10
10
10
Dissertation
Submitted to
UNIVERSITY OF DAYTON
The Degree of
By
Sadra Emami
UNIVERSITY OF DAYTON
Dayton, Ohio
December 2017
DEVELOPMENT OF PROBABILISTIC MODELS FOR LONG TERM RELIABILITY
APPROVED BY:
ii
© Copyright by
Sadra Emami
2017
iii
ABSTRACT
Structural engineers commonly use balsa wood and PVC foam as core materials for
sandwich composite structures. The long-term reliability and damage mechanism of these
composite sandwich structures under severe environmental conditions are still unclear.
These liability concerns prevent most civil structural engineers from considering this
material in infrastructure applications. Also, they must account for the unique nature of the
civil construction industry and offer an advantage over traditional building materials such
as steel and concrete. Introducing a strength reduction factor (ɸ) will increase the chance
approach is a means to reach to this goal. In this project, the long-term effects of
100 days of freeze/thaw exposure (-20oC to 20oC) in the presence of a 3% NaCl saline
solution. The sandwich panels were comprised of using balsa wood (SB100) and foam
iv
core (Airex C70.55) with fiberglass/vinyl ester face sheets, fabricated with vacuum assisted
resin transfer molding (VARTM).Samples were tested for core shear, core compression,
and peel tests. Results confirmed that exposure reduced the balsa wood core properties
significantly, however, PVC foam core shear modulus increased by 25%, and the
compression modulus reduced by 25%. Simulated lifetime core shear fatigue curves were
also developed and evaluated. A substantial reduction in tensile, shear, and compression
properties was observed for the face sheets constructing the sandwich panels. The thermal
cycling events degraded the matrix binding, the warp and fill fibers, thus impairing the
structural integrity of the cross-ply laminate. After testing, a reliability-based approach was
used to examine the “strength reduction” factors (ɸ) of the core materials. Resistance
factors of 0.79 and 0.95 were obtained for balsa and foam core composite sandwich
materials, respectively. The results of this research will ultimately lead to a probabilistic
analysis model that will eventually act as a benchmark to reliably predict the performance
v
This dissertation is dedicated to my beloved family and friends who have provided me
with the support, encouragement, and strength I needed throughout my graduate career.
vi
ACKNOWLEDGMENTS
I sincerely would like to thank my advisor, Dr. Elias Toubia, for giving me the
opportunity to work in this area, and motivating me throughout my Ph.D. program. His
excellent insight into science and engineering, management’s skills, generosity, and
positive attitude have positively affected my academic and social lives. He was extremely
supportive throughout my Ph.D. thesis. Without his technical and funding support, it would
Browning for his support, transferring his extreme knowledge in composites and helps. In
addition, I would like to thank Dr. Donald Klosterman, who was the faculty advisor,
SAMPE chapter president, and my Ph.D. committee member. He was extremely helpful,
nice, and collaborative. Also, I would like thanks, Dr. Kellie Schneider and Dr. Steven
Tobias. And staff members from the University of Dayton Research Institute (UDRI),
specifically to Ms. Mary Galaska, Ms. Marlene Houtz, Mr. Dale Grant, Carl-William O
Sjoblom, and all staffs in the smart lab. And the composite advantage LLC for molding the
sandwich composite panels. Finally, the invaluable contribution made by my dad, mom
vii
TABLE OF CONTENTS
ABSTRACT......………………………………………………………………….…...….iv
DEDICATION………………………………….………………………………...………vi
ACKNOWLEDGMENTS ...................................………………………………...……..vii
LIST OF FIGURES…………………………………………………………..……......….x
LIST OF TABLES……………………………………………………………......……..xiv
CHAPTER 1 INTRODUCTION…………………………………....…….…….…….1
1-1 Problem Statement….………………………….………..............….......4
1-2 Significance of the Research……………………………………….......5
CHAPTER 2 BACKGROUND…………………..…..……………...………….....….7
2-1 Reliability in Civil Engineering Composite Materials………………....7
Traditional and Probabilistic Composite
Structural Design Process…………………………………………8
Steps in Quantifying Randomness……………………………….10
Parameters to Calculate Reliability………………………………11
2-2 Raw Materials Composition and Structure……………………………15
Balsa Wood Structure……………………………………………15
PVC Closed Cell Foam Structure………………………..………18
2-3 Composite Durability…………………………………...……………22
Glass Fiber Laminate Durability…………………………………22
Sandwich Composite Durability…………………………………23
CHAPTER 3 THEORETICAL APPROACH………..………………………....…...26
3-1 Selection of Probability Distribution………………………….…..…..27
3-2 Limit State Functions.……………………..…….…………………….30
viii
CHAPTER 4 EXPERIMENTAL APPROACH………………………………….....35
4-1 Durability Study of Sandwich Composite and Laminate……...……...35
Panels and Laminates Preparation…………………………….....35
4-2 Environmental Experiment Setup………………………….……….....38
4-3 Test Setup………………………………………………………...…...40
GFRP Laminate Testing…............................................................44
4-4 Characterization Methods……………………………………………...46
CHAPTER 5 RESULTS AND DISCUSSION…..………………...………..…….…50
5-1 Balsa Wood Core Sandwich Composites Properties……………...…..50
5-2 Foam Core Sandwich Composite Properties………………….…...….59
5-3 Laminate Tests Results……………………………………….…...…..68
Effect on Shear Properties……………………………………….69
Effect on Tensile Properties…………………………………..….71
Effect on Compressive Properties………………………………..73
Root Cause Failure Analysis……………………………………..74
Damage State in Woven Roving Architecture………………..….78
Matrix and Fiber Damage…………………………………….….80
5-4 Model Results………………………………...……………………….83
Parameters Estimation……………………………………….…..95
CHAPTER 6 SUMMARY OF SIGNIFICANT RESEARCH
RESULTS AND FINDINGS………...………………………….....…98
CHAPTER 7 CONCLUSIONS AND RECOMMENDATIONS…..………….........102
BIBLIOGRAPHY………………….…………...………………..…………………..…104
ix
LIST OF FIGURES
x
Figure 20. Areal weight of balsa wood core sandwich after F/T exposure……….…..50
Figure 21. Molded balsa wood after a) 50 cycles, b) 200 cycles……………………...51
Figure 22. Balsa core shear strength (left) and modulus (right)
-ASTM C393 and D7250……………………………….…………………52
Figure 23. 0 cycle core shear failure mode……………………………………………52
Figure 24. 200 cycles core shear failure mode………………………………….….…53
Figure 25. Flexure results for balsa core sandwich samples……………………….…53
Figure 26. Balsa wood fatigue failure mode……………………………………..……54
Figure 27. Balsa wood core shear fatigue life prediction for exposed
(200 cycles) and unexposed Specimens (0 cycles), R = -0.1 at 4 Hz.......…55
Figure 28. Balsa Wood Texture a) Unexposed b) Exposed Samples (Optical
Microscopy)………………………………………………….………..…..55
Figure 29. Balsa Core Flatwise Compression Strength (Left) and Compression
Modulus (Right)-ASTM C 365 Preliminary Testing at 0 and 200
cycles…………………………………………………………………..…..56
Figure 30. Load-stroke curve for exposed and unexposed specimens………………..57
Figure 31. Balsa wood fracture toughness for unexposed (0 cycles) and
exposed (200 cycles) specimens (ASTM E2004)………………...……….57
Figure 32. a) A specimen under peel test, b) failure modes of exposed and
unexposed specimens……..…………………………………………....….58
Figure 33. Foam core shear strength and shear modulus -ASTM C393
and D7250……………………………………………..……………….......59
Figure 34. Shear four-point test………………………………………..……………...60
Figure 35. Four-point flexure results for foam core sandwich samples …………..….60
Figure 36. SEM: a) unexposed Foam and b) exposed Foam after 200 F/T
cycles, X-Ray CT-Scan: c) foam Core Sandwich
Specimen after F/T Exposure…………………………………………..….61
Figure 37. DMA testing results, a) shear mode testing,
b) 3-point bending mode………………………………………….….……63
Figure 38. Fatigue failure mode of PVC foam………………………………….…….64
xi
Figure 39. Foam core shear fatigue life prediction for exposed (200 cycles)
and Unexposed Specimens (0 cycles), R = -0.1 at 4 Hz………………..…64
Figure 40. Foam core flatwise compression strength and compression modulus
ASTM C 365 preliminary testing at 0 and 200 cycles…………….………65
Figure 41. Load-stroke curve for exposed and unexposed specimens…………..…....66
Figure 42. PVC foam Fracture Toughness for Unexposed (0 cycles)
and Exposed (200 cycles) Specimens (ASTM E2004)………………....…66
Figure 43. a) A specimen in middle of peel test, b) specimens after peel test………..67
Figure 44. Overall results of exposed (200 cycles) and unexposed (0 cycles)
specimens………………………………………………………………….69
Figure 45. Shear failure modes for a) unexposed; and b) exposed specimens……..…70
Figure 46. Percent change in shear modulus and strength………………………….…70
Figure 47. Micrograph of the cross section after failure at the V-notch location a)
unexposed: showing only interlaminar cracks, b) Exposed: showing
interlaminar and intralaminar cracks…………………………..…………..71
Figure 48. Percent change in tensile modulus and strength………………..………….72
Figure 49. Tensile test failure modes; unexposed and exposed specimens…………...72
Figure 50. Longitudinal and transverse Stress-Strain curves for
exposed and unexposed specimens…………………………………...…..73
Figure 51. a) Percent change in modulus and strength, b) Compression failure
modes…………………………………………………………………..….74
Figure 52. DMA results for composite, a) 1st DMA heating cycle, b) 2nd DMA
heating cycle…………………………………………………………….....77
Figure 53. X-Ray computed tomography showing damage in exposed samples a)
surface plies, b) longitudinal cross-section, c) Transverse cross
section……………………………………………………………………...79
Figure 54. Ultrasonic C-scan results (102 mm wide specimens); a) unexposed:
grid pattern, b) exposed: linear pattern…………………………………....80
Figure 55. Local magnified optical image at weave junction showing matrix
failure between damaged warp fibers and fill fibers after exposure…..…..81
Figure 56. SEM micrograph showing degradation of fiber/matrix interface and
fiber fracture…………………………………………………...….…….…81
Figure 57. SEM picture of damaged E-glass fiber after exposure……………...……..82
xii
Figure 58. Shear strength statistical plots of balsa core material…………..…..…..….86
Figure 59. Shear modulus statistical plot of balsa core material………………..….…87
Figure 60. Shear strength statistical Plot of foam core material……............................88
Figure 61. Shear modulus statistical plot of foam core material……….......................89
Figure 62. Compression strength statistical plot of balsa core material…………...….90
Figure 63. Compression modulus statistical plot of balsa core material……...………91
Figure 64. Compression strength statistical Plot of foam core material……...……….92
Figure 65. Compression modulus statistical plot of foam core material……..……….93
xiii
LIST OF TABLES
xiv
CHAPTER 1
INTRODUCTION
Composite sandwich construction consists of two face sheets with a high level of
strength and stiffness, which are adhesively bonded to both sides of a lightweight core
sheet. In general, face sheets are made of composite laminates such as glass fiber or carbon
fiber with resin; the core is typically honeycomb, end-grain balsa wood, or closed-cell
polymer foam. Each component has a different function: face sheets carry bending and in-
plain forces, while core sheets carry the transverse shear load. This function of sheets helps
prevent face-sheet buckling and holds the face sheets in positions far away from the neutral
axis in order to maximize the flexural stiffness of the structure [1]. Figure 1 shows a typical
sandwich composite
1
Figure 1. A typical sandwich composite
transportation structures, and bridge decks where minimum structural weight and
deteriorate at an alarming rate in the U.S. and many other countries. For example, of the
nearly 600,000 bridges in the U.S. transportation system, about 143,000 are older than 50
years and unsuitable to meet the current or projected traffic demands [2]. Therefore, more
applicable to civil engineering techniques. The materials must accommodate the unique
nature of the civil construction industry and must offer an advantage over traditional
building materials such as steel and concrete. Additionally, liability concerns prevent the
employment of less known products if the basis for their technical design data and long-
term durability behavior has not been established. Consequently, for civil applications,
2
composites are deemed to be less reliable than conventional construction materials (steel,
concrete, etc.) for which the design methods, standards, and supporting databases already
exist. These issues are best addressed in the context of probability-based limit state design
[2]. This approach is capable of quantifying the inherent risk of a design and assessing the
service lives. For example, bridge decks or other structures and infrastructures are
subjected to a wide range of repetitive loadings (fatigue) and changes in temperature and
moisture levels. Load carrying ability can be significantly reduced by the presence of
moisture in the polymer foam or balsa core, which is absorbed during use or storage in
the resin, including fiber-matrix debonding and matrix micro-cracking in the composite
and even pitting and crack of fiber [3–5]. Moreover, the presence of deicing salts in winter
must be considered as part of the bridge deck service environment. Cold temperature is
mechanical behavior of materials. In fact, it can be designed with several parameters such
3
as a different choice of resin to fiber ratio, fiber orientation in different lamina or stacking
sequences [7]. Reliability prediction is difficult for these types of materials, as there are a
wide range of failure mechanisms and random variables that influence those mechanisms
which must be taken into consideration while modeling the reliability of composites [8].
This larger number of variables, as compared to isotropic materials, shows that the strength
of composite materials is more uncertain, and due to anisotropic properties, they require
The long-term goal of this research is to assess the feasibility of utilizing sandwich
the first goal is to understand the durability of these materials through the following
uncertainties:
to a freeze-thaw environment?
civil engineering applications is the lack of technical knowledge regarding their long-term
durability; thus, structural engineers do not have access to the design standards for these
materials as they do for traditional building materials such as wood, steel, and concrete.
Therefore, the purpose of this study is to develop probability models that describe the
4
reliability of sandwich composite materials, focusing on freeze-thaw conditions in this
thaw environments?
freeze-thaw conditions?
The National Academy of Engineers has issued a list of grand challenges, which
outlines various initiatives for engineering achievement, including the restoration and
Engineering issued a report card for the U.S. infrastructure, with bridges earning a grade
of C+ [2]. The average age of bridges in this country is 42 years, and it is estimated that an
important that this infrastructure is not only repaired but also improved, with new
technologies that are more beneficial than legacy systems. This work aims to help
applications.
Structures such as bridge decks or other civil infrastructure are subjected to a wide
range of repetitive loadings (fatigue) and changes in temperature and moisture levels. The
load carrying ability of sandwich composites can be significantly reduced by the presence
of moisture in the polymer foam or balsa core, which may be absorbed during use or storage
5
in humid environments [4–5]. The F/T effect, as well as the presence of deicing salts in
winter, must be considered as part of the bridge deck service environment [5].
reinforced concrete deck. A lighter weight (load) for the structure can translate into savings
across the structure, as the size of structural members and foundations can be reduced
accordingly. Fabricating deck panels in a factory and shipping them to the bridge site also
offers several advantages over cast-in-place concrete. For example, quality can be closely
6
CHAPTER 2
BACKGROUND
that is, they cannot be predicted with certainty. Repeated measurement of a physical
phenomenon generates multiple outcomes. Among these many outcomes, some are more
frequent than other. The occurrence of multiple outcomes without any pattern is described
in terms of uncertainty. Engineers always attempt to account for the uncertainty in their
analysis and design. In engineering, satisfactory performance is often given in terms of the
the use of a deterministic approach (that is, the current industry approach) to design and
analyze composite structures may undermine their high utilization and optimization
potential. This may result in higher safety factors. Moreover, the reliability of such an
7
One of the present barriers to the implementation of composite materials in
durability; thus, structural engineers do not have access to the design standards for these
materials as they do for traditional building materials such as wood, steel, and concrete.
Therefore, the purpose of this study is to use probabilistic design in order to predict and
quantify the structural strength and performance of certain FRP composites, particularly
sandwich composites.
deterministic criteria for failure analysis. Their values have been established after many
years of experiments and calibration by judgment; however, they are not suited to new
materials with particular features, such as composites. Safety factors are generated by this
process. They are derived on the basis of past experience but do not guarantee safety or
the way in which the different parameters of the system influence safety. Thus, it is difficult
to design a system on the basis of uniform distribution of safety level. Design safety is
ensured by keeping the resistance of structure (RN) greater than the load on the structure
Nominal SF = RN/SN
Both the resistance and load in this formula are conservative and may fail to convey
8
In the probabilistic design of engineering system, the future is predicted with the
help of information from the past, which includes experience and judgment. New
approaches are required to predict the reliability of composite structures, since strengths
and stresses are random variables. The basic probabilistic approach can be summarized in
the following: (1) quantification of all input variables required for structural analysis, (2)
definition of the resulting stress and strength of the structure for each possible failure mode,
and (3) evaluation of the resulting probability of structural failure. Figure 2 illustrates this
process.
The left-hand side presents the input data required for the determination of the
applied stress distribution, with each stress possibly having a different statistical
distribution. Alternatively, the right-hand side depicts the various capabilities of the
structure. The middle part indicates the associated failure probabilities for each failure
mode, which are applied stress and resistive component strength distribution.
9
Engineers often do not know the underlying distribution of random variables, and
thus, they have to rely on test data and observation for the estimation of the parameters of
the distribution functions. By collecting more information, engineers can reduce statistical
The first step is data collection, for which an appropriate data population is
(COV), and skewness can describe the randomness of a sample population. A complete
histogram. In preparation of the histogram, the number of bins or width of the interval plays
For this purpose, it is necessary to treat discrete and continuous random variables
separately. Probability density function (PDF) and cumulative density function (CDF) are
usually used to calculate the probability of random variables. The nature of structures is
based on the continuous random variables. Commonly used continuous random variables
include the normal or Gaussian distribution, lognormal random variable, and Weibull
random distribution. There are several methods to determine distribution and parameters
from observed data such as probability paper and statistical tests, including the chi-square
10
test and K–S test. In order to estimate the parameters of a distribution, the method of
Instead of using the safety factors for the resistance alone, as in the working stress
method, or for the loads alone, as in the ultimate strength method, it is more rational to
apply the safety factor to both resistance and load, as is done in the case of concrete or steel
load and resistance factor design (LRFD) [56]. This concept has been developed for
The LRFD method depends on the applied load and resistance variables, which are
Applied Load: Loads (applied stress) to the structure are random processes. Civil structures
are often designed for a 50–year lifetime, which means that the maximum loads occurring
over a 50–year period must be taken into consideration. The loads are represented by
stochastic processes and vary continuously with time or take on discrete values. The
instantaneous values of the loads are often considered as the mean values. Permanent loads
include the weight of the structure and immoveable fixtures (walls, carpet, and roof). These
loads have a low variance due to the fact that the permanent load is closely tied to the
Sustained loads change at discrete times, but remain relatively constant between
11
furniture, vehicles, equipment) as well as environmental changes (temperature, humidity,
snow).
Multiple loads are likely to act on the structure at any given time. However, it is
unlikely that multiple types of maximum loads will be experienced at the same time, such
as maximum earthquake load and maximum snow load. Several methods have been used
The Ferry Borges–Castanheta (FB–C) [58] model assumes that each load is
characterized by a sequence of independent load events, with each event having an equal
duration. The service life of the structure is then divided into increments equal to the
duration of the load events. The occurrence rate of the load determines the probability of
the load occurring at an arbitrary time interval. If the load does occur, the probability
distribution of the load magnitudes is applied to the time interval. This process is carried
out for each load type experienced by the structure. The contributions from all the load
effects are combined to predict the maximum combined load that is likely to occur.
In this research project, the main effort is directed toward the development of a
resistance factor for composite materials rather than stochastic load models, as these
models have been already implemented in the ASCE7-10 (Minimum Design Loads for
Buildings and Other Structures standard) [6]. Structural engineers are well acquainted with
12
Component Strength: Calculation of component strength is the most complicated section
probabilistic strength of materials in aircraft applications. They used the β method (the
method with the establishment of target reliability) in order to laminate plates [59–60]. To
implement the reliability concept in civil engineering, the criteria for design strength for
composite sandwiches should be similar in appearance to the design strength for steel and
approached the probabilistic design at the structural civil engineering level, called Load
and Resistance Factor Design (LRFD). LRFD enables designs to be targeted for a desired
design [61–62]. The choice of target reliability level takes into account the possible
consequences of failure in terms of risk to life, injury, potential economic losses, and the
Depending on target reliability, the probability of failure can be calculated with this
formula:
𝛽 = −Ф−1 (𝑃𝑓 )
probabilistic model
13
In order to make the right choice for target reliability, the reference period,
consequences of failure, and cost of safety measures should be analyzed for the specific
case under consideration. It can be calculated with the limit state function, 𝑔(𝑅, 𝑆) [63].
in Figure 3, which indicates the stress and strength along the horizontal and vertical axes
respectively. The line drawn represents the scenarios where stress = strength, or 𝑔(𝑅, 𝑆) =
𝑅 − 𝑆 = 0. This line separates the failure region (g < 0) from the safe region (g > 0). The
function 𝑔(𝑅, 𝑆) is commonly referred to as the limit state function. The probability of
failure is defined as the volume under the surface shown in the failure region where g < 0.
14
where X = {x1,x2,.., xn}T is a vector of random variables that represent uncertain
quantities influencing the state of the structure, fx(X) is the probability density function
(PDF), and g(X) ≤ 0 denotes a subset of the outcome space, where a failure occurs [64].
For mathematical analysis, it is necessary to describe the failure domain g(X) < 0 in an
analytical form, which is widely referred to as the limit state function (LSF).
Due to the difficulty in directly computing this integral, many methods exist for
In 2014, Jawaheri Zadeh and Nanni [67] introduced a method to bypass the loading
variables and weigh the resistance of two structural elements with the same ultimate limit
state directly against each other. They employed the concept of reliability index to develop
comparative reliability. They compared the flexural and shear strength reduction factor for
concrete beams, and the results were the same as available reduction factors.
especially as core materials in sandwich composites. This is due to the relatively high
stiffness to weight and strength to weight ratio [10]. An example of using balsa wood as a
core material is the 56m span Bascule FRP Composite Footbridge – Fredrikstad. The
bridge deck is a sandwich construction with carbon fiber (CFRP) laminates and a balsa
wood core, with embedded heating cables for defrosting during the wintertime. The deck
15
is strong enough to carry a car with up to 2.0–ton axle load. It is the largest moveable bridge
in Scandinavia, which utilizes FRP composites as the main load carrying materials [11].
called tracheids. About 80–90% of the volume is arranged longitudinally; parenchyma are
arranged radially and constitute about 8–15% of the balsa wood volume and sap channels
[12].
Tracheids have an irregular hexagonal cross section, consisting of long tubes. The
materials of these structures chiefly are cellulose, hemicelluloses, and lignin, which
together form the elementary fibrils [13]. The latter, aggregated into larger units by
hydrogen bonds, constitute the microfibrils that are the basic elements of the cell wall
layers. The tracheid cells consist of the primary wall (P) and secondary wall (S) layers, as
presented in Figure 5. The secondary wall is further sub-divided into the S1, S2, and S3
layers. The S2 layer (of approximately 2 lm thickness) accounts for roughly 85% of the
16
secondary wall thickness. The parenchyma rays penetrate the tracheids radially and are
responsible for misalignments of the latter along the natural axis of the tree [14]. Sap
channels, responsible for fluid transport in the tree, have thinner cell walls and are
of density occurs across the trunk due to the different growth of the early/spring wood and
Figure 5. Simplified microstructure of balsa wood, showing middle lamella and tracheid
Balsa wood is a natural product. This indicates that it cannot be controlled in the
manner of industrial materials. Therefore, there are no accurate material properties and
often a single value is utilized for their properties [10]. The density of balsa wood sheets is
usually used to categorize its mechanical properties. Balsa panels are not directly produced
17
from the trunk of the tree. Their production involves joining the small cubes of balsa
together with adhesive. Their fiber direction is perpendicular to panel plate, as presented
in Figure 6.
Figure 6. Panel composed of balsa blocks joined by PVA base adhesive [19]
The density is an average of assembled end-grain blocks with few centimeters size,
each of which may have several different densities. This can induce significant
uncertainties in its properties. [17–19]. Quantifying the uncertainty properties and their
influence on the safety of structures can prove to be a solution for ensuring the quality and
structure that has interconnected small solid struts and or plates, forming an open or closed
cell foam. The bone of humans and animals are examples of foam structures in nature.
18
Styrofoam is an example of artificial foam used for insulation. Additionally, foam panels
are commonly used in surfboards and car bumpers. Structural application of foams as core
materials in wind blades and in bridges has created a considerable amount of demand for
them.
One of the main types of foam structure comes from PVC. Polyvinyl chloride
polymerization of vinyl chloride monomers. The basic repeating unit of the PVC polymer
The PVC polymer contains 57% chlorine (Cl) by weight. The presence of the Cl
atom causes an increase in the interchain attraction and, in turn, increases the hardness and
stiffness of the polymer. Moreover, the combustion of PVC produces dangerous fumes,
which can be fatal on incineration [21]. Initially, the usage of PVC was restricted due to its
brittle nature. Additives, such as plasticizers, were introduced to the polymer, which made
the PVC more flexible and easier to be processed [22]. PVC has a softening point at about
80 ⁰C and is resistant to liquids such as salt water and antifreeze mixtures, which makes it
19
Polymer foams are classified as cellular structures that can either be open or closed
[24]. Open-cell foams have the simplest structure, consisting of beam-like elements, which
define each cell that provides an open grid-like structure (refer to Figure 8a). Closed-cell
foams contain a combination of the beam-like structure of open-cell foams and membranes
that close off the open sections of the cell (Figure 8b). The fraction of polymer in the beam
elements and in the membrane is considered to substantially influence the stiffness of the
foam. For polymer foams, most of the solid plastic is located on the edges.
PVC foams are widely used in sandwich structures, varying from pure insulation
applications to structural core materials used in marine vessels, aerospace structures, and
wind turbine blades. PVC foams dominate the market in terms of polymer foam core
material.
20
In order to produce the PVC foam, a PVC plastisol consisting of isocyanates (for
crosslinking the molecular PVC chains), a blowing agent (for initiating the foaming
process), and a stabilizer are mixed together at temperatures below 100 ⁰C. The mixture is
then placed into the water, which reacts with the isocyanates in order to initiate the cell
nucleation and expansion. The foam is then allowed to cure in a mold in order to form its
final rigid structure. Since the cross-linking and foaming processes occur simultaneously,
Closed-cell foams are made up of polarized molecules that are otherwise known as
dipoles, and have a very low moisture absorption [27]. They have an interpenetrating
polymer network (IPN), a polymer comprising two or more networks that are at least
partially interlaced on a polymer scale but not covalently bonded to each other. The
network cannot be separated unless the chemical bonds are broken [26]. The two or more
networks can be considered to be entangled in such a way that they are concatenated and
cannot be pulled apart, but not bonded to each other by any chemical bond. The mixture of
polyvinyl chloride and polyurea has a good bond strength. Closed-cell PVC foam forms a
solid shape due to its linear structure. However, due to this structure, it is more brittle than
21
2-3 Composite Durability
Glass Fiber Reinforced Plastic (GFRP) composites are extensively being used in
civil engineering applications, namely, pedestrian bridges, piling bumpers, and offshore
and marine structures. This GFRP is fabricated through the Vacuum Assisted Resin
Transfer Molding (VARTM) process, and it is commonly comprised of low cost woven
roving (WR) E-glass fibers infused with a vinyl ester resin. Their end use applications
In cold regions, ice chemicals and rock salts are frequently used on highway and pedestrian
bridges for ice and snow removal. In the last two decades, several researchers have
investigated the F/T cycling effects on composite materials [42–47]. Significant reduction
environment [42–52]. When F/T is combined with salt water, two mechanisms are most
likely to occur. The first one is related to matrix cracking under thermal cycling, and the
second to diffusion through the resin down to the fibers. Resin degradation, fiber-matrix
debonding, and fiber stress corrosion were found to be the major failure mechanisms in
GFRP when immersed in sea water [53]. Experimental data available in the literature
related to WR GFRP under F/T exposure reported the existence of limited mechanical tests.
This study provides a full spectrum of mechanical testing, progressively backed by several
et al. [49], it was felt that a detailed root cause failure analysis is required to assist the
22
development of a mechanistic model on the basis of experiment related to the material
degradation in the observed loss of stiffness and strength. Understanding the cause, nature,
and the way damage is induced by thermal cycling, one would be able to quantify the
reduction and properly predict the structural performance of the WR GFRP construction.
If such root cause failure analysis is properly investigated, specifically by locating the size
and orientation of the defects, then a valid multi-scale durability and damage tolerance
model can be generated and implemented in any finite element software package in order
to predict the material state of composite structures under typical service environments.
effects, the use of a deterministic approach (that is, the current approach) to design and
analyze composite structures may undermine their high utilization and optimization
potential, which results in higher safety factors. Use of probabilistic design to predict and
quantify the structural strength and performance of certain FRP composites, particularly
sandwich composites, can help design engineers and also act as a reference to guide those
applications such as seawalls, piles, bridge decks, tanks, wind turbine blades, and so on.
23
sheets/skins, separated by a lightweight core material infused through the low-cost Vacuum
Assisted Resin Transfer Molding (VARTM) process. The face sheets are commonly
composed of woven roving or knitted e-glass reinforced vinylester composites. The core is
constructed of closed cell polymeric foams or fiber reinforced core materials. End-grain
wood balsa core is still being used in the spar caps and shear webs of offshore wind turbine
blades and other related transportation applications. This type of construction is in demand
in applications where a high strength and stiffness to weight ratio is an essential feature of
the design. In the northern region of the U.S. and Canada, these structures are often treated
by de-icing agents for the purpose of removing snow and ice. These agents contain a
by direct splashing or air-borne debris. The Freeze-thaw (F/T) cycling effect that results
are susceptible to degradation due to severe environmental exposure. Effect of F/T and
Hygrothermal and thermal cycling effects were found to cause damage such as resin
mechanical properties of the FRP materials dropped significantly when the F/T exposure
was combined with the salt water environment [30-32]. Pitting and cracking of glass fibers
A few studies have investigated sandwich construction under the F/T environment.
Zinno et al. [33] studied the effect of F/T thermal cycling on phenolic-impregnated
honeycomb sandwich with glass fiber skin. The authors reported less than 15% reduction
24
in the elastic modulus of the e-glass/face sheets and 20% reduction in the core shear
ultimate strength. Other researchers studied the effect of seawater environment and
mechanical degradation of foam materials (PVC and polyurethane) [32, 34–35]. The
authors reported an increase in the fracture toughness of the foam after exposure. Avilés
and Aguilar‐Montero [34] discovered that face sheets greatly prevent water ingress into the
sandwich structure; however, when moisture diffuses due to the concentration gradient, the
speed of diffusion is accelerated in the foam, along the thickness of the core [34]. This
effect could plasticize the resin and cause cracking. These cracks will weaken the bond at
the interface layer between the face sheet and the core. Consequently, this could affect the
core shear strength and modulus, that is, the most critical mechanical properties of
sandwich structures.
investigated [36–41]. Cantwell et al. [37] claimed that fracture toughness increased after
seawater exposure for balsa wood core sandwiches due to fiber bridging and greater energy
absorption. Singh and Davidson [40] observed that low temperature increased stiffness,
shear strength, and fatigue life of PVC foam core in sandwiches; however, the seawater
effect was minimal and influenced the static failure mode. Recently, Siriruk et. al. [41]
revealed that seawater alone reduced the interfacial toughness between the foam and the
25
CHAPTER 3
THEORETICAL APPROACH
parameters, parameters such as applied loads, material strength, and operational are
analysis model is developed for the entire system and solutions performed to yield failure
probabilities.
Based on the obtained experimental data, the following processes will be assumed
in this research:
anticipated loading conditions: the failure mode can change from location to location with
regard to the structure, with some areas being subject to more than one failure mode
reliability) for each failure mode and/or location on the structure, based on available code
26
3) The application of existing structural analysis methods utilized to model the
internal response (stresses) of the structure on the applied loads and to model the stiffness
of the structure.
4) Statistically, quantify defined random design variables that affect both the
probability of failure at predetermined locations. There are four available methods for this
purpose:
These methods will be examined; and the one that produces the most accurate
location failure probabilities: once failure probabilities of individual locations have been
calculated, the final step involves the determination of an overall probability of failure for
primary task for any probabilistic formulation that requires to be conducted systematically.
The first step of this process entails data collection; this further necessitates having an
appropriate data population. For this process, understanding, mean, variance, and skewness
27
Normal or Gaussian distribution, lognormal random variable, and Weibull random
several methods to determine the distribution and parameters from the observed data, such
as probability papers and statistical tests such as chi-square test and KS test. Method of
moment and method of maximum likelihood constitute the most frequently employed
structure. If the right probability distribution is not employed, the results can vary by over
an order of magnitude. The reason for this pertains to the tail type of different CDFs that
To address the problem of tail sensitivity, several methods are available. For
applications pertaining to civil engineering, LRFD is developed to address this issue. For
example, for steel, a lognormal distribution was utilized for materials properties [56]. For
[57]. However, for engineering wood materials, a Weibull distribution with two parameters
was introduced to describe these materials’ properties [58]. The reference resistance of
wood-based materials for LRFD is also well described in ASTM D5457-15 [59].
the Weibull distribution is first selected and if it was not rejected, it is utilized [60]. If it
was rejected, two other distributions (normal and lognormal) are examined.
28
In this study, normal, lognormal, and two-parameter Weibull distribution were
examined; to test the goodness of fit, several methods exist. As mentioned earlier,
probability distributions are similar in the center and the sensitivity is concentrated in the
tails; the Anderson-Darling test is sensitive in respect to tails [61] and was employed to
𝑛
1−2𝑖
An2 = ∑ [ {ln[𝐹0 (𝑥(𝑖) )] + ln[1 − 𝐹0 (𝑥(𝑛+1−𝑖) )]}] − 𝑛 Equation (4)
𝑖=1 𝑛
F0 (x) = CDF
N = sample size
For this goodness of fit method, the Observed Significance Level (OSL), which
29
Next step would be applying probability plots. A probability plot was formulated
for each category to examine the distribution; further, the R2 results for each one are
offered.
The calculation of the reduction factor, Ф, was performed based on the target
reliability value of 3 as it exhibits a reasonable safety factor for the structure. The target
reliability should not be too small, because the structure will not be safe, or too big as the
cost of the structure will increase accordingly in this case. The limit state of shear strength
is illustrated in equation (5); the variables in this equation are as follows: load (p), width
(b), and sandwich thickness (d). Sallowable constitutes the maximum allowable stress for the
𝑝
𝑔(𝑥) = (2𝑏𝑑) − 𝑆𝑎𝑙𝑙𝑜𝑤𝑎𝑏𝑙𝑒 = 0 )5(
The core shear modulus formula was calculated based on the following formula:
𝑈(𝑑−2𝑡)
𝐺= (6)
b(𝑑−𝑡)2
Where, t is the laminate thickness; U is transverse shear modulus and it can be calculated
𝑃(𝑆−𝐿)
𝑈= P(2𝑆3 −3𝑆𝐿2 +𝐿3 )
(7)
4[𝛥− ]
96𝐷
30
𝐸𝑓 (𝑑3 −𝑐 3 )𝑏
D= (8)
12
The average amount of 3.5 × 106 psi was used for Ef, the facing modulus of glass
The limit state equation used for the shear modulus is depicted in equation (9); G
allowable is the maximum allowable shear modulus for the target reliability of three.
𝑈(𝑑−2𝑡)
𝑔(𝑥) = − 𝐺𝑎𝑙𝑙𝑜𝑤𝑎𝑏𝑙𝑒 = 0 (9)
b(𝑑−𝑡)2
For compression strength calculation, the following limit state equation was used,
where a and b signify the length and width of the sample respectively:
𝑝
𝑔(𝑥) = 𝑎.𝑏 − 𝑠𝑎𝑙𝑙𝑜𝑎𝑏𝑙𝑒 = 0 (10)
following equation:
Where, p and Δ are applied load and recorded deflection for 0.001 and 0.003
deflections. A is the area of the sample. Figure 9 shows the method to calculate the chord
slope.
31
Figure 9. Chord slope calculation method
The limit state equation to calculate the compression shear modulus is based on
reliability and minimize the probability of overdesigning or under designing [65–72], and
the approach should be similar in appearance to the design of wood and steel to be accepted
and implemented in the civil engineering industry. It is specified as Ф𝑅𝑛 ; it constitutes the
The nominal strength depended on several factors and is defined based on the
𝑅𝑛 = 𝐶1 𝐶2 … 𝐶𝑘 𝑅0 (13)
32
Ro is determined based on a laboratory test, mostly from standards such as ASTM
or other international standards test methods. It is based on the 5th percentile of the
µ
Ф = (𝑅𝑅 ) exp[−𝛼𝑅 𝛽𝑉𝑅 ] (14)
𝑛
In this equation, µR signifies the mean of R (in the end-use condition), VR the
equal to 0.6, while 𝛽 constitutes the reliability index. For the calculation of Rn, several
methods related to different kinds of distribution are viable. Further, the equation Ф.Rn =
Using a target reliability value of three, the allowable load was determined by the
First Order Second Moment (FOSM) method to calculate Rn. For Weibull distribution,
ASTMD5457 introduced a formula that is very useful, which is shown in equation 15:
Rn = RP × Ω ×KR (15)
Where,
33
p = percentile of interest expressed as a decimal (for example, 0.05)
The shape (α) and scale (ƞ) parameters of the two-parameter Weibull distribution
can be calculated for shear and compression strength. Parameters’ estimation can be done
through gamma function, but in this research, parameters were calculated through Minitab
software.
The data confidence factor, Ω, accounts for uncertainty associated with data sets.
This factor, which is a function of the coefficient of variation, sample size, and reference
that provides data confidence factors appropriate for lower fifth-percentile estimates [60].
estimate (for example, R0.05) to achieve a target reliability index (β). The reliability
normalization factor comprises the ratio of the computed resistance factor to the specified
resistance factor, adjusted by a scaling factor. These data sets are available in standard
34
CHAPTER 4
EXPERIMENTAL APPROACH
of common core materials utilized in infrastructure applications were selected and molded
was applied to mold all sandwich composite panels. This process is generally employed
for large civil engineering structural applications. The materials selected to manufacture
35
Core materials
Face sheets
36
Figure 10. VARTM process of molding the panels
The materials employed for face sheet consisted of six layers of E-glass plain weave
fabric (813 g/m2) reinforced with Derakane 610C-200 epoxy-vinyl ester resin (Ashland
Inc.). This resin is currently used for several infrastructure applications due to its durability,
toughness, and fire resistant properties. It is comprised of a liquid blend of vinyl ester
monomers in styrene, in which the styrene content is approximately 40–50 wt%. Although
this type of thermosetting polymer resin is often referred to in the industry as an “epoxy
vinyl ester” resin, it does not actually contain any reactive epoxy groups, nor does it cure
through the reaction of epoxy monomers with methacrylic acid that replaces all the epoxide
groups with methacrylate groups. This reaction leads to the formation of a carbon-carbon
double bond (referred to as vinyl) on the ends of each monomer that is attached to the
central monomer through an ester linkage, thus the term “vinyl-ester”. The main purpose
is to facilitate the initiation of the polymerization at room temperature with the utilization
37
carbon double bond. However, the presence of the ester group further facilitates the
The initiator system utilized was Cadox 50 at a level of an amount of 1.25 wt%.
This material primarily comprises a solution of methyl ethyl ketone peroxide (MEKP).
These molded cross-ply laminates were allowed to cure for 30 days under ambient
conditions (23 °C and 50% RH). Nominal properties of composite constituents are listed
in Table 1. The fiber volume fraction was approximately 53% with a total measured
The measured thickness of each face sheet was 3.24 ± 0.1 mm, with a 53% fiber
volume fraction. For compression and facing cleavage tests, 3 plies of E-glass fabrics
constituted the face sheets of the sandwich panels. The facing cleavage test standard ASTM
E2004 was selected in this study, since it targets infrastructural applications, such as
In the presence of a saline water solution (3% salt), 200 cycles of freeze/thaw (F/T)
(– 20 °C to 20 °C) cycles were conducted on all samples (Figure 10). The F/T slab tester
by Qualitest TM was utilized to expose all samples. To facilitate the initiation of the process
of degradation in the core material, all cut specimens were placed in stainless steel pans
inside the test chamber and submerged halfway along the thickness (for laminates was full
immersion). This immersion approach serves to simulate the long-term exposure and saline
38
water diffusion in the core material. Figure 11 depicts the 12-hour period of F/T cycles
employed in this study. The total aging process for the exposed samples lasted for 100
days. Figure 12 exhibits the samples placed inside the F/T machine.
Figure 12. Left: F/T machine with samples inside; Right: flexural samples (top)
39
4-3 Test Setup
Post the process of curing, samples were cut and extracted from panels for the
Table 3.
Standard Test
Method for Facing
Trapezoidal
Cleavage of 6/6 6/6
(see Figure 3)
Sandwich Panels
(ASTM E2004)
Note: Exposed = 200 cycles, unexposed = 0 cycle
40
Table 4. Number of samples and dimension for GFRP laminates
139.7 × 6.35 × 12
Laminate Compression Test (ASTM D6641)
3.24
A four-point bending test fixture illustrated in Figure 13 was employed for the
purpose of this investigation. This fixture restrains any longitudinal rocking motion, while
it serves to fatigue the sandwich samples. All specimens were 406 mm long, 76.2 mm wide,
and 32.7 mm thick, including the core thickness (25.4 mm). Preliminary static testing was
conducted at 0 cycle (ambient specimens) and post 200 cycles (exposed) in accordance to
ASTM C393 [13] and D7250 [14]. The total number of samples is provided in Table 2.
The fatigue testing was performed for R = – 0.1 (R= max. load/min. load), compression-
compression fatigue was performed at a testing frequency of 4 Hz. Since this research
41
Figure 13. Four-point bending fixture for fatigue testing
Test Type Load Max. load (N) Min. load (N) Mean (N) Amplitude (N)
Test Type Load Max Load (N) Min Load (N) Mean (N) Amplitude (N)
42
The compression test was conducted with an MTS machine; Figure 14 depicts a
sample that comprises three layers of laminate on each side under the compression test.
The peel test setup was accomplished according to ASTM E2004, for both balsa
and a foam core specimen; Figure 15 displays the setup condition for this test. To prepare
the samples, two pieces of steel were bonded with an adhesive (LORD 7075) as depicted
in Figure 14a: one side of the sample interface of core/laminate was cut with a saw as
shown in Figure 14b. In Figure 15c, a sample assembled on MTS machine has been
43
a)
b) c)
Figure 15. Peel test a) steel T-shape tabs bonded to the specimen;
b) specimen with interface cut area; c) specimen ready for the test
exposed and unexposed laminate specimens. Two strain gages were placed in the center of
44
the Iosipescu specimens (ASTM D5379-12) to achieve strain values of + 45 and – 45
degrees; the final shear strain was calculated through the average of the absolute values of
individual strain gage results. The shear strain of 0.05% to 0.15% was utilized to calculate
the shear modulus. Figure 16 illustrates a coupon under the test comprising the strain
gauges.
For tensile (ASTM D3039) and compression (ASTM D6641) tests, gages were
installed in the center of the coupon, longitudinal and transverse direction strain values
were calculated with regard to the tensile test. For the compression test, gages were
installed on each side of the specimens to detect any possible buckling while testing. Shear
strain values between 0.05% to 0.25% were utilized for both tests to calculate the shear
modulus. Figure 17 depicts the tensile and compression test setups respectively.
45
a) b)
To study the effect of F/T on the microstructure and physical properties of polymer
and composite, the following characterization experiments were applied prior to and
subsequent to the provision of environmental exposure; the results of analyses were also
compared:
46
Glass transition temperature
Residual curing
Operating parameters for the test were maintained at 25–130 °C, 5 °C/min, cool,
re-reheat; the TMA model was constituted by TA Instrument Q400; the sample size was
determined as 10 x 10 mm.
In this technique, a sinusoidal stress is applied and the strain in the material is measured
that allows the determination of the complex modulus. The temperature of the sample or
the frequency of the stress is often varied, contributing to variations in the complex
was performed. The DMA unit was constituted by a TA instrument Q800. This test was
applied to the balsa wood and foam core materials for core type materials. Two types of
fixture, 3-point bending and shear were utilized; for laminates, merely a 3-point bending
fixture was utilized. The samples were of the following dimensions: 10 mm x 60 mm; the
47
a) b)
The ultrasonic reflector plate with the transducer frequency as 5 MHz, scan/index
length as 15 inches, scan/index increment value 0.02 inches were applied to ascertain,
whether the samples test pieces were free of discontinuities or defects. Figure 19 depicts a
specimen during the performance of the ultrasonic scanning tests in the water tank.
Figure 19. Specimens inside the water tank and an ultrasonic scanning probe
48
This technology possesses the ability to exact three-dimensional cross-sectional
images of the entire part. To perform the X-ray CT-scan Inspection, HMXST high-
performance Real-Time X-ray, X-Tek real-time HMX160 was applied to obtain images
from different cross sections of the specimens. Several samples of sandwich composites
and laminates prior to and subsequent to the environmental exposure were tested.
employed to examine the microstructure, defects, and any kind of degradation on the
surface of samples.
6) Optical Microscopy
The DSRL and the Keyence Optical Microscopy was utilized to observe phase
49
CHAPTER 5
All exposed sandwich samples reported in the following sections were removed
from the environmental F/T chamber and allowed to dry for 72 hours before testing.
The areal weight of balsa wood composite sandwiches shows in figure 20, the
samples reached to saturation point in around 150 cycles of F/T.
12
10
8
Kg/m2
6
4
2
0
0 50 100 150 200
Cycles
Figure 20. Areal weight of balsa wood core sandwich after F/T exposure
50
Figure 21 shows a resin molded balsa wood, in 50 and 200 cycles of F/T, in 200
cycles lot of cracks has been expanded on the surface and inside of balsa wood.
a) b)
Figure 22 below shows the balsa core shear strength and modulus before and after
200 F/T cycles (total of 60 samples). Results are presented in bar charts with a 95%
confidence interval of the mean. Referring to Figure 4, a 13.9% reduction in shear strength
was observed and a coefficient of variation (COV) of 13% at 0 cycle and 21% at 200 cycles.
The shear modulus decreased by only 4% after exposure with larger COV in modulus of
20% at 0 cycles, and 41% at 200 cycles. An analysis of variance (ANOVA) revealed that
the reduction in core shear strength was statistically significant (P-value= 0.008), whereas
the core shear modulus was not statistically significant (P-value= 0.6).
51
Figure 22. Balsa core shear strength (left) and modulus (right) -ASTM C393 and D7250
The failure mode for unexposed and exposed specimens has shown in figure 23 and
24 respectively.
52
Figure 24. 200-cycles core shear failure mode
The core shear failure modes for the unexposed samples formed a 45-degree angle
in the balsa wood core and along the end grain (cells). For the exposed samples, the
fibrillation mechanism of balsa cells was noticeable at the shear span of the four-point bend
test (Figure 24). This large fibrillation is due to the excessive absorption of moisture and
degradation due to F/T cycling. The thermal cycling decreased the shear strength of the
balsa wood core and slightly increased the ductility of the sandwich structure. Figure 25
presents the average trendline of exposed and unexposed specimens when tested according
53
As illustrated in Figure 25, the core shear stress vs. displacement plots show better
toughness for the exposed specimens (200 cycles). Figure 26 shows fatigue failure mode
for balsa wood. Three distinct crack events occurred while cycling the specimens, 1) crack
initiation and propagation just below the load point, 2) core shear along the wood cells
(with fibrillation for the exposed samples) and 3) bottom skin/core delamination.
The simulated lifetime core shear fatigue is shown in Figure 27. It is interesting to
observe that the curve fit trendline of the exposed samples showed an increase in fatigue
life at high-cycle levels. This observation justifies the higher toughness presented in Figure
25. This toughness could be regarded as the energy dissipated due to friction in between
the fibrillated cells while cycling the specimens at 4 Hz. Similar trend was also observed
by other researchers [13] for tension-tension fatigue of wood perpendicular to the grains.
54
Figure 27. Balsa wood core shear fatigue life prediction for exposed (200 cycles) and
A nondestructive technique was sought to justify the increase in core shear fatigue
life of the balsa core. Figure 28 shows the optical microscopy images of two compressed
balsa wood sandwich samples at 0 cycle and 200 cycles. The fibrillation of the wood cell
structures is clearly visible after 200 cycles of F/T exposure (Figure 28 b). It seems that
under F/T cycling, water absorbed by the wood cells (unfilled with resin) through capillary
action caused expansion and contraction inside the cellular wall structure of the balsa. This
sequentially disintegrates the wood cells into several parts and indirectly created a longer
path for the crack to travel along the core while fatiguing the specimen.
Figure 28. Balsa wood texture a) unexposed b) exposed samples (optical microscopy)
55
The results of compression strength on the sandwich specimens with balsa core
before and after F/T exposure is shown in Figure 29. A 36% reduction in compression
strength was observed with a COV of 21% at 0 cycle and 27% at 200 cycles. The
compression modulus decreased by 33% after exposure with a COV of 33% and 30%
respectively. The reduction in balsa core compression strength and modulus were
statistically significant.
Figure 29. Balsa core flatwise compression strength (left) and compression
Figure 30 and 31 show the facing cleavage test results in terms of fracture toughness
(J/m2) for the Balsa wood core samples. A total of 12 samples were tested, with 32%
increase occurred after exposure (200 cycles). The COVs for the unexposed and exposed
were 48% and 30%, respectively. The failure modes experienced in the Balsa core samples
were pure face sheet/core interface failure mode. Due to the large COV and data scatter,
an ANOVA was conducted. Even though the mean increased after exposure (200 cycles),
56
the result (P-value = 0.185) did not support that a strong evidence exists to conclude that
the fracture toughness means are different between the two exposures.
Figure 30. Load- stroke curve for exposed and unexposed specimens
Figure 31. Balsa wood fracture toughness for unexposed (0 cycles) and exposed (200
57
Figure 32 shows a sample under test and failure modes. Not so much different
a)
b)
Figure 32. a) A specimen under peel test, b) failure modes of exposed and unexposed
specimens
58
5-2 Foam Core Sandwich Composite Properties
Figure 33 below shows the core shear strength and modulus before and after 200
F/T cycles for the PVC foam core (Airex C70.55) sandwich samples. The 95% confidence
interval shows 3.2% increase in core shear strength and approximately 30% increase in
core shear modulus. The core shear strength had a covariance of 3.8 % at 0 cycle and 4.8
% at 200 cycles. The shear modulus had a covariance of 7.3 % at 0 cycle and 5.5 % at 200
Figure 33. Foam core shear strength and shear modulus -ASTM C393 and D7250
A specimen after the shear test is shown in figure 34, pure core shear failure mode
was observed in exposed and unexposed (ambient) specimens. The failure occurred at 45-
degree angle in the PVC foam. Typical core shear strength vs. displacement plots (average
of 30 samples) are shown in Figure 35. The 200 cycles (exposed samples) plot clearly
59
indicates an increase in the core shear stiffness. Contrary to the balsa core, no increased in
Figure 35. Four-point flexure results for foam core sandwich samples
To better understand the increase in core shear strength and stiffness after exposure,
microstructure change and variation in the PVC foam network. Figure 36 shows the SEM
60
micrographs of foam core before and after exposure. Examining the foam microstructure
after exposure (Figure 36b), the network shows negligible dimensional change with few
cracks in between the cellular structures of the foam. The X-ray computed tomography
picture shown in Figure 36c quantify the resin thickness absorbed by the foam core. This
Figure shows the absorbed resin confined by the foam cells creating a “key” anchorage
zone that could potentially increase the core shear strength and improve the face sheet core
connection. During the cycling events, the resin absorbed by the foam was continuously in
the glassy state, this in turns restricts chain mobility and segmental motion inside the
Figure 36. SEM: a) unexposed foam and b) exposed foam after 200 F/T cycles, X-Ray
61
The SEM and X-ray computed tomography results were not sufficient to clarify the
increase in stiffness of the PVC foam core. The dynamic mechanical analysis (DMA) in
shear sandwich mode was used next to assess the morphology and viscoelastic transition
response of the foam core. Figure 37 below shows the variation in the storage modulus and
loss modulus with temperature. It is clearly shown that the storage modulus of the foam
increased after 200 cycles of F/T exposure, which is consistent with the results of the
ASTM C393 and D7250 tests. This implies that the freeze/thaw exposure increased the
stiffness of the foam network. Interestingly, the loss modulus also increased from the
freeze/thaw exposure, yet there was no significant change in the glass transition
temperature (as taken from the peak in the E” curve). However, a standard DSC test
conducted on the foam showed that the Tg increased by 4°C after exposure, which could
not fully understood, but it does imply the freeze/thaw exposure produced a structural
change in the foam that could be from chemical and/or physical means. This result is not
which involves a decrease in Tg and increase in the glassy state storage modulus. In the
current study, the Tg remained the same, while both E’ and E” increased.
62
Figure 37. DMA testing results, a) shear mode testing, b) 3-point bending mode
The failure mode of a fatigue test is shown in Figure 38. The failure mode for both
the exposed and unexposed specimen was same with a 45-degree crush. The fatigue plot
for the Airex C70.55 PVC foam is shown in Figure 37. During the fatigue test, three stages
of fatigue crack growth occurred in the foam core specimens. The fatigue life of the
63
Figure 38. Fatigue failure mode of PVC foam
Figure 39. Foam core shear fatigue life prediction for exposed (200 cycles) and
The results of compression strength on foam core samples before and after F/T
exposure has shown in figure 40, 1.5% decrease in compression strength and 2% increase
in compression modulus were recorded after 200 cycles of F/T. Covariance percent for
64
unexposed specimens was 9.42% and 12% for strength and modulus respectively, this
Figure 40. Foam core flatwise compression strength and compression modulus
As the compression and shear strength of foam does not change so much, it means
the average there is no difference in the dissipated energy after F/T effect. These results
can confirm same damping properties and consequently same fatigue life. However, after
F/T, 27% increase in shear modulus happened, means after F/T lower fatigue life is
6 specimens of exposed and 6 unexposed were tested under peel test; the load-
stroke curve is shown in figure 41, in total 8.1% in energy density occurred after exposure
65
to freeze-thaw, the covariance percent for unexposed and exposed was 8.6 and 8.3 percent
Figure 42. PVC Foam fracture toughness for unexposed (0 cycles) and exposed (200
66
Figure 43a shows a specimen in middle of peel test, figure 43b shows the failure
modes of exposed and exposed specimen, no significant change was observed in these
specimens.
a)
b)
Figure 43. a) A specimen in middle of peel test, b) specimens after peel test
67
5-3 Laminate Tests Results
Table 7 shows the average mechanical test results and Coefficient of Variation
(COV) for each standard test (average of 12 samples per standard test). All mechanical
characterization tests were conducted using a calibrated test machine (INSTRON 4208).
The COV for strength and modulus for all specimens were less than 15.3 %.
Average Average
Sample Modulus Strength
Test Direction Modulus Strength
Type COV% COV%
(GPa) (MPa)
Normalized moduli and strengths for shear, tension, and compression are shown in
Figure 44. Large reductions occurred in the tensile and compressive strengths of the GFRP
material. The shear and tensile longitudinal moduli were also reduced after F/T exposure.
In general, the compression strength in WR fabric is lower than the tensile strength; this is
68
due to the already kinked fibers; however, this is not the case after exposure (Table 6, row
6 and 8). The tensile test results are mainly dominated by the fiber properties and clearly,
indicate that the longitudinal fibers and the structural integrity of the WR have been
the mechanism of deterioration of the exposed laminates, is described in the root cause
Figure 44. Overall results of exposed (200 cycles) and unexposed (0 cycles)
specimens
with the v-notched beam test (ASTM D5379-12) [12]. Figure 45 shows typical shear failure
mode that occurred at the mid-span of the V-notched beam. The average reduction in shear
69
modulus and strength after 100 days of exposure is 18.5% and 13.5 %, respectively (Figure
46). Similar reductions were reported by Guzman and Brøndsted [13] after 8 years of
saltwater exposure.
Figure 45. Shear failure modes for a) unexposed; and b) exposed specimens
Following the shear test, failed samples were cut in between the V-notch locations
(across the length of the sample) and dissected using an optical image technique. Figure 46
shows the final fracture states and internal cracks for the unexposed and exposed
specimens. The unexposed failed specimens exhibited interlaminar cracks and separation
in between the resin and warp fibers. The cracks were oriented along the warp direction.
70
For the exposed failed samples, a high density of intralaminar cracks occurred along the
fill fibers as well as their interfaces (Figure 47). This clearly shows that the resin has been
Figure 47. Micrograph of the cross section after failure at the V-notch location a)
intralaminar cracks
In this test, a 23% reduction in the tensile strength and 19.7% in the longitudinal
modulus occurred after exposure (Figure 48). These values are significantly higher than
the COV listed in Table 11. Since the ASTM D3039 [14] test results are mainly dominated
by the fibers, this presumably indicates that the fibers were damaged in the longitudinal
15 % [15]. Lateral strains (90-degree direction) were used to monitor and extract the
stiffness of the laminate in the transverse direction. The transverse modulus decreased by
10 %. The failure modes are shown in figure 49. In addition, it is interesting to note that
71
Guzman and Brøndsted [13] reported similar reductions in transverse tensile modulus and
Figure 49. Tensile test failure modes; unexposed and exposed specimens
72
Figure 50 shows the tensile stress-strain curves in the longitudinal and transverse
directions. For the exposed samples, the transverse strains are more pronounced than the
unexposed samples. These highly deviated curves designate some sort of stiffness
degradation due to a loss of the structural integrity of the laminate, especially at the weave
junctions (nodes) between the 0 and 90-degree fibers of the WR construction. This could
be a direct result of a weakened matrix that binds the warp fiber to the fill fiber. More in-
depth validation of this key failure mechanism is explained in the non-destructive testing
section.
Figure 50. Longitudinal and transverse stress-strain curves for exposed and unexposed
specimens
The combined loading compression test fixture was used to study the effect of
exposure on the compressive properties of the GFRP laminates. A total of 12 samples were
tested in accordance with ASTM D6641 [16]. The exposed samples showed an average
reduction of 17.1% in strength and 12% in modulus. One can notice the high dependency
73
of the matrix degradation and interfacial failure on the compression strength versus the
shear strength observed in Figure 46. Typically experienced failure modes are shown in
figure 51.
Figure 51. a) Percent change in modulus and strength, b) Compression failure modes
To study the internal damage or defects that occurred due to these environmental
conditions, a TMA and DMA tests were conducted on unexposed and exposed laminates.
The DMA results are given in Figure 52. Figure 52a shows the results obtained
from the 1st heating cycle (25-140°C @ 5 C/min). After cooling the sample in the
apparatus, a 2nd heating cycle was run on the same specimen. The reheat cycle results
(Figure 10B) show the condition of the material after being post cured. This analysis is
74
commonly run for thermosets because usually they are not fully cured in the original
manufacturing process.
viscous behavior in response to time-varying stress. The amount of each type of behavior
depends on the state of the polymer structure (chemical composition, crosslink density)
and the temperature. The storage modulus (E’) characterizes the elastic energy stored by
the polymer, while the loss modulus (E”) represents the energy lost to viscous dissipation.
Polymers that are highly crosslinked usually have high values of E’. Also, the shape of the
E’ and E” curves can be used to locate the temperature at which the polymer structure
transitions from a glassy state (at low temperature) to a rubbery state (at higher
temperature).
Referring to Figure 52a, the sample subjected to 200 freeze-thaw cycles exhibited
a 20% decrease in E’ near room temperature. This could be due to a variety of reasons,
degradation of the fiber-matrix interface, etc. Chemical decomposition of the matrix could
be explained by hydrolysis of the ester linkage in the vinyl ester matrix, although hydrolysis
reactions usually slow at the temperatures encountered. Therefore, the change may be due
to physical degradation of the resin and/or fiber-matrix interface from the freeze-thaw
exposure.
The E” curves provide more information about the phase structure of the resin.
Cured vinyl ester resins usually are comprised of two polymers phases: 1) the main
crosslinked structure which is comprised of the vinyl ester-styrene copolymer network, and
75
2) a phase rich in homopolymerized linear polystyrene. The peaks in the E” curve
correspond to the temperature at which each phase transitions from glassy to rubbery state.
“phase 1” peak. The glass transition of the copolymer network is generally observed to be
higher: in this case, it occurs at 115°C and is labeled as “phase 2” in the 0-cycle sample.
It is interesting that the 200-cycle sample does not exhibit the phase 2 transition in the same
location. Most likely, the phase 2 transition was reduced to about 75°C (labeled as phase
2’, which is the shoulder in the E” curve, as well as the steep drop in E’). Absorbed water
would be expected to plasticize the main network because it has some polarity (e.g. ester
bonds), while water should have little or no impact on the highly nonpolar polystyrene
Figure 52B shows the results after post-curing the samples. Any water in the matrix
is expected to be removed in the 1st heating cycle. The value of E’ for the 200-cycle sample
remained lower (17.5%) than the control sample, which indicates that there was some
polystyrene phase (phase 1) was largely unaffected by the post-curing process in both 0
and 200 cycle samples. In the 0-cycle sample, the E” peak of the vinyl ester phase shifted
to a slightly higher temperature (120°C), which can be attributed to residual curing from
the 1st heating cycle. In the 200-cycle sample, the vinyl ester E” peak is seen as a faint
shoulder around the same temperature as the 0-cycle sample. This indicates that the vinyl
ester network was no longer plasticized, but was not as prevalent as the 0-cycle sample.
As with the E’ result, this also indicates some difference in the network or fiber-matrix
76
b
Figure 52. DMA results for composite, a) 1st DMA heating cycle, b) 2nd DMA heating
cycle.
The DSC results for a neat resin samples are summarized in Table 8. This technique
was used to verify that the samples were not fully cured after VARTM processing, and to
measure the Tg of the resin independently of the presence of fibers. The Tg is strongly
related to the nature of the crosslinked network, and therefore it can provide information
about the decomposition of the polymer. The DSC traces for the 1st heat showed a Tg
77
around 67°C for the 0-cycle sample and 63°C for the 200-cycle sample. This difference
was attributed to plasticization of the matrix from the freeze-thaw cycle, as observed in
DMA. However, upon reheating the Tg increased to essentially the same value (96-97°C)
in both cases, due to post-curing experienced in the prior cycle. This implies that there was
primarily to physical degradation of the polymer (cracks) and/or between the polymer and
fiber. The small change to phase 2 detected by DMA is not able to be detected by DSC.
Tg (°C)
# cycles
1st heat 2nd heat
68.0 96.3
0
66.5
62.4 97.4
200
64.3
To investigate the size and location of the internal defects, a 3D X-Ray CT-Scan
test was conducted on the exposed specimens. Figure 53 shows resin cracking at preferred
sites. In addition, intralaminar cracks occurred in the mid-section of the FRP laminate,
specifically in resin rich areas. The damage is shown in Figure 52 (transverse cross-section)
is approximately 1.2 mm in length and 0.7 mm thick. These defects (cracks and fractures)
78
are a source of delamination initiation that further degrades the component strength and
c)
a) b)
Figure 53. X-Ray computed tomography showing damage in exposed samples a) surface
To test the structural integrity of the connections between the fill and warp fibers,
a pulse ultrasonic C-scan technique was performed on two exposed and unexposed
specimens. Figure 54 shows the C-scan images of all tested samples. The grid pattern
(signals at nodes) is clearly shown in the unexposed samples (figure 53a), whereas the
linear pattern is clearly visible in the exposed samples. This change in pattern indicates that
these connections are impaired and ineffective in transferring the loads, including racking
or shear loads. In addition to the shear stiffness loss of the cracked matrix, the deterioration
of these nodes is believed to affect the tensile and shear moduli of the laminate. To interpret
this phenomenon, one could envision the laminate structure like a picture frame (hinged
truss) with some degree of rotational stiffness at the four pins (nodes). Two of these pins
79
could relate to a horizontal spring simulating the lateral stiffness supplied by the matrix. If
a load (tension of shear) is applied to the truss, the capability of the truss to resist load and
deformation (axial or racking) is controlled by the rotational stiffness supplied by the nodes
(bond at the junction of fill and warp fibers) and the axial stiffness of the spring. This
simple mechanistic model helps to understand the basic phenomena governing the behavior
of these systems.
a) b)
Figure 54. Ultrasonic C-scan results (102 mm wide specimens); a) unexposed: grid
pattern, b) exposed: linear pattern
exposure, a high-resolution optical image technique was used to inspect their locations.
Figure 55 shows damaged fibers and matrix bond failure between the fill and warp fibers.
This failure pattern commonly occurred at the kink/weave location between the roving. It
seems that the volume changes due to the thermal cycling initiated high non-uniform
residual stress at the sharpest curvature location and significantly degraded this
node/junction between the fibers. The authors believe that this damage will be recurrent
for any plain weave type no matter what type of reinforcement is used.
80
Figure 55. Local magnified optical image at weave junction showing matrix failure
observed. The mismatch in the coefficient of thermal expansion between the resin and the
E-glass fiber is believed to cause this interfacial failure mode. In addition, fiber fractures
fiber fracture
81
Karbahari et. al. reported that aqueous solution in contact with glass fiber surface
produced free-alkali hydroxide groups that degraded the silica structure of the glass fiber
when exposed to an alkaline media [11]. The effect is through breaking of Si-O bonds in
the glass network and was believed that surface loss and pitting occurred in areas of high
the glass fiber through several processes such as pitting, etching, leaching, and
sample. Several defects (black spots) are observed, including localized areas around these
spots which might be related to coupling agent or polymer swelling. Research on this
82
5-4 Model Results:
Tables 9 and 10 present the statistical values for each of the random variables
Table 9. Statistical values for balsa and foam core shear test of sandwich composites
Test method Statistical Core shear strength, MPa Width (b), beam thickness
value mm (d), mm
0 cycle 200 cycles
D7250)
Modulus
Stan-dev 35.88 70.38 0.03 0.279
D7250)
Modulus
Stan-dev 1.370 1.262 0.462 0.144
83
Table 10. Statistical values for balsa and foam core compression tests of sandwich
composites
Test Method Statistical Core shear strength, MPa Width (b), beam
value mm thickness
0 cycle 200 cycles
(d), mm
Compression strength
Stan-dev 2.261 1.913 1.036 1.041
Compression
Stan-dev 172.28 106.11 1.036 1.041
Modulus
Compression
Stan-dev 0.095 0.192 0.713 0.605
Strength
Compression
Stan-dev 6.245 4.355 0.713 0.605
Modulus
The OSL values for normal, lognormal, and Weibull distribution for each one of
the materials were calculated, and they are provided in Table 11.
84
Table 11. Data results of OSL for normal, lognormal, and Weibull distribution
The results of applying normal, lognormal and Weibull distribution on data has
85
Figure 58. Shear strength statistical plots of balsa core material
86
Figure 59. Shear modulus statistical plot of balsa core material
87
Figure 60. Shear strength statistical plot of foam core material
88
Figure 61. Shear modulus statistical plot of foam core material
89
2. Compression Test Results
90
Figure 63. Compression modulus statistical plot of balsa core materials
91
test
92
Figure 65. Compression modulus statistical plot of foam core material
93
Based on the results of the Anderson-Darling test and probability distribution plot,
Table 12. Selected distributions for balsa core and foam mechanical properties
composite industry; further, it is also recommended by Handbook MIL 17 [61, 64]. This
does not imply that Weibull distribution will necessarily form the most suitable fit for these
composites. However, due to a lack of available data with a sufficient size unlike steel and
concrete, the most common distribution in the composite forms a better alternative for us
and will facilitate the modeling of an appropriate resistance factor and end-use condition
for purposes of practical design. Moreover, the design codes of wood structures are based
on Weibull distribution and will help LRFD similar to it. In this thesis, a distribution was
The distribution for each one of the variables is illustrated in Table 13.
94
Table 13. Selected distributions for variables
p Δ b d
Parameters Estimation
Table 14 demonstrates the results of maximum load, nominal strength, and the
resistance factors for the target reliability value of three and for each type of test methods.
95
Table 14. Resistance factor results
Nominal Max allowable Resistance
Test Method
strength, Rn stress, S (MPa) factor, Ф
Based on the results in Table 14, it is possible to determine the environmental factor
or F/T (Cf-t) for both balsa core and foam core composite sandwiches by the application of
96
Table 15. F/T coefficient introduced in this research
The coefficient of F/T for foam core sandwiches is equal to 1, which means that
freeze-thaw did not have any significant effect on the foam shear strength, but the effect of
F/T on balsa core sandwiches was substantial. These outcomes are in agreement with the
experimental results.
97
CHAPTER 6
The degradation and change in the microstructure of the core materials were
1- The most significant reduction due to F/T cycling combined with saline
and 33%, respectively. In addition, the balsa wood core shear strength decreased
by 14%
2- The F/T cycling increased the PVC foam shear modulus by 25%. This was
storage modulus of the foam after exposure. In addition, the DSC results
98
3- Fatigue life of balsa wood increased after exposure. The optical microscopy
mechanical fatigue from expansion and contraction. The friction in between the
cell walls of the balsa wood increased its damping characteristics; in addition,
while fatiguing the specimens. This behavior was justified by the increase in
loss modulus of the DMA tests performed in both the bending and shear modes.
4- The facing cleavage of sandwich panels test (ASTM E2004) showed minimal
effect on fracture toughness of the balsa wood and PVC foam core materials
composite submerged in saline solution and exposed to 100 days of F–T environment was
studied. State-of-the-art NDI testing techniques were used to investigate the nature of the
mechanistic model describing the failure mechanism and damage initiation and
compared to identical control specimens. The largest reduction was observed in the
tensile properties of the GFRP. A reduction of 23% in strength and 20% in modulus
was detected. The average reductions in shear modulus and strength were around
99
2. The shear failure postmortem optical microscopy images showed delamination and
3. The DMA testing indicated plasticization of the matrix from the exposure. The
water), but there was a permanent decrease in storage modulus that is attributed to
hydrolysis of the resin was not involved. DSC results corroborated these findings.
These results are consistent with the formation of internal cracks inside the GFRP
laminate.
4. The fiber/matrix interface has traditionally been viewed as the main damage due to
thermal cycling. However, when F–T is combined with saline exposure, the warp
fracture.
5. The X-ray computed tomography showed a preferred site location of defects in the
WR construction. The frequency of these defects was mainly located in the resin
6. The ultrasonic C-scan and optical microscopy showed that the volume change due
binding the warp and fill fibers at the weave junction impaired the structural
7. The observed strength and stiffness reductions after accelerated exposure were
similar to what was reported in the literature on GFRP immersed in seawater for up
to eight years. The testing scheme and protocol proposed in this study could be used
100
as an accelerated testing method for future evaluation of durability in composite
materials.
8. The results of this research are expected to apply to plain weave fabric regardless
durability and damage tolerance model would take advantage of the damage types
presented in this paper, including the frequency, sizes, shapes, locations, and
101
CHAPTER 7
such as seawalls, bridge decks, tanks, and wind turbine blades. However, a major concern
in this regard has been whether the sandwich composite, especially the core material, offers
environments. Both novel and existing structures demand more extensive investigation
with regard to the selection of the strength reduction factor (also referred to as safety factor)
or other factors imposed on their design equations. The current values of these factors have
been typically determined based on subjective judgment, and therefore they still require
concrete evidence for validation. This validation can be obtained through a reliability
analysis, which relates the probability of failure to the safety factor and the amount of
tolerated load, providing a basis for their calibration to attain the targeted level of safety.
For this purpose, two types of sandwich core composites were designed and
fabricated using a woven E-glass reinforced vinyl ester resin. This construction included
relatively thick face sheets that promoted core shear failure in all beam specimens. The
effect on core mechanical properties (shear and compression) before and after 100 days of
F/T exposure in saline solution (3% NaCl) was investigated. Shear strength, modulus, and
102
lifetime simulated fatigue testing properties of the balsa wood (SB100) and PVC foam
(Airex C70.55) was assessed.The main conclusions from this study are the following:
The F/T effect reduced the mechanical properties of balsa core sandwich
Say something about whether the use of models was successful, and which
reduction factor (ɸ) will increase the chance for sandwich composites in civil engineering
applications. The probabilistic approach used in this study is a means to reach to this
goal. Currently these factors are not available for designing sandwich composites in civil
engineering applications. This work provided a basis for using this approach and calculated
initial values for consideration. In addition, this study contributed new and detailed
experimental data on the behavior of balsa and foam core composites in freeze/thaw saline
environment, which will augment what little is currently published in the literature on this
subject.
103
BIBLIOGRAPHY
4- Ishai, Ori, Clement Hiel, and Michael Luft. "Long-term hygrothermal effects on
47-55.
459.
6- Sun, Wei, et al. "Damage and damage resistance of high strength concrete under
the action of load and freeze-thaw cycles." Cement and Concrete Research 29.9
(1999): 1519-1523.
104
7- Barbero, Ever J. Introduction to composite materials design. CRC press, 2010.
1997;55(2):171–7.
913
10- Dimitrov, Nikolay, and Christian Berggreen. "Probabilistic fatigue life of balsa
11- http://www.fireco.no/?news=bascule-frp-composite-footbridge-fredrikstad
13- Vural, M., and G. Ravichandran. "Dynamic response and energy dissipation
14- Kahle, E., and J. Woodhouse. "The influence of cell geometry on the elasticity of
238.
16- Soden, P. D., and R. D. McLeish. "Variables affecting the strength of balsa wood."
The Journal of Strain Analysis for Engineering Design 11, no. 4 (1976): 225-234.
105
17- Da Silva, Andre, and Stelios Kyriakides. "Compressive response and failure of
balsa wood." International Journal of Solids and Structures 44, no. 25 (2007): 8685-
8717.
18- Bond, I. P., and Martin P. Ansell. "Fatigue properties of jointed wood composites
Part I Statistical analysis, fatigue master curves and constant life diagrams." Journal
19- Kline, Donald Edgar, Frank E. Woeste, and B. A. Bendtsen. "Stochastic model for
modulus of elasticity of lumber." Wood and fiber science 18, no. 2 (2007): 228-
238.
20- Gibson, Lorna J., and Michael F. Ashby. Cellular solids: structure and properties.
22- Toubia EA, Emami S, Klosterman D. Degradation mechanisms of balsa wood and
28:1099636217706895.
24- Toubia EA, Emami S, durability of sandwich composite structures due to freezing
106
25- Siriruk, Akawut, Dayakar Penumadu, and Y. Jack Weitsman. "Effect of sea
composites under in-plane static and fatigue loads, Materials & Design, Volume
27- Abraham Elmushyakhi and Elias Toubia. 2017. "Axial Fatigue Characterization of
West Lafayette, Indiana, October 23-25, 2017. Lancaster, PA, USA: DEStech
30- Buffel, Bart, Frederik Desplentere, Kris Bracke, and Ignace Verpoest. "Modelling
open cell-foams based on the Weaire–Phelan unit cell with a minimal surface
energy approach." International Journal of Solids and Structures 51, no. 19 (2014):
3461-3470.
31- https://netcomposites.com/guide-tools/guide/core-materials/pvc-foam/
107
32- Alemán JV, Chadwick AV, He J, Hess M, Horie K, Jones RG, Kratochvíl P, Meisel
29. Saenz, Elio. Fatigue and fracture of foam cores used in sandwich composites.
35- Sousa, João M., João R. Correia, Susana Cabral-Fonseca, and António C. Diogo.
36- Ishai, Ori, Clement Hiel, and Michael Luft. "Long-term hygrothermal effects on
(1995): 47-55.
38- Li, Xiaoming, and Y. Jack Weitsman. "Sea-water effects on foam-cored composite
108
39- Gomez, Jose, and Brian Casto. "Freeze-thaw durability of composite materials." In
40- Wu, Hwai-Chung, Gongkang Fu, Ronald F. Gibson, An Yan, Kraig Warnemuende,
and Vijay Anumandla. "Durability of FRP composite bridge deck materials under
freeze-thaw and low temperature conditions." Journal of bridge engineering 11, no.
4 (2006): 443-451.
44- Manujesh, B. J., Vijayalakshmi Rao, and MP Sham Aan. "Moisture absorption and
45- Scudamore, R. J., and W. J. Cantwell. "The effect of moisture and loading rate on
109
46- Cantwell, W. J., G. Broster, and P. Davies. "The influence of water immersion on
47- Neşer, Gökdeniz. "Seawater exposure effect on the fracture toughness of low-
density woods/GRP sandwich systems." Journal of wood science 56, no. 2 (2010):
154-159.
48- Kolat, Koray, Gökdeniz Neşer, and Çiçek Özes. "The effect of sea water exposure
49- Singh, Abhendra K., and Barry D. Davidson. "Effects of temperature, seawater and
50- Siriruk, A., D. Penumadu, and A. Sharma. "Effects of seawater and low
(2012): 25-36.
52- Shi, Jia-Wei, Hong Zhu, Gang Wu, and Zhi-Shen Wu. "Tensile behavior of FRP
53- Gomez, Jose, and Brian Casto. "Freeze-thaw durability of composite materials."
110
54- Wu, Hwai-Chung, et al. "Durability of FRP composite bridge deck materials
55- Lord, Harold W., and Piyush K. Dutta. "On the design of polymeric composite
57- Aniskevich, K., et al. "Mechanical properties of pultruded glass fiber reinforced
(2012): 0731684412462755.
58- Sousa, João M., et al. "Effects of thermal cycles on the mechanical response of
60- Wu L, Murphy K, Karbhari VM, et al. Short-term effects of sea water on E-glass/
111
62- Guzman, V. Alzamora, and Povl Brøndsted. "Effects of moisture on glass fiber-
911-920.
1320-1327.
65- Galambos TV, Ellingwood B, MacGregor JG, Cornell CA. Probability based load
66- MacGregor, James G. "Load and resistance factors for concrete design." Journal
67- Gromala DS, Sharp DJ, Pollock DG, Goodman JR. Load and resistance factor
design for wood: the new US wood design specification. InProc., Int. Timber
69- ASTM D5457 – 15, Standard Specification for Computing Reference Resistance
Factor Design
112
70- Handbook-MIL-HDBK M. 17-3F: Composite Materials Handbook, Volume 3-
73- King RL. Statistical methods for determining design allowable properties for
74- American Society of Civil Engineers. Minimum design loads for buildings and
76- C. Lind, "The design of structural design norms." Journal of Structural Mechanics
77- Ellingwood, Bruce R. "Toward load and resistance factor design for fiber-
(2003): 449-458.
78- Nowak, Andrzej S., and Kevin R. Collins. Reliability of structures. CRC Press,
2012.
79- Ditlevsen, Ove, and Henrik O. Madsen. Structural reliability methods. Vol. 178.
113
80- M.Chiachio, J.Chiachio and G.Rus, ‘Reliability in Composites - A Selective
43 (2012), 902–13
81- Zadeh, Hany Jawaheri, and Antonio Nanni. "Reliability analysis of concrete
114