Sei sulla pagina 1di 18

Durability of GFRP Pultruded Profiles made of Vinylester

Resin

Joo Pedro Giro Meireles de Sousa

M.Sc. Dissertation Extended Abstract

October 2011

J.Sousa

DURABILITY OF GFRP PULTRUDED PROFILES MADE OF


VINYLESTER RESIN
Joo Pedro Giro Meireles de Sousa
Summary: This paper presents new findings of an ongoing research project developed in
partnership by Instituto Superior Tcnico and Laboratrio Nacional de Engenharia Civil (ISTLNEC). It essentially focuses on the deterioration of glass fiber reinforced polymer (GRFP)
made of vinylester resin and E-glass fibers, when exposed to different environmental agents.
Keywords: GFRP; E-Glass; Vinylester; Durability; Ageing environments; Experimental tests

INTRODUCTION

The need of higher construction speed in civil engineering applications, together with the
durability problems experienced by traditional materials, such as steel and reinforced concrete,
have been fostering the development of new structural solutions [1]. The structural use of glass
fiber reinforced polymer (GFRP) as pultruded profiles in civil engineering applications has
grown significantly in the last two decades. This is due to their several advantages when
compared to traditional materials, namely, high strength, lightness, good insulation properties
and resistance to corrosion [2]. Until recently, the effect of fluids on composites has often been
considered to be of secondary importance in glass-fiber reinforced polymers because, to date,
the detrimental effects have been overcome by significant overdesign [3,4]. However, the use of
inordinately high factors of safety on this account cannot be easily justified and there is a urgent
need to deepen the understanding of mechanisms leading to moisture-associated degradation,
especially as related to vinylester systems for which there is still a significant lack of data.
Together with this fact, the relatively low elasticity modulus, a brittle behaviour and material
costs (which are still not competitive for mainstream applications) are delaying the widespread
use of GFRP profiles [1,5].
This paper presents results of an ongoing experimental research on the physical, chemical,
aesthetical and mechanical changes suffered by GFRP pultruded profiles made of vinylester
resin and E-glass fibers, following accelerated exposure to moisture, water as well as natural
ageing. This study also aims to understand the influence of moderate temperature cured systems
(dried systems) in order to access reversibility effects, as well as coating of unprotected
specimen parts (isolated systems) into deterioration of the GFRP profiles, measured by the same
changes as described above.

J.Sousa

MATERIALS

The material under study was obtained from one commercial GFRP pultruded tubular profile
(50 mm x 50 mm, thickness of 5 mm, with common usage as part of ladders and handrails,
produced by ALTO Perfis Pultrudidos Lda, Portugal). This material consists of alternate layers
of unidirectional E-glass fiber rovings and strand mats embedded in vinylester resin.

METHODS

3.1

Exposure environments

In order to study the potential degradation of the GFRP profiles in typical civil engineering
applications environments, test specimens were subjected to the exposure conditions described
in Table 1. The latter also indicates the batches of aged material already tested.
Table 1 Exposure aging conditions.

Type of exposure
Group I

Duration

Conditions

(represents the ongoing study related to prior researches)


(W-20)

Immersion in
demineralised water

Immersion in saltwater

Temperatures: 20 (2) C,

(W-40)

40 (1) C e 60 (1) C

(W-60)

3,6,9,

(S-20)

12,18,24 months (a)

Composition: 35g/l NaCl

(S-40)

Temperatures 20 (2) C,

(S-60)

40 (1) C e 60 (1) C
In the roof of the LNEC building,

Natural Environment

(NE)

1, 2, 5, 10 years (b)

where temperature, relative


humidity and UV radiation are
continuously monitored

Group II

(represents the effects of isolated (I) and dried(D) systems)


(WI-20)

Immersion in

(WI-40)

demineralised water

(WD-20)

6, 12, 18 months (c)

Temperatures: 20 (2) C,
40 (1) C

(WD-40)
Continuous

(CCI-40)

Condensation

(CCD-40)

6, 12, 18 months (c)

Temperatures: 40 (2) C
Relative humidity: 100%

Batches of aged material previously tested: (a) 3, 6 and 9 months [6,7].


Batches of aged material tested in this study: (a) 12, 18 and 24 months; (b) 1 and 2 years;
(c) 6 and 12 months

3.2

Experimental procedures

Before exposure, Group II batches were prepared using the following techniques:

J.Sousa

(i)

Isolated systems: unprotected specimen parts (namely its side parts) derived from their
cut and preparation. Therefore, the epoxy resin Icosit K 101 N provided by Sika was
applied and cured one week at 50 C onto those unprotected parts.

(ii)

Dried systems: test specimens were subjected to moderate temperature cure, until no
significant mass change was reported according to ASTM D 5229 standard [8].

After exposure to the different ageing conditions described in Table 1, the aged batches were
subjected to the following characterization techniques:
(i)

Sorption behaviour: Mass changes for the control specimens (with similar geometry to
that of specimens used in dynamic mechanical analysis) were recorded using an
electronic scale, removing them periodically from immersion and continuous
condensation exposures. To complement this data, apparent diffusion coefficients were
calculated assuming Fickian sorption behaviour.

(ii)

Aesthetical characterization: Samples aged at Natural Environment conditions were


tested in comparison with unaged and QUV accelerated tested material:

Colour changes: specimens with similar geometry to that used in tensile


properties were tested in accordance with parts 1 and 2 of ISO 7724 standard
[9,10], using the colour system CIE 1976 [9]. Tests were carried out using a
Macbeth Coloreye 3000 colorimeter.

Gloss: specimens with similar geometry to the one used in tensile properties were
tested in accordance to ISO 2813 standard [11]. 20 , 60 and 85 incidence
angles were measured with Novo-Gloss Statistical Glossmeter.

(iii)

Dynamic mechanical analysis (DMA): DMA technique has been used to analyse the
viscoelastic response of materials such as the GFRP, as well as to assess the glass
transition temperature ( ), in accordance with part 1 and 5 of ISO 6721 standard [12,13].
Three-point bending type clamp specimens with 5 x 15 x 60 mm were tested at a constant
frequency of 1 Hz and strain amplitude of 15 m, using a Q800 model of TA Instruments.
The analyses were conducted from room temperature up to 200 C, at a rate of 2 C/min.
Three replicates were tested for each duration and ageing condition.

(iv)

Mechanical behaviour: Five samples for each duration of ageing condition were
submitted to mechanical tests in the longitudinal direction:

Flexural properties: three-point bending flexural tests were performed according


to ISO 14125 standard [14] in rectangular test specimens with 5 x 15 x 150 mm
in a 100 mm span. Tests were carried out at a loading rate of about 2 mm/min,
using a system from Seidner Form Test, constituted by a hydraulic press with a
10 kN load capacity.

J.Sousa

Tensile properties: tensile tests were conducted according to parts 1 and 4 of ISO
527 standard [15,16] in rectangular test specimens with 5 x 25 x 300 mm, without
end tabs. Tests were carried out in an Instron universal testing machine with a
load capacity of 100 kN at a 2 mm/min loading rate.

Shear properties: interlaminar shear tests were carried out in accordance with
ASTM D2344 standard [18] in rectangular test specimens with 5 x 15 x 30 mm in
a 20 mm span. Tests were carried out at a loading rate of 1 mm/min, with the
Seidner Form Test system.

Excluding the study of the mass changes, after being removed from the different exposure
environments and prior to further testing, specimens were placed inside polyethylene bags.
These were hermetically closed, trying to maintain the moisture content of the material, and
then placed inside a room with temperature controlled at 20 (2) C. Prior to testing, specimens
were removed from the polyethylene bags and immediately tested without any further
conditioning.

RESULTS AND DISCUSSION

4.1

Initial characterization

To fully understand the ongoing changes in the properties of the GFRP pultruded profiles,
results of their initial physical and mechanical characterization are listed in Table 2.
Table 2 - Physical-chemical properties of GFRP profiles (un-aged) [1].

Property
Glass fiber
content (%)
Density
(C)

Method
Calcination
Immersion
DMA

Flexure
Mechanical
properties

Results

Tension
Interlaminar
shear

68,7 0,4
(g/cm3)
Einitial

2,03 0,052
98,6 7,0

tan

126,9 2,3

tu (MPa)

393 51

Et (GPa)

38,9 4,1

fu (MPa)

537 73

Ef (GPa)

28,4 3,4

u (MPa)

39,2 4,2

With regard to the mechanical behaviour, in all characterization tests (tension, flexure and
interlaminar shear) the GFRP profile exhibited a well-defined linear elastic behaviour up to
failure, which is one of the main features of this material.

J.Sousa

4.2

Sorption behaviour

Figure 1 illustrates the mass changes that occurred during the ageing exposure to immersions
(both in demineralised and salt water) for each temperature (20 C, 40 C and 60 C). Figure 2
shows the evolution of the same characteristic considering fully (FI) and partially coated (I)
specimens compared with unprotected ones, in demineralised water and continuous
condensation, for each temperature (20 C and 40 C) in order to evaluate coating sorption
effects.
0,80

Mass change (%)

0,70

W-20

0,60

W-40

0,50
0,40

W-60

0,30

S-20

0,20
S-40

0,10

S-60

0,00
0

2000

4000

6000

8000

10000

12000

14000

16000

18000

Time (h)

Figure 1 Mass changes for different hygrothermal ageing conditions.


1,20

Mass change (%)

1,00

WD-20
WI-20

0,80

WFI-20

0,60

WD-40
WI-40

0,40

WFI-40

0,20

CCD-40
CCI-40

0,00
0

2000

4000

6000

8000

CCFI-40

Time (h)

Figure 2 Mass changes for coated specimens at different hygrothermal ageing conditions.

The overall results, in terms of maximum percentage weight gains and apparent diffusion
coefficient (D) are listed in Table 3. It should be noted that mass changes showed, roughly, a
Fickian response, with rates of mass uptake increasing with temperature, especially at a short
time, which are concordant with Karbhari and Zhang [18]. After 8700 hours saturation point has
been already achieved, however some mass reduction is noted at higher temperatures suggesting
post-cure effects, which were also noted by Liao et al. [19] although at an earlier point in time.
Moreover, for similar temperatures, mass uptakes of specimens immersed in salt water were
always lower compared to specimens immersed in demineralised water. Although mass changes

J.Sousa

were not as low as expected in coated specimens, possibly due to some previous saturation of
unprotected specimens, diffusion coefficient results indicate decreasing values as the protection
degree increases and therefore showing sorption retention by the protective epoxy. All these test
results are consistent with mass changes from other batches having a 2% variation, which points
out to the system reliability.
Table 3 Maximum percentage weight gain and apparent diffusion coefficient.
Group I

4.3.1

Mm (%)

D (x10-7 mm2/s)

WD-20

0,64

11,1

1,22

WI-20

0,68

3,52

2,26

WFI-20

0,74

3,22

0,34

1,47

WD-40

1,04

2,70

S-40

0,31

1,65

WI-40

0,96

2,46

S-60

0,45

6,14

WFI-40

0,64

2,30

CCFI-40

0,77

2,67

CCD-40

0,44

5,55

CCI-40

0,91

3,00

Mm (%)

D (x10 mm /s)

W-20

0,77

1,69

W-40

0,64

W-60

0,77

S-20

Environment

4.3

Group II
-7

Environment

Aesthetical characterization
Colour changes

Two years old natural ageing specimens were subjected to testing for aesthetical
characterization. Table 4 lists the values of colour space system coordinates including their
colour variation E*.
Table 4 Values of colour space system coordinates after natural ageing.
L*

a*

b*

L*

a*

b*

E*

Unaged

74,15

-0,88

1,13

NE (2 years old)

72,31

-1,70

7,94

-1,85

-0,82

6,82

7,11

Specimens

These results show that there is a loss of coloration mainly due to the UV radiation. The values
of L*, a* and b* have shown that the material becomes more yellow (roughly 7,5% increase),
which corresponds to significant colour changes, easily identifiable. These results can correlate
to QUV accelerated ageing (measured by Carreiro [6]), differing 1,4% when compared to 2000h
of ageing and showing significant changes after 2 years of exposure (as described also by
Bogner and Borja [20]).
4.3.2

Gloss

Gloss testing was made on the same specimens at 20 , 60 and 85 light incidence angles. The
results obtained are listed in Table 5.

J.Sousa

Table 5 Gloss values for profiles after natural ageing.


Specimens

Surface

Unaged
NE (2 years old)

Light incidence angle


20

60

85

4,3

27,7

27,4

Sun

0,9

2,0

1,4

Shade

3,7

21,2

19,6

All light incidence angles show significant gloss loss, especially to sun oriented surfaces which
only retained 7% at 60 after being exposed for two years. Comparing again with QUV
accelerated ageing values are nearly the same as the ones obtained by Carreiro [6] at 3000h of
exposure. Significant gloss loss is attained when exposed to prolonged UV radiations.

4.4

Dynamic mechanical analysis

Figures 3 and 4 plot Tg variation (mean value and standard deviation) against time, representing
all ageing environments of Groups I and II, respectively.
145

Tg ( C)

140
135
W-20

130

W-40

125

W-60

120

S-20

115

S-40
S-60

110
0

12

16

20

24

Time (months)

Figure 3

variation of Group I ageing conditions.

145

Tg ( C)

140
135

WD-20

130

WD-40

125

CCD-40

120

WI-20
WI-40

115

CCI-40

110
0

12

16

20

24

Time (months)

Figure 4

variation of Group II ageing conditions.

Attained results through DMA analysis show that higher Tg loss is initially observed in higher
temperature environments. In addition, it is also possible that this exposure accelerates post-cure

J.Sousa

phenomena, which are the most accentuated at W-60 environment. This was also suggested by
Chin et al. [21]. After two years of exposure, Group I environments results show little
difference between them, consisting of 5-7 C when compared to initial values. Additionally,
results show that demineralised immersions and salt-water environments cause similar changes
in Tg parameter. Lowest Tg values were noted at 9 months of exposure never decreasing more
than 10%. Group II environments show that in most of them Tg is likely susceptible to higher
temperature and moisture content. It is also noted that both moderate temperature cured systems
and protective coatings translate into a positive factor, since they attenuate the degrading
process resulting even in Tg regains (more highlighted on the dried state). Consequently, it is
suggested that moderate temperature cured systems helps to reach the post-cure status in a faster
way. Protective coating specimens presented different individual results compared to the other
environments, due to the different Tg of the epoxy coating.

4.5
4.5.1

Mechanical behaviour
Flexural properties

Results obtained from flexural tests, namely the flexural strength and modulus as a function of
time for the different exposure environments, are presented in Figures 5 to 8.

Flexural Strength [MPa]

600

500

W-20
W-40

400

W-60
S-20
S-40

300

S-60
NE

200
0

12

16

20

24

Time (months)

Figure 5 Flexural Strength of Group I ageing conditions.

Results show an overall tendency of flexural strength and modulus degradation over time.
However, there are signs of post-cure effects, resulting in property gain, consistent with
moisture sorption mass reductions. Salt immersed environments tend to accentuate post-cure
effects. All Group I environments show significant flexural strength loss after two years of
exposure (apart from natural ageing which caused negligible loss).

J.Sousa

Flexural Strength [MPa]

600

500
WD-20
WD-40

400

CCD-40
WI-20

300

WI-40
CCI-40

200
0

12

16

20

24

Time (months)

Figure 6 Flexural Strength of Group II ageing conditions.

Flexural Modulus [GPa]

30

25

W-20
W-40

20

W-60
S-20
S-40

15

S-60
NE

10
0

12

16

20

24

Time (months)

Figure 7 Flexural Modulus of Group I ageing conditions.

Flexural Modulus [GPa]

30

25
WD-20
WD-40

20

CCD-40
WI-20

15

WI-40
CCI-40

10
0

12

16

20

24

Time (months)
Figure 8 Flexural Modulus of Group II ageing conditions.

These losses vary between 17,8 % (S-20) and 47,7% (W-60). Both higher temperatures and
moisture potentiate flexural strength decrease and salt water has less influence than
demineralised water. The latter have been also noted by Liao et al. [19]. Group I flexural
modulus show its greatest losses at 12 months of exposure, where W-60 dropped 45,4%,
experiencing a recovery process afterwards. Group II environments show that coated systems

J.Sousa

10

experience greater flexural strength loss at 12 months than on cured systems and nearly the
same flexural modulus. Compared with Group I, only slight differences are found, presenting
slightly less mechanical deterioration.
4.5.2

Tensile properties

Results obtained from tensile tests, namely the tensile strength and modulus as a function of
time for different exposure environments, are presented in Figures 9 to 12.

Tensile Strngth [MPa]

450
400

W-20
W-40

350

W-60

300

S-20
S-40

250

S-60
NE

200
0

12

16

20

24

Time (months)

Figure 9 Tensile Strength of Group I ageing conditions.

Tensile Strngth [MPa]

500

450
WD-20
WD-40

400

CCD-40
WI-20

350

WI-40
CCI-40

300
0

12

16

20

24

Time (months)

Figure 10 Tensile Strength of Group II ageing conditions.

As in flexural properties, results show an overall tendency of tensile strength and modulus
degradation over time. The same post-cure related gains are noticeable especially in Group I
lower temperature environments. The biggest tensile strength reductions are shown at higher
temperatures with a 38,1% loss in W-60 environment. Two years of exposure immersion in
demineralized water shows bigger deterioration effects when compared to salt water
environments.

J.Sousa

11

Tensile Modulus [GPa]

55
50
W-20

45

W-40

40

W-60
S-20

35

S-40

30

S-60
NE

25
0

12

16

20

24

Time (months)

Figure 11 Tensile Modulus of Group I ageing conditions.

Tensile Modulus [GPa]

45

WD-20

40

WD-40
CCD-40

35

WI-20
WI-40
CCI-40

30
0

12

16

20

24

Time (months)

Figure 12 Tensile Modulus of Group II ageing conditions.

Natural ageing specimens show only 3,3% loss of tensile strength after one year of exposure.
Tensile modulus show some initial oscillations, maintaining a decrease tendency until 12
months of exposure, subsequently stabilizing over time. The maximum reduction of this
property is 16,2% in S-60 environment after two years of exposure. Group II environments
show little change in tensile strength and a slow decreasing tensile modulus. The maximum
Group II tensile loss occurred in CCI-40 environment with a 8,13% decay after one year of
exposure, where higher temperatures and increased moisture are relevant to this deterioration.
However, both Group II systems helped slowing overall deterioration in test specimens.
Chu et al. [22] described that the same reversibility system on GFRP vinylester pultruded
profiles presented a 6,9-27,3% improvement in tensile strength, after 28 days of exposure in
23 C demineralised water.
4.5.3

Interlaminar shear strength

Figures 18 and 19 illustrate the variation of the interlaminar shear strength as a function time,
for different exposure environment conditions.

J.Sousa

12

Interlaminar shear stength [MPa]

45
40
W-20

35

W-40

30

W-60
S-20

25

S-40

20

S-60
NE

15
0

12

16

20

24

Time (months)

Inetrlaminar shear strength [MPa]

Figure 13 Interlaminar shear strength of Group I ageing conditions.


45
40
WD-20

35

WD-40
CCD-40

30

WI-20

25

WI-40
CCI-40

20
0

12

16

20

24

Time (months)

Figure 14 Interlaminar shear strength of Group II ageing conditions.

Interlaminar shear strength results point to the same tendencies as the other mechanical tests. It
generally decreases over time and Group I environments show possible post-cure effects due to
strength gains after 9 months of exposure, which has also been noted by Karbhari [5]. However,
it is noticed the existence of a plateau until 18 months, followed by a slow downward deflection
in strength. Losses after two years of exposure, highlight high variations - from 52,3% at higher
temperatures and only 6,6 % at lower ones. This fact suggests that interlaminar shear strength
degradation is highly temperature dependant. Over time demineralized water seems to have
more deteriorative effects than salt water, however, final results present almost no differences
between these two types of environments. Natural ageing results point towards a slight gain of
interlaminar shear strength at 12 months of exposure. However, after two years, deteriorative
effects are noticed resulting in a 19,7% loss, when compared to the initial data. On the other
hand, group II environments present minor changes after 6 months of exposure and some
significant changes in continuous condensation and high temperature environments after 12
months, where CCI-40 decayed 35,2%. Nevertheless, Group II environments show less
deterioration when compared with the same Group I conditions at 6 month of exposure. Group

J.Sousa

13

II environment do not show strength recovery, in spite of some cases where this phenomena
occurred for longer exposure times.

Conclusions

The main objective of this study was successfully achieved, leading to the determination of the
mechanical, physical and aesthetical variations suffered by GFRP pultruded profiles as result of
time and under different ageing conditions.
Regarding moisture sorption it was confirmed that water molecules incorporation is temperature
dependant, especially in the initial states. High initial sorption rates, slowing down to a plateau
over time indicates approximate Fickian behaviour. Higher temperatures environments showed
higher initial sorption rates. However, demineralised water immersions at 60 C highlighted
mass losses after 8000h of exposure. Mass loss at high temperature is strongly related to
hydrolysis. Group I environments presented 0,77% maximum mass change in two years and in
Group II the WD-40 environment absorbed higher amounts of water (1,04%). Apparent
diffusion coefficients of both groups present a decrease pattern with lower temperatures and
with the use of protective coating (lowest on fully protected ones). Moreover, the latter supports
the fact that temperature increases sorption rates and the protective coating decreases it and acts
as a sorption wall. All these test results are consistent with mass changes of other batches
having a 2% variation, which, in turn, confirms the system reliability.
Colour and gloss variation showed significant variations. Initially grey, two years old natural
aged specimens presented themselves 7,5% more yellow, which could be visually confirmed on
them and showed only 20% gloss retention at 20 , being these changes similar to accelerated
QUV tests.
DMA analysis revealed that the biggest variations occurred in environments where moisture and
temperature play an important role. Group I environments presented biggest Tg changes at 9
months of age in both immersion types. However, Tg loss never surpassed 10%. It was noted
that demineralised water immersion and salt water caused, at different temperatures, identical T g
variations taken from tan peak over time. Salt water presented the same tendency as
demineralised water but with fewer variations. Decreased Tg values over time indicate
plasticization phenomena, usually found when polymeric materials contact with water
molecules. However, Tg decrease seems to slow after a given exposure period and even
presenting some gains, especially at higher temperatures, which suggests post-cure effects on
the vinylester matrix. The applied coating helped in maintaining Tg value over time (slowing
deterioration effects) - protected specimens showed no significant loss during one year.
Moderate temperature cured system effects showed Tg gains over time, which are subsequently
related to the material reversibility process.

J.Sousa

14

All mechanical tests showed linear elastic behaviour until rupture. Mechanical behaviour
revealed significant performance decrease especially in tensile (maximum 38,1% loss) and in
interlaminar shear strength (52,3% loss). Temperature and moisture content effects highly
influence the material deterioration degree. Different testing offer complementary data as tensile
properties are strongly correlated to fiber performance, and interlaminar shear strength is related
with matrix and its interface deterioration. The biggest interlaminar shear strength reductions
suggest performance loss not only due to fiber degradation but general composite deterioration.
All test showed, at some time, property gains implying a possible post-cure effect in the
vinylester matrix that can be related to incorrect cure state achieved during production. In a
global sense, moderate temperature cured and protective coated systems tests went as expected,
showing less overall deterioration over time when compared to other testing environments. This
can be a decisive factor to improve the real material performance when in service conditions,
which can even result in reconsideration of the safety coefficients related to material behaviour.

REFERENCES

[1] S. Cabral-Fonseca, J.R. Correia, R. Costa, A. Carreiro, M. Paula Rodrigues, M. Isabel


Eusbio, F.A. Branco, Environmental degradation of GFRP pultruded profiles made of
polyester and vinylester resins, 15th International Conference on Composite Structures,
(Editor: A.J.M Ferreira), pp. 1-5, Porto, 2009.
[2] J.R. Correia, GFRP Pultruded Profiles in Civil Engineering: hybrid solutions, bonded
connections and fire behaviour, PhD Thesis in Civil Engineering, Instituto Superior Tcnico,
Universidade Tcnica de Lisboa, Lisboa, 2008, 420p.
[3] Y. J. Weitsman, Effects of fluids on polymeric composites - a review, Report
MAES98-5.0-CM, Office of Naval Research, Mechanical and Aerospace Engineering and
Engineering Science, University of Tennessee, Knoxville, 1998
[4] V.M. Karbhari, Durability of composites for civil structural applications, Woodhead
publishing, July 2007.
[5] V. M. Karbhari, E-Glass/Vinylester Composites in Aqueous Environments: Effects on
Short-Beam Shear Strength, Journal of Composites for Construction, 8:2, 2004, 148-156
[6] A. Carreiro, Durabilidade de perfis pultrudidos de vinilster reforado com fibras de vidro
(GFRP), MSc Dissertation in Civil Engineering, Instituto Superior Tcnico, Lisboa, Maio
2010, 122p.

J.Sousa

15

[7] R. Costa, Durabilidade de perfis pultrudidos de polister reforado com fibras de vidro
(GFRP), MSc Dissertation in Civil Engineering, Instituto Superior Tcnico, Lisboa, Novembro
2009, 108p.
[8] ASTM D 5229 Standard Test Method for Moisture Absorption Properties and Equilibrium
Conditioning of Polymer Matrix Composite Materials, American Society for Testing and
Materials, West Conshohocken, PA, 2004
[9] ISO 7224-1, Paints and varnishes Colorimetry Part 1: Principles, International
Organization for Standardization, Genve, 1997.
[10] ISO 7224-2, Paints and varnishes Colorimetry Part 2: Colour management,
International Organization for Standardization, Genve, 1997.
[11] ISO 2813, Paints and varnishes Determination of specular gloss of non-metalic paint
films at 20, 60 and 85, International Organization for Standardization, Genve, 1994.
[12] ISO 6721-1 Plastics Determination of dynamic mechanical properties Part:1 General
Principles, International Organization for Standardization, Genve, 1994.
[13] ISO 6721-5 Plastics Determination of dynamic mechanical properties Part:5 Flexural
vibration - Non-ressonance method, International Organization for Standardization, Genve,
1996.
[14] ISO 14125, Fibre-reinforced plastic composites Determination of flexural properties,
International Organization for Standardization, Genve, 1998.
[15] ISO 527-1, Plastiques Dtermination des proprieties en traction Partie 1: Principes
gnraux, International Organization for Standardization, Genve, 1993.
[16] ISO 527-4, Plastics Determination of tensile properties Part 4: Test conditions for
isotropic and orthotropic fibre-reinforced plastic composites, International Organization for
Standardization, Genve, 1997.
[17] ASTM D 2344. Standard Test Method for Short-Beam Strength of Polymer Matrix
Composite Materials and Their Laminates, American Society for Testing and Materials, West
Conshohocken, PA, 2000.
[18] V.M. Karbhari, S. Zhang, E-Glass/Vinylester Composites in Aqueous Environments I:
Experimental Results, Applied Composite Materials, 10, 2003, 19-48.

J.Sousa

16

[19] K. Liao, C.R. Schultheisz, D.L. Hunston, Effects of environmental aging on the properties
of pultruded GFRP, Composites: Part B, 30, 1999, 485-493.
[20] B.R. Bogner, P. P. Borja, Strenght retention of pultruded composites after exposure to
ultra-violet (UV) light, BP Amoco research center, Naperville Illinois, EUA.
[21] J. W. Chin, W.L. Hughes, A. Signor, Elevated temperature aging of glass fiber reinforced
vinyl ester and isophthalic polyester composites in water, salt water and concrete pore
solution Proceedings of the 16th ASC Conference, Blacksburgh, 2001, 12pp
[22] W. Chu, L. Wu, V.M. Karbhari, Durability evaluation of moderate temperature cured Eglass/vinylester systems, Composite Structures. 66, 2004, 367-376.

Potrebbero piacerti anche