You are on page 1of 57

Fatigue of Foam and Honeycomb Core

Composite Sandwich Structures:


A Tutorial

NITIN SHARMA, RONALD F. GIBSON* AND EMMANUEL O. AYORINDE


Department of Mechanical Engineering, Advanced Composites Research
Laboratory, Wayne State University, Detroit, MI 48202, USA

ABSTRACT: This article is intended to be a tutorial on the subject of fatigue of


foam and honeycomb core composite sandwich structures. First, several different
analytical models for predicting the fatigue life of sandwich composites are
presented. Then representative publications which have reported on the major
failure modes in sandwich beams under dynamic fatigue loading are summarized,
along with several related publications dealing with static and impact loading.
Papers dealing with the effects of loading frequency, environmental factors,
and block loading on the fatigue life of sandwich composites are discussed.
Finally, recent research on different types of non-destruction evaluation (NDE)
techniques employed for failure investigations during fatigue testing of sandwich
structures is reviewed. Conclusions and generalizations that can be drawn from
the literature are presented along with discussions of areas in which further
research is needed.

KEY WORDS: composite sandwich, foam core, honeycomb core, fatigue.

INTRODUCTION

ANDWICH STRUCTURES CONSIST of two thin face sheets having high


S stiffness and high strength which are adhesively bonded to both sides of
a lightweight core sheet. The face sheets are normally made of composite
laminates and the core is typically made of end grain balsa wood or closed
cell polymer foams or honeycombs. The face sheets carry the tensile and
compressive loads, while the core transmits shear loads and serves to hold
the face sheets in positions far away from the neutral axis so as to maximize
the flexural stiffness of the structure. Sandwich structures are extensively

*Author to whom the correspondence should be addressed. E-mail: gibson@eng.wayne.edu


Figure 14 apperas in color online: http: //jsm.sagepub.com

Journal of SANDWICH STRUCTURES AND MATERIALS, Vol. 8—July 2006 263


1099-6362/06/04 0263–57 $10.00/0 DOI: 10.1177/1099636206063337
ß 2006 SAGE Publications

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


264 N. SHARMA ET AL.

used these days in structural applications such as components in spacecrafts,


aircraft structures, marine vessels, transportation structures, tanks, refrig-
erator containers, bridge decks, and car body shells where minimum
structural weight and maximum stiffness/strength is important. In such
applications, composite sandwich structures are often subjected to repetitive
loading which may lead to fatigue failure (e.g., the repetitive loading of
waves on the hull of a ship, the aeroacoustic excitation of a turbine engine
housing by the rotating turbine blades, or the repeated loading of motor
vehicle traffic over a bridge deck). The life-limiting events in sandwich
composite materials and structures under long-term cyclic loading condi-
tions form a complex process. Several damage mechanisms may grow
and interact to alter the state of material, change the stress distribution,
and define the life of the structure.
Although the literature on fatigue of composite sandwich materials is
extensive, there seems to be a need for an article that pulls together
representative literature in a systematic way and acts as a tutorial to
researchers interested in working in this field. Thus, the primary objective
of the present article is to review and analyze representative work done
by different researchers in the area of fatigue of sandwich composites.
The scope is limited to uniform sandwich structures with foam and
honeycomb cores, as these are most widely used in structural application.
Firstly, different analytical models for predicting the fatigue life of sandwich
structures are presented. These models are typically based on either
stress versus no. of cycles (S–N) curves, strength degradation, stiffness
reduction, cumulative damage modeling, or some combination of these
approaches. Then the different failure modes observed in static, impact, and
dynamic fatigue loading in foam cored sandwich structures and honeycomb
cored sandwich structures are discussed. The fatigue life of sandwich
structures is also influenced by various factors such as: loading frequency,
environmental factors such as temperature and moisture, and the nature
of the repetitive loading. Thus, previous work on the characterization
of these effects on the fatigue life of sandwich composites is discussed.
Lastly, different types of nondestructive evaluation (NDE) techniques that
are employed for failure investigation during fatigue testing are discussed
and compared.

ANALYTICAL MODELING

Analytical models for many aspects of mechanical behavior of sandwich


structures and materials (except for fatigue) are described in detail in the
book by Vinson [1]. Perhaps, one reason for this is that the vast majority
of the fatigue-related work on composites has focused on fiber-reinforced

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 265

composites such as composite laminates, not sandwich structures. Such


results would be relevant to fatigue of sandwich structures only when the
mode of failure is face-sheet tension or compression, but the primary
mode of fatigue failure in sandwich structures appears to be core shear.
For example, Degrieck and van Paepegem have recently published a review
article on the fatigue of fiber-reinforced composites [2], but there is no
mention of composite sandwich structures.
From the limited amount of work done on analytical modeling of fatigue
of composite sandwich structures that has been reported in the literature,
most of the analytical approaches seem to be based on a simple lifetime
prediction using S–N curves [3–5], while in other research, fatigue damage
has been evaluated by characterizing residual stiffness, or residual strength
[6–8]. Some work has also been done on cumulative damage modeling [9],
but all of the analytical methods require a large amount of experimental
fatigue test data.

Basic S–N Approach

The models based on this approach extract information from the S–N
curves or Goodman-type diagrams plotted with the help of experimental
data. The approach is to develop some empirical expression which best fits
the S–N curve for a given material. Kanny and Mahfuz [3] derived a simple
expression for predicting the fatigue life of foam core sandwich beams.
Flexural fatigue tests were performed on sandwich beams made up of a glass
fiber skin and a polyvinyl chloride (PVC) core of varying densities. Figure 1
shows the log–log form of S–N diagram (applied bending stress vs. no. of
cycles to failure) for two different sandwich beams referred as H130 and
R260 based on the density of the foam core.
The final equation for the straight line on the S–N curve was derived as:

1
log  ¼  log N þ C  ð1Þ
m
where 1/m is the slope of the straight line S–N curve in a log–log plot.
C  ¼ (log C)/m is the material constant.  is the stress range given by:

 ¼ max  min ¼ ð1  RÞ  max ð2Þ

and

min
R¼ ð3Þ
max

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


266 N. SHARMA ET AL.

200

R260 3Hz
100
Stress (MPa)

80

60
H130 3Hz

40

20
1E+005 1E+006
Number of cycles (N)
Figure 1. Log–log representation of S–N data for sandwich beam tested at 3 Hz [3].

Burman and Zenkert [4,5] also proposed a simple curve fitting S–N
approach using a two-parameter Weibull function and found reasonable
agreement between experimental and analytical results. Two sandwich
configurations were used in experimental investigations; one with a
Divinycell H100 core and the other with a Rohacell WF51 core. The faces
on H100 consisted of four layers of DBL-850 fabric and vinylester 8084. The
WF51 had faces of four layers of epoxy impregnated quasiisotropic glass
fiber prepregs. Four point bend tests were carried out according to ASTM
C393-62 [10] to investigate the shear strength of the samples. The samples
were tested at different loading ratios R ¼ 1, 0.5, 0.1, 0.25, and 0.5, where
R is defined as the ratio of minimum stress to maximum stress in the load
cycle i.e., R ¼  min/ max. The failure mode observed in the experimental
studies was shear failure of the core. Analytically, the fatigue life
representation based on a two-parameter Weibull function was written as:

b
ðN Þ ¼ th þ ð^  th Þe logðN=aÞ ð4Þ

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 267
1
0.9
0.8
0.7
0.6
t /t

0.5
0.4
0.3 R= −1
0.2 R= 0.1
0.1
0
1E+00 1E+01 1E+02 1E+03 1E+04 1E+05 1E+06 1E+07 1E+08
Log N
Figure 2. S/N diagram for beam with Divinycell H100 core and DBL-850 fabric/vinylester
skins. Lines are curve fits according to Equation (1) [4].

where  is the shear stress in the beam for a given number of load cycles
to failure,  th is the fatigue threshold (or endurance limit), ^ is the static
ultimate stress, N is number of cycles to failure, and a, b are the curve fitting
parameters found by minimizing the quadratic error between the test results
and the function. The fatigue threshold  th is defined as the shear stress level
below which no damage will initiate or if damage has already formed, no
further growth or propagation will take place. In the experimental studies
 th can be calculated by setting a limit on the number of load cycles and
monitoring the damage. The load cycle limit here was set to 5  106 cycles.
Fatigue results for one of the experiments in their study are shown as a
standard S–N diagram in Figure 2, but with the load normalized with
respect to static failure load.
There seems to be a reasonable agreement between experimental results
and predictions for the materials used in the study, but this approach
requires a large amount of specific experimental data to calculate  th for
each type of material and at each loading level R and also to determine the
curve fitting parameters a and b. Also, it is not known how robust this
approach would be in the prediction of fatigue failure in sandwich structures
having other geometries or other failure modes.

Strength Degradation Approach

Models based on this approach describe the deterioration of the


initial strength during fatigue life. Sendeckyj [6] briefly reviewed the
residual strength degradation fatigue theories for composite materials.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


268 N. SHARMA ET AL.

Based on similar lines, Dai and Hahn [7] developed a wear-out model,
based on the concept of strength degradation to assess the fatigue behaviors
of sandwich beams. The model requires only two parameters to describe
the strength degradation in fatigue. One parameter represents the rate of
strength degradation and the other a relative fatigue life. A new approach
was used to determine these parameters from static strength distributions
and the fatigue stress–life relations of the sandwich beams.
As per this model, the probability distribution of static strength  s is
described by a two-parameter Weibull distribution:
  
s
Pðs Þ ¼ exp  ð5Þ

where  is a scale parameter and  is a shape parameter.
In constant amplitude fatigue at a maximum fatigue stress  a, the residual
strength  r after n cycles is related to the initial static strength  s, by a
deterministic equation, a wear-out model:
"  #s
r 1=s
s ¼ a þðn  1Þf ð6Þ
a

where s and f are experimental parameters.


The parameter s is the absolute value of the asymptotic slope at long life
on a log–log plot of the S–N curve. So s can describe the rate of strength
degradation. Now fatigue failure occurs when the residual strength
decreases to the maximum fatigue stress (i.e., when  a ¼  r). Thus the
relationship between static strength ( s) and fatigue stress ( a) is given by:

s ¼ a ½1 þ ðn  1Þf s ð7Þ

The resulting fatigue life distribution is then


   a 
a ½1 þ ðn  1Þf  s
PðnÞ ¼ exp  ð8Þ


The residual strength distribution after n cycles also follows from


Equations (5) and (6) as
( "    #s   )
  r 1=s a 1=s s 
r
P ¼ exp  þ fðn  1Þ þ ð9Þ
n   

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 269

Experimentally, the static strength distribution is determined from the


ranked static strength data using the median rank as

i  0:3
Pðsi Þ ¼ 1  ð10Þ
M þ 0:4

where  si is the ith strength and M is the total number of data.

PARAMETER DETERMINATION
Dai and Hahn [7] also proposed a new approach to determine the
degradation parameters s and f. In this approach, any fatigue point is
defined by a set of three numbers ( a,  r, n): the fatigue stress, residual
strength, and number of cycles applied, respectively. A failure in fatigue
is represented by the condition  a ¼  r, and n becomes the number of cycles
to failure. All data points are ranked so that the ith data point ( a,  r, n)i
is given a probability of survival Pi given by Equation (10). With a proper
selection of s and f, an equivalent static strength  ei corresponding to this
data point is calculated by Equation (6). For the given Pi, the corresponding
static strength  si follows from the static strength distribution given by
Equation (5). The strength degradation parameters s and f are then varied
until the mean square difference between  ei and  si for the entire set of data
is minimized. The values of s and f for each possible ranking can be
conveniently obtained by fitting S–N (stress vs. no. of cycles) data using
Equation (7). Once the degradation parameters s and f are known, fatigue life
in terms of the number of cycles to failure can be estimated using Equation
(7). Figure 3 shows the comparison between model predictions and
experimental data for a sandwich beam with Divinycell H100 core at
R ¼ 1. The static strengths and the maximum fatigue stresses were
normalized by the scale parameter. In Figure 3, the actual strength
distribution is based on Equation (5) while the calculated strength
distribution makes use of Equation (6). In case of fatigue, the actual data
are the experimental data and the calculated points are obtained from
Equation (7). The model seems to work quite well in representing the data in
both cases. The implication from the experimental data is that the mode of
failure is core shear except when defects such as face sheet/core disbonds are
present, especially in tension–compression fatigue. The robustness of this
approach in the case of other failure modes is not clear.

Stiffness Reduction Approach

Many investigators have examined the effectiveness of stiffness reduction


approaches in the prediction of fatigue life of composites. One of the reasons

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


270 N. SHARMA ET AL.

0.6

Normalized fatigue stress


0.5

0.4

0.3

0.2

0.1

0
0 1 2 3 4 5 6 7
Log N
Actual Calculated

Figure 3. Fatigue stress–life relation comparison between experimental and model


predictions for a sandwich beam with H100 PVC core [7].

for this is that the residual stiffness is a parameter that can be monitored
nondestructively, and therefore can be related to residual strength and
fatigue life of the specimen. For example, much important work on the
residual stiffness approach as applied to fiber-reinforced composites has
been done by Reifsnider et al. [11]. Wu et al. [12] also selected the residual
stiffness as a parameter to describe the degradation behavior and to predict
the fatigue life. Philippidis and Vassilopoulos [13] concluded that in the
stiffness degradation approach, only a limited amount of data is needed for
obtaining reasonable results. Whitworth [14] also suggested that the stiffness
reduction approach is an accurate way to predict the fatigue life of
composites. However, the applicability of such approaches to sandwich
structures seems to be limited to face sheet tension or compression fatigue,
whereas core shear seems to be the dominant mode of fatigue failure in
foam core sandwich structures.
Along similar lines with work done on fatigue of composites [11–14],
El Mahi et al. [8] developed a model based on the stiffness reduction
approach for sandwich composites and compared the results with experi-
mental data from three-point bending fatigue tests on sandwich material
made of PVC foam core and E-glass/epoxy skins. Fatigue modulus or
stiffness is defined as the ratio between the applied stress and the resulting
strain at a given number of cycles. This modulus is a function of loading
cycles n and applied stress level r. The rate of decrease of fatigue modulus
from an initial static value can be expressed as:

Gf ðnÞ ¼ Go for n  nif


ð11Þ
Gf ðnÞ ¼ Go  Aeðnnif ÞC for n  nif

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 271

1.0

Load (Fmax /F0 max) 0.8

0.6

0.4

0.2

0.0
0.0 0.2 0.4 0.6 0.8 1.0
Number of cycles (N /Nf)
Figure 4. Normalized load vs normalized number of cycles in sandwich composites under
displacement control fatigue with logarithmic fit [8].

where nif is the number of cycles to initiate damage.


Two different equations were developed for experiments in displacement
control and load control. But only the model for displacement control
experiments will be discussed here. The model for load control experiments
can be derived along similar lines [8]. In these tests the displacement control
dmean is the static mean displacement (midspan deflection in the three-point
bend test) and dam is the amplitude of the applied sinusoidal waveform.
Decrease in load (stiffness) according to the number of cycles is recorded.
For example, Figure 4 is for a mean displacement dmean ¼ 0.5 du, where du is
the value of the failure displacement in the static tests and for an amplitude
dam ¼ 1.75 mm. The load has been normalized with respect to the maximum
applied load in the first cycle and the number of cycles has been normalized
with respect to cycles to failure.
The results in Figure 4 show a three-step load loss: (i) an initial stage
characterized by rapid load reduction; (ii) an intermediate stage in which
an additional load reduction occurred at a much slower rate; and (iii) a final
stage, in which rapid load reduction is observed as specimen failure is
approached. Load reduction is related to the decrease of the flexural fatigue
modulus.
The load reduction can be expressed as a logarithmic function:

Fmax
¼ 1  Ad lnðnÞ ð12Þ
Fo max

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


272 N. SHARMA ET AL.

where Fo max is the maximum applied load in the first cycle and Ad can be
described according to different load levels r by a power function as:

Ad ¼ a0d  rad ð13Þ

where a0d and ad are the parameters that depend on the material properties
and on the loading conditions. The load expression according to the number
of cycles and the applied displacement level r can be written as:

Fmax
¼ 1  a0d  rad ð14Þ
Fo max

The parameters a0d and ad can be determined experimentally.


This expression for load reduction is valid for only the first two stages of
the curve shown in Figure 3. Degrieck and van Paepegem [2] have also
mentioned that the stiffness reduction models are not always valid in the
third stage of final failure. Thus the model is not valid up to complete failure
and the failure condition for predicting fatigue life can be predefined
by a certain percentage of load (stiffness) reduction generally referred to as
N where  is the percentage of load reduction, for example N3, N5, N10
criteria.
Hence the expression for predicting fatigue life in displacement control
corresponding to a load reduction of % can be written as:
 
1
N ¼ exp
a0d  rad ð15Þ

where  ¼ 1 
100

A limitation is that this concept cannot be used for displacement levels r


near unity, where the failure occurs very rapidly, and specimen fracture
occurs before significant reduction in load or increase in displacement.

PARAMETER DETERMINATION
Fatigue tests in displacement control at different load levels were carried
out and the results were plotted as load versus the number of cycles as
shown in Figure 5. According to the Equations (12)–(14), logarithmic
functions were fitted into experimental results (shown as solid bold lines).
From these experimental results the parameters used in Equations (12)–(14)
were determined.
Figure 6 shows the coefficient Ad as a function of the level of loading r.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 273

1.0 rd = 60%

rd = 70%
0.8
Load (Fmax /F0 max)

rd = 80%
0.6

0.4
rd = 95%

0.2 rd = 90%

0.0
1 100 10000 1000000
Number of cycles (N)
Figure 5. Stiffness reduction according to the number of cycles for different loading levels
in displacement control [8].

0.020

0.016
Coefficient (Ad)

0.012

0.008

0.004

0.000
0.5 0.6 0.7 0.8 0.9 1.0
Applied displacement level (rd)
Figure 6. Coefficient of interpolation Ad as a function of loading level r in displacement
control [8].

From the value of Ad the values of coefficients a0d and ad can be


determined for different loading levels using Equation (13). Thus, once all
the parameters have been determined, the fatigue life can be predicted using
Equation (15). Experimental and analytical model results were plotted

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


274 N. SHARMA ET AL.

1.0

Applied displacement level (rd)


0.8

0.6

0.4

0.2 Experimental results


Analytical results

0.0
1 100 10000 1000000
Fatigue life (N10)
Figure 7. S–N curves of sandwich composites in displacement control [8].

(Figure 7) with N10 criteria. Strong correlation was found between


experimental and analytical results.
One of the limitations of this work is that the failure initiation and
progression has not been observed experimentally and also there is no
mention about the final failure mode. So again there is concern about the
validity of the approach when the failure mode changes.

Cumulative Damage Modeling

Many cumulative damage theories have been developed for metallic


materials, the most widely used approach being the Palmgren–Miner
criterion, or linear damage rule [15], in which damage is assumed to be
linearly proportional to the fractional life used up at different cyclic stress
levels and fatigue failure is assumed to occur when the sum of these fractions
exceeds unity. But this approach has not been successful for composite
materials because the behavior of composite materials under fatigue loading
is more complex than that of metallic materials (i.e., because there are
many more factors that influence the fatigue crack growth in composites,
including the matrix material, fiber material, volume fraction, and fiber
orientation). However, there have been attempts by different researchers to

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 275

modify the Palmgren–Miner model so that it can be used for composites.


Sendeckyj [6] briefly reviewed the available fatigue accumulation models
and compared the applicability and accuracy of each model, including
the life prediction models proposed by Hashin and Rotem [16], Wang et al.
[17], and Broutman and Sahu [18]. More recently, Epaarachchi and
Clausen [19] have also developed a new cumulative damage model for
glass fiber-reinforced plastic (GFRP) composites. All of these models
are specifically for composite laminates, which can be used as skins in
sandwich structures.
By contrast, there seem to be few publications on the use of cumulative
damage models for composite sandwich materials. Clark et al. [9] reported
on cumulative damage modeling of sandwich beams based on the stiffness
degradation approach under two-step loading. In the experimental
investigation, the core material used in the specimens was Airex C70.130,
a thermoset cross-linked cellular foam, and the skin materials consisted
of hybrid glass/Kevlar/epoxy balanced 0/90 woven construction. Flexural
fatigue tests were carried out by loading the beam at eight equally spaced
points along its length in order to simulate a uniform load. The observed
failure mode was brittle core shear cracking. Firstly, a general damage
model defining the fatigue damage parameter, D, was defined assuming
constant frequency and environmental conditions.

D¼0 @ n ¼ 0,
ð16Þ
D¼1 @ n ¼ Nf ,

For a sequence of ‘m’ loadings:

D ¼ 0 @ n ¼ 0,
Xm
ð17Þ
D¼ Di ¼ 1 @ n ¼ Nf ,
i¼1

where Di is the damage experienced at load level i, n is the number of cycles,


Nf is the number of cycles to failure. The total damage is the summation of
all the damage components at each loading level. Thus for two-step loading,
the residual life of a beam can be determined from the residual damage, Dr.
This is represented schematically in Figure 8. If N1 and N2 are the expected
lives under the first and second loads, respectively, then the remaining
damage can be expressed as

Dr ¼ 1  D12, ð18Þ

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


276 N. SHARMA ET AL.

1.0
Damage

r1 r2

D1=D12

n1 N1 n12 N2

nf
Number of cyles
Figure 8. Schematic illustration of determination of residual life: two-step loading [9].

where D12 is the level of damage experienced under the first stress level and
equated to an amount of damage at the start of loading at the second stress
level. The remaining life is therefore the number of cycles to failure under
the second stress level N2, minus the number of cycles under the second
load that equates to the already accumulated damage level D12 under the
first load.
Different forms of the cumulative damage parameter, D, can be chosen
depending on the degree of linearity of the degradation response. Three
different models were proposed. The first model is linear, based on ‘number
of cycles’, whereas the second model is based on changes of ‘modulus’ and
the third model is based on changes of ‘strain’. Damage is assumed to
initiate when fatigue damage is first observed, i.e., at n ¼ nif. At n ¼ Nf, the
damage is equal to one. For the purpose of all cumulative damage models
investigated, it was assumed that:

DðnÞ ¼ 0, where n  nif ,


0  DðnÞ  1, where nif  n  Nf ð19Þ

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 277

Model 1: The first model was based on basic Palmgren–Miner linear


cumulative damage model. According to this model, the amount of damage
at a given cyclic stress level is equal to the ratio of the number of cycles at
a given stress level to the number of cycles required to cause fatigue failure
at that stress level. In this case, the damage model occurs after the initiation
of damage and can be expressed as:

n  nif
DðnÞ ¼
, where n  nif ð20Þ
Nf  nif
Model 2: The damage function was defined in terms of the fatigue
modulus as:
Go  Gf ðnÞ
DðnÞ ¼
ð21Þ
Go  Gf Nf

where Gf (n) and Go are defined as the transient fatigue modulus and
instantaneous static modulus, respectively, and Gf (Nf) is the fatigue
modulus at failure. The relation between Gf (n) and Go is:

Gf ðnÞ ¼ Go , where n  nif ,


ð22Þ
Gf ðnÞ ¼ Go  Aeðnif ÞC , where n  nif ,

where A and C are material constants to be determined from experimental


data.
Using Equation (22), Equation (21) was modified as:

eðnnif ÞC
DðnÞ ¼ where n  nif ð23Þ
eðNf nif ÞC

Model 3: In this case the damage function was defined in terms of shear
strain as:
ðnÞ  ð0Þ
DðnÞ ¼
ð24Þ
 Nf  ð0Þ
Again using Equation (22) and stress–strain relationships, the damage
function can be modified as:

h r i eðnnif ÞC 
DðnÞ ¼ where n  nif ð25Þ
1  r B  eðnnif ÞC

where B ¼ Go /A.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


278 N. SHARMA ET AL.

1.0
Model I: r=0.7 & r=0.4
0.9 Model II: r=0.7
0.8 Model III: r=0.7
Model II: r=0.4
Damage, D 0.7 Model III: r=0.4

0.6
0.5
0.4
0.3
0.2
0.1
0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Normalized life after initiation of fatigue damage
Figure 9. Cumulative damage models [9].

Damage curves were plotted for each of the three models as shown in
Figure 9.
As observed from the damage curves, the first model is linear, whereas
the other two models are nonlinear. Although there seem to be no direct
comparisons of predicted and measured fatigue life in this study, the
exponential nature of the increasing damage near the end of the fatigue life
is more consistent with the experimentally observed behavior and Model 3
best describes such behavior.
All of the aforesaid models for predicting the fatigue life of sandwich
materials require extensive fatigue data on a specific specimen, and the
results appear to be useful for that specific specimen only. Also it is not
clear that these models will be valid if the failure mode changes. It is hence
concluded that there is a need for further research in this area and
specifically some models should be developed that can predict fatigue failure
based on a more general knowledge of the material and geometrical
properties of the sandwich specimen.

FAILURE MODES

Sandwich composites can fail in several ways including tension or


compression failure of the facings, shear failure of the core, wrinkling failure
of the compression facing, local indentation, debonding of the core/facing
interface, and global buckling. The initiation, propagation, and interaction
of failure modes depend on the type of loading, constituent material
properties, and geometrical dimensions. Hence, this part of the article
emphasizes the study of various failure modes in sandwich composites under
static loading, impact loading, and fatigue loading. The objective of this

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 279

section is to help the reader understand different failure modes in sandwich


structures under different loading conditions and how transitions occur
from one failure mode to another. Different equations are given which can
be used to predict failure loads for a specific failure mode once beam
parameters and dimensions are known. Although the focus of this article is
on fatigue loading, it is also useful to briefly review static failure modes as
well. Under static loading, failure modes in both foam-cored sandwich
composites and honeycomb-cored sandwich composites are discussed.

Static Loading

FOAM-CORED SANDWICH COMPOSITES


A substantial amount of work has been performed on failure modes of
foam-cored sandwich panels under static loading. For example, Allen [20]
and Zenkert [21] have reviewed the basic failure modes in sandwich
structures, Daniel et al. [22] have studied the failure modes in sandwich
beams, Gdoutos et al. [23] have characterized the indentation failure
and Gdoutos et al. [24] have reported on compression face wrinkling
as well. Steeves and Fleck [25,26] studied in detail the failure mechanisms
in sandwich beams made of woven glass-epoxy skins and Divinycell PVC
foam core under static loading. Four main modes of failure for sandwich
beams in three-point and four-point bending were identified: (i) face
yield or face microbuckling, (ii) wrinkling of the compressive face sheet,
(iii) core shear, and (iv) indentation beneath the loading rollers, as shown
in Figure 10.
Based on the geometry of the beam in three-point loading (Figure 11), the
expressions given below for predicting collapse loads for face yield/face
microbuckling, for core shear, and for face wrinkling were developed by
Zenkert [21], while for indentation only the expression was developed by
Steeves and Fleck [25].
In Figure 11, the midpoint of the beam deflects by a transverse
displacement u due to the applied load P of the midroller. L is the beam
length between the supports, H the overhang at each end, b the width of the
beam, c the core thickness, and tf the face thickness. The relevant mechanical
properties of the isotropic core are the Young’s modulus Ec, shear modulus
Gc, compressive strength  c, and shear strength  c; for the face sheets, the
pertinent properties are the axial compressive strength  f and Young’s
modulus Ef, and d ¼ c þ tf.
The predicted collapse load for face yielding or microbuckling is given as:

4f btf d
P¼ ð26Þ
L

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


280 N. SHARMA ET AL.

P P

Core shear Microbuckling

P/2 P/2 P/2 P P/2


P

Indentation Face wrinkling

P/2 P/2 P/2 P/2


Figure 10. Failure modes of a sandwich beam in three-point bending [26].

P,u
tf Ef,σf

c Gc,Ec,σc,Tc

P/2 P/2
L

H H
Figure 11. Geometry of a sandwich beam in three-point bending [26].

whereas the corresponding load for core shear failure is

P ¼ 2c bd ð27Þ

while the load for face wrinkling is

2btf d p
3
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P¼ Ef Ec Gc ð28Þ
L

and the load for indentation is


 2 2 1=3
 c Ef d
P ¼ btf ð29Þ
3L

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 281
0.25
Experiment - indentation failure
Experiment - core shear failure
Experiment - core crush failure
0.2
Line H
Line B
0.15 Line D Core shear
tf /c

Line C Line G
0.1
Line J
Line E
Line A Line F
0.05

Line I
Indentation
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35

c/L
Figure 12. Failure mode map for sandwich beam with H100 PVC foam core and GFRP
faces [26].

Thus, based on these equations, the failure load, P, can be predicted for
each of the failure modes once the beam parameters and dimensions are
known.
To understand the effect of sandwich beam geometry upon the collapse
mode, the failure mode map shown in Figure 12 was constructed. The map
takes as axes the ratio of core thickness c to span L, and the ratio of face
sheet thickness tf to core thickness c.
There are ten lines, labeled A–J, representing trajectories over which the
functional relationship between failure load and geometric parameters has
been explored. Lines A–H are paths of varying beam length, line I is a path
of varying thickness, and line J is a path of varying core thickness. The
observed failure modes for this sandwich beam in this mode map are
indentation beneath the central roller and core shear. It was observed
that microbuckling and face wrinkling occurred with denser cores like
Divinycell H200.
Konsta–Gdoutos and Gdoutos [27] also studied load and geometry effects
on the failure of sandwich structures. Two types of beams, simply supported
and cantilever, were studied and the material system was carbon/epoxy
facings with high-density PVC foam. The failure modes observed were com-
pression facing wrinkling and core shear failure. From elementary beam theory,

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


282 N. SHARMA ET AL.

Table 1. Values of C for different loading and boundary conditions [27].

P q q

C=2 C=4 C=3√3


q
P

C=1 C=2 C=3

it was shown that the failure mode transition from core shear failure to
compression facing failure occurs when

L Ff
¼C ð30Þ
hf Fcs

where L is the length of simply supported or cantilever beam, hf is the facing


thickness, Ff is the facing strength in compression, Fcs is the core shear
strength, and C is a constant that depends on the form of the beam and
type of the applied load. Values of C for several loading conditions and
boundary conditions are given in Table 1.
When the left hand term of Equation (30) is smaller than the right-hand
term, failure occurs by core shear, whereas in the reverse case failure occurs
by facing compression. Failure mode maps according to Equation (30) are
shown in Figure 13.
Figure 14 shows the variation of the critical load, Pcr, versus span length
for the initiation of core shear failure and compression facing wrinkling for a
simply supported beam loaded by a concentrated load P. The curves are the
theoretical predictions based on material and geometrical properties of the
specimen using Equation (30). The critical load of the beam is the smaller of
the two values predicted by the two failure modes. The two curves of the
critical load versus beam span intersect at a critical span at which the
predicted transition from one failure mode to the other takes place. For
beam spans smaller than the critical span, failure initiation in the beam takes
place by core shear failure; while for beam span greater than the critical
span, predicted failure starts by compression facing wrinkling. Similarly
results were compiled for both simply supported and cantilever beam for
different loading conditions. Table 2 presents the values of the critical beam

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 283

P
P

L L

Core shear
Ff / Fcs

Ff / Fcs
failure
Core shear failure

1 Facing
Facing compression
1
1 compression failure
2 failure
45°
26.56°

L/hf L/hf
Figure 13. Failure envelopes for failure mode transition from core shear failure to
compression facing failure for a simply supported and a cantilever beam subjected to a
uniform load [27].

8
H250
Pcr=2.46/L P Facing wrinkling
6
Pcr=6.45 kN 25.4 (mm)
Pcr (kN)

25.4 (mm)
4 Core shear

H100 Pcr=1.28/L
2
Pcr=2.1 kN

0
0.38 0.62
0 0.2 0.4 0.6 0.8 1
L (cm)
Figure 14. Critical load vs span length for failure initiation for a cantilever sandwich
subjected to a concentrated load [27].

span for which transition from core shear failure to compression facing
wrinkling takes place.
Thus, it was concluded that initiation of a particular failure mode depends
on the constituent material properties, geometry, and type of loading and
that there is a critical beam span at which transition of failure from one

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


284 N. SHARMA ET AL.

mode to another occurs for different loading conditions. It is important


to note that this was a theoretical study only, and experimental validation
of the predictions was not provided.

HONEYCOMB-CORED SANDWICH COMPOSITES


Honeycomb core is a complex structure and failure modes in sandwich
structures with honeycomb cores are different than those in foam-cored
sandwich composites. This is particularly true for the core shear failure
mode. Lee and Tsotsis [28] investigated the indentation failure behavior of
honeycomb panels and Pan et al. [29] studied the shear failure process of
honeycomb cores in sandwich structures. Petras and Sutcliffe [30] studied
in detail different failure modes in honeycomb sandwich beams and also
developed failure mode maps. Sandwich beams made of GFRP laminate
skins and Nomex honeycomb core were used in experimental investigation.
Three-point bend tests were carried out as per ASTM C393-62 [10] standard.
Observed failure modes were divided into two categories: skin failure and
core failure.
Skin failure modes Include: (a) face yielding, (b) intra-cell dimpling, or
(c) face wrinkling as shown in Figure 15, and as characterized by the
equations below from Petras and Sutcliffe [30].

Table 2. Critical values of beam span for failure mode transition from core
shear failure to compression facing wrinkling [27].

Simply supported beam

Divinylcell H100 Lcr ¼ 0.62 m Lcr ¼ 1.24 m Lcr ¼ 1.67 m


Divinylcell H250 Lcr ¼ 0.38 m Lcr ¼ 0.76 m Lcr ¼ 0.99 m

Cantilever beam

Divinylcell H100 Lcr ¼ 0.31 m Lcr ¼ 0.62 m Lcr ¼ 0.93 m


Divinylcell H250 Lcr ¼ 0.19 m Lcr ¼ 0.325 m Lcr ¼ 0.57 m

(a) Face yielding (b) Intra-cell dimpling (c) Face wrinkling


Figure 15. Failure modes in the skin [30].

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 285

(a) Face Yielding: Failure occurs in the top skin due to face yielding when
the axial stress in either of the skins reach the in-plane strength  fY of
the face material for loading along the beam axis,
fx ¼ fY ð31Þ

(b) Intra-cell dimpling: A sandwich with a honeycomb core may fail by


buckling of the face where it is unsupported by the walls of the
honeycomb. Simple elastic plate buckling theory can be used to derive
an expression for the in-plane stress  fi in the skins at which the intra-
cell buckling occurs as
 2
2Efx 2t
fi ¼ 2
ð32Þ
1  fxy 

where  is the cell size of the honeycomb, t is the thickness of skin,


Efx and fxy are the elastic modulus and Poisson’s ratio for the skin
for loading in axial direction, respectively.
(c) Face wrinkling: Face wrinkling is a buckling mode of the skin with
a wavelength greater than the cell width of the honeycomb. Buckling
may occur either into the core cells or the core or outwards, depending
on the stiffness of the core in compression and the adhesive strength.
In practice, with three-point bending, inward wrinkling of the top skin
occurs in the vicinity of the central load. The expression that gives the
critical compressive stress  fw that results in wrinkling of the top skin
was given by Allen [20] as

3
fw ¼
E1=3 2=3
fx E3 ð33Þ
2 2 1=3
12ð3  cxz Þ ð1 þ cxz Þ
where cxz is the out-of-plane Poisson’s ratio and E3, the out-of-plane
Young’s modulus of the honeycomb core.
Core failure modes include: (a) core shear and (b) indentation by local
crushing, as shown in Figure 16.

(a) Core shear (b) Local indentation


Figure 16. Failure modes in the core [30].

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


286 N. SHARMA ET AL.

(a) Core shear: The mean shear stress in the core is given by

W
cxz ¼ ð34Þ
2d

where W is the central load per unit width and d is the distance between the
midplanes of top and bottom skin.
Now core shear failure occurs when the applied shear stress  cxz equals
the shear strength  cs of the honeycomb core.

cxz ¼ cs ð35Þ

As compared to the core shear mechanism in case of a foam core


(discussed in detail in ‘Fatigue Loading’), where core shear initiates with
a crack which propagates leading to failure, no crack is observed in the
honeycomb core and failure occurs as a global shear deformation without
a crack.
(b) Local Indentation: Failure of sandwich panels in three-point bending
can occur at the load point due to local indentation. Failure is due to the
core crushing under the indenter. A simple empirical approach has been
used to model this failure mechanism. Assuming that is the length
of contact between the central roller and the top skin and further
assuming that the load is transferred uniformly to the core over this
contact length, the out of plane compressive stress  z is given by

W
z ¼ ð36Þ

Again W is the load per unit width at central roller. Failure is predicted
when this compressive stress equals the out-of-plane compressive strength
 cc of the honeycomb core, i.e., when

z ¼ cc ð37Þ

Impact Loading

Impact behavior of sandwich composites has been widely studied,


particularly for the case of low velocity impact. For example, Abrate has
published a book [31] and a review article [32] on the subject. It has been
found that impact loading can induce damage to the facings, the core

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 287

material, and the core–facing interface. Low velocity impact damage on


sandwich beams and plates with carbon-epoxy facings, and honeycomb
cores is typically confined to the impacted facing and the core, while the
opposite facing is usually undamaged. In honeycomb cores, damage consists
of crushing or ‘buckling’ of cell walls in a region surrounding the impact.
Lim et al. [33] investigated the transverse impact failure modes in foam core
sandwich beams. The impact tests based on the ASTM D 5942 standard [34]
were performed using specimens with fixed core density and four different
face thicknesses to verify the effect of the face thickness. Also, to determine
the effect of the core density, the specimens with the fixed face thickness
and four different core densities were tested. The material used was made
with woven glass fiber epoxy prepreg as the skin and Divinycell HT
grade foam core. It was found that failure modes under impact loads were
limited to only core shear failure and face fracture mode. It was found
that the thin face and high density core leads to the face compressive
fracture because the stress of the face is high in the case of thin face and
high-density core.
Failure mode transitions with respect to face thickness for the fixed
core density and vice versa were studied, and are shown in Figure 17.
In Figure 17(a), when the core density is 97 kg/m3, the transition skin
thickness corresponding to the failure mode change from core shear to
face failure occurs between 0.72 and 0.87 mm. Also, from the result of
Figure 17(b), the sandwich specimen with a fixed face thickness of 1.02 mm
appears to have a transition of failure modes from core shear to face failure
at a core density between 97 and 117 kg/m3.

Fatigue Loading

Failure modes in fatigue are often similar to those observed in static and
impact loading. But under cyclic loading, sandwich beams are particularly
prone to core shear failure. This is due to the fact that cyclic loading appears
to reduce the residual shear strength of the foam core. This has been
observed in various studies in the literature, and Harte et al. [35] confirmed
it in their study of the fatigue failure of sandwich beams with an aluminum
alloy foam core. Three modes of failure were observed: face-sheet yield,
core shear, and indentation. Face-sheet yield and indentation failure were
limited only to specimens with very low face-sheet strength and thickness.
And it was concluded that the predominant failure mode under cyclic
fatigue is core shear. It was observed that due to the cyclic loading, the shear
strength of the Alporas (aluminum) foam decreased quite significantly,
implying that sandwich beams are particularly prone to core shear under
cyclic loading.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


288 N. SHARMA ET AL.

Face thickness

1.02 mm
Core shear

0.87 mm

0.72 mm

0.58 mm Face failure

(a)

Core density

54 kg/m3 Core shear

70 kg/m3

97 kg/m3

117 kg/m3 Face failure

(b)
Figure 17. Experimental results of transverse impact failure modes with respect to:
(a) face thickness (fixed core density of 97 kg/m3) and (b) core density (fixed thickness
of 1.02 mm) [33].

Kulkarni et al. [36] studied fatigue failure mechanisms and found core
shear to be the mode of failure. The complete core shear crack growth
mechanism was studied in sandwich beams made of glass/epoxy and PVC
foam which will be discussed here in detail. Sandwich panels were
manufactured using the coinjection resin transfer molding (CIRTM) process
to infuse the resin simultaneously in both top and bottom face sheets.
Fatigue tests were carried out in three-point bending. Damage in the core
was analyzed at a microscopic level and the crack growth mechanism was
studied. It was observed that three distinct damage events took place before
the specimen failed.
Damage event (1) was the crack initiation and propagation on the
compression side just below the top face-sheet–core interface. It was
noticed that the crack was located about 1–1.5 mm below the interface.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 289

Figure 18. Damage event (1) during crack growth in fatigue loading of sandwich
beams [36].

Figure 19. Microscopic view of damage event (1) [36].

Careful examinations revealed that the resin from the facings penetrated
into the core material by this depth and the resin-soaked cells, and the dry
cells below the actual core–skin interface created a subinterface. The crack
initiated from this subinterface and propagated parallel to the beam axis
as shown in Figure 18. Damage event (1) occupied about 85% of the fatigue
life. Figure 19 shows the microscopic view of damage event (1).
Damage event (2) was core shear, which followed the event (1). The
propagated crack in the event (1) kinked at a certain distance depending
on the load level and sheared through the core thickness as shown in
Figure 20. This was a rapid event and occupied only about 7–8% of the
fatigue life. The core shear angle was also found to be dependent on the
stress level. The crack reached the bottom face-sheet–core interface at
the end of the event. A microscopic view of damage event (2) is shown in
Figure 21.
Damage event (3) as shown in Figure 22, follows the core shear, and
consists of delamination at the bottom face sheet–core interface causing the
separation of the core from the face sheet. As the core shear propagated
at a faster rate, the energy at the crack tip became sufficiently high that the
crack crossed the subinterface and reached the much stiffer face sheet where

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


290 N. SHARMA ET AL.

Figure 20. Damage event (2) during crack growth in fatigue loading of sandwich
beams [36].

Figure 21. Microscopic view of damage event (2) [36].

Figure 22. Damage event (3) during crack growth in fatigue loading of sandwich
beams [36].

it was deflected along the core–skin interface, rather than along the
subinterface. This was also a rapid event and occupied the remaining 7–8%
of the fatigue life. Figure 23 shows the microscopic view of the damage event
(3). The specimen ultimately failed after this and overall photograph of the
failed specimen is shown in Figure 24. In damage event (3), the face–core
interface on the tension side distinctly debonded from each other, which
also explains the rapid crack growth in this final stage of failure.
Similar core shear failure mode and crack growth has been observed
by Kanny and Mahfuz [3], Burman and Zenkert [4,5], Clark et al. [9],

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 291

Figure 23. Microscopic view of damage event (3) [36].

Figure 24. Failed specimen after cyclic loading [36].

and Shenoi et al. [37] in their studies on fatigue and failure of composite
sandwich materials.
In conclusion, it appears that there are a number of possible failure modes
in sandwich beams under static and impact loading, and the mode of failure
depends on the material properties and loading conditions. However, core
shear failure seems to be the predominant failure mode in cyclic fatigue
loading of foam core sandwich composite materials.

EFFECT OF LOADING FREQUENCY

Several publications have reported that increasing excitation frequency


has a detrimental effect on the fatigue life of sandwich composites due to the

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


292 N. SHARMA ET AL.

corresponding increase in core temperature. The increase in temperature


leads to deterioration of core shear properties, which has been documented
by Challis et al. [38] for a variety of structural foams in the mid-density
range. It was shown that a loss of approximately 3% in core shear properties
resulted from a core temperature increase of 10 C, rising to a 10% loss
in core shear properties for a temperature increase of 20 C. Since core
shear is the main mode of fatigue failure in foam core sandwich structures,
as was concluded in the previous section, the fatigue life of these
structures can be expected to decrease with increasing frequency. This has
also been observed in studies by Burman and Zenkert [4] and Shenoi
et al. [37]. Sharma et al. [39] also supported this view with the help of
flexural test experiments in their fatigue studies of sandwich structures
made of polyurethane foam (PUF) core and fiber-reinforced plastic
(FRP) outer skins. Three types of specimens: epoxy/glass-PUF-epoxy/
glass (EPE), polyester/glass-PUF-polyester/glass (PPP), and epoxy/glass-
PUF-polyester/glass (EPP) were examined. Three-point bend tests according
to ASTMC393-62 [10] were performed at frequencies of 1, 3, and 5 Hz. The
S–N curves were plotted for the three panels at the three test frequencies,
as shown in Figure 25.
It is evident from the plots that there exists a transition point (TP)
that defines the transition between the steady region (SR) and the
deteriorating region (DR). The transition point is marked by a sudden
change in the slope in all the cases. For a given frequency, the TP for
all three types of material was nearly the same. But as the frequency
increased, the TP shifted to lower N (number of loading cycles to failure).
At 1 Hz the TP for all three cases was observed at 7.5  105 cycles.
At higher fatigue frequencies the TP was observed at lower cycles,
i.e., 1.76  103 and 150 cycles for 3 and 5 Hz respectively. After the TP,
the sandwich specimens displayed significant cracking between the foam
core and the skin.
It was concluded that during the cyclic loading, the outer skin
and the foam rubbed against each other at the interface, leading
to an increase in temperature at the mating surfaces. The interface
temperature increased with increases in frequency of fatigue cycles
due to increased sliding and inadequate time for heat dissipation.
In addition to the degradation of the core, the adhesive bond strength at
the interface between the skin and foam also decreases with increases
in temperature at the mating surface; hence stiffness degradation of
all the specimens increases with increases in frequency of fatigue cycles,
leading to premature failure of the specimens. Thus, it was observed
that fatigue life decreases with increases in frequency over the frequency
range 1–5 Hz.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 293
0.4

Fatigue bending stress (MPa)


Frequency = 1 Hz EPE

0.35 EPP

PPP
0.3 Transition point

0.25

0.2
SR DR
0.15
0.4
Frequency = 3 Hz
Fatigue bending stress (MPa)

0.35

0.3
Transition point
0.25

0.2

0.15
SR DR
0.1
0.4
Frequency = 5 Hz
Fatigue bending stress (MPa)

0.35

0.3 Transition point

0.25

0.2

0.15

0.1 SR DR
0.05
1.00E+00 1.00E+01 1.00E+02 1.00E+03 1.00E+04 1.00E+05 1.00E+06 1.00E+07
Fatigue cycles (N)
Figure 25. S–N curves of sandwich panels at different frequencies [39].

Frequency Effect – Contradictory View

Although the publications referred to in the previous section have shown


evidence that the fatigue life of composite sandwich structures decreases
with increasing frequency, Kanny and Mahfuz [3] have published a contra-
dictory view. Extensive experimental work was done on the flexural fatigue
characteristics of sandwich structures at different loading frequencies.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


294 N. SHARMA ET AL.

(a) 200 (b) 200

R260 3Hz R260 at 15Hz

100 100
Stress (MPa)

Stress (MPa)
80 80

60 60
H130 3Hz

40 40 H130 at 15Hz

20 20
1E+005 1E+006 1E+005 1E+006
Number of cycles (N) Number of cycles (N)
Figure 26. Log–log S–N curves for sandwich beams at 3 and 15 Hz, R ¼ 0.1[3].

In this study, glass fiber/vinylester sandwich constructions with closed cell


PVC foam cores of four different densities of core materials were used.
Cyclic flexure three–point bend tests as per ASTM C393-62 [10] were
performed at a stress ratio R ¼ 0.1 and at frequencies of 3 and 15 Hz.
The temperature increase in the core was monitored by using a digital
thermometer with two external probes. One probe was attached to the
core on the free surface directly below the load point and other probe was
used to measure the surrounding air temperature. It was observed that
on increasing the frequency from 3 to 15 Hz, the temperature of the core
increased by around 10 C. But a different effect on the fatigue life was found
as compared to the previous works [4,37–39].
For a given applied bending stress level, the number of cycles to failure N
at 15 Hz was greater than the corresponding number of cycles to failure at
3 Hz. S–N plots for two types of beams tested at 3 and 15 Hz are shown
in Figure 26.
The explanation given for the increase in the number of cycles to failure
at 15 Hz was based on the concept that energy input is equal to energy
dissipated plus energy stored. It was suggested that work or energy input at
any given stress level is constant for both 3 and 15 Hz. The temperature
increase measured in the core of the sandwich beams tested at 3 Hz was
2  1 C while at 15 Hz it was 12  2 C. Thus, some of the energy dissipated
was converted into heat, leaving less available stored energy to fuel the
damage process, and this was proposed to be the reason that beams tested
at 15 Hz survived a greater number of cycles.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 295

The failure mode observed was core shear and numerous minute cracks
were initiated in the core just below the loading point. The cracks grew
together forming a single dominant crack that propagated on the
compression side of the beam, parallel to beam until they reached some
critical length. After reaching this critical length, the crack kinked at an
angle of approximately 45 toward the tension side of the beam. It then
propagated in the core on the tension side delaminating toward the edge of
the beam. The crack growth was monitored and it was concluded that the
crack growth rate was faster at 3 Hz than at 15 Hz. One possible explanation
was attributed to the fact that the heat generated in the core of beams cycled
at higher frequency makes the core more compliant and hence effectively
blunts the crack tip thereby reducing the fatigue crack growth rate.
Thus in the literature, two contradicting views exist on the effect of
loading frequency on the fatigue life of foam core sandwich structures
with each view having possible explanations and supportive experimental
work. Although there is considerable evidence in the literature that there
is a reduction in fatigue life with increasing frequency, the work showing
the opposite effect cannot be ignored and thus more research in this field
is required.

EFFECT OF ENVIRONMENTAL FACTORS

Composite sandwich structures are used extensively in aerospace,


military, and marine applications. In such applications they are often
subjected to varied and extreme environmental conditions. These environ-
mental factors include exposure to variable temperatures and moisture
conditions, especially in the case of marine applications. Thus, the effects of
these two factors on the long-term fatigue life of the composite sandwich
structures is of great importance. This section of the article will focus on the
effect of temperature and moisture on the fatigue life of sandwich composite
structures.

Effect of Temperature

Berkowitz and Johnson [40] studied the effect of temperature on the


fracture toughness and fatigue life of sandwich composite structures
comprised of a Nomex honeycomb core with graphite/epoxy skins.
Specimens were tested in a double cantilever beam (DCB) configuration.
ASTM D5528-94a was used as guideline for fracture toughness testing.
Fatigue testing was performed at 4 Hz and at an R-ratio (Pmin/Pmax) of 0.1
in displacement control. Tests were performed at three different tempera-
tures, hot temperature (HT) (77 C (170 F)), room temperature (RT)

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


296 N. SHARMA ET AL.

(21 C (70 F)), and cold temperature (CT) (54 C (65 F)). The data
showed clearly that test temperature had an effect on fracture toughness.
The CT tests resulted in the highest fracture toughness, and the hot
temperature tests yielded the lowest fracture toughness. The failure was
observed to be always in the honeycomb core. The fracture toughness
results are shown in Table 3.
Fatigue crack growth was also observed at three different temperatures
and results are shown in Figure 27. Although there is an overlap in the data,
conclusions can be drawn that cold temperature resulted in slower crack
growth rates than room temperature. As shown in the figure, HT the data
shows considerable overlap with the RT data. The point averages suggest
that HT resulted in faster growth, but lack of separation of the scatter
bands on the figure causes any conclusions to be unclear.
Erickson et al. [41] studied the effect of temperature on the low velocity
impact behavior of composite sandwich panels. The materials used in the

Table 3. Fracture toughness of graphite/epoxy/Nomex sandwich structure


at different temperatures [40].

Specimens Values Gc,avg(J/m2)

Room temperature (21 C) 6 56 1180


Cold temperature (54 C) 3 62 1620
Hot temperature (77 C) 2 42 1160

Fatigue crack growth rate vs. strain energy release rate


1.0E-03
da/dN (mm/cycle)

1.0E-04
RT (21°C)
CT (−54°C)
HT (77°C)

1.0E-05

Frequency=4 Hz
R=0.1
1.0E-06
100 1000

∆G (J/m2)
Figure 27. Fatigue crack growth data for graphite/epoxy/Nomex sandwich structure at three
different temperatures [40].

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 297

study were made of cross-woven E-glass with 0/90 weave and an epoxy
matrix. Four different core material-thickness combinations were used to
make the sandwich samples. They were 2.54 and 1.43 cm thick cardboard
honeycomb filled with a low density foam, and 2.54 and 1.43 cm thick plain
cardboard honeycomb. The impact tests were carried out at three different
temperatures (25, 25, and 75 C). It was found that low temperature
samples exhibited delamination of the back face sheet, the RT samples
showed delamination as well as fiber breakage on the back face sheet, and
the high temperature samples were completely penetrated.
Kanny et al. [42], specifically investigated the effects of elevated
temperature on the fatigue behavior of foam core sandwich structures.
Specimens were made of two types of PVC cores; one was linear HD foam
and other a cross-linked H foam. Face sheets made of glass/vinylester
composite were used in both types of specimens. Three-point bend flexure
tests were conducted on these sandwich structures at RT, 40 and 80 C.
The tests were conducted on a servohydraulic testing machine equipped
with an environmental chamber. During static tests it was found that the
strength of the sandwich beams decreased with an increase in temperature.
Close inspection revealed that damage occurred mainly in the foam cores
and was more severe in HD beams. The temperature sensitivity was
measured by observing the load displacement responses shown in Figure 28.
It was concluded that linear foam cores are more temperature sensitive than
the cross-linked foam cores.
Fatigue tests were performed at a stress ratio R ¼ 0.1 and a frequency
of 3 Hz and S–N diagrams are shown in Figure 29, where ‘stress’ is the
flexural stress. For both types of beams, the number of cycles to failure at
RT was considerably higher than at elevated temperatures and the fatigue
life decreased with increased temperature.
800 800

600 HD130 at RT 600 H130 at RT


Load (N)

Load (N)

H130 at 40°C
400 400
HD130 at 40°C

200 200
HD130 at 80°C H130 at 80°C

0 0
0 0.004 0.008 0.012 0.016 0.02 0 0.002 0.004 0.006 0.008 0.01
Displacement (m) Displacement (m)
Figure 28. Load–displacement curves at RT 40, and 80 C for: (a) HD and (b) H beams [42].

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


298 N. SHARMA ET AL.

1 1

0.9 0.9
Stress/StressULT

Stress/StressULT
H130 at RT
0.8 RT 0.8
40°C 40°C
80°C
80°C
0.7 0.7

0.6 0.6

0.5 0.5
1000 10000 100000 1000000 10000 100000 1000000
Number of cycles (N) Number of cycles (N)
Figure 29. S–N curves at RT, 40, and 80 C for: (a) HD sandwich beam and (b) H sandwich
beam [42].

When beams were tested at room temperature in cyclic tests, no visible


damage was observed for most of the fatigue life. Just before failure
however, numerous small cracks formed in the core on the compression side
of the beam at some distance from the core–skin interface. These cracks
coalesce into a larger dominant crack, which propagates parallel to the
core–skin interface. Thereafter, the crack branches at approximately 45
propagating to the tension side where it propagates as a debond crack at
the core–skin interface to reach the beam edge. While analyzing the failure
behavior at elevated temperatures, it was observed that the core degraded
quite significantly at high temperatures. The cells close to upper and lower
face sheets collapsed which even led to a reduction in the height of the beams
and hence a volume change. Correspondingly, the deflection in the central
part of the beam increased, leading to failure by gradual collapse of the
foam core. No cracks were observed in the foam core as observed at RT.
At an even higher temperature, plastic yielding of the core occurred and
the resin in the interface area decomposed. This resulted in face/core
delaminations. Though the failure occurred in the core only at both RT and
elevated temperatures, there was a slight change in the mechanism of core
failure as discussed earlier. So the decrease in fatigue life at an elevated
temperature can be partially attributed to the change in the failure mode.
Thomas et al. [43] analyzed dynamic response of sandwich composites at
sub-ambient temperatures, with emphasis on the through-the-thickness
compressive strength. Four closed cell PVC foam cores of variable densities,
75, 130, 260 kg/m3 (4th core also had a density of 130 kg/m3 but a lighter
degree of cross-linking, almost linear) were used as core materials and three
layers of dry plain weave S2-glass fiber preforms with vinylester as resin

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 299

Figure 30. Comparison of room and sub-ambient response: (a) R75 and (b) H130
specimen [43].

were used as skins. The dynamic response of the sandwich specimens was
evaluated at both sub-ambient and RT. In order to achieve sub-ambient
temperatures, specimens were first submerged in liquid nitrogen and sealed
for 30 min. Afterwards, each specimen was removed from liquid nitrogen
(196 C) and immediately placed on the test fixture. Dynamic compres-
sion tests were performed in a Split Hopkinson Pressure Bar apparatus.
A comparison of two responses for two different specimens at similar
rates is depicted in Figure 30. In each category of sandwich specimens,

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


300 N. SHARMA ET AL.

an increase in compressive strength was achieved at sub-ambient conditions.


This increase in strength ranged from 18% for the cross-linked sandwich
specimen and 30% for the linear sandwich specimen. Thus, the conclusion
of this study is that sandwich structures perform better under dynamic
compression at low temperatures than they do at RT.
By comparison with high temperature effects, the literature on low
temperature fatigue behavior of sandwich composites is very scarce, and
there is a need for more research in this field.

Moisture Effects

Sandwich structures used in marine applications are exposed to seawater


for long periods of time. Under longtime exposure, polymers and polymer
composites absorb moisture. It is known that moisture absorption causes
a reduction in the glass transition temperature and corresponding softening
of a polymer. In addition, moisture absorption generates hygroscopic
expansional strains, which may further degrade material properties. For
example, Smith and Weitsman [44] investigated the immersed fatigue
response of polymer composites and found that fatigue life was reduced
drastically when saturated specimens were fatigued in an immersed
environment. Weitsman and Elahi [45] also published a review article on
the effects of fluids on deformation, strength, and durability of polymeric
composites in which similar results were reported.
Along similar lines, Scudamore and Cantwell [46] studied the effects of
moisture on the mechanical behavior of sandwich structures. The material
used in the investigation had E-glass/epoxy as skins and an aluminum
honeycomb core. It was found that prolonged seawater exposure in an
aluminum honeycomb structure caused degradation of the bond between
the epoxy matrix and the aluminum core, facilitating crack advance along
the skin-core interface. Li and Weitsman [47] also investigated the sensitivity
of the material properties of sandwich structures to long-term exposure to
seawater. Divinycell H100 and H200 closed-cell PVC foams were employed
in the study and sandwich specimens with glass/vinylester skins were used.
The fracture toughness of the foam core material and the face/core
debonding fracture toughness in the case of sandwich specimens were
measured. Fracture toughness of wet and dry H100 and H200 PVC foams
was characterized employing single edge notched bending specimens
following ASTM Standard D5045-99 [48]. The specimens were precracked
using a band saw and subsequently sharpened by a knife blade. The wet
H100 specimens were immersed in simulated seawater for 2500 h and the
H200 for 3000 h prior to fracture testing in a servohydraulic testing machine.
Table 4 lists the fracture toughness for the H100 and H200 PVC foams in

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 301

Table 4. Wet and dry fracture toughness of H100 and H200 PVC foams [47].

GIC(J/m2)

H100 H200

Wet 786 1218


Dry 598 1128

wet and dry cases. It was observed that absorption of seawater increased the
toughness of the foam materials by 31 and 8% for the H100 and H200 PVC,
respectively. This may be attributed to the softening and enhanced ductility
of the wet polymeric foam cells as the glass transition temperature was
reduced.
Similarly, for investigating the effects of seawater on face/core debonding
fracture, sandwich panels made of glass/vinylester skins and H100 and
H200 PVC cores were used. Pre-cracks of length a ¼ 50 mm were cut along
the top face/core interface and the samples were loaded. The specimens
were tested after 3000 h of immersion in seawater, when relative weight
gain reached equilibrium level. From the experimental results, it was
concluded that fracture toughness at the core/facing interfaces showed
degradations of about 36% for H100 and 17% for the H200 sandwiches,
respectively, attributed to the presence of seawater.
Ishai et al. [49] also investigated the long-term hygrothermal effects
on damage tolerance of composite sandwich panels made of carbon-
fiber-reinforced plastic (CFRP) and GFRP skins and a syntactic
foam core. The panels were immersed in water at two different tempera-
tures 25 and 50 C for long periods of time. During hygrothermal
exposure, moisture absorption versus time was recorded by weight
measurements to an accuracy of 0.1 mg. After various exposure periods,
samples were subjected to impact testing. It was concluded that there
is a significant strength reduction with moisture content for syntactic
foam specimens. Although impact damage size seems to be affected
only slightly, open-edge sandwich beams exposed to long-term hygro-
thermal conditions tend, in severe cases, to fail prematurely by core shear
failure.
Ishiaku et al. [50] also concluded that mechanical properties of sandwich
composites are degraded by absorption of ambient moisture. No work
on the effects of moisture or seawater exposure specifically on the fatigue life
of sandwich structures was found in the literature and thus more research
in this field is also needed.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


302 N. SHARMA ET AL.

EFFECT OF REPETITIVE LOADING

Most of the fatigue testing work in laboratories involves constant


amplitude sinusoidal loading in three-point or four-point bend tests. But
composite structures are rarely subjected to uniform constant amplitude
loads in service. The load could fluctuate randomly according to a
distribution, creating a load spectrum, or the load could vary in sequential
steps (i.e., block amplitude loading) or the loading could consist of various
combinations of these types of loading. An example could be the gust load
spectrum applied to a composite aircraft wing during flight, or the effects
of ice motion or sea waves on composite sandwich ship hull materials.

Block Loading

As with environmental effects, most of the work on block loading has


been done on composite materials, but not much has been done on sandwich
structures. For example, Harris et al. [51] studied the effects of block loading
conditions on the fatigue life of carbon-fiber-reinforced laminates. It was
shown by the four-unit block loading experiments that a combination of
tension and compression loading led to significant reductions in total fatigue
life. Gamstedt and Sjogren [52] also investigated the sequence effect in block
amplitude loading of carbon-fiber/epoxy cross-ply laminates. In their
experimental investigation it was found that a sequence of high–low
amplitude levels results in shorter lifetimes than a low–high order.
Clark et al. [9] investigated the fatigue behavior of sandwich beams under
two-step and block loading regimes. The core material used in the specimens
was Airex C70.130, thermoset cross-linked cellular foam, and the skin
materials consisted of hybrid glass/Kevlar/epoxy balanced 0/90 woven
construction. Fatigue tests were carried out on beam specimens, which were
loaded at eight equally spaced points along the length, in order to simulate a
uniform load. A frequency of 0.5 Hz was used and applied loads were chosen
between 30 and 80% of the ultimate static load. In some cases, residual
strength tests were undertaken in four-point bending on a separate rig. The
failure mode for the beams was brittle core shear cracking. A combination
of low–high loads and then high–low loads were carried out to investigate
the influence of load sequence. The beams were loaded for 50% of their
average fatigue life at each respective load. Most of the beams did not fail
and were tested for their residual static strength under static loading
conditions. It was found and concluded from the two-step loading tests
that because no fatigue failures occurred, the beams were stronger than
the previous batch that had been used for single loading tests. Some of the
results are shown in Table 5. From the results of the residual strength of

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 303

Table 5. Two-step loading tests; static failure load ¼ 9661 kg [9]: a low/high
load combination and b high/low load combination.

1st load 2nd load Residual


Beam no. (N/mm) (N/mm) 1st cycle 2nd cycle strength (kg)

S01a 28.21 35.48 1,50,000 20,000 9462


S05a 35.39 42.22 20,255 1498 9677
S09a 28.64 42.47 15,00,000 1500 10,044
S20b 42.89 34.89 1500 20,000 9138
S22b 42.41 28.55 1580 150,000 9527
S28b 35.3 28.05 20,000 150,000 7198

the specimens, it was concluded that load sequence affects the fatigue life
and a high/low load combination is more damaging than a low/high load
combination.

Hydromat Test System

To simulate more realistic test conditions, especially for marine structures


where sandwich composites are used in hull materials, Bertelsen [53,54]
developed the Hydromat test system (HTS), a new test system for applying
more realistic distributed loading to two-dimensional sandwich panels, and
which has been approved as ASTM standard D 6416-99 [55]. Ahtonen and
Sikarskie [56] also have done extensive work on analyzing the effectiveness
of the HTS test system and concluded that HTS is both an effective and
efficient test device for sandwich panels. A schematic drawing of the system
is shown in Figure 31.
The heart of the test is a self-contained pressure bladder called a
hydromat, which is made of two square pieces of tough industrial belting
sealed at the edges. The prototype holds about 5 gallons of water and works
with 24  2400 square sandwich panel specimens. A special fixture supports
the panel by its edges in a standard servohydraulic test machine. With the
power on, the actuator begins to squeeze the Hydromat against the surface
of the test panel. One sensor records the deflection of the center of the panel
while the other monitors the contact pressure. The belting fabric follows the
contour of the bulging test panel so that it experiences a nearly uniform,
distributed load, even at significant deflections. The Hydromat system can
be used for both static and fatigue testing, but so far most of the applications
have been to static testing. Bertelsen [53] has presented some results of
block-cycle fatigue tests of sandwich panels using the Hydromat, but so far
there are no published comparisons of Hydromat fatigue test results with

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


304 N. SHARMA ET AL.

Load frame crosshead

Load cell

Upper panel
support frame

LVDT

Journal bearings Corner bolts

Test panel Lower panel


support frame

HTS pressure bladder


Clamps

Load frame platen

Pressure transducer
Bladder support slab

Figure 31. Schematic diagram of the Hydromat test system [56].

those from three-point or four-point bending fatigue tests, and there is


a need for further research in this area.

NONDESTRUCTIVE EVALUATION

NDE for detection of damage initiation and propagation, is extremely


important during fatigue testing. This is particularly true for composite
materials such as sandwich structures, which may exhibit different modes of
failure, each of which may have a different NDE signature. In general, there
is no single NDE method that will adequately cover all aspects of fatigue
damage, and in most cases a combination of NDE methods is necessary for
complete characterization of such damage. This part of the article deals with
the NDE methods that have been used most frequently on sandwich
structures, as described in the following sections.

Vibration Methods

The use of NDE methods for thin composite structures is much


more abundantly documented than for sandwich and thick constructions.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 305

Nylon (2,2) (3,1-1,3)


filament Accelerometer
Impulse PCB 303 A03
hammer

(3,1+1,3) (3,2)

Force transducer
PCB 208 A02 Test plate
(2,3) (1,4)

Conditioning Conditioning
amplifier amplifier
PCB 480A PCB 480A
(4,1) (3,3)

FFT analyzer
HP3582A

(2,4-4,2) (2,4-4,2)
Desk computer Printer
HP9000 Model 332 Epson FX850

Figure 32. Schematic of the vibration identification system, and the first ten modes of plate
vibration [62].

In the latter group, which calls vibration testing more into question by virtue
of the thickness, Ayorinde [57] advanced a method for the elastic
characterization of thick composite plates, employing an inverse method
based on a Timoshenko–Mindlin formulation for the vibration. Gibson [58]
summarized recent research on using the modal vibration response
measurements to characterize composite materials and structures. Liew,
Xiang, and Kitiponchai [59] reviewed vibration work on thick composite
plates, including NDE applications. Salawu [60] summarized vibration-
based damage detection and evaluation research that focused on frequency
change methods. Palozotto et al. [61] used vibration NDE to evaluate
impact damage to sandwich composite structures reinforced in the thickness
direction.
As shown in Figure 32, the vibration identification method involves the
use of impulsive excitation of a plate specimen and fast Fourier transform
analysis of the response to identify modal frequencies and mode shapes,
which can be combined with an analytical model of the vibrating plate to
determine the intrinsic stiffness of the plate. Internal damping of the plate
can also be determined from this method. It has been found that changes
in the plate material or geometry due to damage or degradation result in
corresponding changes in the modal parameters (damping is particularly
sensitive), so this method can be used for NDE. Although most of the work

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


306 N. SHARMA ET AL.

with this method has been concerned with laminated composite plates, there
is no reason why it could not be used on sandwich panels as well. Thus,
Ayorinde et al. [57,62], and others as cited, have shown that both thin and
thick composite plates can be elastically characterized by rapid testing of
samples in completely free vibration modes.
Vinson [1] showed that in cylindrical sandwich shells under asymmetric
loads, inter-laminar shear stresses are restricted to the ‘bending boundary
layer’ (BBL) regions near load discontinuities, with only in-plane membrane
stresses existing elsewhere. This suggests that in using NDE to detect skin/
core debonding, only the BBL regions need to be inspected, and that for
vibration methods, only the higher modes could detect such damage. Zou,
Tong, and Steven [63] reviewed vibration model dependent damage
evaluation methods. Liu and Chen [64] applied the genetic algorithm to
vibration responses of sandwich plates for flaw detection. Kim and Hang [65],
from examinations of skin–core debonding with nondestructive instrumen-
ted impact tests, concluded that natural frequencies were lowered, damping
ratios were enhanced with increasing debond length, and there exists
a critical debond length beyond which the natural frequencies are degraded
disproportionately. Thus, vibration NDE appears to be viable low-cost
method applicable to thick laminate and sandwich composite structures.

Ultrasonic Methods

Ultrasonic investigation of material properties has been widely researched


for several composite materials. Some applications have been made, even in
the less tractable area of sandwich composites. Akay and Hanna [66] used
ultrasonic C-mode scanning to assess damage in Nomex honeycomb and
Rohacell foam. Sachse et al. [67] reviewed works in quantitative ultrasonic
NDE, and summarized new approaches for active and passive methods in
composite materials measurements. Gupta and Sankaran [68] established the
quality of foam core sandwich specimens and checked skin/core debonds
with ultrasonic C-scans. More recently, Legendre, Goyette, and Massicot [69],
among others, utilized wavelet theory in the ultrasonic NDE of composites.
Gür [70] ultrasonically investigated the microstructure of aluminum/silicon
carbide metal matrix composites, deducing that in general ultrasonic
velocity increased with SiC content, but at larger Al/SiC particle size ratio
and higher SiC volume fraction, the velocity was decreased by increased
porosity, thus showing the usefulness of ultrasonic techniques for structural
characterization. Hosur et al. [71] utilized ultrasonic scans to characterize
thick-section woven composites that underwent ballistic impacts and
different repair methods, observing clear distinctions in the damage
characteristics. Roth [72] utilized a wet ultrasonic approach to do many

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 307

(a) Ultrasonic transducer

Water
FS B1 B2

L Ultrasonic M′
beam

Lucite
d Sample support
L′

Reflector plate

(b) Initial transducer pulse


M′′
FS B1 B2
M′
Voltage

0 Time (s)
2t1 2τ 2t2 ∆t
Time delay:

Corresponding
to echo
traveling
distance 2L 2d 2L′

Figure 33. Schematic of ultrasonic pulse–echo immersion testing and resulting wave-
forms [72].

measurements with a single pulser/receiver sensor, including a very effective


correction for uneven thickness corrections. In the single transducer
thickness-independent velocity imaging methodology shown in Figure 33,
according to Roth’s work, pulse-echo ultrasonic velocity measurements can
made with a transducer directly contacting the sample (labeled the contact
mode) or separated by a liquid coupling medium (immersion mode system).
In Figure 33 a schematic of an immersion-type pulse–echo testing system is
shown, along with resulting waveforms when a reflector plate is positioned
underneath, and apart from, the sample. Apparent velocity values are
obtained using the time delay between the first front surface signal and first
back surface echo, or between two successive back surface echoes. The
velocity in a material sample can also be measured by using echoes off of
the reflector plate and the front and/or back surface echoes. With the sample
present between the transducer and the reflector plate as shown in Figure 33,

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


308 N. SHARMA ET AL.

the pulse that travels from the transducer through the sample to the reflector
plate and back to the transducer is observed at a certain time. The times for
the pulse–echo time delays of the ultrasonic pulse from the transducer face
to the sample front surface, from the sample front surface to the sample
back surface, and from the sample back surface to the reflector plate front
surface, respectively are obtained. The sample is removed, and the pulse that
travels from the transducer to the reflector plate and back to the transducer
is also timed. In the resulting equations, sample thickness is not a variable,
and thus, this method does not need a prior knowledge of sample thickness.
When extended to multiple measurements across the sample, effects of the
variation of sample thickness are eliminated in the image obtained. This is
readily applicable to sandwich composites.

Infrared/Thermal Wave Methods

Infrared (IR) thermography NDE has also been used on composite


materials. Favro et al. [73,74] developed a fast IR thermal wave method for
wide-area inspection of structures. Guazzone and Danjoux [75] used thermal
waves to inspect new aero engines, and Dattoma et al. [76] utilized thermal
waves to acquire color-coded temperature profiles. The authors asserted
the impossibility of using thermal waves to capture accurate dimensions
of defects, and the suitability of the method as a good complementary
technique to other NDE approaches. Mandelis [77] used a spatial white
noise description to remove surface roughness effects in thermal wave NDE
and considered many practical application examples. Mandelis, Munidasa,
and Nicolaides [78] applied laser infrared photothermal radiometry to many
thermal wave inverse-problem NDE applications.
Figure 34 is a schematic diagram of the experimental arrangement used
by Favro et al. [73]. A low frequency (20 kHz) ultrasonic transducer injects
a short (50–250 ms) pulse of sound into the sample. The surface temperature
of the area under inspection is imaged by an IR video camera that is coupled
to a fast frame grabber in a computer. The ultrasonic transducer is placed in
a location that facilitates a suitable contact surface. The camera digitizes the
video images and transmits them to the computer as a stream of data.
The metal tip of the ultrasonic gun can be placed in direct contact with the
sample, or if essential, to prevent surface scars, a softer material like paper
is used between the gun and the sample. In many cases, energies of 10 J or so
are adequate to image the cracks. With such ultrasonic excitation energies,
we observe temperature increases of the order of a few degrees to a few tens
of degrees in the vicinity of surfacebreaking cracks in metal samples. It has
also been shown by the same authors that no visible change in the length of
the crack occurs from fatigue effects solely due to the ultrasonic excitation.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 309

Ultrasonic Sample
transducer

Crack

IR camera
Figure 34. Schematic of the setup for sonic IR inspection [74].

The state of the thermal wave NDE research shows that this method is
readily applicable to practically all types of materials and architectures.
These include layered materials and sandwiches, as thermal conduction
delays across thickness, are reckoned into the calculations.

Plate and Surface Waves Methods

Acoustic loading has also been used in other ways to investigate materials.
Wendler and Grigoryan [79] examined shear vertical Lamb waves and shear
horizontal Love waves at interfaces of sandwich layers, obtaining dispersion
and spatial velocity distribution curves. Osmont, Devillers and Taillade [80]
used Lamb waves to typify damages in sandwich plates. Lee [81] showed
that, using effective plate stiffness, it is possible to obtain very good values
for the wave velocities in composite materials in various directions.
A typical experimental setup for plate wave measurements is shown in
Figure 35, from the work of Lee [81]. This particular system consisted
of a function generator, an ultrasonic analyzer, a plotter, and a waveform
analyzer that performs time and frequency domain wave analyses. The
excitation of the sponge-pad supported specimen was by a gated three-cycle
sine pulse from the function generator at varying frequencies. The initial
spacing of 60 mm between the sender and the receiver was increased to
80 mm after measurement, and the resulting time delay used to obtain the
wave speed. The inset graph shows wave speed curves for the symmetric
(full lines) and asymmetric (broken lines) wave modes. The analytical
results tally well with experimental values. Thus analytical models could be

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


310 N. SHARMA ET AL.

Data 6000
wave analyzer
20
a1 s1 s2

Wave speed (km/s)


15
Wavetek Panametrics
function generator ultrasonic analyzer
10 s0

5
Sender Receiver
a0
Specimen 0
0 200 400 600 800 1000
Sponge pad
Frequency (kHz)
Figure 35. Experimental setup for plate wave measurements and dispersion curves for U-D
laminates [81].

generated that adequately predict acoustic speeds and dispersion character-


istics in composite materials of various architectures, including sandwich
composite materials.

Acoustic Absorption/Transmission Methods

Chen, Lee, and Chiang [82] examined the role of surface shapes
and perforations on the acoustic absorption of porous materials.
Forest, Gibiat, and Hooley [83] used acoustic parameters like the reflection
coefficient over the frequency domain to characterize granular and other
materials.
Figure 36 shows the schematic diagram of the instrumentation system
used in this referenced work, and some results. Using the two-microphone
impedance tube system, shown in the figure, the complex reflection
coefficient R(!) can be measured and determined by the equation

Ar ð!Þ H21 ð!Þ  eika ðz1 z2 Þ i2ka z1


Rð!Þ ¼ ¼ e ð38Þ
Ai ð!Þ eika ðz1 z2 Þ  H21 ð!Þ

where R is the acoustic reflection coefficient, ! is the radian frequency,


H is the transfer coefficient,pffiffiffiffiffiffiffiffiffi
k isffi the wave number, z is the specific
acoustic impedance, and I ¼ ð1Þ. The acoustic properties of sandwich
composites of various material and geometric constructions could thus be
easily found.
It may be concluded from the foregoing that with multiparam-
eter filtering, laminates and other composite structures can be

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 311

PC

Frequency
analysis system

Absorber
Microphones sample

Z2

Z1

Speaker
Figure 36. The two-microphone impedance tube system [82].

characterized for damage, using acoustic absorption and transmission


measurements.

Acoustic Emission Methods

Acoustic emission (AE) NDE has been used on many material systems,
including composites. Earlier reviews of applications include those of
Yamaguchi et al. [84] and Hamstad [85]. Various models and analyses of the
AE phenomenon have been proposed, such as by Dzenis and Qian [86] and
Aberg [87]. Damage quantification with AE has been attempted by some
workers, including Ma and Takemoto [88], Kim and Weiss [89], and Paul
and Fowler [90]. It has been suggested from the above research that of all the
better known NDE methods, AE has proven to be the most sensitive to the
smallest flaw sizes.
Two approaches, labeled the parametric AE analysis and transient AE
analysis respectively, have been developed for analysis, with the former

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


312 N. SHARMA ET AL.

Rise time Decay time


0.32
0.28
0.24

Amplitude
0.2 Count
0.16
0.12
0.08
0.04
0
−0.04
−0.08 Threshold
−0.12
−0.16
−0.2
−0.24
−0.28

Duration
Figure 37. A typical acoustic emission signal [86].

being used for the vast majority of the research to date on damage evolution
in materials. This method is based on the extraction of a number of
parameters (almost three dozen) from individual AE signals. A typical AE
signal is shown in Figure 37. Some of the AE parameters are defined in this
figure, including, signal amplitude, duration, rise time, decay time, and
AE counts. Other possibilities include signal duration, cumulative counts,
average frequency, energy, etc. In a basic, parametric AE system, an
ultrasonic wave caused by the damage event is detected by a piezoelectric
AE sensor, which converts the mechanical vibration into an analog signal.
The signal is amplified by a preamplifier and digitized by the AE system.
The system electronically extracts a number of parameters for each AE
event. These AE parameters are recorded into a parametric AE file, together
with some additional information like time of arrival, and some external
parameters, such as current load, etc. The AE signal itself is not further
utilized in the parametric AE approach. A key advantage of the parametric
analysis method is its simplicity. Modern AE systems provide
powerful analysis utilizing filtering, statistical deductions, location, etc.
The transient analysis capitalizes on utilizing the waveform of the AE signal,
transforming it into various domains and forms. The literature cited have
shown the efficacy of both approaches, and suggest that the best results
would probably be obtained by a judicious combination of both. The AE
approach is particularly suitable for sandwich composites because of their
very complex failure regimens, which make it very hard to follow by many
of the other NDE approaches.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 313

Since the method depends on the propagation of transient elastic waves


in solids, it is well adapted to tests of composite sandwiches.

CONCLUSIONS

. The analytical models to predict the fatigue life of sandwich composites


can be classified into four major categories: basic S–N approach, strength
degradation models, stiffness reduction models, and cumulative damage
models. The main drawback of all these models is their dependency on
large amounts of experimental input for each material, lay-up, and
loading conditions. There is a need for further research in this field;
specifically for developing a model that can predict fatigue failure based
on a more general knowledge of material and geometrical properties of
the sandwich specimen.
. There are a number of possible failure modes in sandwich beams under
static and impact loading like face yielding, face wrinkling, core shear,
indentation, and intra-cell dimpling. The initiation and progression of the
mode of failure depends on the material properties, geometrical
dimensions of the specimen and loading conditions. However, core
shear failure seems to be the predominant failure mode in cyclic fatigue
loading of sandwich composite materials.
. Two contradicting views are found in literature on the effect of loading
frequency on the fatigue life of sandwich structures with each view having
possible explanations and supportive experimental work. With increase
in frequency of loading, the temperature of the core increases. According
to one view, stiffness of the beam decreases and the interface bond
strength also decreases with increase in temperature and this leads to
premature failure. Another view is that with the increase in temperature,
energy dissipated gets converted into heat leaving less available stored
energy to fuel the damage process, so beams tested at higher frequencies
survive a greater number of cycles to failure. It cannot be concluded
at this point as to which view is correct and more theoretical and
experimental work is needed in order to better understand the effect of
frequency on fatigue life.
. Temperature and moisture are major environmental factors affecting
fatigue life of sandwich structures. At elevated temperatures the
degradation of the core increases, thus reducing the fatigue life. At sub-
ambient temperatures it has been observed that the through thickness
compressive strength improves slightly but no work has been published
so far on the effect of sub-ambient temperatures on the fatigue life
of sandwich structures. So further research in this area is required.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


314 N. SHARMA ET AL.

Moisture and contact of seawater reduces the fracture toughness of the


core–skin interface but again there is hardly any work available in
the literature that concentrates on evaluating the effect of moisture on the
fatigue life of sandwich composites and possibility of further research
in this area also exists.
. Three-point and four-point flexure tests do not neccesarily simulate the
practical conditions of fatigue loading on practical sandwich structures
that are subjected to repetitive loading conditions. Block loading in steps
is one way of overcoming this limitation and it has been observed that
fatigue life is reduced when a sandwich structure is subjected to varying
loads in steps as compared to a uniform load. Also a sequence of high–
low amplitude levels results in shorter lifetimes than a low–high order.
A recently developed Hydromat test system (HTS) simulates more
realistic test conditions, especially for marine structures where sandwich
composites are used as structural material in the hull of ships and vessels.
But the area of fatigue behavior of sandwich panels using the HTS
system is underexplored and there is a need for further research in
this field.
. High-technology NDE methods do exist now (like scanning with various
types of electron microscopes, etc.) which can be used to monitor the
damage during fatigue testing of sandwich composites, but they tend
to be very expensive and more laboratory based. However, it has been
shown that more common approaches like vibration, acoustic absorp-
tion, acoustic emission, and ultrasonic and thermal wave NDE have
been well developed to the level of being very useful and adequate for
quite a wide variety of modern NDE applications including fatigue
testing of sandwich composite beams and panels.

ACKNOWLEDGMENTS

The authors gratefully acknowledge the support of the Low Temperature


Navy Research Center at Wayne State University, which is sponsored by
the US Office of Naval Research. The ONR program manager is Dr Kelly
Cooper.

REFERENCES

1. Vinson, J.R. (1999). The Behavior of Sandwich Structures of Isotropic and Composite
Materials, Technomic Pub. Co., Lancaster, PA.
2. Degrieck, J. and van Paepegem, W. (2001). Fatigue Damage Modeling of Fibre-reinforced
Composite Materials: Review, Applied Mechanics Review, 54(4): 279–299.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 315

3. Kanny, K. and Mahfuz, H. (2005). Flexural Fatigue Characteristics of Sandwich Structures


at Different Loading Frequencies, Composite Structures, 67(4): 403–410. Figures [1] and
[26] reprinted with permission from Elsevier.
4. Burman, M. and Zenkert, D. (1997). Fatigue of Foam Core Sandwich Beams – 1:
Undamaged Specimens, Int. J. Fatigue, 19(7): 551–561. Figure [2] reprinted with permission
from Elsevier.
5. Burman, M. and Zenkert, D. (1997). Fatigue of Foam Core Sandwich Beams – 2: Effect of
Initial Damage, Int. J. Fatigue, 19(7): 563–578.
6. Sendeckyj, G.P. (1990). Life Prediction for Resin-matrix Composite Materials,
In: Reifsnider, K.L. (ed.), Fatigue of Composite Materials, 431–480.
7. Dai, J. and Hahn, H.T. (2004). Fatigue Analysis of Sandwich Beams using a Wear-out
Model, Journal of Composite Materials, 38(7): 581–589.
8. Mahi, El A., Farooq, M.K., Sahraoui, S. and Bezazi, A. (2004). Modeling the Flexural
Behavior of Sandwich Composite Materials under Cyclic Fatigue, Materials & Design,
25(3): 199–208. Figures [4], [5], [6] and [7] reprinted with permission from Elsevier.
9. Clark, S.D., Shenoi, R.A. and Allen, H.G. (1999). Modeling the Fatigue Behavior of
Sandwich Beams under Monotonic, 2-step and Block Loading Regimes, Composite Science
and Technology, 59(4): 471–486. Figures [8] and [9] and Table 5 reprinted with permission
from Elsevier.
10. ASTM Standard, C393-62. Standard Test Methods for the Flexural Properties of Flat
Sandwich Constructions.
11. Reifsnider, K.L., Schulte, K. and Duke, J.C. (1983). Long Term Behavior of Composite
Materials, In: Long Term Behavior of Composites, ASTM STP, 813: 136–159.
12. Wu, W.F., Lee, L.J. and Choy, S.T. (1996) A Study of Fatigue Damage and Fatigue
Life of Composite Laminates, Journal of Composite Materials, 30(1): 123–137.
13. Philippidis, T.P. and Vassilopoulos, A.P. (1999). Fatigue of Composite Laminates under
Off-axis Loading, International Journal of Fatigue, 21(3): 253–262.
14. Whitworth, H.A. (1998). A Stiffness Degradation Model for Composite Laminates under
Fatigue Loading, Composite Structures, 40(2): 95–101.
15. Miner, M.A. (1945). Cumulative Damage in Fatigue, Journal of Applied Mechanics, 12(3):
159–164.
16. Hashin, Z. and Rotem, A. (1978). A Cumulative Damage Theory of Fatigue Failure,
Journal of Material Science Engineering, 34(2): 147–160.
17. Wang, A.S.D., Chou, P.C. and Alper, J. (1981). Effect of Proof Test on the Strength
and Fatigue Life of a Unidirectional Composite, Fatigue of Fibrous Composite Materials
ASTM STP, 723, 116–132.
18. Broutman, L.J. and Sahu, S.A. (1972). New Theory to Predict Cumulative Fatigue
Damage, In: Fiberglass Reinforced Plastics, Composite Materials: Testing and Design
(Second Conference), ASTM STP, Vol. 497, pp. 170–188.
19. Epaarachchi, J.A. and Clausen, P.D. (2005). A New Cumulative Fatigue Damage Model
for Fiber Reinforced Plastic Composites under Step/Discrete Loading, Composites: Part A,
36(9): 1236–1245.
20. Allen, H.G. (1969). Analysis and Design of Structural Sandwich Panels, Paragon Press,
London.
21. Zenkert, D. (1995). An Introduction to Sandwich Construction, Chameleon, London.
22. Daniel, I.M., Gdoutos, E.E., Wang, K.A. and Abbot, J.L. (2002). Failure Modes of
Composite Sandwich Beams, International Journal of Damage Mechanics, 11(4): 309–334.
23. Gdoutos, E.E., Daniel, I.M. and Wang, K.A. (2002). Indentation Failure in Composite
Sandwich Structures, Experimental Mechanics, 42(4): 426–431.
24. Gdoutos, E.E., Daniel, I.M. and Wang, K.A. (2003). Compression Facing Wrinkling
of Composite Sandwich Structures, Mechanics of Materials, 35(6): 511–522.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


316 N. SHARMA ET AL.

25. Steeves, C.A. and Fleck, N.A. (2004). Collapse Mechanisms of Sandwich Beams
with Composite Faces and a Foam Core, Loaded in Three-point Bending. Part I:
Analytical Models and Minimum Weight Design, International Journal of Mechanical
Sciences, 46(4): 561–583.
26. Steeves, C.A. and Fleck, N.A. (2004). Collapse Mechanisms of Sandwich Beams
with Composite Faces and a Foam Core, Loaded in Three-point Bending. Part II:
Experimental Investigation and Numerical Modeling, International Journal of Mechanical
Sciences, 46(4): 585–608. Figures [10], [11] and [12] reprinted with permission from
Elsevier.
27. Konsta–Gdoutos, M.S., Konsta and Gdoutos, E.E. (2005). Load and Geometry Effect on
the Failure of Sandwich Structures, In: Proc. Society for Experimental Mechanics Conference
on Experimental and Applied Mechanics, Portland, Oregon, Paper No. 34 (on CD ROM).
Figures [13] and [14] and Tables 1 and 2, reprinted with the permission of the Society for
Experimental Mechanics, Inc., copyright 2005.
28. Lee, S.M. and Tsotsis, T.K. (2000). Indentation Failure Behavior of Honeycomb Sandwich
Panels, Composites Science and Technology, 60(8): 1147–1159.
29. Pan, S.D., Wu, L.Z., Sun, Y.G., Zhou, Z.G. and Qu, J.L. (2004). Longitudinal Shear
Strength and Failure Process of Honeycomb Cores, Composite Structures, 72(1): 42–46.
30. Petras, A. and Sutcliffe, M.P.F. (1999). Failure Mode Maps for Honeycomb Sandwich
Panels, Composite Structures, 44(4): 237–252. Figures [15] and [16] reprinted with
permission from Elsevier.
31. Abrate, S. (1998). Impact on Composite Structures, Cambridge University Press.
32. Abrate, S. (1997). Localized Impact on Sandwich Structures with Laminated Facing,
Applied Mechanics Reviews, 50(2): 83–96.
33. Lim, T.S., Lee, C.S. and Lee, D.G. (2004). Failure Modes of Foam Core Sandwich Beams
under Static and Impact Loads, Journal of Composite Materials, 38(18): 1639–1662.
34. ASTM D 5942 (1999). Determining Charpy Impact Strength of Plastics, American Society
of Testing and Materials.
35. Harte, A.M., Fleck, N.A. and Ashby, M.F. (2001). The Fatigue Strength of
Sandwich Beams with an Aluminum Alloy Foam Core, International Journal of Fatigue,
23(6): 499–507.
36. Kulkarni, N., Mahfuz, H., Melanie, S. and Carlson, L.A. (2004). Fatigue Failure
Mechanism and Crack Growth in Foam Core Sandwich Composites under Flexural
Loading, Journal of Reinforced Plastics & Composites, 23(1): 83–94.
37. Shenoi, R.A., Clark, S.D. and Allen, H.G. (1995). Fatigue Behaviour of Polymer
Composite Sandwich Beams, Journal of Composite Materials, 29(18): 2423–2446.
38. Challis, K.E., Hall, D.J. and Paul, D.B. (1986). A Novel Method for Determining
the Temperature Dependence of Shear Properties of Structural Foams, Cellular Polymers,
5(1): 91–101.
39. Sharma, S.C., Murthy, H.N.N. and Krishna, M. (2004). Interfacial Studies in Fatigue
Behavior of Polyurethane Sandwich Structures, Journal of Reinforced Plastics &
Composites, 23(8): 893–903.
40. Berkowitz, C.K. and Johnson, W.S. (2005). Fracture and Fatigue Tests and
Analysis of Composite Sandwich Structure, Journal of Composite Materials, 39(16):
1417–1431.
41. Erickson, M.D., Kallmeyer, A.R. and Kellogg, K.G. (2005). Effect of Temperature on the
Low-velocity Impact Behavior of Composite Sandwich Panels, Journal of Sandwich
Structures and Materials, 7(3): 245–264.
42. Kanny, K., Mahfuz, H., Thomas, T. and Jeelani, S. (2004). Temperature Effects on the
Fatigue Behavior of Foam Core Sandwich Structures, Polymers and Polymer Composites,
12(7): 551–559. Figures [28] and [29] reprinted with permission from the journal of
Polymers and Polymer Composites.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 317

43. Thomas, T., Mahfiz, H., Kanny, K. and Jeelani, S. (2004). Dynamic Compression of
Sandwich Composites at Sub-ambient Temperatures, Journal of Composite Materials,
38(8): 641–653.
44. Smith, L.V. and Weitsman, Y.J. (1996). The Immersed Fatigue Response of Polymer
Composites, International Journal of Fracture, 82(1): 31–42.
45. Weitsman, Y.J. and Elahi, M. (2000). Effects of Fluids on Deformation, Strength
and Durability of Polymeric Composites, Mechanics of Time-Dependent Materials, 4(2):
107–126.
46. Scudamore, R.J. and Cantwell, W.J. (2002). The Effect of Moisture and Loading Rate
on the Interfacial Fracture Properties of Sandwich Structures, Polymer Composites, 23(3):
406–417.
47. Li, X. and Weitsman, Y.J. (2004). Sea-water Effects on Foam-cored Composite Sandwich
Lay-ups, Composites: Part B, 35(6–8): 451–459. Table 4 reprinted with permission from
Elsevier.
48. ASTM D 5045-99 (1999). Standard Test Methods for Plane-strain Fracture Toughness and
Strain Energy Release Rate of Plastic Materials, American Society for Testing and Materials
49. Ishai, O., Hiel, C. and Luft, M. (1995). Long-term Hygrothermal Effects on Damage
Tolerance of Hybrid Composite Sandwich Panels, Composites, 26(1): 47–55.
50. Ishiaku, U.S., Hamada, H., Mizoguchi, M., Chow, W.S. and Ishak, Z.A. (2005). The Effect
of Ambient Moisture and Temperature Conditions on the Mechanical Properties of Glass
Fiber/Carbon Fiber/Nylon 6 Sandwich Hybrid Composites Consisting of Skin-core
Morphologies, Polymer Composites, 26(1): 52–59.
51. Harris, B., Gathercole, N., Reiter, H. and Adam, T. (1997). Fatigue of Carbon-fibre-
reinforced Plastics under Block-loading Conditions, Composites Part A, 28(4): 327–337.
52. Gamstedt, E.K. and Sjogren, B.A. (2002). An Experimental Investigation of the Sequence
Effect in Block Amplitude Loading of Cross-ply Composite Laminates, International
Journal of Fatigue, 24(2–4): 437–446.
53. Bertelsen, W.D. (1992). The Hydromat System: An Experimental Technique for the Static
and Fatigue Testing of Sandwich Panels, In: Proceedings Second International Conference
on Sandwich Construction, Gainesville, Florida.
54. Bertelsen, W.D. (2000). Development and Certification of the ASTM D 6416-99 Plate
Flexure Test and its Implications for the Future of Composite Panel Design and
Construction, In: Proceedings MACM 2000, 8th International Conference on Marine
Applications of Composite Materials, Melbourne, Florida.
55. ASTM D 6416-99 (1999). Standard Test Method for Two-dimensional Flexural Properties
of Simply supported Sandwich Composite Plates Subjected to Distributed Load,
American Society for Testing and Materials.
56. Ahtonen, P.J. and Sikarskie, D.L. (1998). An Analytical and Experimental Comparison
of Orthotropic Sandwich Panels using the Hydromat Test System, Composites Part B,
29(6): 705–714. Figure [31] reprinted with permission from Elsevier.
57. Ayorinde, E.O. (1995). Elastic Constants of Thick Orthotropic Composite Plates, Journal
of Composite Materials, 28(8): 1025–1039.
58. Gibson, R.F. (2000). Modal Vibration Response Measurements for Characterization of
Composite Materials and Structures, Composite Science and Technology, 60(15): 2769–
2780.
59. Liew, K.M., Xiang, Y. and Kitiponchai, S. (1995). Research on Thick Plate Vibration:
A Literature Survey, Journal of Sound and Vibration, 180(1): 163–176.
60. Salawu, O.S. (1997). Detection of Structural Damage through Changes in Frequency:
A Review, Engineering Structures, 19(9): 718–723.
61. Palozotto, A.N., Gummadi, L.N.B., Vaidya, U.K. and Herrup, F.J. (1995). Low Velocity
Impact Damage Characteristics of Z-fiber Reinforced Sandwich Panels – An Experimental
Study, Composites Structures, 43(4): 275–288.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


318 N. SHARMA ET AL.

62. Zheng, Z., Ayorinde, E.O. and Gibson, R.F. (2001). A Thick Beam Library Solution
Method for Vibration-Based Characterization of Thick Composite Plates, Journal
of Vibrations and Acoustics, 123(1): 76–83. Figure [32] reprinted with permission from
ASME.
63. Zou, Y., Tong, L. and Steven, G.P. (2000). Vibration-based Model-dependent Damage
(delamination) Identification and Health Monitoring for Composite Structures — A
Review, Journal of Sound and Vibration, 230(2): 357–378.
64. Liu, G.R. and Chen, S.C. (2001). Flaw Detection in Sandwich Plates based on Time-
harmonic Response using Genetic Algorithm, Computer Methods in Applied Mechanics and
Engineering, 190(42): 5505–5514.
65. Kim, H. and Hwang, W. (2002). Effect of Debonding on Natural Frequencies and
Frequency Response Functions of Honeycomb Sandwich Beams, Composite Structures,
55(1): 51–62.
66. Akay, M. and Hanna, R. (1990). A Comparison of Honeycomb-core and Foam-core
Carbon-fibre/Epoxy Sandwich Panels, Composites, 21(4): 325–331.
67. Sachse, W., Castagnede, B., Grabec, I., Kim, K.Y. and Weaver, R.I. (1990). Recent
Developments in Quantitative Ultrasonic NDE of Composites, Ultrasonics, 28(2): 97–104.
68. Gupta, N. and Sankaran, S. (1999). On the Characterization of Syntactic Foam Core
Sandwich Composites for Compressive Properties, Journal of Reinforced Plastics &
Composites, 18(14): 1347–1357.
69. Legendre, S., Goyette, J. and Massicott, D. (2001). Ultrasonic NDE of Composite
Material Structures using Wavelet Coefficients, NDT & E International, 34(1): 31–37.
70. Gür, C.H. (2003). Investigation of Microstructure–Ultrasonic Velocity Relationship in
SiCp-reinforced Aluminium Metal Matrix Composites, Materials Science and Engineering
A, 361(1–2): 29–35.
71. Hosur, M.V., Vaidya, U.K., Myers, D. and Jeelani, S. (2003). Studies on the Repair of
Ballistic Impact Damaged S2-glass/Vinylester Laminates, Composite Structures, 61(4):
281–290.
72. Roth, D.J. (1997). Using a Single Transducer Ultrasonic Imaging Method to Eliminate the
Effect of Thickness Variation in the Images of Ceramic and Composite Plates, Journal
of Nondestructive Evaluation, 16(2): 101–120. Figure [33] reprinted with permission from
Springer Science and Business Media.
73. Favro, L.D., Jin, H.J., Wang, Y.X., Ahmed, T., Wang, X., Kuo, P.K. and Thomas, R.L.
(1993). IR Thermal Wave Tomographic Studies of Structural Composites, In: Thompson,
D.O. and Chimenti, D.E., (eds), Review of Progress in Quantitative Nondestructive
Evaluation, NDT & E International, 28 Jul.–2 Aug. 1991; Vol. 26(2), p. 99, Plenum Press,
Brunswick, Maine (United States) 11A: 447–451.
74. Favro, L.D., Thomas, R.L., Han, X., Ouyang, Z., Newaz, R.L. and Gentile, G.D.
(2001). Sonic Infrared Imaging of Fatigue Cracks, International Journal of Fatigue,
23(Supplement 1): S471–S476. Figure [34] reprinted with permission from Elsevier.
75. Guazzone, L.J.-P. and Danjoux, R. (1996). Inspection of New Aeroengine Components by
Thermal Wave, NDT & E International, 29(6): 392.
76. Dattoma, V., Marcuccio, R., Pappalettere, C. and Smith, G.M. (2001). Thermographic
Investigation of Sandwich Structure made of Composite Material, NDT AND
E International, 34(8): 515–520.
77. Mandelis, A. (2001). Diffusion-wave Laser Radiometric Diagnostic Quality-control
Technologies for Materials, NDE/NDT, NDT & E International, 34(4): 277–287.
78. Mandelis, A., Munidasa, M. and Nicolaides, L. (1999). Laser Infrared Photothermal
Radiometric Depth Profilometry of Steels and its Potential in Rail Track Evaluation,
NDT & E International, 32(8): 437–443.
79. Wendler, L. and Grigoryan, V.G. (1988). Acoustic Interface Waves in Sandwich Structures,
Surface Science, 206(1–2): 203–224.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016


Fatigue of Foam and Honeycomb Core Composite Sandwich Structures 319

80. Osmont, D., Devillers, D. and Taillade, F. (2001). Health Monitoring of Sandwich Plates
based on the Analysis of the Interaction of Lamb Waves with Damages, In: Proceedings
of SPIE – The International Society for Optical Engineering, 432(7): 290–301.
81. Lee, J. (1999). Plate Waves in Multi-directional Composite Laminates, Composite
Structures, 46(3): 289–297. Figure [35] reprinted with permission from Elsevier.
82. Chen, W.H., Lee, F.C. and Chiang, D.M. (2000). On the Acoustic Absorption of Porous
Materials with Different Surface Shapes and Perforated Plates, Journal of sound and
Vibration, 237(2): 337–355. Figure [36] reprinted with permission from Elsevier.
83. Forest, L., Gibiat, V. and Hooley, A. (2001). Impedance Matching and Acoustic
Absorption in Granular Layers of Silica Aerogels, Journal of Non-Crystalline Solids,
285(1–3): 230–235.
84. Yamaguchi, K., Oyaizu, H., Johkaji, J. and Kobayashi, Y. (1991). In: Sachse, W.,
Roquet, J. and Yamaguchi, K. (eds), Acoustic Emission: Current Practice and Future
Directions, ASTM STP 1077, American Society for Testing and Materials, Philadelphia,
PA, p. 123.
85. Hamstad, A. (1996). A Review: Acoustic Emission, a Tool for Composite Material Studies,
Experimental Mechanics, 26(1): 7–13.
86. Dzenis, Y.A. and Qian, J. (2001). Analysis of Microdamage Evolution Histories in
Composites, International Journal of Solids and Structures, 38(10–13): 1831–1854. Figure
[37] reprinted with permission from Elsevier.
87. Aberg, M. (2001). Numerical Modeling of Acoustic Emission in Laminated Tensile Test
Specimens, International Journal of Solids and Structures, 38(36–37): 6643–6663.
88. Ma, X.Q. and Takemoto, M. (2001). Quantitative Acoustic Emission Analysis of Plasma
Sprayed Thermal Barrier Coatings subjected to Thermal Shock Tests, Material Science and
Engineering A, 308(1–2): 101–110.
89. Kim, B. and Weiss, J. (2003). Using Acoustic Emission to Quantify Damage in Restrained
Fiber-reinforced Cement Mortars, Cement and Concrete Research, 33(2): 207–214.
90. Paul, H. and Fowler, T. (2003). Fiber Reinforced Vessel Design with a Damage Criterion
Approach, Composite Structures, 61(4): 395–411.

Downloaded from jsm.sagepub.com at PENNSYLVANIA STATE UNIV on September 19, 2016

You might also like