You are on page 1of 21

DAMSTRUC'2000

2° Congresso Internacional Sobre o Comportamento de Estruturas Danificadas


2nd International Conference on the Behaviour of Damaged Structures

RENEWAL OF CIVIL INFRASTRUCTURE USING FRP


COMPOSITES – EFFICIENT USE OF MATERIALS AND
PROCESSES FOR REHABILITATION

Vistasp M. KARBHARI
Ph.D., Associate Professor and Vice-Chair, Department of Structural Engineering,
University of California San Diego, La Jolla, Ca 92093-0085

ABSTRACT
Fiber reinforced polymer (FRP) composites are increasingly being used for the renewal of
civil infrastructure since they provide enhanced properties, potentially long-term
durability, and fabrication/construction options often not possible with conventional
materials. Initially used extensively for seismic retrofit, these materials are now being
used in a variety of forms for repair and strengthening of beams, girders, and slabs, and
for the development of new bridge, building, and industrial structural systems. This paper
provides an overview of some applications in the area of repair and strengthening,
emphasizing the importance of, and the effects of, choice of materials and processes.
Aspects relating to durability, the current state of knowledge in this area, and test
standards are also discussed. Opportunities and challenges pertaining to the development
of new materials and processes designed specifically for this sector of the FRP industry
are also discussed.

1. INTRODUCTION

Initially applied to the seismic retrofit of columns and in cables/tendons as alternative


reinforcement in concrete, fiber reinforced polymer (FRP) matrix composites are rapidly
being accepted by the civil engineering and construction communities in applications
ranging from replacement for steel components in applications such as rebar and cables,
to use in repair and retrofit of concrete components, and even in new structural systems
for bridges and industrial facilities. The tailorability of these materials derived from the
inherent isotropy predicated by the arrangement of the fiber reinforcement in the polymer
matrix, in addition to the corrosion resistance, lightweight, high strength- and stiffness-to-
weight ratios enable their use in ways that are not available to conventional materials.
These unique characteristics provide significant impetus for their use in rehabilitation and
restoration of historic structures without causing significant changes to the aesthetic
features or geometric configurations of the original structures. Similarly, the performance
attributes, in conjunction with their light-weight attributes, enable their use in
strengthening of severely degraded structural elements, as well as in the modification of
existing structures without egress on available head-room or open space. Further, when
used in new structural systems they provide the designer with the ability to truly blend
form and function.

Despite the increasing use of FRP composites in civil infrastructure applications


worldwide, there are a number of significant challenges that have still to be faced before
this class of materials is widely accepted as belonging to the palette of construction
materials. Chief among the challenges are those related to (a) the verification and
validation of long-term durability (75+ years), (b) the development of lower cost, but high
quality, manufacturing processes, (c) the development of appropriate codes, standards,
specifications and guidelines, and (d) the education of civil engineers and designers in the
intricacies of composites while simultaneously educating the composites community in
the requirements and constraints of civil infrastructure. The last two challenges are
further heightened since short-term properties and performance attributes, or a subset of
these, can be arrived at through the use of a range of materials-configuration-process
alternatives, making down-selection and specification a concern, especially in light of the
close interaction between performance and processing details.

Notwithstanding the large number of demonstration projects completed in the US, Japan,
and Europe, and the tremendous increase in commercial projects related to rehabilitation
over the past three years, the widespread acceptance of FRP composites will be
contingent on the development of appropriate products and methods of use in the field at
costs that are competitive with current costs using conventional materials and
technologies. Although life-cycle durability is a strong motivation for the use of FRP
composites, by itself this facet is not sufficient to overcome cost differentials between
FRP composite products and otherwise equivalent conventional ones. However, savings
accruing at the systems level due to faster construction, thereby causing less distress to
the community, lower dead weight requiring smaller and lighter substructure as well as
lighter construction equipment, and the ability to achieve renewal in cases where
conventional technologies have only the option of complete rebuilding, present strong
motivation for the use of this class of materials.

It must be stressed at the very outset that there are a number of myths regarding FRP
composites, chief among which are that:
• Composites are a panacea for the deterioration of civil infrastructure;
• Composites do not degrade and are not affected by environmental influences;
• All composites have well documented data bases;
• Variation in field conditions, incoming raw materials, or manufacturing processes
do not cause any changes in properties and performance, and
• Performance attributes of FRP composites used in civil infrastructure, at present,
are comparable to those of aerospace grade prepreg based autoclave cured
composites.
In reality, however
• FRP composites do provide, in a variety of cases, options for rehabilitation that
are not possible with conventional materials, although they cannot, and in fact
should not, be used to repair all deficiencies in existing structures
• The environmental durability of FRP composites is largely predicated on the
appropriate selection of materials and process combinations, and although they do
not rust, FRP composites can degrade due to a combination of harsh
environmental influences. In most cases, however, this degradation can be
guarded against through good detailing and process control.
• FRP composites likely to be used in civil infrastructure applications have
significant differences from the prepreg based, autoclave cured, well-characterized
systems conventionally used in the aerospace industry. A number of these
systems are based on the use of polyester and vinylester resin systems which have
significantly different cure characteristics, moisture sensitivity, and creep
characteristics than the higher temperature cure epoxies used in the aerospace
world. Although well defined and validated data bases do not exist for the newer
versions, there is sufficient data and anecdotal evidence to suggest that very good
environmental durability and long life under working conditions can be obtained
with these materials.
• Variations in incoming raw materials, process conditions and characteristics and
local environmental conditions can cause variations in the properties of FRP
composites. In a number of cases, especially as related to the rehabilitation of
damaged concrete structural components, FRP composites have to be fabricated in
the field using manual processes and there can be significant batch-to-batch
variation in these properties.
Notwithstanding these aspects, in reality there is a great deal of information that provides
a very strong argument for their increased use in civil infrastructure, building on the
successes in the area of marine structures using similar materials and processes, and using
the knowledge gleaned from the decades of use in the aerospace area, where aspects of
structural safety are as critical as those for civil structures. This paper provides a brief
overview of specific applications of FRP composites in the areas of structural
rehabilitation of concrete structures as a means of emphasizing the immense potential of
these materials. Aspects related to durability, materials form, manufacturing and design
are also discussed with a view to outlining critical needs to be met as FRP composites are
used to face the challenges of the renewal of civil infrastructure in the 21st century.

2. REHABILITATION AND STRENGTHENING


Degradation due to corrosion of steel reinforcement, extensive cracking of concrete due
carbonation and/or freeze-thaw action or other deterioration, and rapidly changing traffic
needs (both in terms of intensity and load levels) have created a critical need for methods
of strengthening of soffits of bridge decks. In the case of buildings, besides materials
degradation, changing needs of building occupancy (e.g. residential to office) as well as
upgrades to facilities (e.g. cutting of floor slabs to facilitate the construction of elevator
shafts for the installation of elevators in older buildings), necessitates the strengthening of
existing slabs and beams. Fiber reinforced polymer matrix composites provide a cost-
effective, efficient, and rapid alternative to conventional methods of repair/strengthening,
and can often facilitate renewal in cases where it would not have been possible through
conventional means. Since the composite element is bonded onto the concrete substrate,
the efficacy of the method depends on the combined action of the entire system with
emphasis on the integrity of the bond and the interface layers. Thus the composite-
adhesive/resin-concrete system has to be considered as a complete system, and materials
aspects of each of the constituents, and interactions thereof between themselves and with
the external environment could have a significant effect on overall efficiency, renewal
capacity, and durability.
The ‘plate’ (or external reinforcement) can itself be fabricated in three generic
ways namely (a) adhesive bonding of prefabricated elements, (b) wet layup of fabric, and
(c) resin infusion. Of these, the pre-manufactured alternative shows the highest degree of
uniformity and quality control, since it is fabricated under controlled conditions. The
composite strip/panel/plate is prefabricated and then bonded onto the concrete substrate
using an adhesive under pressure. Application is rapid, but the efficiency in use is
predicated by the use of an appropriate adhesive and through the achievement of a good
bond between the concrete substrate and the composite adherend. Care must be taken to
ensure that the adhesive is chosen to match as closely as possible both the concrete and
composite vis-à-vis their elastic moduli and coefficients of thermal expansion, while
providing an interlayer to reduce mismatch induced stresses. The efficiency and durability
of the system rests on the ability of the adhesive to transfer load, and hence is dependent
on the integrity of the bond which is constructed in the field and on the durability of the
adhesive in a harsh and widely varying environment which could include excessive
moisture (or immersion in water, in the case of a flood plain) and large temperature
gradients. Typically the adhesives used to date in such applications are cured under
ambient conditions and are either polyurethanes or epoxies. It should be noted that
polyurethanes generically contain moisture sensitive components and may undergo
reversion in the presence of heat and/or moisture. Although they have good properties at
low temperatures they typically show significant reductions in shear performance at
elevated temperatures (Figure 1). Epoxies in comparison have a higher range and show
less creep and shrinkage, although they have a lower overall peel strength (Figure 2).
Unfortunately, to date, very little attention has been paid to the temperature and moisture
related deleterious effects on bond performance.

40 0.7
Polyurethanes
35 Polyurethanes
0.6
Tensile Shear Strength (MPa)

30 Epoxies
0.5
Peel Strength (MPa)

25
Phenolics 0.4
20
Polyimides
0.3 Toughened Epoxies
15

0.2 Acrylics
10

5 0.1
Higher Temperature
Phenolics
Adhesives
0
0
-50 0 50 100 150 200 250
0 100 200 300 400 500 600
Temperature (C)
Maximum Use Temperature (C)
Figure 1: Typical Working Temperature Figure 2: Relative Values of Peel Strength
Ranges for Structural Adhesives of Structural Adhesives as a Function of
Maximum Service Temperature

Although strips and plates can be premanufactured using a variety of processes,


commercially available strips are currently fabricated using unidirectional carbon fiber
reinforcement, which is pultruded to preset thicknesses and widths. The use of a
unidirectional form does provide the maximum theoretical efficiency of the fiber, but
necessitates the use of thicknesses in excess of those required on the basis of mechanical
properties in order to avoid longitudinal splitting during handling, and can suffer from
premature interlaminar cracking between layers of filaments. Figure 3 shows an example
of this wherein the failure surface has moved from one within the concrete to one within
the composite strip itself. In this case the change in failure path is due to the use of a
pultruded unidirectional carbon/epoxy strip as the external reinforcement, with the
composite having a fiber volume fraction of about 70%. This translates to a relatively
low resin content, resulting in microscopic dry spots and/or voids between individual
filaments creating weak zones along which fracture can easily propagate. Further, the
resin used was itself fairly brittle and hence the energetics of fracture favor a path of
propagation through this region once a crack has reached the adhesive-composite
interface (as from a flexural crack in the concrete that carries through the adhesive).

Figure 3: Divergence of Failure Surface from Concrete into the FRP Composite

The wet lay-up process is perhaps the most used currently and gives the maximum
flexibility for field application, and is probably also the cheapest alternative. However, it
presents the most variability, and necessitates the use of excessive resin, and could result
in the wrinkling or shear deformation of the fabric used, decreasing its designed
strengthening efficiency. The process entails the application of resin to the concrete
substrate followed by the impregnation of layers of fabric, which are bonded onto the
substrate using the resin itself. Both the composite and the bond are formed at the same
time in the field. The process affords the maximum flexibility in the field but has the
disadvantage of field mixing and fabrication and the potential absorption of moisture
and/or inclusion of impurities. Also, the process inherently bears with it the potential for
non-uniform wet-out of the fabric, and wrinkling/shearing of the fabric, both of which
affect final properties. Also the process carries with it the intrinsic entrapment of air
voids, and the resulting potential for deterioration with time. The composite formed in
this case is generally cured under ambient conditions. At present both plain-weave and
unidirectional fabrics are available commercially, with specific types of unidirectionals
with paper backing having been developed in Japan specifically to facilitate ease of use in
rehabilitation in the field, mimicking the application of wall paper.

The in situ resin infusion method is a fairly new variant and is capable of achieving
uniformity and good fabric compaction, while making it easier for the reinforcement to be
placed without excessive unintended deformation. However, this scheme is difficult to
apply over large areas and necessitates application of vacuum, which may be difficult to
hold on a severely cracked surface. In the resin infusion process, the reinforcing fabric is
first formed into a preform, which is attached to the substrate using a vacuum bag. Resin
is infused into the fibrous assembly under vacuum to form the composites. As in the wet
layup process, the composite and bond are formed at the same time. In a variant of the
process, the outer layer of the preform assembly is actually impregnated and partially
cured prior to placement in order to provide a good outer surface finish. The process is
far slower and has the same disadvantages as the wet layup process in terms of ambient
cure effects on durability. The process also has the potential to leave dry spots or
otherwise not completely fill the preform due to local irregularities, surface conditions, or
the inability to hold full vacuum.

As noted earlier, the efficacy of the external composite is largely dependent on the bond
between the composite and the concrete. Although significant research has been
conducted elsewhere on the critical aspect of surface characterization and treatment,
unfortunately very little attention, beyond the level of sand blasting and abrasion, has
been paid to this mechanism as related to concrete strengthening. Figures 3(a)-(c)
schematically depict possible variations and interfaces for the three cases of wet layup,
adhesive bonding, and resin infusion, respectively.
(a) Wet Layup

STRUCTURAL
RESPONSE

Structural Element

1. Putty / Paste / Filler 1. Primer


2. Primer 2. Resin Layer
3. Resin Layer 3. Composite Internal Influences
4. Composite

(b) Adhesive Bonding


Durability
Interfacial
External Influences
Influences

τ Interface
Characteristics

δ0

Material
1. Putty / Paste / Filler 1. Putty / Paste / Filler 1. Adhesive
σ} = {d} { ε}
{σ Constitutive
2. Adhesive 2. Adhesive 2. Composite
3. Adhesive 3. Composite
Theory
4. Composite

(c) Resin Infusion


Concrete
Materials
Adhesive / Resin Level
Composite

MATERIALS
RESPONSE

Figure 4: Levels of Integration Needed To


Link Materials Level Characteristics to
1. Putty / Paste / Filler
2. Composite
1. Resin (in cracks)
2. Composite
1. Resin
2. Composite
Structural Response (Interface
Characteristics Will Depend on the Choice
Figure 3: Schematic of Concrete- of Process As in Figure 3
CompositeBond Configurations with
Interphases
In each case the order of constituents likely to be used, in order from concrete surface
outwards, is provided. As seen in the schematics, in most cases local irregularities and
cracks in the concrete itself could be filled with a putty, paste or filler. This is likely to be
a “softer” and more compliant layer vis-à-vis crack propagation and fracture. In the case
of resin infusion, the infusing resin can itself act in fashion similar to crack injection,
filling all cracks, and hence significantly altering the substrate material. It should be
noted that if the surface irregularities are deep or of nonuniform shape/configuration,
some will not be filled completely when an adhesive or primer layer is the first one
brushed or rolled onto the concrete surface, thereby potentially trapping air voids, hollows
or gaps which can substantially affect both durability and fracture behavior. Substantial
fundamental research is still needed in this area, especially as related to effects on
durability in terms of moisture rich zones at the concrete-adhesive/resin interface, which
can cause delamination, or separation through freeze-thaw cycling, or even dramatic
temperature cycling in hot climates (affecting shear transfer efficiency of the adhesive,
emphasized through local discontinuities). In the wet layup and resin infusion methods
discussed above, the function of the adhesive is taken by the resin itself with the bond to
the concrete substrate being formed simultaneously with the fabrication of the composite.
This is both an advantage and a disadvantage, since the elimination of third phase in the
system, the adhesive, results in the formation of fewer interfaces at which failure could
occur, and also eliminates the use of a more compliant layer. The formation of cracks at
these interfaces and/or the changes in properties of the adherents as a function of time and
environment can result in premature failure of these rehabilitation strategies. Although
effects of various phenomena in this regard have been determined experimentally, the set
of results does not provide an overall basis for design guidelines or even the
establishment of a lifetime predictability-based design philosophy. The development of
comprehensive computational simulations that would encompass both structural response
and changing properties of the adhesive layer and interfaces as a function of time and
environment would go a long way in providing simple design procedures and solutions.
A schematic of the levels of integration needed to provide for a comprehensive solution
from the materials to the systems phase is shown in Figure 4.

Significant research has already been conducted on the use of these materials for the
flexural strengthening of beams and design guidelines, although not universally accepted,
have been proposed by a number of investigators and professional organizations. In
general the main points of divergence relate to concepts of required initial strength of the
concrete, development of bond, and level of strengthening to be allowed beyond the
original capacity of the existing structure. In addition, there are still significant concerns
related to the use of these materials as related to the long-term integrity of the adhesive
bond, methods of termination of the composite plates, and the methodology to be used for
either shear- or combined flexural and shear strengthening. It should be noted that at
present most of the analyses follow a fairly straightforward analysis of forces across the
adherends following the general approach of elastic plate approach rather than the non-
linear approaches of based on materials response or on the basis of energetics of fracture
and interphase mechanics. It is envisaged that the extension of the above work to the
plating problem would provide significant insight not only to failure mechanisms but also
to the design of the composite plate/strip and the form of adhesive used.

Notwithstanding the availability of some preliminary guidelines, it is important that


specific aspects related to design philosophy be highlighted herein. Since the
strengthened structural system is inherently dependent on the condition of the substrate
concrete, it is essential that there be sufficient shear transfer capacity in the concrete.
This can most efficiently be determined through the use of the pull-off test (ACI 503R)
using a 50 mm diameter disk. Based on experimental investigations, it is recommended
that a minimum value of 1.4 MPa (200 psi) be used. Any structure showing concrete
pull-off strength lower than this should not be strengthened using just external bonding
since there is a high degree of probability that the composite with a layer of concrete
attached would delaminate from the rest of the structure due to its inherent weakness.
Similarly it is essential that the composite is itself bonded strongly to the concrete so as to
preclude debonding at the composite-concrete interface region. A minimum value of 1.7
MPa (250 psi) is recommended for a pull-off test to ensure that failure will always be in
the concrete (since the composite-concrete pull-off value is higher than that of the
concrete, at minimum). Acknowledging that the external composite can itself be
deteriorated due to heat, impact, environmental influences, and accidental damage, and
that the adhesive/resin bond can also degrade, it is essential that concrete components
being strengthened have a minimum capacity in the unstrengthened state so as to provide
a factor of safety against collapse. In the lack of substantial data to determine this limit it
is recommended that a factor of safety similar to that provided by ACI guidelines for load
testing be used. The ACI load test value of 0.85 [1.4 (Dead Load) + 1.7 (Live Load)] can
be closely approximated by calling for the ultimate strength of the unstrengthened system
to exceed service load level by 20-25%, which would result in an implied factor of safety
of 1.33-1.39 for flexure (using a reduction factor for reinforced concrete as 0.9), and an
implied factor of safety of 1.41-1.47 for shear (using a reduction factor for reinforced
concrete as 0.85). In order to ensure a ductile, rather than catastrophic, failure when the
external strengthening fails in overload it is essential that there exist sufficient steel to
provide yield capacity. One possible criterion is to compare the idealized moment-
curvature plot for the strengthened component with that for a hypothetical member
without external strengthening but reinforced to a level of 0.75 ρbal, where ρbal is the steel
reinforcement ratio which would cause simultaneous failure of concrete in compression
and of steel yielding under flexural load. This condition is analogous to the ACI
requirement for ductile failure of conventional reinforced concrete flexural members.
Further strength reduction factors for the composite itself need to be applied during
design to take into account variability in raw material, process and environmental related
aspects.

As with the use of composites for seismic retrofit of columns, aspects related to design
detailing have perhaps not received as much attention as they should have. A common
example of the misuse of ease of conformance is in the use of composites for the shear
strengthening of I-beams, through the covering of the beam on its sides and bottom by
FRP fabric. As can be noted from detailing practice used for conventional concrete
construction, shear stirrups are not curtailed in the beam region, but are carried over into
the slab section to be anchored in the compression zone and to provide the sought after
truss mechanism. When the composite is curtailed in the vertical section below the slab
there is a distinct possibility that it will debond or delaminate along the vertical edges
above the neutral axis, significantly reducing reliability and safety margins. For good
detailing, the composite should be continued directly into the slab and then anchored in
that region, or at the minimum, to provide ease of practical use, be continued for an
appropriate distance over the horizontal soffit section of the flange so as to provide
sufficient development length for full transfer without fear of peel. Similarly, the
continuation of the composite directly around the bulbs of the lower flange can result in
local peel and debonding along corners. This again is an example of detailing practice at
odds with that used in conventional steel reinforcement. Similar errors are often made in
the application of strips to repair/strengthen slabs. Irrespective of actual material needed,
strips need to be spaced such that local punching shear failure cannot take place in large
unreinforced gaps between the external strips. Correct procedure would be to place the
strips closer together similar to the placement of internal steel rebar, or to use fabric over
the entire area leaving open sufficient space for evaporation of moisture directly from the
concrete and to allow the concrete to “breathe.”

As a general rule, external rehabilitation measures on concrete structures should emulate


conventional reinforcement detailing. In cases where this is not practical, strict
limitations on allowable increase in capacity should be applied. Care should also be
taken to address the durability and efficiency of the resin/adhesive system used to attach
the composite to the concrete surface. Conditions such as moisture entrapment at the
interface between the adhesive and the concrete, or at the composite interface present
challenges for the civil designer as related to the placement of ‘weep’ areas for removal
of moisture and water that seeps through the deck surface, and raises questions related to
long term durability assurance due to the potential for plasticization of the adhesive/resin.
It should be noted that adhesives generically are susceptible to severe deterioration after
combined exposure to moisture and temperature variation. Results of a modified peel test
to determine the effect of environmental exposure on critical energy release rates of
fracture on different wet layup based composite systems used in strengthening of concrete
beams show that of all effects, exposure to sub-zero temperatures (including freeze-thaw)
result in significant changes to values of mode I and II critical energy release rates,
emphasizing the strong effect on the interphase between the concrete and the composite.
Study of the energetics of fracture based on energy release rates provides a good
indication of potential failure sites and of changes in mechanisms based on environmental
exposure. These aspects need to be evaluated further using a fracture-mechanics approach
so as to understand the role played by adhesion between the dissimilar surfaces and for
the characterization of failure and propagation within the entire system.

2.1 Application to Cut-Outs

The use of FRP composites as externally applied reinforcement for the repair and
strengthening of concrete structural elements is very attractive due to the ease of
application of the extremely light additions to gain tremendous increases in load carrying
capacity. This technique is of special interest in enabling the modification of existing
structures through the addition of elevators, escalators, service facilities through the
cutting of holes in existing slabs. In order to provide these facilities, areas of existing
slabs, including the steel reinforcement, are cut necessitating the use of additional
columns and walls to support the weakened slab. FRP composite strips can, however be
used around the cut-out to externally reinforce and slab and carry load locally,
redistributing it back into the remaining structure, without the need for additional
construction. This not only saves space that would otherwise have been needed for
construction of additional supporting members around the cutout but also saves cost.

In a recently completed series of tests, the response of unstrengthened and strengthened


full-scale slabs, of size 6 m x 3.2 m with a central cut-out of size 1 m x 1.6 m, was
investigated. The objective of the test was to externally strengthen the slab, after the
cutout was made, using prefabricated carbon/epoxy pultruded strips that were adhesively
bonded to the tension face of the slab. Equivalent point loads were applied in the
longitudinal and transverse directions on either side of the cut-out in two separate sets of
tests (unstrengthened and strengthened) for transverse and longitudinal loading, to
observe biaxial bending behavior as well as shear behavior in the slab. Strengthening
schemes were based on determination of external reinforcement area required for
rehabilitation through finite element analysis. Prefabricated (pultruded) FRP composite
strips of 1-mm thickness and 100 mm and 50 mm width were used. For loading along the
transverse direction a total of four 50 mm width strips of length 250 cm and two 100 mm
width strips of length 350 cm were placed in the longitudinal direction centered on either
side of the cutout, with one 50 mm width strip placed on either side of the cutout in the
transverse direction over the width of the slab at 25.4 cm from the edge of the cutout.
Figures 5 and 6 show a comparison of results between the unstrengthened and
strengthened slabs, tested in the transverse and longitudinal directions, respectively.

200 250.0
Sika Strengthened
Slab without cutout
160 200.0

Slab without cutout Sika Strengthened


Load (KN)

120

Load (KN)
150.0

Slab with cuto ut


Slab with cutout
80 100.0

40 50.0

0 0.0
0.0 5.0 10.0 15.0 20.0 0.0 5.0 10.0 15.0 20.0 25.0
Displacement (cm) Displacem ent (cm )

Figure 5: Load-displacement Response Figure 6: Load-displacement Response


under Transverse Loading under Longitudinal Loading

As can be seen, the use of the FRP strips results in the slab regaining its initial capacity,
albeit with less deformation capability. A comparison of load envelope and deflection
profile of an unstrengthened slab and one strengthened using the adhesively bonded strips
and loaded in the transverse direction is shown in Figures 7 and 8.
120.0 0
-3 -2 -1 0 1 2 3
100.0 -5

80.0 Scaf olding


Deflection (cm)
Load (KN)

removed
-10 22 KN
60.0
33 KN
40.0 -15 44.5 KN

55 KN
20.0
-20
66 KN
0.0
0.0 5.0 10.0 15.0 20.0 -25
Displacem ent (cm ) Distance from Mid-span (m)

Figure 7a: Load-deformation Envelope Figure 7b: Deflection Profile for the
for an Unstrengthened Slab Loaded in Unstrengthened Slab Loaded in the
the Transverse Direction Transverse Direction
180.0
0
160.0 -3 -2 -1 0 1 2 3
140.0 -2

120.0
Load (KN)

-4

Deflection (cm)
100.0

80.0
Scaffolding
-6 Removed
60.0
22 Kn

40.0 -8 45 Kn

20.0 68 Kn
-10
0.0 100 Kn
0.0 5.0 10.0
-12 164 Kn
Displacement (cm)
Distance from Mid-span (m)
Figure 8a: Load-deformation Envelope Figure 8b: Deflection Profile for the
for an Unstrengthened Slab Loaded in Unstrengthened Slab Loaded in the
the Transverse Direction Transverse Direction
It is noted that in addition to enabling load distribution through the strips resulting in a
recovered load capacity, the FRP composite also restrains crack growth especially in
areas of local stress concentration formed through the construction of the cutout. Figures
9(a) and (b) depict configurations of the final failure mode of strip debonding with the
horizontal crack being either within the cover concrete or within the FRP composite itself.
Although the initiation of failure is approached in fairly linear fashion, there is significant
prior warning through local debonding and concrete cracking. It is also noted that after
debonding of the FRP composite the slab response follows that of the original slab with
the cutout, thereby emphasizing that if appropriately designed, failure would not be
catastrophic.

Figure 9a: Initial Debonding of Strips is Figure 9b: Peeling and Debonding of
at the Corner Strips from the Concrete Substrate

2.2 Application to Rehabilitation of Punching Shear

In a number of cases bridge decks subjected to high traffic loads and having deficiency of
reinforcement in the slabs in one direction show distress in the form of punching shear.
Direction of Rolling Load

Stage 1 Stage 2 Stage 3


Cracks develop Cracks develop Punching failure
along transverse perpendicular to develops along
reinforcement transverse existing flexural
slab strips cracks

Figure 10: Schematic of Slab Failure Model

This form of distress follows the schematic shown in Figure 10, wherein for the case of
deficiency in longitudinal reinforcement, transverse spanning decks first crack in
longitudinal bending along the transverse soffit reinforcement since longitudinally very
little flexural capacity exists and transverse cracks can open wide. Subsequently the
effect width in the transverse direction is limited to the spacing between transverse cracks
resulting in significant flexure and shear overloads. As a consequence, longitudinal
cracks develop which spread along the entire deck length due to the moving nature of the
wheel loads. Finally the local shear capacity is exceeded over a shear area limited by the
existing flexural cracks in both directions.

In an ongoing project, FRP composites are being applied to the soffit of the deck slab of
the Byron Road bridge, a 103 m long 5 span bridge in northern California built in 1964 as
a cast-in-place reinforced concrete T-girder bridge that has longitudinal deck
reinforcement that is only 57% that of the transverse reinforcement and has started
showing advanced signs of distress as shown in Figures 11(a) and (b).

Figure 11(a): Distributed Crack Pattern Figure 11(b): Local Punching Shear
in a Bay Failure
Failure assessment studies were performed using a simple slab demand model for deck
cracking and it was ascertained that maximum moment demand is essentially the same in
both the longitudinal and transverse directions, indicating that the slab was trying to
support wheel loads in both directions, close to the point of load application. Using a
simple punching shear model, damage levels and capacities were assessed and the
following rehabilitation criteria were established:
• To avoid the sequential crack opening, first in the transverse and subsequently
in the longitudinal deck slab directions, the longitudinal deck slab bottom
reinforcement was required to be brought up to a level equivalent to that of the
transverse reinforcement, through the addition of external FRP composite
strips
• To maintain soffit crack widths of less than 1 mm for sufficient aggregate
interlock the bottom soffit strains were required to be limited to less than
0.75%. This also ensures that a nominal concrete shear capacity of more than
2(fc’)1/2 could be assumed.
• To prevent punching shear cones from forming, clear spacing between the
externally applied FRP strips was required to be less than the spacing for the
observed punching shear diameter (300-mm).
Following these criteria, the soffits were rehabilitated using a combination of pultruded
strips that were adhesively bonded to the concrete, and fabric strips which were applied
using the wet layup process directly onto the prepared concrete substrate. In areas where
substantial cracking had already taken place and punching shear distress was deemed to
be imminent, transverse strips at 600-mm spacing were added in addition to the
longitudinal strips placed on all slabs. Examples of the rehabilitation are shown in
Figures 12(a) and (b).

Figure 12(a): Application of Figure 12(b): Application of


Longitudinal and Transverse Strips in a Longitudinal Strips in a Bay with Use of
Bay Bracing for Temporary Support

Rehabilitation was conducted in continuous fashion with traffic being limited to a single
lane so as to decrease effects of vibration and movement since a number of bays were
seen to be at the point of severe distress. Since the rehabilitation was conducted in winter
months during which temperatures were generically lower than those required for
appropriate cure of the adhesive and resin, and since significant rainfall is seen in winter
in the geographical vicinity of the bridge, work was carried out using rolling scaffolds
that were encapsulated in plastic sheets as shown in Figure 13 with hot air being blown in
through an external heater in order to maintain a constant temperature in bays that were
being rehabilitated.
Figure 13: Use of Enclosed Scaffolds with Hot Air Blown in to Reduce Effects of
Varying Environmental Conditions. The Setup Also Warded Against Any Potential
Contamination of the Water in the Aqueduct
2.3 Application to Strengthening of PCCP

Large diameter prestressed cylinder pipe (PCCP) is often used for the transport of water.
These pipes consist of an internal liner of steel and an external layer of prestressing
strands in helical close fashion with significant thickness of concrete between the two
layers. Small thicknesses of concrete mortar separate the inner steel liner from the
flowing water on the inside, and the prestressing wires from the surrounding environment
on the outside. Due to a variety of factors including defective steel used in the
prestressing wires, corrosion of the wires, and loss of prestress, the sections of these pipes
have recently been seeing dramatic failures resulting not just in disruption of water supply
but also significant damage to the surrounding area. Since, in a number of cases, the
PCCP are reaching the end of their initial design lifetimes, there is a critical need to
develop rapid and cost effective methods for the repair of these life-lines. The application
of FRP composites as an inner liner as shown in Figure 14 provides such a solution for
preventive maintenance without reduction in overall cross-section.
}
Dense Cement Mortar
High Tensile Prestressing Wire
High Strength Concrete Core
Existing Pipe
Steel Cylinder
High Strength Concrete

}
Polymer Liner
Composite Laminate
Adhesive / Resin
Retrofit

Figure 14: Schematic of the Rehabilitation Concept

A series of experiments conducted over the past two years has shown that liners of both
glass and carbon fiber reinforced composites can be used to provide capacity in excess of
that provided by the initial prestressing. Figure 15 emphasizes the effect of a 5mm thick
inner layer of carbon/vinylester composite applied after the prestressing strands were cut
to simulate damage. As can be seen, the liner not just enhances capacity, but also
preserves the integrity of the cross-section during loading, thereby ensuring that even
with complete loss of prestress the PCCP would not collapse under the effects of
overlying soil/earth pressure. It is noted that the application of the FRP composite can be
conducted rapidly either through the use of the wet layup process or through the adhesive
bonding of prefabricated strips or sections of FRP composite. After completion of the
process of application a coating of a polyurethane or appropriate epoxy can be applied to
the FRP composite liner to serve as a barrier for moisture diffusion and as a protective
layer against abrasive damage from floating debris and/or sand within the water flow
itself.

500 0.52" _ dR=0.43"

400 0_
0.
23. dR=0.24"
Load (kips)

300
45.

Load (kips)
200 dR=0.39"
0
100 68.

100 200
Repair (Prediction) 300
Repair (Test) 400
As-Built (Test) Original Radius = 60"
90.
0 451
I I
0 0.1 0.2 0.3 0.4 0.5 0 0.53"
Displacement (in)

Figure 15(a): Comparison of Load Figure 15(b): Deflection Profile of a


Capacity of Repaired Section (With Quadrant of the Rehabilitated PCCP
Prestress Cut) With the As-Built Section
3. DURABILITY
Since most structures used in civil infrastructure are expected to perform as designed over
long periods of time, there is some concern related to the durability of FRP composites
considered for these applications. Civil applications, at present, are more likely to (a) use
processes such as wet lay-up, pultrusion and resin infusion than autoclave molding, (b)
fiber and resin as separate constituents rather than in the form of preimpregnated material,
and (c) resin systems such as polyesters, vinylesters, phenolics and lower temperature
cure epoxies rather than the higher temperature curable epoxies and thermoplastics.
Further, there is likely to be extensive use of processes under ambient conditions in the
field, rather than fabrication in factory controlled environments. Thus, the civil
engineering environment not only brings with it new challenges for the control of quality
and uniformity of composites, but also makes it difficult (if not impossible) to use the
well established databases generated by DoD sponsored research (such as those for
AS4/3501-6 or T300/5208 based systems). Environmental conditions likely to be faced
by these composites range from extremes in temperature (heat as in Saudi Arabia, cold as
in Alaska), significant variation in humidity levels, potential for routine submersion or
immersion (such as for structures in flood plains), exposure to chemicals, including road
salt, and widely varying load conditions. It is important that these issues be kept in mind
while selecting materials and processes for the fabrication of FRP composites for civil
infrastructure.

The use of processes such as wet lay-up, pultrusion, wet winding, RTM, and resin
infusion require that the resins have a sufficiently low viscosity so as to be able to infuse
the reinforcing structure of fibers cost-effectively. Both polyesters and vinylesters are
likely to be widely used with preference being given to vinylesters because of their
greater durability. The common vinylesters used to date in applications such as bridge
decks, rebar, prefabricated plating, and profiles belong to the generic classification of
thermosetting resins and are comprised of low molecular weight polyhydroxyether chains
with reactive groups at chain ends. Styrene in monomeric form is used as a diluent in the
resin in quantities between 20-60%. It is important to note, however, that the increase in
styrene content results in (a) an increase in hydrophobicity, thereby effectively decreasing
the level of moisture absorption, and (b) an increase in shrinkage to levels of 5-10% by
volume which can result in significant microcracking in resin rich areas and high residual
stresses in composites having high volume fractions. In comparison to polyesters, which
have double bonds at about 250 g/mol level, vinylesters have reactive double bonds at
about every 500-1000 g/mol. The increased distance between cross-linkages in the
polymer results in a network that has greater fracture toughness. It is important to note
that irrespective of the cure mechanism used, vinylesters do not completely polymerize,
generally reaching a level of cure higher than 95%, with the last part of cure continuing
very slowly. Incomplete cure can result due to environmental conditions, incorrect
stoichiometry of resin system components, or the failure to reach a sufficient temperature
of cure. This state can affect mechanical properties, moisture absorption, and
susceptibility to moisture induced degradation of the resin, and the fiber-matrix
interphase. Figure 16 schematically depicts the relation between degree of cure (or cross-
link density) and specific characteristics in these materials.
Tensile Modulus
Fatigue
Resistance

Performance Characteristic
Tensile Strength

Hardness

Degree of Cure
Figure 16. Schematic Showing Effect of Degree of Cure on Material Characteristics

The aspect related to change in properties with time is important not just from a
performance point of view for thermosetting resin based composites, but especially for
composites used in civil infrastructure which in a large number of cases will be cured
under ambient conditions and then either post cured over a prolonged exposure to sun
light and ambient conditions, or through a short period of exposure to elevated
temperatures.

Although carbon fibers would generically be preferred on the basis of their inertness to
most environmental conditions, cost and performance considerations (such as the need for
higher levels of strain to failure, coefficients of thermal expansion in composite form to
match that of concrete etc.) entail that E-glass fibers will be used in a majority of current
applications. It should be noted further that current sizings/finishes on carbon fibers are
not compatible with vinylesters and polyesters, especially when loads are to be carried in
compression or shear. Although fiber-resin compatibility is not as great an issue with E-
glass fibers, the susceptibility of E-glass to moisture and alkali induced damage creates
the necessity for the appropriate selection of the resin system to serve not just as a binder
but also as a protective layer that would reduce diffusion and UV induced degradation.
Due to the high potential for use in the field of processes such as wet lay-up and resin
infusion in the presence of moisture, the use of anhydride cured epoxies would not be
recommended.

Although changes in glass transition temperature due to moisture absorption are


considered in other application areas, the effect is perhaps more striking in the civil
infrastructure area due to the low initial Tg itself (due to the predominance of ambient
cure, especially in applications involving structural rehabilitation). It should be noted that
although the Tg has been shown to recover after drying, the actual situation in a large
number of geographical areas is that a high level of humidity and/or moisture will exist
continuously with very little chance of the composite decreasing in moisture content.
Further, effects of continuous load on structural elements are likely to result in
exacerbating the overall detrimental effect.

In a large number of cases, the fiber reinforced composite component will be used in
contact with or adjacent to a concrete substrate. In such cases one must consider that
concrete is a porous and chemically active material with pH of pore water being as high
as 13.5. Existing concrete can also contain high levels of chlorides, carbonates, and
sulfates, all of which can be brought to the concrete-composite interface through moisture
that diffuses through the concrete. Figure 17 emphasizes the various levels of interaction
that need to be considered in using composites in applications such as external
strengthening, wherein a relatively thin composite layer is adhesively bonded to the
surface of "active" concrete.

Internal Influences
• Chemical activity
• Electrochemical activity
• Alkali content and pH level
• Stress
• Moisture infiltration
• Transport of solutions, salts...

Interfacial Influences
• Moisture entrappment
External Influences • Moisture diffusion
• Selective transport of chemicals
• Humidity
• Thermal and elastic mismatch
• Moisture
• Temperature
• Temperature Cycling (daily, seasonal, annual)
• Aggressive natural and manmade agents
• UV
• Oxygen (related to steel)

Figure 17: Systems Level Interactions

Interactions at these levels need to be carefully considered since degradation of any one
of the elements, FRP composite, bond-line, and concrete could result in premature failure
of the entire systems. Environments of specific concern in this regard are freeze-thaw
and effects of solutions migrating through the concrete to the surface of the adhesive or
resin layer between the FRP composite and the concrete.

4. CHALLENGES AND OPPORTUNITIES


As the use of FRP composites in civil infrastructure rehabilitation increases there is a
need to optimize materials, that heretofore had been directly derived from commercial
systems initially developed for other application areas such as aerospace, marine or
automotive, for use in this new applications area. Chief amongst these aspects is the need
for development of materials systems and product forms that lend themselves to use in a
varying and harsh environment wherein quality control attributes cannot be expected to
reach levels expected in the aerospace industry and where routine maintenance and
inspection of the type used on FRP composites in the aerospace industry will be rare.

Since the civil infrastructure area is likely to be largely driven by cost constraints there is
some reluctance to use the higher cost carbon fibers instead of the lower cost glass fibers.
Carbon fibers do have significant higher levels of stiffness with the standard varieties
having a modulus equivalent to structural steel and the more advanced varieties having
moduli 3-5 times that of structural steel. However, the levels of strain-to-failure are fairly
low and infact the levels guaranteed by glass fibers are significantly higher. This aspect f
glass fibers is an advantage as is the higher impact tolerance and general damage
tolerance. There is thus the need for the development of newer types of fibers that have
the stiffness and the inertness of the carbon fiber, but with the higher strain-to-failure,
impact- and damage tolerance of the glass fiber (in FRP composite form) at a cost equal
to or lower than that of the glass fiber. Similarly, with the increasing use of field
fabrication and of processes such as resin infusion and pultrusion there is a need for the
development of low viscosity, stable, moisture-resistant resin systems that are not only
easy to process and cure under ambient or moderately elevated (akin to a heated room)
temperatures, but which also have excellent resistance to moisture both during cure and
after formation of the composite, with fire-resistant properties similar to or exceeding
those of phenolics.

With the increasing success in the application of FRP composites in the field through
demonstration projects and through slow penetration in the commercial market, the use of
FRP composites by conventional large- and small-contracting firms is likely to increase.
The inherent tailorability of FRP composites, through various permutations of raw
materials, fabric forms and processing techniques, presents a concern vis-à-vis quality
control and specification. The prepreg materials form decreases these concerns to an
extent through the use of a well-controlled and factor-certified materials system
analogous to that of a grade of steel in plate form. However, the additional cost of the
prepregging operation is initself prohibitive, in addition to the increased costs associated
with conventional autoclave cure methods of processing. There is need for the
development of lower cost prepreg forms that can be cured through non-autoclave, lower
temperature processes, as well as for tailored materials systems that can be “snap-cured”
through the use of techniques such as radiation bursts (E-beam and microwave sources),
ultrasonic scans, and even progressive introduction of catalysts in pressure sensitive films
such that the system and the catalyst can be applied in sheet form and then reacted
through application of roller derived pressure.

There is significant ongoing research into the determination of durability of FRP


composites for use in civil structures. However, due to the use of lower-cost newer
systems there is a lack of complete knowledge about long-term durability in a civil
engineering environment. In fact a recently completed HITEC-CERF/MDA study
concluded that although there is significant anecdotal evidence to suggest that existing
FRP composites could be used for a range of applications in the renewal of civil
infrastructure, there are still a large number of critical gaps that need to be addressed,
especially as related to long-term environmental durability under varying environments.
Application Area Continuous Intermittent Synergistic
Exposure Exposure Effects
A B A B A B
Int. Reinforcement
Rebar 6 6 8
Ext. Reinforcement
Beams 10 10 10 8 10 10
Slabs 10 10 10 10 10 10
Columns 10 10 10 10 10 10
Seismic Retrofit
Columns, piers 8 8 6 6 8 8
Walls 8 10 6 8 10 10
Deck Systems
Conventional beams/girders 10 8 6 6 10 8
Integral/composite 10 10 6 8 10 10
beams/girders
Structural Elements
Wall panels, profiles 6 6 8

Table 1: Overall Ranking of Gaps for Moisture/Solution Effects

Application Area Sustained Pure Fatigue Fatigue Fatigue & Fatigue &
Stress Loading And Temp. Moisture / Creep
Loading Salt
A B C A B C A B C A B C A B C
Internal Reinforcement
Rebar 6 8 4 7 8 4 10 10 2 8 10 2 10 10 2
External Reinforcement
Beams 5 7 3 6 9 6 10 10 2 10 10 2 10 10 2
Slabs 5 7 3 5 9 6 10 10 2 10 10 2 10 10 2
Columns 7 9 4 7 9 4 10 10 2 10 10 2 10 10 2
Seismic Retrofit
Columns, piers 6 9 5 7 8 4 8 10 2 8 10 2 8 10 2
Shear Walls 6 9 5 5 8 4 8 10 2 8 10 2 8 10 2
Deck Systems
Conventional 6 9 9 9 10 10 2 10 10 2 10 10 2
beams/girders
Integral/composite 6 9 6 9 10 10 2 10 10 2 10 10 2
beams/girders
Structural Elements
Wall panels, profiles 5 9 5 8 9 5 10 10 2 10 10 2 10 10 2

Table 2: Overall Priority of Data for Fatigue Effects

Tables 1 and 2 provide examples of gap analysis for effects of moisture-solution and
fatigue, respectively, wherein a ranking of 10 denotes an area in which there is a critical
need for data. In both the tables “A” denotes effects on the FRP composite itself, “B” is
the effect at the interphase (such as the adhesive) level, and “C” denotes the effect on the
substrate. Keeping in mind the need for consideration of the original structure and the
rehabilitation scheme as a complete system, it is emphasized that further research is
needed at specific levels of materials, systems, and field implementation focussing on
combined effects of stress and environmental exposure. A schematic of needs from the
constituent to the applications levels is shown in Figure 18.
Material Level Model

• Statistics of Failure
• Type of Fiber • Creep
• Environmental susceptibility • Relaxation
• Surface Energy Profile
• Sizing/Finish Type
Time-dependent
Fiber Response

• Gradients
• Residual stresses
• Bond efficiency
• Environmental susceptibility

Interphase Processing Life-Cycle


Method Predictions

• Cure temperature
• Degree of consolidation
• Method of infusion
Resin
Environmental
Degradation
• Stoichiometry
• Cure mechanisms
• Environmental susceptibility • Susceptibility to environment
• Moisture diffusion
• UV, ozone degradation
• Resin based changes

Figure 18: Schematic for Overall Evaluation of Durability

5. SUMMARY

FRP composites are rapidly being accepted as viable alternatives to conventional


materials for the rehabilitation (retrofit, repair and/or strengthening) of civil
infrastructure, and in fact provide the means to renew/extend the life of structures for
which life-extension without significant new construction would not have been possible
with conventional materials. Increasing efforts are being made to develop design and
construction approaches, guidelines, and specifications for the use of these materials.
However, aspects related to education and instruction across the boundaries of civil
engineering (as related to design requirements and environmental considerations) and
materials engineering (vis-à-vis the optimization of materials systems and consideration
of effects of the environment on performance thereof) still need to be further encouraged.
Till this cross-fertilization is assured, FRP composites will provide an attractive, but not
widely accepted, alternative to conventional materials. Notwithstanding this challenge
which has to be considered at levels of both university level and continuing education and
training, it is likely that these materials will provide civil engineers with the means to not
only extend the functionality of our critical life-lines, but will also enable the restoration
of historical structures, and the construction of new structural systems that combine form
and function at levels only dreamt of so far.

ACKNOWLEDGEMENTS
The author acknowledges the thought-provoking discussions with a number of his
colleagues and students at UCSD, and the interactions with Prof. Protasio Castro initiated
during Prof. Castro’s sabbatical in 1994.

You might also like