, M.M., Ling, F.F., 2008. On the thermodynamics of Varvani-Farahani, A., Haftchenari, H., Panbechi, M., 2007. An energy-
degradation. Proc. R. Soc. A 464, 2001–2014. based fatigue damage parameter for off-axis unidirectional FRP
Clarebrough, L.M., Hargreaves, M.E., Head, A.K., West, G.W., 1955a. Energy composites. Compos. Struct. 79, 381–389.
stored during fatigue of copper. Trans. Am. Inst. Mining Metallurg. Voyiadjis, G.Z., Faghihi, D., 2011. Thermo-Mechanical strain gradient
Eng. 203, 99–100. plasticity with energetic and dissipative length scales. Int. J. Plasticity,
Clarebrough, L.M., Hargreaves, M.E., West, G.W., Head, A.K., 1955b. The doi:10.1016/j.ijplas.2011.10.007.
Energy Stored in Fatigued Metals. Proc. R. Soc. A 242, 160–166.

You might also like