2015 Effects of Hygrothermal Conditioning On Epoxy Adhesives Used in FRP Composites

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Construction and Building Materials 96 (2015) 679–689

Contents lists available at ScienceDirect

Construction and Building Materials


journal homepage: www.elsevier.com/locate/conbuildmat

Effects of hygrothermal conditioning on epoxy adhesives used


in FRP composites
B. Paige Blackburn 1, Jovan Tatar ⇑, Elliot P. Douglas 2, H. R. Hamilton 3
University of Florida, Department of Civil and Coastal Engineering, Gainesville, FL, USA

h i g h l i g h t s

 Six epoxies were tested to determine curing kinetics in hygrothermal environments.


 Competing effects of added cure and plasticization were observed in all epoxies.
 Significant decrease in Tg from theoretical values was observed in all clear epoxies.
 Average Tg of tested epoxies was below the design temperature for some U.S. bridge locations.

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

Article history: Durability of FRP composites and their bond to concrete is essential to structural integrity of an FRP
Received 11 March 2015 repair. Epoxy adhesives are used to form FRP composites, and as a bonding medium between the FRP
Received in revised form 30 July 2015 and concrete substrate. Susceptibility of epoxy to the negative effects of water and high temperature
Accepted 9 August 2015
affects the longevity of FRP repairs in hygrothermal environmental conditions. The presented work
investigates the effects of such environments on the curing kinetics of epoxy. Competing effects of
plasticization and post-cure during accelerated conditioning are discussed; recommendations targeting
Keywords:
researchers, practitioners, and manufacturers are made based on the research findings.
FRP
Epoxy
Ó 2015 Elsevier Ltd. All rights reserved.
Concrete
Bond
Durability
Hygrothermal
Cohesive
Adhesive
Accelerated aging

1. Introduction bond critical environments, the reliability of the FRP strengthen-


ing/repair scheme is compromised. Effects of environmental expo-
FRP composites are deemed a prominent method for strength- sure, primarily moisture, on the integrity of epoxy–concrete bond
ening and repair of aged structures. However, their application in is shown graphically in Fig. 1: in dry ambient conditions external
harsh environments is still inhibited by the poor resistance of load is completely transferred into the concrete substrate, allowing
FRP–concrete bond to such conditions. Due to the changing for distribution of damage in the substrate (Fig. 1a) at critical load-
mechanical properties of adhesives (loss of stiffness) and their ing; however, when stiffness of epoxy and chemical bonding are
adhesion properties (loss of chemical bonds), when exposed to adversely affected by the presence of moisture, full bond capacity
cannot be attained, as the bond fails prematurely along the
epoxy–concrete interface (Fig. 1b). Development of accelerated
⇑ Corresponding author at: University of Florida, Department of Civil and Coastal
conditioning protocols (ACP) to assess the durability of FRP–con-
Engineering, 365 Weil Hall, Gainesville, FL 32611, USA.
crete bonded systems is necessary in establishing more reliable
E-mail addresses: pblackburn182@gmail.com (B.P. Blackburn), jtatar@ufl.edu
(J. Tatar), edoug@mse.ufl.edu (E.P. Douglas), hrh@ce.ufl.edu (H.R. Hamilton). design factors. Due to multiple degradation mechanisms existing
1
U.S. Air Force, 365 Weil Hall, Gainesville, FL 32611, USA. in polymers when exposed to different environments, the selection
2
University of Florida, Department of Materials Science and Engineering, 156 of appropriate ACP is crucial in determining the durability of the
Rhines Hall, Box 116400, Gainesville, FL 32611, USA. bonded system.
3
University of Florida, Department of Civil and Coastal Engineering, 365 Weil Hall,
Gainesville, FL 32611, USA.

http://dx.doi.org/10.1016/j.conbuildmat.2015.08.056
0950-0618/Ó 2015 Elsevier Ltd. All rights reserved.
680 B.P. Blackburn et al. / Construction and Building Materials 96 (2015) 679–689

Decreased External Load Loss of


epoxy stiffness chemical bonds
External Load Mechanical key
Proper bond:
damage distributed in formed at the
concrete substrate concrete surface

Epoxy Epoxy

Concrete Chemical Bonds Concrete

(a) (b)
Fig. 1. Failure mode of epoxy–concrete bond: (a) in dry ambient conditions; (b) following exposure to moisture.

2. Motivation cross-linking between the polymer chains to form additional cova-


lent bonds. The described process is termed ‘‘curing of epoxy”;
Current industry testing standards for FRP composites used in with subsequent cross-linking a non-crystalline hardened molecu-
civil infrastructure are prescribed by multiple internationally rec- lar structure is formed [22]. At ‘‘full-cure” no additional cross-
ognized technical organizations: AASHTO FRPS-1 [8], ACI 440.2M linking between the polymer chains is possible; at this point epoxy
[10], and ICC AC125 [26]. To determine the durability properties has reached its full mechanical properties. The cross-linking den-
of FRP, the composite is to be tested in direct tension following a sity is usually described in terms of conversion that takes values
prescribed ACP. ACP incorporate a range of exposure environ- from 0 to 1.0, where a value of 1.0 signifies a fully-cured epoxy
ments, among which some of them dictate a temperature of 60 ° [16].
C, which can be slightly above the glass transition temperature of The bond between epoxy and concrete is established through
cold-cured epoxies used to form the composite. chemical interactions and mechanical bonding [30]. While the
In addition to testing the durability of FRP composites, multiple exact nature of the chemical bond is not known, it is thought to
researchers ([1–3,21,33]) have concentrated their efforts on consist of mostly hydrogen bonding between the surface molecules
three-point bending beam test [24] to determine the durability of concrete and epoxy [19]. Integrity of epoxy–concrete bond is
properties of FRP–concrete bond. Three-point bending test method considered to be primarily due to mechanical interlocking
is currently a candidate for standardization, under the auspices of ([19,27]; Stewart, 2011; [30]). Stiffness of epoxy matrix, and conse-
ASTM and ACI 440 subcommittee K – FRP Material Characteristics. quently the integrity of its mechanical bond, can be compromised
The work of ACI 440 subcommittee L – Durability was focused by two main degradation mechanisms: its transition to rubbery
primarily on specification of appropriate ACP to determine the state at temperatures higher than the T g , and plasticization of
durability of bond in three-point bending test. The main obstacle epoxy matrix.
in this process was the lack of information in literature on deter- Once the T g is exceeded, the covalent bonds between the poly-
mining the appropriate conditioning temperature for this purpose. mer chains are capable of rotating, but remain intact. Therefore,
While higher temperature is deemed to accelerate the degradation the general shape of the epoxy structure in glassy state is main-
processes that take place at FRP–concrete bond, the main concern tained, but its stiffness and strength are reduced. The value of T g
is that if the conditioning temperature is higher than the T g , the is dependent on polymer chain mobility. The lower the cross-
mechanisms leading to loss of bond may be different than the linking density of the epoxy molecular structure, the less thermal
mechanisms that occur during field exposure at lower tempera- energy is required for transition from a glassy state to a rubbery
tures. The purpose of this study is to provide a better understand- state [22,23]. Amounts of free volume and cross-linking (measured
ing of evolution of degradation mechanisms in epoxy with time by conversion) affect polymer chain mobility and consequently
when exposed to hygrothermal ACP at temperatures lower than change the epoxy T g . At room temperature conditions for two or
and higher than the measured T g of epoxy adhesive. more weeks, the conversion of most commercially available struc-
tural epoxies ranges from 0.8 to 1.0, which means that 80–100% of
possible covalent bonds are formed [34]. Exposure to temperatures
3. Background above those experienced during initial curing will cause additional
cross-linking [16]. Cross-linking restrains polymer chain mobility
Structural epoxies used with FRP composites are available in by interconnecting individual chains together; therefore, the
two forms: clear epoxy, and more viscous paste epoxy. The differ- higher the cross-linking density (more covalent bonds), the less
ence in viscosity between the two types of adhesives comes from chain mobility, and the higher T g .
additives (such as silica particles) that are not found in clear epox- Chain mobility is also affected by the amount of free volume in
ies. Epoxy adhesives used in construction are generally based on the polymer structure. Free volume is the available space within
bisphenol A molecules hosting an epoxide functional group at both the polymer chain network on a microscopic level [22,23]. With
ends, forming the monomer in the epoxy structure [29]. Epoxy increase in free volume, the mobility of polymer chains increases
hardener is composed of amines that react with the epoxide groups as there is more available space for their movement; hence, less
to form covalent bonds. The amines bond the monomers together thermal energy is required to convert the epoxy into a rubbery
in a linear fashion to form polymer chains. They also allow for state. Water absorption can lead to increases in free volume in
B.P. Blackburn et al. / Construction and Building Materials 96 (2015) 679–689 681

the epoxy. Epoxy curing results in the formation of hydroxyl An informal survey, conducted by ACI Committee 440, found that DMA and DSC
are the two most commonly utilized methods for T g measurements in the industry
groups in the network. Upon exposure to water, the water mole-
[15]. DSC was selected to determine the T g values of epoxy in this study. DSC does
cules are ‘‘drawn” into the epoxy and interact with the hydroxyl not depend upon mechanical manipulation to detect thermal transitions; it is the
groups via hydrogen bonding [23,16]. This hydrogen bonding cre- only method that measures T g using a nonmechanical thermal process. The DSC
ates more space between the cross-linked polymer chains, thereby purely thermal calculation of T g relates well to the epoxy property state over the
enlarging the free volume available. Water absorption into epoxy is course of exposure, without incorporating effects of mechanical loading. Another
advantage of calculating T g with the DSC is the small sample size. To test samples
referred to as plasticization [25], and will result in a decrease of T g
in the DSC, only 5–10 mg epoxy specimens can be used. Such small DSC samples are
[23,29,16]. Complicating this issue further, exposing epoxy to much more representative of epoxy that forms mechanical interlocks on the con-
water at temperatures above the T g can cause an increase in the crete surface, compared to relatively large bar samples that are used in other test
rate of water absorption due to the greater chain mobility at those methods, such as the DMA and TMA. Moreover, in an interlaboratory study on con-
temperatures. struction epoxies, Bakis et al. [15] noted remarkably high repeatability of T g values
obtained by DSC method.
According to Zhou and Lucas [31], two types of water bonding
EXSTAR6000 DSC station from Seiko Instruments Inc. equipped with a 220CU
may exist in an epoxy matrix. One type of bonding is responsible module and automatic gas cooling unit was used. Epoxy specimens were prepared
for plasticization by breaking the inter-chain Van der Wals forces, by casting epoxy directly into the sample pans to maximize the heat transfer
while the other water bonding type is deemed to allow for between the pan and epoxy, as described by Choi and Douglas [17]. T g measure-
secondary cross-linking (pseudo cross-linking) by forming ments were obtained from the first run at a heating rate of 10 °C/min. All T g values
in this study are reported as midpoint temperature (T m ) on DSC thermal curve, as
hydrogen bonds between two polymer chains. The former is defined in ASTM E1356 [12].
associated with lower desorption activation energy (approxi- An additional set of samples were simultaneously tested in Fourier Transform
mately 10 kcal/mol), and water sorption process that takes place Infrared Spectrometry (FTIR) to determine the conversion and absorbed water con-
in low-temperature hygrothermal conditions; the other type of tent in epoxy samples. The epoxy thin film samples, ranging from 0.23 to 0.69 mm
in thickness, were cast between glass plates; Teflon sheets were used to prevent
water bonding in such conditions is considerably less. On the
adhesion between epoxy and glass. FTIR Nicolet Magna 760 from Thermo
other hand, presence of the latter type of water bonding is Electron Cooperation was utilized to record near-infrared spectra over the range
associated with hygrothermal conditioning at higher tempera- of 3800–6800 cm1 in an inert atmosphere. Scanning in near-infrared range was
tures and longer exposure times, and approximate desorption conducted because a single absorption band for water exists in this range of wave-
activation energy of 15 kcal/mol. lengths; the absorption band for water in the mid-infrared range is at approxi-
mately 3400 cm1; however, this band is broad and contains contributions from
Multiple researchers examined the effects of hygrothermal con- other hydrogen-bonded species present in the epoxy, making the interpretation
ditioning on structural epoxies used in construction. Moussa et al. of data challenging. Absorption data for each specimen were collected with 32
[28] showed that strength and T g of epoxy strongly depends on scans at a 4 cm1 resolution implementing a white light source, a CaF beam splitter,
curing temperature as well as the duration of curing. They also and an MCT light detector.
found that decreasing the testing temperature increased epoxy
stiffness. Changes in T g , conversion and water absorption of two 5. Test matrix
construction epoxies, during the 0–28-day hygrothermal condi-
tioning, were reported by Choi [16]. It was found that effects of DSC and FTIR tests were conducted on five different epoxies
plasticization and additional cross-linking during exposure were provided by three independent manufacturers, and one model
in competition with each other causing changes in T g throughout epoxy system of known chemical composition (epoxy D).
the exposure. In all test groups, by 28 days of conditioning, Properties of tested epoxies, as reported by the corresponding
increases in T g ranged from 1 to 14 °C. Stewart [29] found that both manufacturers, are summarized in Table 1; properties of epoxy D
modulus and strength of ambient-cured epoxies decreased due to were determined by Stewart [29]. Main known constituents of
hygrothermal conditioning, mostly due to plasticization. Similar each epoxy are listed in Table 2. Epoxies were categorized into
findings were reported by Quan et al. [4], Benzarti et al. [6], and two groups based on their viscosity and transparency, clear and
Lattieri and Frigione [5]. paste. Clear epoxy was nearly translucent with viscosity similar
to honey. Paste epoxies are opaque with a significantly higher vis-
4. Experimental methods cosity. Epoxy C had a viscosity similar to that of clear epoxies, but
was opaque and was categorized as clear-opaque.
Several techniques are available to determine the T g of polymers, among them Exposure conditions in this study were chosen to match the
the most commonly employed ones are: dynamic mechanical analysis (DMA),
work of Stewart [29] on effects of hygrothermal conditioning on
thermo-mechanical analysis (TMA), differential scanning calorimetry (DSC), and
heat deflection method. There is no particular standard method preferred among mechanical properties of epoxy, and the study by Tatar and
researchers; each method is recognized and covered by a respective ASTM standard. Hamilton [3] on examination of effects of hygrothermal condition-
Glass transition occurs over a temperature region and each method registers a dif- ing on FRP–concrete bond. The DSC test matrix presented in
ferent temperature within the region to indicate the transition. No conversion of T g
Table 3 summarizes the selected conditions. Immersion samples
values between the methods is officially exercised; manufacturers, engineering
design codes, and researchers employ different test methods for T g , therefore T g test
were immersed in covered water tanks at the specified tempera-
data must be accompanied by the testing technique and specification of curing ture. Humidity samples were placed in the air space above the
history. water in the water tanks; humidity was monitored with a

Table 1
Epoxy properties as reported by manufacturers.

Epoxy Category Cure conditions Tg (°C) (ASTM) Tensile strength (MPa) Tensile modulus (MPa)
ASTM D638 [32] ASTM D638 [32]
A Clear 3 days, 60 °C 82 (D4065 [13]) 72.4 3180
B Clear 5–14 days, 23 °C, 50% R.H. 46 (N.S.) 55 1724
C Clear-opaque Time N.S., 20 °C, 40% R.H. 71 (N.S.) 55.2 3034
D Clear 56 days, ambient 54 (ASTM E1356 [12]) 74 2440
E Paste 7 days, 23 °C, 50% R.H. 47* (ASTM D648 [11]) 24.8 4482
F Paste Time N.S., 20 °C, 40% R.H. 50 (N.S.) 27.6 N.S.

N.S. = Information not specified by manufacturer.


R.H. = Relative Humidity.
*
Heat deflection temperature.
682 B.P. Blackburn et al. / Construction and Building Materials 96 (2015) 679–689

Table 2
Main constituents of epoxy resin and hardener.

Epoxy Resin Hardener


A DGEBA Polyoxyropylenediamine; Polyetheramine
B DGEBA Polyoxypropylenediamine
C DGEBA, Oxirane, mono[(C12-14-alkyloxy)methyl] derivs., Oxirane, alpha-(2-Aminomethylethyl)-omega-(2-aminomethylethoxy)-poly(oxy(methyl-1,2-
2,20 -[(2,2-dimethyl-1,3-propanediyl)bis(oxymethylene)]bis-; ethanediyl)); 3-aminomethyl-3,5,5-trimethylcyclohexylamine; 1,6-Hexanediamine,
ethylbenzene C,C,C-trimethyl-; Oxirane, 2,20 -[(1-methylethylidene)bis(4,1-
phenyleneoxymethylene)]bis-, homopolymer; Benzyl alcohol; imidazole
D DGEBA 5-phenoxy-2,3-pentadienylamine
E Mineral additives; DGEBA; 1,4-bis(2,3 epoxypropoxy)butane) Quartz; Benzyl alcohol; 3,6-diazaoctanethylenediamin;
dimethylaminomethylphenol; methylamine
F DGEBA, crystalline silica; ethylene glycol 2-piperazin-1-ylethylamine; Phenol, 4-nonyl-, branched; crystalline silica; ethylene
glycol

Table 3
tested in FTIR because it was not feasible to prepare thin film
DSC test matrix. epoxy specimens that were transparent to light.

Exposure Moisture Temp. Exposure Number of


(°C) time (weeks) specimens 6. Results and discussion
Control Atmospheric, 23–25.5 12 6
20–28% RH 6.1. Conversion and water absorption
Immersion Water 30 2 6
4 6 To obtain conversion and water content data, areas under char-
8 6 acteristic peaks of FTIR spectra were analyzed. Natural frequencies
12 6
of interest in this study included characteristic water peak at
60 2 6
4 6 5230 cm1, phenyl peak at 4622 cm1, and the epoxide functional
8 6 group peak at 4530 cm1 (Fig. 2). From the collected FTIR spectra
12 6 conversion can be calculated as:
Humidity 100% R.H. 30 2 6
AðtÞ
4 6 a¼1 ð1Þ
8 6 Að0Þ
60 2 6
4 6 where A(0) is epoxide area at zero cure (conversion of 0.0) sample,
8 6 and A(t) is the epoxide area of a partially cured sample, at time of
curing t. To account for differences in epoxy thin film thickness
(thicker samples absorb more light), area of epoxide peaks was nor-
malized to the phenyl peak (which is consistent throughout sample
Table 4
FTIR test matrix. life, regardless of conversion). This allowed for direct comparison of
results between different samples. An unknown peak was observed
Exposure Moisture Temp. Exposure Number of
at approximately 4560 cm1 (Fig. 2) that overlaps the epoxide peak.
(°C) time (weeks) specimens
Area of this peak, measured on a fully-cured sample, was subtracted
Control Atmospheric, 23–25.5 12 4
from the total area of epoxide peaks to obtain absorbance corre-
20–28% RH
sponding only to the epoxide groups. The base lines under each
Immersion Water 30 2 4
8 4
60 2 4
0.5
Humidity 100% RH 30 2 4
8 4
60 2 4 Phenyl 4622 cm-1
8 4 0.4

Water 5230 cm-1

humidity meter, and was determined to be consistently in 95–


Absorbance

0.3
100% R.H. range. Note that each epoxy had one sample tested for
each exposure condition, yielding a total of 6 test specimens per
condition. Exposure times were selected to correspond to typical
conditioning times used in conjunction with the three-point bend- 0.2
ing tests ([3,21]). Conditioning temperatures were chosen to be
lower than (30 °C) and higher than (60 °C) the typical T g of control Unknown 4560 cm-1
group of tested epoxies. 0.1
FTIR testing (Table 4) utilized the same exposure conditions
Epoxide 4530 cm -1
that were applied to DSC samples, but shorter (2- and 8-week)
exposure times were needed to gain understanding of specimen
0
behavior. Only 2-week samples were tested for conditioning tem- 5400 5200 5000 4800 4600 4400
perature of 60 °C. This decision was based on prior research that Wavenumber (cm-1)
reported plateaus in conversion and water content under this con-
dition after only 4 days of exposure [16]. Paste epoxies could not be Fig. 2. Typical FTIR spectra for epoxy.
B.P. Blackburn et al. / Construction and Building Materials 96 (2015) 679–689 683

peak for area calculation were determined as per Dannenberg [20]. exposure temperature was above the T g , the increased mobility
Reliability of the quantitative method for calculation of epoxy cure of the network allowed additional cross-linking opportunities to
for construction epoxies was confirmed by Choi and Douglas [17]. become available between polymer chains that were not accessible
Relative amount of water within each sample was calculated in in the glassy state. By the end of 8-week exposure period, all epoxy
a similar fashion as conversion. Area under the characteristic water systems were close to their fully cured state.
peak at 5230 cm1, normalized by area of phenyl peak, provided an Epoxies A, B, and D had similar conversion values prior to expo-
indirect measure of the relative amount of water in the sample. sure ranging from 0.88 to 0.91. Epoxy C, however, had a larger con-
Choi and Douglas [17] verified the reliability of this method for trol conversion of 0.96, possibly because its curing activation
tracing the relative changes in water content of conditioned epoxy energy is lower than that of other tested epoxy systems.
samples. Consequently, epoxy C reached a fully cured state, for all ACs, fol-
FTIR analysis of epoxy A, B, C, and D conversions are summa- lowing only 2 weeks of conditioning.
rized in Fig. 3. Data from a previous study was included too, where Based on the presented data it appears that the curing kinetics
applicable [16]. The plots indicate increases in the degree of poly- of epoxy systems are independent of the moisture conditioning
merization throughout the 8 week exposure period despite varia- method (immersion vs. humidity). Slight differences that were
tions in moisture and temperature conditioning. These results observed in 30 °C samples, between immersion and humidity sam-
were expected since the epoxy samples were initially cured under ples, indicate that at lower temperatures the method of moisture
standard laboratory conditions (20–24 °C), which was 6–40 °C conditioning may play a role in the rate of polymerization of the
cooler than temperatures experienced during accelerated condi- epoxy matrix; however, assuming that conversion measurement
tioning (AC). A very small negative slope, between 2 and 8 weeks, errors are on the order of ±1%, the recorded values could be pre-
in plots corresponding to epoxies A and D for 60 °C samples is scribed to the measurement errors.
likely due to measurement error. Reduction in cross-linking would The additional data by Choi [16] revealed that all 60 °C samples
only be possible by burning the epoxy structure [22], which does reached a conversion value close to 1.0 after only 2 days of expo-
not occur during AC. Test results also revealed that 60 °C samples sure. This trend agrees well with the findings from this study,
consistently achieved higher (epoxies A, B, and D) conversion than regardless of whether samples were immersed or exposed to
30 °C samples at 2 weeks of exposure. Such behavior demonstrates humidity. Inconsistency between the initial (control) conversion
the important influence of temperature on the curing kinetics of values from the two studies was noted in epoxy B and epoxy C;
epoxy adhesives. AC temperature, above that experienced during the source of such discrepancy was determined to be a slight dif-
the initial cure, provided additional energy for cross-linking within ference in the curing temperatures. Samples from previous
the epoxy structure, thereby increasing conversion. In addition, if research were stated to have been cured under ambient conditions,

1 1
0.98 0.98
0.96 0.96
0.94 0.94
Conversion

Conversion

0.92 0.92
0.9 0.9
Immersion 30 C
0.88 0.88 Humidity 30 C
0.86 Immersion 30 C 0.86 Immersion 60 C
0.84 Humidity 30 C 0.84 Humidity 60 C
Immersion 60 C Immersion 30 C (Choi)
0.82 Humidity 60 C 0.82 Immersion 60 C (Choi)
B
0.8 0.8
0 2 4 6 8 0 2 4 6 8
Exposure Time (weeks) Exposure Time (weeks)

(a) (b)

1 1
0.98 0.98
0.96 0.96
0.94 0.94
Conversion

Conversion

0.92 0.92
0.9 0.9
Immersion 30 C Immersion 30 C
0.88 Humidity 30 C 0.88 Humidity 30 C
0.86 Immersion 60 C 0.86 Immersion 60 C
0.84 Humidity 60 C 0.84 Humidity 60 C
Immersion 30 C (Choi) Immersion 30 C (Choi)
0.82 Immersion 60 C (Choi) 0.82 Immersion 60 C (Choi)
C D
0.8 0.8
0 2 4 6 8 0 2 4 6 8
Exposure Time (weeks) Exposure Time (weeks)

(c) (d)
Fig. 3. Conversion results from FTIR data for: (a) epoxy A; (b) epoxy B (includes [16]); (c) epoxy C (includes [16]); (d) epoxy D (includes [16]).
684 B.P. Blackburn et al. / Construction and Building Materials 96 (2015) 679–689

suggesting a constant 20 °C room temperature [16], while in the increasing water diffusion into the epoxy network which explains
present research, samples were cured at standard laboratory tem- the quicker saturation of 60 °C samples vs. 30 °C samples.
perature, which proved to range from 20 °C to 25 °C during the Water content of each sample was measured to investigate the
cure period, resulting in higher control conversions. possible correlations between the decrease in T g and plasticization.
Water absorption of specimens was traced by measuring char- Based on the characteristic water content measurements, it can be
acteristic water content from FTIR data. Characteristic water con- assumed that moisture equilibrium was maintained throughout
tents of epoxies A, B, C, and D are summarized in Fig. 4. Data by the 2–8-week exposure period.
Choi [16] is included, where applicable.
All samples exhibited an increase in characteristic water con- 6.2. DSC results
tent during exposure, as Fig. 4 denotes. It appears that, within
the first 2 weeks of exposure, all specimens reached a fully satu- T g of 6 different epoxies was measured for varying ACPs at spec-
rated state despite the epoxy used. Data in Fig. 4 illustrates no sig- ified time increments by DSC. Generally clear epoxies exhibited
nificant difference in water absorption between immersed samples similar behavior over time, while clear-opaque and paste epoxies
and those exposed to humidity. Within 2 weeks of exposure, all had varying behavior. Exemplary DSC curves, corresponding to
samples had reached equilibrium moisture content. Conditioned the model epoxy system (epoxy D), for varying exposure times
samples, on average, had characteristic water content approxi- and conditions are shown in Fig. 5. DSC curves exhibited distinct
mately 2.68 times larger than that of their equivalent control sam- glass transition regions in the majority of tested samples; some
ples. Some samples presented a slight decrease in water content of the samples that were exposed to immersion at 60 °C, however,
between 2 week and 8 week samples. This behavior was unex- had a wider glass transition region. It is recognized that in such
pected and it was accredited to measurement errors, drying of samples the reported midpoint temperature could vary approxi-
the sample prior to the test, or both. mately ±2 °C due to the subjectivity of the analysis process.
Water content data from prior research was included where
applicable in Fig. 4, to illustrate water absorption behavior at rela- 6.2.1. Clear epoxies
tively small time increments between 0 and 2 weeks of condition- Clear epoxies (A, B, and D) all shared similar behavior over the
ing [16]. It is clear that during the first week of exposure, water 12 week exposure (Fig. 6). Conditioning at 30 °C caused insignifi-
absorption occurred at a higher rate in samples conditioned at cant variations in T g from the control value in all epoxies, which
60 °C when compared to those conditioned at 30 °C. Full saturation could be well within the error of the measurement. By the end of
was achieved within 1–4 days of exposure in 60 °C conditions, and the 12 week exposure, the final T g was greater than the original
not until about 2 weeks in 30 °C conditions. Higher conditioning control T g (prior to exposure) for most 60 °C exposure conditions.
temperatures provide more energy to water molecules, thereby Conditioning temperature proved to have a strong influence on

7.5 7.5
Characteristic Water Content

Characteristic Water Content

6.5 6.5

5.5 5.5

4.5 4.5
Immersion 30 C
Humidity 30 C
3.5 Immersion 30 C 3.5 Immersion 60 C
Humidity 30 C Humidity 60 C
2.5 Immersion 60 C 2.5 Immersion 30 C (Choi)
A
Humidity 60 C B
Immersion 60 C (Choi)
1.5 1.5
0 2 4 6 8 0 2 4 6 8
Exposure Time (weeks) Exposure Time (weeks)

(a) (b)
7.5 7.5
Characteristic Water Content

Characteristic Water Content

6.5 6.5

5.5 5.5

4.5 4.5
Immersion 30 C Immersion 30 C
Humidity 30 C Humidity 30 C
3.5 Immersion 60 C 3.5 Immersion 60 C
Humidity 60 C Humidity 60 C
2.5 Immersion 30 C (Choi) 2.5 Immersion 30 C (Choi)
C
Immersion 60 C (Choi) D
Immersion 60 C (Choi)
1.5 1.5
0 2 4 6 8 0 2 4 6 8
Exposure Time (weeks) Exposure Time (weeks)

(c) (d)
Fig. 4. Water absorption results from FTIR data for: (a) epoxy A; (b) epoxy B (includes [16]); (c) epoxy C (includes [16]); (d) epoxy D (includes [16]).
B.P. Blackburn et al. / Construction and Building Materials 96 (2015) 679–689 685

Immersion at 30 C Immersion at 60 C
-4000 -4000
Control Control
2 weeks 2 weeks
-6000 -6000
4 weeks 8 weeks
8 weeks 12 weeks
-8000 12 weeks -8000
Heat Flow ( W)

Heat Flow ( W)
-10000 -10000

-12000 -12000

-14000 -14000

-16000 -16000

-18000 -18000

-20000 -20000
10 30 50 70 90 100 10 30 50 70 90 100
Temperature ( C) Temperature ( C)

Humidity at 30 C Humidity at 60 C
-4000 -4000
Control Control
2 weeks 2 weeks
-6000 -6000
8 weeks 8 weeks

-8000 -8000
Heat Flow ( W)

Heat Flow ( W)

-10000 -10000

-12000 -12000

-14000 -14000

-16000 -16000

-18000 -18000

-20000 -20000
10 30 50 70 90 100 10 30 50 70 90 100
Temperature ( C) Temperature ( C)

Fig. 5. DSC curves for epoxy D.

70 70 70

65 65 65

60 60 60
Tg ( C)

Tg ( C)

Tg ( C)

55 55 55

Immersion 30 C Immersion 30 C
50 50 Humidity 30 C 50 Humidity 30 C
Immersion 30 C
Immersion 60 C Immersion 60 C
Humidity 30 C Humidity 60 C
45 45 45 Humidity 60 C
Immersion 60 C Immersion 30 C (Choi) Immersion 30 C (Choi)
Humidity 60 C Immersion 60 C (Choi) Immersion 60 C (Choi)
40 40 40
0 2 4 6 8 10 12 0 2 4 6 8 10 12 0 2 4 6 8 10 12
Exposure Time (weeks) Exposure Time (weeks) Exposure Time (weeks)

(a) (b) (c)


Fig. 6. Tg values over 12 weeks of conditioning: (a) epoxy A; (b) epoxy B (includes [16]); (c) epoxy D (includes [16]).

the T g of all clear epoxies. By 12 weeks, all samples exposed to 60 ° at 60 °C allowed more pseudo cross-linking from hydrogen bond-
C temperatures had a greater T g than the corresponding 30 °C sam- ing of water molecules to the network to occur than conditioning
ple. This behavior occurred at 8 and 12 weeks despite the moisture at 30 °C. Moreover, additional cross-linking that occurred in 60 °C
conditioning method. It is possible that hygrothermal conditioning samples due to the increased mobility above the glass transition
686 B.P. Blackburn et al. / Construction and Building Materials 96 (2015) 679–689

may have contributed to the relatively higher T g , too. A discrep- early in the exposure, as shown in Fig. 7. At 12 weeks of exposure,
ancy between data reported by Choi [16] and the present study epoxy C recovered to its control T g temperature that existed prior
was noted in epoxies B and D. In epoxy B, this was likely due to to exposure (Fig. 3). One of the reasons for such behavior may be
the dissimilar curing histories between the two sets of samples the relatively high conversion in control sample, when compared
prior to the exposure that caused initial T g values by Choi [16] to to clear epoxies. Since post-cure cure was limited, plasticization
be lower, resulting in more pronounced competing effects of plas- was deemed the controlling factor on T g early in the exposure,
ticization and added cure. In epoxy D, the discrepancy in T g evolu- hence, reducing the T g of exposed samples below the control T g .
tion plots is believed to be due to the fact that epoxies came from By 8 weeks of exposure, Epoxy E samples of all ACP converged
different batches, and might have had slightly differing stoi- to the control T g value. Therefore, over time the competitive effects
chiometries depending on the mixing procedures. The range of of post-cure and plasticization were neutralized for 8 week expo-
measured T g values, however, is similar between Choi [16] and sure. Prolonged exposure to water immersion, however, caused
the present study for epoxy D. This indicates that both groups of an increase in T g , above the measured control value.
samples demonstrated a similar response to the effects of added Epoxy F exhibited the largest change in T g of all the epoxy sys-
cure and plasticization. tems, with a drop of 19 °C from control value, under water immer-
There is some evidence that samples exposed to humidity were sion conditions (Fig. 7). The presented data indicates extreme
not affected by plasticization during the first 2 weeks of condition- sensitivity of epoxy F to plasticization. Interestingly, the T g of
ing, as much as immersed samples. Looking at each epoxy, mea- epoxy F humidity samples stayed at, or slightly above the control
sured T g of every 2-week humidity sample was higher than the T g . These data suggest that epoxy F was more susceptible to plas-
respective immersed sample. This suggests that effects of post- ticization due to immersion in water than due to humidity. Higher
cure overpowered the effects of plasticization in humidity samples, T g values in humidity samples at 30 °C when compared to samples
while the opposite was true for immersed specimens. Although conditioned at 60 °C could be due to effects of mineral fillers on the
there were significant differences in T g between immersed and moisture diffusion under different conditioning temperatures.
humidity samples at 2 weeks of conditioning, they decreased as Unfortunately, due to their inability to pass light, paste epoxies
exposure continued. By 8 weeks the difference in T g between were not analyzed in FTIR; therefore, the comparative water
humidity and immersed samples was considerably reduced; this contents of epoxy F samples were not available.
indicates slight recovery of T g of immersed samples over the
course of the exposure, and effects of post-cure. Interestingly,
6.3. Master plots
water absorption plots for epoxies A, B, and D (Fig. 4) do not appear
to correlate well with changes in T g (Fig. 6). This suggests that
Since changes in T g during hygrothermal exposure are depen-
measured water content may not be as proportionally related to
dent on two different competing influences (post-cure and plasti-
effects of plasticization on the T g , primarily because the amount
cization), master plots were developed to isolate decrease in T g
of pseudo cross-linking from hydrogen bonding of water molecules
caused by the effects of water only [17]. Master plots show the
to the network cannot be discerned. A strong correlation of 0.89
dependence of T g on the level of conversion, in dry ambient condi-
was observed between the data corresponding to the two condi-
tions (Fig. 8). With these theoretical curves, the T g of exposed sam-
tioning temperatures at 2-weeks conditioning. For 8-week condi-
ples could be compared to that of unexposed samples of equal
tioning, however, no correlation between data corresponding to
conversion, to determine the decrease in T g (DT g ). By plotting T g
the two conditioning temperatures was observed. This suggests
based on sample conversion, the effects of post-cure are minimized,
that, early in the exposure, behavior of tested epoxies was gov-
leaving any differences in T g to be the result of plasticization [16].
erned by the same mechanism, while the degradation mechanism
Theoretical curves were created based on the following equation:
changed later in the exposure depending on the conditioning
temperature.  
DC p1
ð1  aÞT g0 þ DC p0
T g1
Tg ¼   ð2Þ
DC p1
6.2.2. Clear-opaque and paste epoxies ð1  aÞ þ DC p0
a
The clear-opaque epoxy C and the paste epoxies (E and F) each
exhibited unique behavior in terms of T g . Changes in epoxy C T g where a is conversion (0–1.0); T g0 and T g1 are the T g values of the
over the exposure time appeared to be dominated by plasticization monomer (0% cure) and the fully cured epoxy network,

70 70 70
Immersion 30 C Immersion 30 C
Humidity 30 C Humidity 30 C
65 Immersion 60 C 65 Immersion 60 C
60
Humidity 60 C Humidity 60 C
60 Immersion 30 C (Choi) 60
Immersion 60 C (Choi) 50
Tg ( C)

Tg ( C)

Tg ( C)

55 55
40
50 50
Immersion 30 C
Humidity 30 C 30
45 45
Immersion 60 C
Humidity 60 C
40 40 20
0 2 4 6 8 10 12 0 2 4 6 8 10 12 0 2 4 6 8 10 12
Exposure Time (weeks) Exposure Time (weeks) Exposure Time (weeks)

(a) (b) (c)


Fig. 7. Tg values over 12 weeks of conditioning: (a) epoxy C; (b) epoxy E (includes [16]); (c) epoxy F (includes [16]).
B.P. Blackburn et al. / Construction and Building Materials 96 (2015) 679–689 687

Maximum Service Temperature (ºC)


80

70 Min. Allowed Service Temperature


Max. Allowed Service Temperature
60

50

40

30

20

10

0
A B C D E F A B C D E F
ACI 440.2R AASHTO FRPS-1

Fig. 10. Maximum and minimum service temperatures calculated from experi-
mental data; shaded region represents typical maximum design temperature range.

immersion samples would likely have higher water absorption than


Fig. 8. Master plot construction. the humidity samples, that is not the case (Fig. 4); FTIR water
absorption measurements may not be directly proportional to plas-
ticization, as the amount of pseudo cross-linking caused by hydro-
27
gen bonding of water molecules is not clear.
2-week exposure
24 8-week exposure
Reduction in Tg ( Tg)

21 6.4. Maximum service temperature


18
ACI 440.2R [9] does not recommend application of composites
15
at service temperatures that are equal to or higher than Tg – 15 °
12 C. AASHTO FRPS-1 imposes a more conservative limit on maximum
9 design temperature of Tg – 22 °C. Based on the previously pre-
sented T g values of representative epoxy adhesives used in civil
6
infrastructure, the allowed maximum service temperature ranges
3
are presented in Fig. 10, based on ACI 440.2R and AASHTO FRPS-
0 1 requirements. It is apparent that maximum service temperatures
ABCD ABCD ABCD ABCD determined from measured T g values for epoxies used in this study
Immersion Immersion Humidity Humidity are within, or lower than the typical maximum design range of
30ºC 60ºC 30ºC 60ºC
temperatures for concrete bridges in the U.S. (40–50 °C) [7].
Fig. 9. Reduction in Tg for 2- and 8-week conditioning.
7. Summary and conclusions

respectively; DC p0 and DC p1 are the heat capacity changes at T g0 A study was conducted to enhance the understanding of how
and T g1 , respectively. This semi-empirical equation was developed accelerated conditioning in hygrothermal environments affects
by Couchman and Karasz [18] to establish the relationship between the properties of structural epoxies; throughout the exposure to
T g and cross-linking (conversion) in polymers. water immersion and humidity at temperatures below (30 °C)
To generate the theoretical curves, T g and conversion values and above (60 °C) the Tg of control group, evolution of Tg for six dif-
corresponding to the same curing conditions were experimentally ferent epoxy formulations were measured by DSC. Additionally,
obtained for unexposed (control) conditions. The parameter conversion and characteristic water contents were measured by
DC p1 =DC p0 was determined from the best line fit for experimental means of FTIR on four epoxies that were transparent to light.
T g values corresponding to increasing conversion. In this study, Results from this research program yielded the following
only three points were used to fit the theoretical curve; three conclusions:
points were deemed sufficient given the fact that the area of inter-
est was a very narrow region of the theoretical curve (correspond- 1. Clear epoxy samples experienced increase in conversion
ing to 0.9–1.0 conversion). The approach taken in this study, was throughout accelerated conditioning. This increase was depen-
compared to a more rigorous curve fitting approach (based on 19 dent on accelerated conditioning temperature; moisture condi-
experimental data points) performed by Choi and Douglas [17] tioning (immersion or humidity) had insignificant effect on
for the same epoxy system; both approaches yielded similar conversion.
results for the narrow region that was of interest in this study. 2. Moisture equilibrium was attained in all epoxy samples after
Fig. 9 shows the decrease in T g of samples after 2 and 8 weeks of 2 weeks for both immersion and humidity, regardless of expo-
conditioning, from their respective theoretical curves, thus indicat- sure temperature.
ing the effects of plasticization. There appears to be little difference 3. An increase in T g by the end of accelerated conditioning, when
between the reductions in T g of specimens conditioned in water compared to control (unconditioned) samples, was noted in
immersion and humidity at 30 °C. Conversely, while samples condi- clear epoxies. Samples conditioned at 60 °C experienced a sig-
tioned in water immersion and humidity at 60 °C exhibited similar nificant increase in T g above the control value, while the
conversion levels, the immersion samples experienced significantly increase in 30 °C samples was insignificant. The increase in T g
higher decrease in T g . Even though this evidence suggests that 60 °C is explained by the effects of post-cure, primarily.
688 B.P. Blackburn et al. / Construction and Building Materials 96 (2015) 679–689

4. No direct correlation between decrease in T g and characteristic Beams, The Composites and Advanced Materials Expo (CAMX 2014), Orlando,
FL, 2014.
water content measured from FTIR was observed. Even though
[3] J. Tatar, H.R. Hamilton, Bond durability factor for externally bonded CFRP
characteristic water content stayed constant throughout the systems in concrete structures, ASCE J. Compos. Constr. (2015), http://dx.doi.
accelerated conditioning, T g was fluctuating. This is explained org/10.1061/(ASCE)CC.1943-5614.0000587.
[4] Y. Quan, G. Xian, V.M. Karbhari, Hygrothermal ageing of and epoxy adhesive
by varying amounts of pseudo cross-linking induced by hydro-
used in FRP strengthening of concrete, J. Appl. Polym. Sci. 107 (2008) 2607–
gen bonding of water molecules. 2617.
5. Clear-opaque epoxy experienced a significant drop in T g from [5] M. Lattieri, M. Frigione, Effects of humid environment on thermal and
mechanical properties of a cold-curing structural epoxy adhesive, Constr.
the control value early in the exposure; however, it recovered
Build. Mater. 30 (2012) 753–760.
its initial T g value by the end of conditioning. [6] K. Benzarti, S. Chataigner, M. Quiertant, C. Marty, C. Aubagnac, Accelerated
6. All clear epoxies experienced a significant decrease in T g from ageing behavior of the adhesive bond between concrete specimens and CFRP
overlays, Constr. Build. Mater. 25 (2) (2011) 525–538.
their theoretical values due to exposure to hygrothermal condi-
[7] AASHTO, LRFD Bridge Design Specifications, American Association of State
tions, ranging from approximately 9–25 °C. The most severe Highway and Transportation Officials, Washington, D.C., 2015.
decrease in T g was observed in specimens conditioned by [8] AASHTO, Guide Specifications for Design of Bonded FRP Systems for Repair and
Strengthening of Concrete Bridge Elements, AASHTO FRPS-1, American
immersion at 60 °C.
Association of State Highway and Transportation Officials, Washington, D.C.,
7. One of the tested paste epoxies demonstrated high susceptibil- 2010.
ity to degradation in properties when exposed to water immer- [9] ACI Committee 440, Guide for the Design and Construction of Externally
sion. The T g value of exposed samples was measured at Bonded FRP Systems for Strengthening Concrete Structures, ACI 440.2R-08,
American Concrete Institute, Farmington Hills, MI, 2008.
approximately 25 °C, corresponding to a drop from control T g [10] ACI Committee 440, Specification for Carbon and Glass Fiber-Reinforced
of about 19 °C. Polymer (FRP) Materials Made by Wet Layup for External Strengthening of
8. None of the tested epoxies are suitable for application in Concrete and Masonry Structures, ACI 440.8M-13, American Concrete
Institute, Farmington Hills, MI, 2013.
bridges (maximum design temperature range 40–50 °C), as [11] ASTM Standard D648, Standard Test Method for Deflection Temperature of
per ACI 440.2R and AASHTO FRPS-1. Plastics under Flexural Load in the Edgewise Position, ASTM International,
West Conshohocken, Pa, 2007.
[12] ASTM Standard E1356, Standard Test Method for Assignment of the Glass
8. Recommendations Transition Temperatures by Differential Scanning Calorimetry, ASTM
International, West Conshohocken, Pa, 2008.
[13] ASTM Standard D4065, Standard Practice for Plastics: Dynamic Mechanical
8.1. Epoxy–concrete bond tests Properties: Determination and Report of Procedures, ASTM International, West
Conshohocken, Pa, 2012.
Based on the findings from this research, accelerated condition- [14] C. Au, O. Buyukozturk, Peel and shear fracture characterization of debonding
in FRP plated concrete affected by moisture, J. Compos. Constr. 10 (1) (2006)
ing of FRP-concrete bond specimens shall be performed by water 35–47.
immersion for 8 weeks at the maximum design temperature [15] C.E. Bakis, L.A. Bisby, M.M. Lopez, S.E. Witt, T. Alkhraji, Interlaboratory
expected in the field. Eight weeks of conditioning is deemed evaluation of Tg of ambient-cured epoxies used in civil infrastructure, in:
Proceedings of The 7th International Conference on FRP Composites in Civil
sufficient to allow the stabilization of competing influences of
Engineering (CICE 2014), International Institute for FRP in Construction (IIFC),
post-cure and plasticization. Moreover, 8 weeks was found to be Vancouver, British Colombia, Canada, 2014.
sufficient to allow the full saturation of FRP–concrete bondline [16] S. Choi, Study of Hygrothermal Effects and Cure Kinetics on the Structure–
by previous researchers [14]. Property Relations of Epoxy-Amine Thermosets: Fundamental Analysis and
Application, Master’s Thesis. University of Florida, Gainesville, FL.
Accelerated conditioning of epoxy caused changes in T g , from [17] S. Choi, E.P. Douglas, Complex hygrothermal effects on the glass transition of
the initial value, measured prior to exposure. Even though acceler- an epoxy-amine thermoset, Appl. Mater. Interfaces 2 (3) (2010) 934–941.
ated conditioning may be conducted at temperatures well below [18] P.R. Couchman, F.E. Karasz, A classical thermodynamic discussion of the effect
of composition on glass transistion temperatures, Macromolecules 11 (1)
the T g reported by the manufacturer, it should be confirmed that (1978) 117–119.
T g of exposed samples did not drop below the accelerated condi- [19] F. Djouani, C. Connan, M. Delamar, M.M. Chehimi, K. Benzarti, Cement paste–
tioning temperature during the conditioning. It was found, in this epoxy adhesive interactions, Constr. Build. Mater. 25 (2) (2011) 411–423.
[20] H. Dannenberg, Determination of functional groups in epoxy resins by near-
study, that decrease in T g due to moisture conditioning can be infrared spectroscopy, SPE Trans. 3 (1963) 78–88.
quite dramatic. [21] C.W. Dolan, J. Tanner, D. Mukai, H.R. Hamilton, E. Douglas, Design Guidelines
for Durability of Bonded CFRP Repair/Strengthening of Concrete Beams,
NCHRP Web-Only Document 155, 2008.
8.2. Field application [22] E. Douglas, Introduction to Materials Science and Engineering: A Guided
Inquiry, Pearson Higher Education Inc, Upper Saddle River, NJ, 2013.
[23] M. Frigione, M.A. Aiello, C. Naddeo, Water effects on the bond strength of
Following hygrothermal conditioning the T g of epoxy E reduced concrete/concrete adhesive joints, Constr. Build. Mater. 20 (2006)
to 25 °C (77 °F), and in the other epoxies it fell in the range 957–970.
between 45 and 54 °C (113–129 °F). These T g values are well [24] A. Gartner, E. Douglas, C. Dolan, H. Hamilton, Small beam bond test method for
CFRP composites applied to concrete, J. Compos. Constr. 10 (6) (2011) 52–61.
within the temperature range experienced in many field applica- [25] E.H. Immergut, H.F. Mark, Principles of plasticization. Plasticization and
tions. It is evident that structural epoxies with T g values near or plasticizer processes, Adv. Chem. 48 (1965) 1–26.
lower than the design temperature conditions should not be used [26] International Code Council (ICC), Acceptance Criteria for Concrete and
Reinforced and Unreinforced Masonry Strengthening using Externally
where water immersion and/or high levels of humidity are Bonded FRP Composite Systems, ICC AC125, ICC Evaluation Service, Whittier,
expected. Designer should not rely on the T g values reported by CA, 2012.
the manufacturer, as it was shown that T g is very sensitive to the [27] W.S. Kim, I.H. Yun, J.J. Lee, H.T. Jung, Evaluation of mechanical interlock effect
on adhesion strength of polymer–metal interfaces using micro-patterned
conditioning environment. The design T g value is to be determined surface topography, Int. J. Adhes. Adhes. 30 (6) (2010) 408–417.
experimentally for epoxy samples cured in conditions representa- [28] O. Moussa, A.P. Vassilopoulos, D. Castro, T. Keller, Mechanical recovery of
tive of the actual design environment. epoxy adhesive subsequent to exceeding glass transition temperature, in: 4th
International Conference on Durability of Composites for Construction (CDCC
2011), University of Sherbrooke, Quebec, Canada, 2011.
References [29] A. Stewart, Study of Cement-Epoxy Interfaces, Accelerated Testing, and
Surface Modification, Ph.D. Dissertation, University of Florida, Gainesville, FL,
2012.
[1] R.A. Atadero, D.G. Allen, O.R. Mata, Long-term Monitoring of Mechanical
[30] J. Tatar, P. Blackburn, C. Weston, H.R. Hamilton, Direct Shear Adhesive Bond
Properties of FRP Repair Materials, Colorado Department of Transportation,
Test, in: Proceedings of the 11th International Symposium of Fiber Reinforced
Report No. CDOT-2013-13, 2013.
Polymer for Reinforced Concrete Structures (FRPRCS-11), University of Minho,
[2] Z. Karim, D. Mela, M. Di Benedetti, A. Nanni, Durability Evaluation of Shear
Guimaraes, Portugal, 2013.
Bond Strength of Carbon and Glass FRP Laminates Using Small Plain Concrete
B.P. Blackburn et al. / Construction and Building Materials 96 (2015) 679–689 689

[31] J.M. Zhou, J.P. Lucas, Hygrothermal effects of epoxy resin. Part I: The nature of [33] S. Choi, A.L. Gartner, N.V. Etten, H.R. Hamilton, E.P. Douglas, Durability of
water in epoxy. Part II: Variations of glass transition temperature, Polymer 40 concrete beams externally reinforced with CFRP composites exposed to
(20) (1999) 5505–5522. various environments, J. Compos. Constr. 16 (2012) 10–20.
[32] ASTM, Standard D638, Standard Test Method for Tensile Properties of Plastics, [34] B.P. Blackburn, Effects of Hygrothermal Conditioning on Epoxy Used in FRP
ASTM International, West Conshohocken, Pa, 2014. Composites. M.S. Thesis, University of Florida, Gainesville, FL, 2013.

You might also like