You are on page 1of 11

Construction and Building Materials 31 (2012) 94–104

Contents lists available at SciVerse ScienceDirect

Construction and Building Materials


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

Review

Towards a durability framework for structural elements and structures made of


or strengthened with high-performance fibre-reinforced composites
Viktor Mechtcherine ⇑
Institute of Construction Materials, Technische Universität Dresden, Helmholtzstrasse 10, 01069 Dresden, Germany

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

Article history: High-performance fibre-reinforced cement-based composites (HPFRCCs) have great potential for applica-
Received 18 August 2011 tion in structures exposed to severe mechanical or environmental loading. This paper presents an over-
Received in revised form 5 December 2011 view of the current knowledge on the durability of two prominent representatives of this new group of
Accepted 23 December 2011
concrete materials: strain-hardening cement-based composites (SHCCs) and textile-reinforced concrete
Available online 21 January 2012
(TRC). Hereby not only the durability of these composites and their components are considered but spe-
cial focus is centred on the protection of steel reinforcement when it is used in combination with SHCC, as
Keywords:
in the case of R/SHCC structures or RC structures repaired with a layer of SHCC, or TRC, as in the case of
Strain-hardening cement-based composites
Textile reinforced concrete
the strengthening or repair of RC structures using TRC.
Durability concept Based on such considerations, transport properties in the cracked state, long term strain capacity, and
Multiple cracking resistance to aggressive environments have been identified as critical parameters. Current knowledge
Transport properties indicates that HPFRCCs show high long-term strain capacity and favourable transport properties in the
Corrosion resistance cracked state. The composites also exhibit superior resistance to aggressive environments, as compared
to ordinary concrete. However, there is little information available on the effects of aggressive environ-
ments on the mechanical properties of the materials.
Since SHCC and TRC are new materials, there is no information available on their long-term perfor-
mance in the field. To be able to utilise the superior qualities of these new materials fully, it will be nec-
essary to develop a realistic and reliable performance-based durability design concept for structures
made of or strengthened with high-performance fibre-reinforced cement-based composites.
Ó 2011 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
2. Characteristic loads and exposures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
3. Basics of durability design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
3.1. General remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
3.2. Protection of steel reinforcement from corrosion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
3.3. Durability of the matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
3.4. Fibre durability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
3.5. Fibre–matrix bond durability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
4. Characteristic material properties to predict long-term durability and service life. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
4.1. General remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
4.2. Transport properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
4.2.1. Air permeability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
4.2.2. Water-transport properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
4.2.3. Chloride migration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
4.2.4. Other transport properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
4.3. Strain capacity of HPFRCCs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
4.3.1. Long-term strain capacity and ageing behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

⇑ Tel.: +49 351/4 63 3 63 11; fax: +49 351/4 63 3 72 68.


E-mail address: mechtcherine@tu-dresden.de

0950-0618/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.conbuildmat.2011.12.072
V. Mechtcherine / Construction and Building Materials 31 (2012) 94–104 95

4.3.2. Cyclic loading and fatigue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101


4.3.3. Behaviour under sustained loads. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
4.4. Self-healing of cracks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
5. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

1. Introduction during transport and mounting is an accomplished fact. However,


considerable laboratory research has been performed in the last
Adding fibres to cement-based construction materials is not few years on the behaviour of structural members made of SHCC,
new. The first patent for steel fibre-reinforced concrete dates from and the number of structural applications is increasing. There are
the year 1872, and many experiments were undertaken with dif- also a number of examples of the successful use of TRC for
ferent types of fibres in the 20th century. However, the first conclu- strengthening existing RC structures. Based on these first insights,
sive, scientifically substantiated concepts specifically related to the some major tendencies may be deduced already.
design of fibre-reinforced concrete composite materials have only From the current point of view, fibre reinforcement is to be
been developed in the last few years. In combination with modern combined with conventional steel reinforcement in the majority
construction material analysis and process technology, these ap- of structural elements made of SHCC (R/SHCC), cf. Fig. 1a. Thus,
proaches make the target-oriented development of high-perfor- the protection of steel from corrosion must be considered with re-
mance, fibre-reinforced, cement-based composites (HPFRCCs) gard to the durability of such structures [5]. This also holds true for
possible. Using these new types of concrete could revolutionise the repair of reinforced concrete structures using SHCC, cf. Fig. 1b,
the planning, development, dimensioning, design, and construc- as well as their strengthening or repair by means of TRC, cf. Fig. 1c.
tion as well as the renovation of building structures in special areas Fig. 1c illustrates also the case when a TRC layer is used to build an
of application such as structures exposed to severe mechanical or integrated formwork.
environmental loading. This article focuses on the durability of As an example, a practical application for the strengthening of
structural elements made of or strengthened with two prominent existing structures using textile reinforced concrete should be
representatives of high-performance fibre-reinforced cementitious presented here briefly. This measure was carried out in 2006
materials: strain-hardening cement-based composites (SHCCs) and for the retrofitting of an RC roof shell at the University of Ap-
textile reinforced concrete (TRC). plied Sciences in Schweinfurt, Germany. The project was carried
SHCC exhibit strain-hardening, quasi-ductile behaviour due to out with the technical support of the Collaborative Research
the bridging of fine multiple cracks by short, well-distributed fibres Centre SFB 528 ‘‘Textile Reinforcement for Structural Strengthen-
usually made of polyvinylalcohol (PVA) or polyethylene (PE). The ing and Retrofitting’’ of the TU Dresden. The 80 mm-thick RC
favourable mechanical properties of this material offer many pos- hyparshell has a lateral length of approximately 27 m and a
sible applications in both new structures and in the strengthening maximum span of approximately 39 m, cf. Fig. 2. Large deforma-
and repair of structural elements made of reinforced concrete or tions (up to 200 mm drop) in the shell’s cantilevered wings had
other traditional materials [1,2]. been measured; the design tensile stress levels in the upper steel
TRC is a composite construction material in which textile rein- reinforcement layer of the shell had been exceeded significantly,
forcement consisting of polymer, glass, or carbon fibres is embed- such that strengthening measures were absolutely necessary. A
ded into a fine concrete matrix. Since textile reinforcement TRC-layer 15 mm thick, containing three layers of textile sheets
generally needs no minimum concrete covering as a protection made of carbon fibre was applied to these particular areas in or-
against corrosion, it is possible to manufacture thin-walled con- der to increase the load-carrying capacity and to prevent further
struction components and layers. The fibre material can be increase in deformations [6]. A few questions arise with regard
exploited very effectively by adjusting the volume proportion to this strengthening measure. How durable is the strengthening
and orientation of the long fibres in the direction of tensile stress. layer? Or, how well will this layer protect the reinforcement in
Textile reinforced concrete can be employed in the manufacture of the hyparshell from corrosion?
thin-walled façade elements, load-bearing integrated formwork, Depending on the target applications, durability requirements
tunnel linings, or in the strengthening of existing reinforced con- for different structures made of SHCC and TRC may vary signifi-
crete (RC) structures [3,4]. cantly. To meet these requirements best, the material composition
Most current applications of HPFRCCs are non-structural. Fibre of these composites may, in turn, vary considerably as well. This
reinforcement is primarily used in controlling shrinkage cracks as implies that a general framework for dealing with the problem of
well as cracks induced by temperature changes. Furthermore, the durability of such structural elements and structures is needed.
control of cracks in elements of SHCC or TRC due to overloading This paper is an attempt to create such a framework.

Fig. 1. Considered combinations of HPFRCC and steel reinforcement or RC structures.


96 V. Mechtcherine / Construction and Building Materials 31 (2012) 94–104

Fig. 2. (a) Damaged steel reinforced concrete hyparshell at the FH Schweinfurt (Germany), and (b) strengthening with TRC [6].
Chloride concentration C(x,t)

C(x,t) In most cases, cracking is inevitable in concrete structures and


µ=0 is generally essential for the effective use of reinforcement, both
µ=1
by steel bars and fibres. A great advantage of HPFRCCs is that this
group of materials display self-controlled crack width, which is
very small in comparison to conventionally reinforced structures.
This enables, at least theoretically, the full decoupling of the steel
reinforcement from crack control. Steel reinforcement is used only
to ensure adequate load capacity, while SHCC or TRC controls
f(C(x,t)) cracking behaviour. TRC can be simultaneously used for both pur-
poses: increasing the load-carrying capacity of structural members
Depth x and the related crack control. These constellations open new
opportunities for structural design, thereby allowing a more effec-
R/SHCC tive use of such hybrid reinforcements.

Cl¯
3. Basics of durability design

3.1. General remarks


sea cover depth
water While discussing the basics of durability design for SHCC struc-
tures, reference is made to existing durability design rules for
Fig. 3. Fuzzy random profile of chloride concentration for a steel-reinforced SHCC steel-reinforced concrete structures. The prescriptive design rules
member exposed to chloride ions (courtesy of F. Altmann).
of current standards are unlikely to be suitable for taking full
advantage of the new class of materials discussed in this report.
A performance-based durability design approach will likely be
2. Characteristic loads and exposures much more appropriate. While such an approach exists for crack-
free ordinary concrete (DuraCrete 2000 [9], fib bulletin 34 [10],
The wide variety of potential applications for HPFRCCs means Schießl and Gehlen [11]); significant efforts will be required to de-
that the range of mechanical and environmental loads, as well as velop a similar approach suitable for cracked and crack-free SHCC
their possible combinations, may be very broad. Particular perfor- and TRC.
mance requirements may vary even for different parts of one struc- Recently Altmann et al. [12] and Altmann and Mechtcherine
tural member, cf. for example [7]. [13] took a first step towards such a design concept by developing
In considering durability of structures, the mechanical loads are a fuzzy-probabilistic approach to forecast chloride ingress into
of particular interest with regard to cracking since cracks dramat- SHCC, cf. Fig. 3. This approach allows the quantification of param-
ically increase the transport of fluids and gases in concrete and eters based on limited data and expert knowledge. More research
consequently the corrosion process. Apart from the mode of load- will be required to develop and expand this approach further until
ing, e.g., tension, shear, and/or bending, all of which affect the crack it allows a reliable durability design for cracked and crack-free
pattern and the crack width, the loading type should be considered SHCC under environmental exposures. Only then will it be possible
since, for instance, single overloading, fatigue loading, or high sus- fully to utilise the inherent durability of SHCC for the construction
tained loading can affect the cracking behaviour of HPFRCCs (crack of more sustainable structures. The same approach is followed by
width and crack distribution); see, for example, [8]. the author with regard to the durability of TRC as well as RC struc-
However, crack formation and propagation often begin before tures strengthened by TRC.
any crack-inducing mechanical loads are applied. Cracks resulting In order to ensure structural durability, the durability of the ap-
from thermal movements, restrained drying shrinkage, or autoge- plied materials should first be assured. Since we are considering a
nous shrinkage can eventually be exacerbated by fatigue loading, combination of steel reinforcement and SHCC or RC and SHCC or
for instance. These load combinations for R/SHCC, RC/SHCC, or TRC here, the following requirements arise with regard to the dura-
RC/TRC structures are generally the same as for steel-reinforced bility of individual system components and consequently of the
concrete structures. system as a whole:
V. Mechtcherine / Construction and Building Materials 31 (2012) 94–104 97

(1) Protection of the steel reinforcement from corrosion. concrete to spalling. For ordinary concrete spalling resistance
(2) Durability of the SHCC or TRC matrix. increases with increasing thickness of the concrete layer (Mül-
(3) Fibre durability. ler et al. [18]). In the case of SHCC, a much higher spalling resis-
(4) Durability of the fibre–matrix bond properties. tance than that of ordinary concrete can be achieved with only a
very thin overlay.
3.2. Protection of steel reinforcement from corrosion – The lack of coarse aggregates means that there are no large aggre-
gate–matrix interfaces with higher porosity than aggregate-free,
Ordinary steel reinforcement requires anti-corrosion protection, hardened cement paste, which might serve as ‘‘highways’’ for the
which the alkaline environment of the cementitious matrix pro- transport of fluids and gases. However, due to their higher binder
vides. If aggressive agents such as chloride or carbon dioxide pen- content, the overall porosity of SHCC or TRC may be more ele-
etrate the matrix and reach the reinforcement, however, the steel vated than that of ordinary concrete. However, this depends on
may corrode. In established codes and regulations (e.g., Eurocode the mix’s water-to-binder ratio, which is usually relatively low
2, DIN EN 1045), three measures are demanded to protect steel for HPFRCCs. To which extent the presence of short fibres influ-
reinforcement: limitation of the maximum crack width, sufficient ences the transport properties must still be investigated in more
concrete cover, and appropriate concrete quality as defined by cor- detail. However, for TRC it was shown that multifilament yarns
responding threshold values (minimum compressive strength, promote water transport actively [15].
maximum water-to-cement ratio, minimum cement content). In – SHCC and TRC usually possess higher binder content than ordin-
principle, these requirements also hold true for R/SHCC structures ary concrete. The associated higher ability to bind aggressive
or RC structures with a repair or strengthening layer made of SHCC substances such as chlorides and carbon dioxide is advanta-
or TRC; however, the weighting and the thresholds values change geous with regard to preventing steel corrosion.
accordingly.
Li and Stang [14] suggested two levels of protection of struc- In conclusion, it can be stated that for the same exposure classes
tural elements with regard to the steel corrosion. The first level re- (e.g., XC carbonation, XD de-icing salt, or XS seawater according to
lies on tight crack-width control of the SHCC, i.e., <100 lm, which CEN EN 206-1) the threshold values for SHCC or TRC would prob-
delays the penetration of aggressive agents in their movement to- ably differ significantly from the corresponding values for ordinary
ward the steel reinforcement. According to Mechtcherine and Lie- concrete if the established approach for durability design would be
bold [15] crack widths in TRC specimens under tension remain applied to R/SHCC, RC/SHCC or RC/TRC structures.
clearly below 100 lm as well in the operating strain range up to
0.5% at least. Hiraishi et al. [16] showed that the corrosion rate in 3.3. Durability of the matrix
an R/SHCC beam is significantly lower than that of a reference or-
dinary RC beam preloaded to the same level. The second level of Frost attacks, chemical attacks by aggressive substances, and, in
protection afforded by SHCC is the prevention of radial crack for- some cases, abrasion must all be considered with regard to the
mation, even if the steel rebar corrodes and expands. This anti- durability of the SHCC or TRC matrix. The basic effects of matrix
spall ability of SHCC has been demonstrated in experiments by composition on concrete resistance against such exposure are rea-
Kanda et al. [17]. sonably well known for ordinary concrete. However, they cannot
Comparing this approach to existing codes, the following obser- be simply extrapolated to predict the behaviour of SHCC or TRC be-
vations can be made: cause of their very particular mixture composition, i.e., the absence
of coarse aggregates, the high binder content, and the presence of
– Intrinsically small crack widths in R/SHCC, RC/SHCC or RC/TRC fibres. The fine, well distributed fibres might hinder the deteriora-
structures are much smaller than the threshold values allowed tion of the composite even if the matrix is severely loaded. The first
by codes for RC structures. The question arises as to whether or results on the durability of SHCC under frost and chemical attack
not adjustments should be made as regards the other two pro- were characterised by sufficient resistance of the matrix [5] if it
tection measures of the steel reinforcement, namely the thick- has an appropriate composition, especially concerning the water-
ness of the concrete cover, and concrete quality. to-binder ratio and the binder type.
– The requirement of a particular minimum value of concrete As an example, a study by Brüdern and Mechtcherine [19]
cover is based on numerous considerations, including the car- should be cited here. The authors showed that for the same
bonation of concrete, the transport of aggressive agents through water-to-binder ratio of 0.3 an SHCC with a very high fly ash con-
the concrete layer to steel reinforcement, and resistance of the tent (fly ash/binder = 0.70) suffered much more damage after 28

(a) (b)

20 mm 20 mm

Fig. 4. Condition of specimen surfaces for two different SHCC compositions after completion of the CDF-test: (a) SHCC with a very high fly ash content (fly ash/binder = 0.70),
and (b) SHCC with a more moderate fly ash content (fly ash/binder = 0.55), adapted from [19].
98 V. Mechtcherine / Construction and Building Materials 31 (2012) 94–104

freeze–thaw cycles (mass loss about 1400 g/m2, cf. also Fig. 4a) is very expensive, PVA fibre has been used in much research work
than an SHCC with a more moderate fly ash content (fly ash/bin- on SHCC as well as in most applications.
der = 0.55, mass loss of approximately 500 g/m2). The mass loss Eternit Corp. began delivering PVA fibre-reinforced roofing
could be decreased further to a level clearly below 200 g/m2 by slates in 1983. This experience has provided reliable data on the
the addition of 2 kg of superabsorbent polymers (SAPs) per cubic longevity of PVA fibres [22]. Furthermore, it was discovered that
meter SHCC as well as some extra water (in order to retain work- from the determination of tensile strength, crystallinity, and
ability) to the mix with fly ash/binder ratio 0.55 (cf. also Fig. 4b). molecular weight, no sign of degradation could be observed among
The positive effect was attained here due to the well distributed the fibres, which were extracted from fibre cement sheets exposed
system of SAP voids, which act similarly as entrained air voids. to outdoor weathering [23].
Fig. 4 shows conditions of the specimen surfaces after completion From accelerated ageing tests in cementitious alkaline water
of the CDF-test [20]. The difference in the maximum loss is clearly environments, PVA fibre tensile strength is predicted to be main-
visible here. However, it should be noted that due to the anchoring tained at more than 95% over 100 years [24]. Apart from investiga-
function of the fibres, not all the degraded material could be sepa- tions on their behaviour in highly alkaline environments, nothing
rated from the specimen during the cleaning of the test surface in is known on the chemical resistance of PVA fibres in aggressive
an ultrasonic bath. environments. However, PVA materials are generally considered
The investigations performed by the group around the author to be resistant against chemical attacks. Therefore, it can be rea-
into the freeze–thaw resistance of TRC showed similar tendencies, sonably assumed that PVA fibre itself has good durability in ordin-
also basically known for ordinary concrete, i.e., the frost resistance ary chemical environments.
of TRC decreases with increasing substitution of Portland cement Alkali-resistant (AR) glass fibre and carbon fibre are materials
by fly ash and particularly blast furnace cement. used most often in the manufacture of TRC. There is also research
With regard to abrasion the results of laboratory experiments on the use of polymer and natural fibre for this purpose. Carbon fi-
are contradictory, with the suitability of the matrix and thus the bre is well known for its high chemical resistance; no degradation
composite SHCC for heavy roadway traffic depending not only on of properties could be observed due to usual chemical environ-
the mixture composition but also on the test method. Nonetheless, ments. The situation with fibre made of AR-glass is however much
the results in field applications are encouraging. Furthermore, it more complex and should be addressed here in more detail.
was found that the mechanical performance of SHCC remained Standard AR-glass contains 16–20% zirconium dioxide by mass
practically unaltered [21] on being exposed to frost attack. How- and thus shows significantly enhanced resistance in highly alkaline
ever, a chemical attack might very well affect the mechanical prop- environments as compared to conventional glasses. The corrosive
erties of the matrix, which would then lead to a change in SHCC or damage of AR-fibre occurs non-uniformly due to different element
TRC strength and, especially, in their ductility. A substantial concentrations near the filament surface. This deterioration in-
amount of research is still required to understand the behaviour creases with increasing pH value of the pore solution and with
of these materials under various exposures and to confirm statisti- temperature [25]. Organic polymer sizes applied to filament sur-
cally observed characteristic material performance. faces during the production of glass yarns can delay such corrosion
Other issues are the ageing process and the associated hydra- significantly. Scheffler et al. [26] showed that the corrosion of AR-
tion, altering the mechanical performance of HPFRCCs by changing fibres in a cement solution in accelerated tests is characterised by
the composition of the hardened binder paste phases, and the the formation of holes of different size and depth, cf. Fig. 5. Accord-
microstructure and the properties of the bond between matrix ing to Butler et al. [27] flaws in the size and inconsistencies in the
and fibres (Section 3.5). composition of the bulk glass as well as high pH values of the pore
solution are necessary prerequisites for local damage to fibre made
of AR-glass. Corrosion of glass filaments embedded in fine-grained
3.4. Fibre durability
concrete could be observed only as rare exceptions. The authors
conclude that stress corrosion (delayed failure) at nanoscopic de-
In principle different types of fibre can be used for the produc-
fects on the filament surface becomes an important mechanism
tion of SHCC. As yet the best results have been obtained by using
for damage if, at least at some areas of filament surface, the size
high-density polyethylene (HDPE) and PVA fibres. Since HDPE fibre
is widely removed. However, in the multifilament yarns used in
TRC the failure of individual filaments due to corrosion can likely
be compensated by the redistribution of the load to neighbouring
filaments. This means that the failure of a few filaments cannot af-
fect the load-bearing capacity of the multi-filament yarn signifi-
cantly [27]. Hence, other mechanisms seem to be additionally
responsible for pronounced ageing behaviour of TRC made with
AR-glass.

3.5. Fibre–matrix bond durability

The effects of ageing and environmental exposure to hot and


humid environments on the interfacial properties of SHCC have
been examined using hot water immersion [28]. It was found that
the chemical fibre–matrix bond increased while the apparent fibre
strength decreased. This reduces the overall strain capacity of the
composite, which nonetheless generally remains at a level high en-
ough to ensure the desired performance.
For TRC with textiles made of AR-glass it was found [29] that
the extent of the observed performance losses with increasing age-
Fig. 5. Formation of holes on the surface of an AR-glass fibre after 8-days’ storage in ing duration depends primarily on the matrix’s alkalinity, which
cement solution at 80 °C [26]. influences the formation of solid phases in the fibre–matrix
V. Mechtcherine / Construction and Building Materials 31 (2012) 94–104 99

Fig. 6. ESEM images of AR-glass filaments in matrix made of (a) Portland cement (here after 28 days of accelerated ageing, and (b) blast furnace cement and pozzolana (here
after 360 days of accelerated ageing).

interface and on the filament surface. The thickness and brittleness


of these shell-like crusts increase with increasing potential for
portlandite formation in the matrices. Also amplified precipitation
of Ca(OH)2 between the filaments can be observed with increasing
alkalinity of the matrix, cf. Fig. 6a. These filament incrustations re-
duce the slip of filaments in the vicinity of the cracks, so that fail-
ure strain of filaments is reached at smaller crack widths.
Furthermore, straining the filaments causes local spalling in and
on the shells as well as the splitting of the Ca(OH)2 crystals. The
resulting singularities and lateral pressure caused by the wedged
crystals act as notches, leading to premature failure of the fila-
ments. However, when the Portland cement clinker content of
the binder is low and pozzolanic additives are used (low alkalinity), Fig. 7. Volume flow rate through SHCC as a function of differential pressure for
a stratum of thin C-S-H phases leading to good bonding character- selected specimens with different residual strain levels (c.l. is the cumulative crack
istics dominates the fibre–matrix interface, cf. Fig. 6b. Such thinly length) [34].
walled coatings of the outer filaments enable good, flawless bond-
ing with the matrix. In the vicinity of matrix cracks a partial deb-
with high water-to-cement ratios. There is great need for further
onding and slip of filaments is possible, leading to ductile overall
research on the changes of fibre–matrix bond properties due to
behaviour of crack bridging yarn [29].
chemical attack.
Furthermore, during continued hydration and ageing a slow and
gradual process of deposition of hydration products between the
filaments may change the nature of bonding by increasing the 4. Characteristic material properties to predict long-term
sleeve-to-core ratio [30]. This results in a stronger bond, which is durability and service life
a favourable effect but may also lead to embrittlement, mainly
when brittle fibres such as glass are used [31,32]. This means that 4.1. General remarks
the bonding changes over time and thus also the composite
properties and durability. Based on the considerations regarding the durability of individ-
Cohen and Peled [30] described two degradation mechanisms ual components outlined in Section 3, the following characteristic
observed during ageing, both due to matrix penetrability into the properties of the composite material seem to be decisive with re-
bundle: (i) changing the telescopic pullout derived by cement pen- gard to long-term durability and service life of R/SHCC, RC/SHCC
etrability, as more fibres fracture after ageing, leading to more brit- and RC/TRC structures:
tle behaviour; and (ii) chemical deterioration of the glass fibres
themselves as more fibres are in direct contact with the cement – Transport properties of HPRFCC material (SHCC or TRC) in a
matrix, reducing composite strength. The authors found that fa- cracked state.
voured filler systems are those that moderate cement hydrates’ – Strain capacity of HPFRCCs.
penetrability between the bundle filaments over time, keeping – Resistance of HPFRCCs in aggressive environments.
the telescopic pullout mode and composite ductility, and at the
same time protecting the core filaments from chemical attack, thus On the material level these issues are addressed in previous sec-
retaining their high performance and the overall tensile strength of tions. The following sections provide a brief overview as well as
the composite. special considerations for elements and structures made of SHCC
There is little specific information on the changes of the bond- and TRC.
ing properties of SHCC or TRC under chemical attack. Kabele
et al. [33] found that in the case of calcium leaching in SHCC, the 4.2. Transport properties
reduction in the bond quality may be so severe that multiple crack-
ing of the material can no longer be guaranteed. However, it should The key question is what effect a considerable number of fine,
be noted that those experiments were carried out on specimens well distributed cracks have on the transport properties of SHCC
100 V. Mechtcherine / Construction and Building Materials 31 (2012) 94–104

or TRC. Average crack width has been used routinely as a reference through cracked TRC could be described by a simple model based
parameter. However, for materials with multiple cracks of self-lim- upon Hagen–Poiseuille’s Law. However, the model should be re-
iting width an approach using the cumulative crack length seems fined in future investigations in order to enable a more realistic
to be more appropriate [34]. Where this is not practical, average description of the transport mechanisms in TRC.
strain may serve as a useful surrogate. However, it can be assumed Pourasee et al. [37] described a series of experiments in which
that the fine fibres in combination with a fine grain matrix also af- fluid transport was measured using X-ray radiography in cracked
fect the condition of crack profiles and the continuity of the entire cement paste samples that were reinforced with different commer-
crack system, which of course influences the transport of fluids and cially available fabrics made with monofilament and multifilament
gases through the material. There is significant need for research yarns that were both coated and uncoated. Results showed that
into reliable and practical methods to describe the crack system fabrics made of multifilament yarns without coating may be prob-
in SHCC and TRC both in laboratory as well as in field applications. lematic from a durability point of view. However, when multifila-
ment yarns were coated, the fabric behaved as a monofilament
4.2.1. Air permeability system. The transport behaviour was observed to be highly depen-
There is very little information available on the air permeability dent on the quality of the coating. The authors concluded that this
of cracked SHCC, but it seems that the air permeability of un- fact should be considered in addition to mechanical performance
cracked SHCC is well within the range of typical values for ordinary in assessing the type of fabrics that should be used for a specific
concrete. application.
For material strained up to 0.5%, an increase of almost two or- Lieboldt et al. [38] investigated the transport mechanisms of
ders of magnitude has been observed [34], cf. also Fig. 7. While this water in and through composite concrete specimens made of a
may no longer be comparable to sound concrete, it should be noted cracked ordinary concrete (OC) as a substrate and textile reinforced
that the specimens were tested at an age of 45–50 days and the concrete (TRC) as a cover layer for its strengthening and repair. The
mix had a high content of fly ash. Thus, it can be expected that TRC cover layer was assessed with regard to its efficiency as a pro-
in field applications with similar mixes ongoing hydration and tective layer against the ingress of water. Since in real applications
self-healing will have a mitigating influence over time. such TRC layers may actually be or presumed to be cracked, there-
The coefficient of oxygen permeability in crack-free TRC speci- by activating the load-carrying function of the textile reinforce-
mens was found to be influenced clearly by the polymer coating ment, the TRC layer was cracked for purposes of this study. The
of the textile but was only marginally dependent on the fineness water transport in the OC specimens without TRC layer was used
of multifilament yarns in the textile [15]. The corresponding per- as reference. Neutron radiography served as the main testing tech-
meation coefficients were reduced by up to 65% if coated textiles nique. In ordinary concrete, quick and deep ingress of water
were embedded into the composite instead of uncoated textile. through relatively wide macro-cracks (100–200 lm), followed by
In situ permeation tests on cracked TRC subjected to uniaxial transport through the capillary pore system, caused saturation of
tensile loading revealed a pronounced non-linear increase in the large areas in a rather short time. TRC applied to the OC surface re-
transport rates of oxygen through the TRC (perpendicular to the duced the ingress of water to a large extent. Its small crack widths
composite surface) with increasing strain, corresponding to larger of approximately 20 lm modified the suction behaviour funda-
crack widths [15]. A higher degree of fineness in the textile rein- mentally. In the cracked substrate of ordinary concrete, capillary
forcement, i.e., a higher degree of reinforcement, and the polymer suction was obviated, and transport through the pore system of
coating of the multifilament yarns led to a larger number of cracks the matrix became the prevailing transport mechanism, based on
in the TRC specimens at given strain levels, but these cracks were the suction force characteristic of capillary pores. Not only was
considerably finer. As a result, the permeability for oxygen through the mechanism altered, but the transport of water deep into inner
the TRC made with such reinforcement was lower. This effect was regions was significantly retarded as well.
distinctly apparent at strain levels above 0.2%.
4.2.3. Chloride migration
4.2.2. Water-transport properties So far, very little information is available on the chloride migra-
The water-transport properties of various cracked, cementitious tion in and through SHCC and TRC. Miyazato and Hiraishi [39] ob-
materials have been studied by a number of researchers, cf. [35]. served no macro-cell corrosion in cracked R/SHCC specimens with
These investigations demonstrate the significance of crack width steel corrosion due to chloride penetration and a penetration depth
and cumulative crack length in controlling the permeability and that was significantly lower than for cracked reinforced mortar
water absorption of cracked SHCC. A comparison with values for specimens. More recently Sahmaran et al. [40] conducted immer-
ordinary concrete shows that for strain levels up to 0.5% the water sion and ponding tests on SHCC and reinforced mortar specimens.
absorption of SHCC is comparable to that of sound concrete. It The chloride penetration depth and diffusion coefficient observed
should be noted, however, that the reported crack widths in SHCC both in cracked and uncracked SHCC specimens were much lower
are typically between 50 and 100 lm. While capillaries of this than those of the mortar specimens. The authors found the chlo-
width may not transport water very far into the material, the pro- ride diffusion coefficient to be proportional to the square of the
cess is rather fast. In conjunction with the natural porosity of SHCC crack width and, in the case of multiple cracking, linearly propor-
and submicro-cracks [36], the water may be transported deep into tional to the number of cracks. The observed higher resistance of
the material over time. Further research in this area is required. cracked and uncracked SHCC against the ingress of chloride ions
Mechtcherine and Liebold [15] reported that if measured paral- can be attributed to a higher content of cementitious materials, a
lel to the reinforcing textile layers, water absorption of TRC in- low water-to-cement ratio, the self-limiting crack width and, as a
creases with increasing fineness of multifilament yarns and result, self-healing of the cracks. While this is promising, more re-
decreases with an increasing amount of polymer used to coat the search is required. This is particularly true of textile reinforced
yarns. Both these observations can be traced back mainly to differ- concrete also, since no publications are known on the chloride
ences in the capillary system formed in multifilament yarns. For migration in and through TRC.
water permeability, dependencies between the crack system and
the volume of transported media were found to be similar to those 4.2.4. Other transport properties
described above for oxygen permeability. It was also shown that For sound, ordinary concrete numerous correlations were
the influence of imposed strain on the permeation of water experimentally and in part also theoretically derived between
V. Mechtcherine / Construction and Building Materials 31 (2012) 94–104 101

Fig. 8. Tri-linear representations of stress–strain curves for TRC made with (a) Portland cement matrix, and (b) blast furnace cement matrix with addition of pozzolana; after
demoulding all specimens were stored for 6 days in water (20 °C); subsequently, the reference specimens were subjected to storage at 20 °C/65% RH until the testing in the
age of 28 days; the other part of the specimens was accelerated aged in a fog room (40 °C/99% RH); the duration of the fog room storage was 28, 56, 90, 180 and 360 days,
respectively [29].

different transport mechanisms (diffusion, permeation, absorp- strain level [44]. Matsumoto et al. [45] performed fatigue tensile
tion) and for different transport media (e.g., water, air, CO2) tests with a number of cycles up to 500,000. It was found that
[41,42]. For cracked concrete, however, such comparisons have the bridging stress degradation developed gradually at low fatigue
not yet been performed, as is the case for SHCC or TRC. The knowl- loading cycles, and the rate of stress reduction increased when the
edge of possible correlations is essential for the prediction of the number of cycles increased. No systematic studies on fatigue
durability of structures made of or strengthened with HPFRCCs, if behaviour of TRC are known thus far.
it is based on the results of just one or two particular test methods. There is little information available on the fatigue behaviour of
There is a great need for research in this area. SHCC in field applications. Zhang and Li [46] investigated SHCC in
high-cycle fatigue scenarios, such as highway repairs. Both SHCC/
4.3. Strain capacity of HPFRCCs concrete and concrete/concrete overlay specimens were tested in
flexural fatigue. The tests showed that the load capacity of SHCC/
4.3.1. Long-term strain capacity and ageing behaviour concrete specimens was twice that of concrete/concrete speci-
Since high strain capacity is a decisive property with regard to mens, the deformability of SHCC/concrete specimens was signifi-
the durability of SHCC and TRC, these materials as well as members cantly higher, and the fatigue life was extended by several orders
or structures made of them can only be considered as truly durable of magnitude. Kunieda and Rokugo [47] found that while SHCC
if their strain capacities do not change negatively at a substantial has a longer fatigue life than ordinary FRC at higher fatigue stress
rate over time. In laboratory experiments SHCC long term strain levels, its fatigue life at lower stress levels is shorter. This can be
capacities of 3% compared to initial ultimate strain levels of 5% attributed to the propagating of only a few cracks in the specimen
were observed [43]. While both values are far above the deforma- at low stress levels. Thus, the cracking state under fatigue load is an
tion demand imposed by many applications, average strain capac- important factor for fatigue resistance of SHCC.
ities in large scale SHCC production outside the laboratory are
likely to be significantly lower. Furthermore, the binder composi-
4.3.3. Behaviour under sustained loads
tion, the type of cement and the cement replacing materials as well
as hygral and thermal ambient conditions should be considered. Creep of SHCC consists of matrix creep and time-dependant
More research is required in order to be able to predict the long crack development accompanied by gradual delamination of fibres
term mechanical performance of SHCC. from the matrix, leading to subsequent pullout. As is the case with
Butler et al. [29] investigated the effect of matrix composition ordinary concrete, high sustained loads will lead to time-depen-
by varying the hydration kinetics and alkalinity of the binder mix dent failure (Boshoff [8], Jun and Mechtcherine [48]).
on the ageing of TRC. The results of tensile tests on accelerated
aged specimens made of TRC showed a pronounced decrease in
the tensile strength and strain capacity for TRC whose matrix
was most alkaline (Portland cement was used exclusively as binder
in this composition); see Fig. 8a. The performance of TRC made
with a modified, alkali-reduced matrix composition was to a great
extent unaffected by exposure to accelerated ageing, cf. Fig. 8b. The
authors showed that reductions in the toughness of TRC could be
attributed mainly to the observed disadvantageous new formation
of solid phases in the fibre–matrix interface, cf. also Section 3.5.

4.3.2. Cyclic loading and fatigue


The available information on the behaviour of SHCC under cyc-
lic loading indicates that cracking behaviour is comparable to
behaviour under monotonic loading. The material stiffness de-
creases significantly with an increasing number of load cycles, Fig. 9. Time-dependent reduction of the water transport trough cracked TRC in
but there is no evident influence of cyclic loading on the ultimate unloaded state [15].
102 V. Mechtcherine / Construction and Building Materials 31 (2012) 94–104

There is little information on the creep behaviour of SHCC in out by Mechtcherine and Liebold [15] in the unloaded, saturated
field applications. Rokugo et al. [49] reported increasing crack state of TRC specimens at time intervals of 7 days. Fig. 9 shows
widths of up to 0.12 mm after 24 months for the patch repair of the time-dependent reduction of the transport rates of samples
a concrete retaining wall suffering from ASR. However, it is unclear with different reinforcement textiles. After storing the cracked
whether or not this was due to increasing crack widths in the sub- specimens in water for 14 days, the flow rate was reduced to less
strate, relaxation in the SHCC, or both. than 50% of the value obtained on the same specimen before stor-
ing in water. No measurable volume flow occurred after continued
water exposure of 21–35 days, the particular duration depending
4.4. Self-healing of cracks
on the type of the textile reinforcement used. The samples contain-
ing uncoated AR glass yarns with low yarn fineness (1280 tex)
The self-healing of cement-based materials is a well-known
needed a longer time until a negligible volume flow was reached,
phenomenon. The main causes of such self-healing are twofold:
which can be traced back to the initially wider cracks in these spec-
On the one hand non-hydrated binder particles (cement or pozzo-
imens. The specimens containing coated carbon textile layers
lana) can hydrate further on if the material comes in contact with
showed the fastest self-healing, which could be expected since
water again. The newly formed calcium–silicate–hydrate (C–S–H)
they had the finest cracks.
phases can effectively bridge fine cracks. On the other hand, cal-
Using an ESEM (Environmental Scanning Electron Microscope),
cium hydroxide (CH) can be deposited inside the cracks and can
additional investigations of cracks before and after exposure to
be converted to calcium carbonate as a result of carbonation (or,
water revealed the formation of new crystal structures which
if pozzolanic materials are involved, to C–S–H phases). CH and cal-
closed the cracks to transport, cf. Fig. 10. Fig. 10a shows the origi-
cium carbonate are rather weak from a mechanical point of view,
nal crack with a width of 20 lm before water contact. Fig. 10b
making the crack bridging less effective, but they can fill and seal
illustrates the condition of a similar crack after the water perme-
cracks.
ability tests with duration of 21–35 days, indeed largely closed
In ordinary concrete the crack widths are rather large and the
by deposits of calcium carbonate (calcite) [15].
water-to-cement ratio is relatively high; the deposit of CH and cal-
cite dominates the self-healing process. Larger cracks cannot be
sealed completely. In contrast, SHCC or TRC possess very small 5. Summary
crack widths because of their tendency to multiple cracking. For
these kinds of cementitious composites a complete sealing of There is great potential for the application of high-performance
cracks and even a partial restoration of stiffness seem to be attain- fibre-reinforced cement-based composites in structural elements
able to a great extent. The restoration of the stiffness is due to the and structures exposed to severe mechanical or environmental
bridging of fine cracks by the newly developed C–S–H phases. loading. Present knowledge indicates that SHCC and TRC show a
Yang et al. [50] used transverse resonant frequencies in charac- high long-term strain capacity and favourable transport properties
terising the self-healing of SHCC with short PVA fibres. Specimens in the cracked state. Beyond that, the composites exhibit in many
featuring crack widths below 50 lm were completely healed, and instances superior resistance to aggressive environments when
samples with crack widths between 50 and 150 lm only partly compared to ordinary concrete. However, with the exception of
recovered. At specimens with crack widths exceeding 150 lm no exposure to alternating freezing and thawing, which does not alter
recovery was observed. In uniaxial tension tests after damage the mechanical properties significantly, there is no information
and self-healing, high strain capacity was measured (1.8–3.1%) available on the effects of aggressive environments on the mechan-
even in samples which had undergone pre-loading up to a strain ical properties of the materials.
of 2% or 3%. Microscopic observation identified cracks propagating Since SHCC and TRC are new materials, there is no information
both through the healed material area and the heretofore un- available on the long-term field performance of elements and
cracked matrix. The authors concluded that at least 4–5 cycles structures made from or strengthened with SHCC or TRC. However,
(submerged in water at 20 °C for 24 h and dried in air at 21 °C first field applications indicate that the superior performance in a
for 24 h) were needed for a meaningful recovery of stiffness during laboratory environment as discussed in previous chapters and
the healing cycle. Kan et al. [43] proved two distinct types of heal- summarised above, does indeed translate to superior field
ing products in SHCC materials: In wide cracks primarily large cal- performance.
cite grains were formed, in small cracks mainly fibre-like C–S–H There is significant research needed to understand the long-
phases could be observed. term behaviour of SHCC, TRC, R/SHCC, RC/SHCC and RC/TRC ele-
In order to investigate the effect of self-healing on the transport ments under complex combined loads as encountered in the field.
properties of TRC, water permeation measurements were carried To be able to utilise the superior qualities of this new material

(a) (b)

20 20
µm µm

Fig. 10. Self-healing effect in TRC: (a) crack before and (b) after water exposure [15]; both images show the same specimen.
V. Mechtcherine / Construction and Building Materials 31 (2012) 94–104 103

fully, it will be necessary to develop a realistic, reliable, perfor- [20] RILEM recommendation. RILEM TC-117 FDC – CDF test – Test method for the
freeze-thawresistance of concrete with sodium chloride solution, vol. 29.
mance-based durability design concept for such structures. This
Mater Struct; 1996. p. 523–8.
paper is a first step towards a theoretical framework for this [21] Li VC, Lepech M. Engineered cementitious composites: design, performance
concept. and applications. In: Mechtcherine V, editor. Ultra-ductile concrete with short
fibres – development, testing, applications. Stuttgart: ibidem Verlag; 2005. p.
99–120.
[22] De Lhoneux B, Akers S, et al. Durability study of PVA fibres in fibre-cement
Acknowledgements products. In: Proceedings of the IVth international symposium on concrete for
a sustainable agriculture, Agro-, Aqua and community applications, Ghent,
Belgium: Ghent University; 2002. p. 275–84.
A considerable part of the work presented in this paper was ini- [23] Horikoshi T, Ogawa A, Saito T, Hoshiro H. Properties of polyvinylalcohol fibre
tiated in the Collaborative Research Centre SFB 528 ‘‘Textile Rein- as reinforcing materials for cementitious composites. In: Proceedings PRO 49
of international RILEM workshop on HPFRCC in structural applications,
forcement for Structural Strengthening and Retrofitting’’ financed
Honolulu: RILEM Publications S.A.R.L.; 2006. p. 145–53.
by the German Research Foundation (DFG). The author would like [24] Ogawa A, Hoshiro H. Durability of fibres. In: Van Zijl GPAG, Wittmann FH,
to acknowledge with gratitude the foundation’s financial support. editors. Durability of strain-hardening fibre-reinforced cement-based
composites (SHCCs). RILEM State of the Art Reports: Springer; 2011. p.
81–8.
[25] Yilmaz V, Glasser F. Reaction of alkali-resistant glass fibres with cement. Part
References 2: Durability in cement matrices conditioned with silica fume. Glass Technol
1991;32(4):38–47.
[26] Scheffler C, Förster C, Mäder E, Heinrich G, Hempel S, Mechtcherine V. Aging of
[1] Fischer G, Li VC, editors. High performance fiber reinforced cementitious
alkali-resistant glass and basalt fibres in alkaline solutions: evaluation of the
composites (HPFRCCs) in structural applications. In: Proceedings PRO 49 of
failure stress by Weibull distribution function. J Non-Cryst Solid 2009;355(52–
international RILEM workshop on HPFRCC in structural applications, Honolulu,
54):2588–95.
US: RILEM Publications S.A.R.L.; 2006.
[27] Butler M, Mechtcherine V, Hempel S. Experimental investigations on the
[2] Mechtcherine V, editor. Ultra-ductile concrete with short fibres –
durability of fibre-matrix interfaces in textile-reinforced concrete. Cem Concr
development, testing, application. Stuttgart: Ibidem-Verlag; 2005.
Compos 2009;31(4):221–31.
[3] Brameshuber W, editor. Textile reinforced concrete. RILEM state-of-the-art
[28] Oh BH, Kabele P. Durability under chemical loads. In: Van Zijl GPAG, Wittmann
report. Rep036, RILEM Technical Commitee 201-TRC; 2006.
FH, editors. Durability of strain-hardening fibre-reinforced cement-based
[4] Butler M, Lieboldt M, Mechtcherine V. Application of textile-reinforced
composites (SHCCs). RILEM State of the Art Reports: Springer; 2011. p. 41–58.
concrete (TRC) for structural strengthening and in prefabrication. In:
[29] Butler M, Mechtcherine V, Hempel S. Durability of textile-reinforced concrete
Advances in cement-based materials. London: Taylor & Francis Group; 2010.
made with AR glass fibre – effect of the matrix composition. Mater Struct
p. 127–36.
2010;43:1351–68.
[5] Van Zijl GPAG, Wittmann FH, editors. Durability of strain-hardening fibre-
[30] Cohen Z, Peled A. Controlled telescopic reinforcement system of fabric–cement
reinforced cement-based composites (SHCCs). RILEM state of the art reports,
composites – durability concerns. Cem Concr Res 2010;40:1495–506.
vol. 4. Springer; 2011.
[31] Zhu W, Bartos P. Assessment of interfacial microstructure and bond properties
[6] Weiland S, Ortlepp R, Hauptenbuchner B, Curbach M. Textile reinforced
in aged GRC using novel microindentation method. Cem Concr Res
concrete for flexural strengthening of RC-structures – Part 2: Application on a
1997;27(11):1701–11.
concrete shell. In: ACI SP-251-3 design and applications of textile-reinforced
[32] Bentur A, Diamond S. Aging and microstructure of glass fiber cement
concrete. S. 41–58.
composites reinforced with different types of glass fibers. Durab Build Mater
[7] Inaguma H, Seki M, Suda K, Rokugo K. Experimental study on crack-bridging
1987;4:201–26.
ability of ECC for repair under train loading. In: Proceedings PRO 49 of
[33] Kabele P, Novák L, Nemeček J, Kopecký L. Effects of chemical exposure on bond
international RILEM workshop on HPFRCC in structural applications, Honolulu,
between synthetic fiber and cementitious matrix. In: Proceedings PRO 50 of
US: RILEM Publications S.A.R.L.; 2006. p. 499–508.
ICTRC2006 – 1st international RILEM conference on textile reinforced
[8] Boshoff WP. Time-dependant behaviour of ECC [dissertation]. South
concrete, Aachen, Germany: RILEM Publications S.A.R.L.; 2006. p. 91–9.
Africa: University of Stellenbosch; 2007.
[34] Mechtcherine V, Lieboldt M, Altmann F. Preliminary tests on air-permeability
[9] European Brite-Euram Programm. Probabilistic performance based durability
and water absorption of cracked and uncracked strain hardening cement-
design of concrete structures. DuraCrete final technical report. Report No.:
based composites. In: Proceedings of the international RILEM workshop on
BE95-1347/R17; 2000.
transport mechanisms in cracked concrete, Ghent, Belgium: RILEM
[10] fib bulletin 34. Model code for service life design. International Federation for
Publications S.A.R.L.; 2007. p. 55–66.
Structural Concrete (fib), Lausanne; 2006.
[35] Van Zijl GPAG. Durability under mechanical load – micro-crack formation
[11] Schießl P, Gehlen Ch. New approach to service life design of concrete structure.
(ductility). In: Van Zijl GPAG, Wittmann FH, editors. Durability of strain-
In: Proceedings of an international workshop on combined mechanical and
hardening fibre-reinforced cement-based composites (SHCCs). RILEM State of
climatic loads (CMCLs), Qindao, China: Aedificatio Publishers; 2005. p. 3–14.
the Art Reports: Springer; 2011. p. 9–39.
[12] Altmann F, Mechtcherine V, Reuter U. A novel durability design approach for
[36] Wittmann FH, Wilhelm T, Beltzung F, Grübl P. Multi-crack formation in strain
new cementitious materials: modelling of chloride ingress in strain-hardening
hardening cement-based composites. In: Proceedings PRO 53 of fifth
cement-based composites (SHCCs). In: Alexander MG, Beushausen F, Dehn P,
international workshop on high performance fiber reinforced cement
Moyo P, editors. International conference on concrete repair, rehabilitation
composites (HPFRCC5), Mainz, Germany: RILEM Publications S.A.R.L.; 2007.
and retrofitting II – ICCRRR 2008, Cape Town, South Africa, London: Taylor &
p. 125–34.
Francis Group; 2009. p. 117–8 [complete version on CD only].
[37] Pourasee A, Peled A, Weiss J. Fluid transport in cracked fabric-reinforced-
[13] Altmann F, Mechtcherine V. Modelling the influence of cracking on chloride
cement-based composites. J Mater Civil Eng, ASCI 2011;23(8):1227–38.
ingress into strain-hardening cement-based composites (SHCCs). Advances in
[38] Lieboldt M, Schröfl C, Mechtcherine V. Water transport through textile
cement-based materials. London: Taylor & Francis Group; 2010. p. 193–201.
reinforced concrete. In: Wittmann FH, Mercier O, editors, Proceedings of an
[14] Li VC, Stang H. Elevating FRC material ductility to infrastructure durability. In:
ASMES international workshop basic research on concrete and Applications,
Proceedings PRO 39 of BEFIB, Varenna, Italy: RILEM Publications S.A.R.L.; 2004.
Freiburg: Aedificatio Publishers; 2011. p. 21–34.
p. 171–86.
[39] Miyazato S, Hiraishi Y. Transport properties and steel corrosion in ductile fiber
[15] Mechtcherine V, Liebold M. Permeation of water and gases through cracked
reinforced cement composites. Torino, Italy: Paper 4484 of compendium of
textile reinforced concrete. Cem Concr Compos 2011;33:725–34.
papers CD-ROM, ICF 11; 2005.
[16] Hiraishi Y, Honma T, Hakoyama M, Miyazato S, editors. Steel corrosion at
[40] Sahmaran M, Li M, Li VC. Transport properties of engineered cementitious
bending cracks in ductile fiber reinforced cementitious composites. In:
composites under chloride exposure. ACI Mater J 2007;104(6):303–10.
Proceedings of the JCI symposium on ductile fiber reinforced cementitious
[41] Hilsdorf HK, Schönlin K, Tauscher F. Dauerhaftigkeit von Betonen (in
composites (DFRCCs), Tokyo, Japan: Japan Concrete Institute; 2003 [in
German). Düsseldorf: Beton-Verlag GmbH; 1997.
Japanese].
[42] Lawrence CD. Measurements of permeability. In: Proceedings of 8th
[17] Kanda T, Saito T, Sakata N. Tensile and anti-spalling properties of direct
international congress on the chemistry of cement, vol. V, Brazil FINEP: Rio
sprayed ECC. J Adv Concr Technol 2003;1(3):269–82.
de Janeiro; 1986. p. 29–34.
[18] Müller HS, Mechtcherine V, Geis M, Hewener A. Concrete degradation due to
[43] Kan LL, Shi HS, Sakulich AR, Li VC. Self-healing characterization of engineered
corrosion of steel reinforcement – numerical investigations. In: Proceedings of
cementitious composites (ECCs) materials. ACI Mater J 2010;107:619–26.
15th international conference on building materials research (ibausil),
[44] Jun P, Mechtcherine V. Behaviour of strain-hardening cement-based
Weimar, Germany: F.A. Finger-Institut für Baustoffkunde, Bauhaus-
composites (SHCCs) under monotonic and cyclic tensile loading. Part 1:
Universität Weimar; 2003. p. 1457–67.
Experimental investigations. Cement Concr Compos 2010;32:801–9.
[19] Brüdern A-E, Mechtcherine V. Multifunctional use of SAP in Strain-hardening
[45] Matsumoto T, Chun P, Suthiwarapirak P. Effect of fibre fatigue rupture on
Cement-based Composites’. In: International RILEM conference on use of
bridging stress degradation in fibre reinforced cementitious composites. In: Li
superabsorbent polymers and other new additives in concrete, Lyngby:
VC, et al., editors. Proceedings of fracture mechanics of concrete structures, Ia-
Technical University of Denmark; (RILEM proceedings PRO 74, 2010). p. 11–
FraMCos; 2004. p. 653–60.
22.
104 V. Mechtcherine / Construction and Building Materials 31 (2012) 94–104

[46] Zhang J, Li VC. Monotonic and fatigue performance in bending of fiber performance fibre reinforced cementitious composites HPFRCC5, RILEM
reinforced engineered cementitious composite in overlay system. J Cem Concr Publications S.A.R.L.; 2007. p. 97–104.
Res 2002;32(3):415–23. [49] Rokugo K. Applications of strain hardening cementitious composites with multiple
[47] Kunieda M, Rokugo K. Recent progress on HPFRCC in Japan. J Adv Concr cracks in Japan. In: Mechtcherine V, editor. Ultra-ductile concrete with short fibres –
Technol 2006;4(1):19–33. development, testing, applications. Stuttgart: ibidem Verlag; 2005. p. 121–33.
[48] Jun P, Mechtcherine V. Behaviour of strain hardening cement-based [50] Yang Y, Lepech MD, Yang EH, Li VC. Autogenous healing of engineered
composites (SHCCs) under repeated tensile loading. In: Reinhardt HW, cementitious composites under wet-dry cycles. Cem Concr Res
Naaman A, editors. Proceedings PRO 53 of RILEM-symposium on high- 2009;39:382–90.

You might also like