You are on page 1of 16

SP-250—6

Dimensioning and Application of


Textile-Reinforced Concrete
by J. Hegger, H.N. Schneider, S. Voss, and I. Bergmann

Synopsis: Textile-reinforced concrete (TRC) is a composite material consisting of textile reinforcement


and fine-grained concrete. Because of different bond performance and material properties of textiles
compared to reinforcing steel, the load-bearing behavior of TRC differs from that of ordinary reinforced
concrete. As a consequence, the design methods for ordinary reinforced concrete cannot be directly
applied to TRC structures. This paper investigates the load-bearing behavior of TRC. Design methods
for TRC are derived considering different textile properties and different loading conditions such as
tension, bending, and shear. Based on the extensive theoretical and experimental investigations
conducted, design methods have been derived that allow the dimensioning of TRC structures. Further
investigations are aimed at identifying appropriate applications for TRC. The results of theoretical and
experimental investigations conducted have been put into practice by producing TRC elements for
applications such as façade elements, formworks, and tanks.

Keywords: AR glass fibers; carbon fibers; dimensioning; textile-reinforced concrete

69
70 Hegger et al.
Josef Hegger is Professor at the Institute of Structural Concrete, RWTH Aachen University, Germany since 1993.
He obtained his Ph.D. in 1985 from the TU Braunschweig. From 1985 until 1993 he was employed as a project
manager in a construction company. His main research interests are bond behavior, shear-carrying capacity, high-
performance concrete and composite structure. He is also an ACI member.

Hartwig N. Schneider is Professor at the Chair of Building Construction and Design, Department of Architecture,
RWTH Aachen University, Germany since 1999. He studied at the University of Stuttgart in Germany and the
Illinois Institute of Technology in Chicago. Since 1988 he is partner of an Architectural Consulting Office in
Stuttgart and member of the German Association for Architects since 1993.

Stefan Voss is research assistant at the Institute of Structural Concrete of the RWTH Aachen University, Germany
since 2002. He obtained his Diploma Degree in the field of structural engineering from the RWTH Aachen
University in 2002.

Ingo Bergmann is research assistant at the Chair of Building Construction and Design, Department of Architecture,
RWTH Aachen University, Germany since 2001. He obtained his Diploma Degree in the architecture from the
RWTH Aachen University in 1996.

RESEARCH SIGNIFICANCE

TRC is a new composite material which opens up new fields of application for concrete as a building material. In
recent years, the most favored fiber reinforcement materials have been alkali-resistant glass and carbon. Their
different material properties result in differences in the load-bearing behavior of the composite. These have to be
considered when deriving applicable design methods for TRC structures, which are important for a successful
application.

INTRODUCTION

In present architecture there is a distinctive trend towards more and more lean structures of high-quality materials
which continuously increases the requirements placed in the construction materials and which demands a continuous
development of their properties. Instead of the usual concrete reinforcement of steel, non-metallic reinforcement
materials of a high tensile strength are used which open new fields of application for concrete as a building material.
Currently, the main application areas for the building material textile reinforced concrete are in the field of façade
construction but also increasingly in the field of bearing structures for minor vertical loads. The established models
known from reinforced concrete construction cannot just be transferred to the dimensioning of textile reinforced
concrete members because the bearing behaviour of textile reinforcements differs from that of concrete reinforcing
steels due to the specific material and bond properties. Therefore, the development of dimensioning models for
textile reinforced concrete is the topic of systematic investigations whose essential results as well as examples of
application are presented in this paper.

MATERIAL PROPERTIES

The textile reinforcement is usually a two-dimensional fabric or a three-dimensional spacer fabric. The fabrics are
made out of rovings which themselves consist of some hundreds to thousands of single fibers (filaments). The
currently most favorable fiber materials for manufacturing textile reinforcement fabrics are alkali resistant glass (AR
glass), carbon and aramid.

In Table 1 the properties of the fabrics used for the investigations presented in this paper are provided. The fabrics
have been produced out of AR glass and carbon rovings, respectively, both having a cross-sectional area of 0.89
mm² (0.001 in.²). Producer of the AR glass roving is Saint Gobain Vetrotex and the carbon roving has been
produced by Tenax Fibers.

The tensile strength of the fabrics in 0°- and 90°-direction was measured in tensile tests with 125 mm (4.92 in.) long
parts of the rovings taken from the fabrics. A description of this test method is given in (Roye and Gries 2005). The
strength of the rovings is much lower than those of the single filament. Reasons are e.g. the damaging of the rovings
Textile-Reinforced Concrete 71
during the textile production process, the non-uniform loading of the filaments due to their waviness and other
complex effects described in (Vorechovsky et al. 2005)

The properties of the textile reinforcement lead to special demands on the concrete mixture. In order to enable the
penetration of the fabrics the maximum grain size is limited to about 2 mm (0.08 in.) and a high flowability is
needed. Furthermore, the concrete should have a high strength to enable an early removal of the formworks because
TRC is mainly meant for the use in the precast industry. Based on these requirements the mixture shown in Table 2
was developed (Brameshuber et al. 2005). The fine grained concrete has a compressive strength of about 85 MPa
(12.3 ksi) and a tensile strength of 4.4 MPa (0.64 ksi) (Table 3). The elastic modulus of the fine grained concrete is
lower and it has a higher strain at failure of about 5‰ compared to 2.5‰ for normal concrete.

The previous investigations showed that in the composite material TRC the tensile strength of the filaments cannot
be fully exploited. The main reason is that there is a decrease in bond-performance from the outer filaments towards
the core of the roving. Several models have been established taking this effect into consideration. In classical
idealizations the cross-sectional area of the roving is subdivided into core filaments and sleeve filaments according
to the model of Ohno and Hannant (Ohno and Hannant 1994).

TENSILE STRENGTH OF THE COMPONENT

Based on the results of tensile tests on textile-reinforced concrete elements described in (Hegger and Voss 2005) a
factor for the efficiency can be calculated, which is the ratio between the tensile strength of the filament and the
measured average strength of the reinforcement in the concrete. The efficiency is governed by the bond properties of
the rovings and is therefore influenced by the filament diameter, the type of the binding and the roving thickness.
These parameters affect the roving geometry and the penetration of the matrix into the roving. Moreover, the tensile
strength of the component is influenced by lateral tensile stresses as biaxial tensile tests on discs made of TRC
showed. The cracking of the concrete in reinforcement direction led to damaging of the rovings and therefore caused
a decreasing strength of the component (Hegger and Voss 2004).

Another very important effect is the orientation of the reinforcement in respect to the load direction. The different
orientation of reinforcement and loading leads to significant losses of tensile strength. The main effects which are
responsible for this loss of strength are the transversal actions and the bending stresses of the filaments at the crack
edges. A detailed description of the effects is given in (Hegger et al. 2005). Based on the results of the investigations
the presumption has been made that the tensile strength of the reinforcement decreases linearly with increasing slope
of the reinforcement. This interrelationship is described by the reduction factor k0,α which is defined in Equation 2.
The results of other investigations concerning the effects of sloped loading are described in (Molter 2005; Jesse
2005). There, the detected loss of strength was higher.

Considering the described effects the tensile strength of the textile reinforced component can be calculated with
Equation (1):

Fctu = At ⋅ f t ⋅ k1 ⋅ k 0 ,α ⋅ k 2 (1)

with At cross-sectional area of the reinforcement


ft tensile strength of the reinforcement
k1 efficiency factor (Hegger et al. 2005)
α
k0,α factor for orientation of the reinforcement in respect to the load direction: k 0 ,α = 1 − (2)
90°
k2 factor for biaxial loading (Hegger and Voss 2004)

BENDING CAPACITY

The bending capacity of TRC elements was tested in four-point bending tests on I-shaped beams with a length of
one meter (3.3 ft.). The beams are 120 mm (4.72 in.) high, their flanges are 110 mm (4.33 in.) wide and the
reinforcement is placed in the center of the components (Fig. 1). The reinforcement ratio of the flange in the tensile
zone was varied between one to three layers of fabrics whereas the web was always reinforced with two layers. In
72 Hegger et al.
order to guarantee the precise position of the textile reinforcement the fabrics were clamped and prestressed with
very low tensile force before concreting. While the vertical loading was applied with deformation-control the
deflections and the strain of the components tension zone were measured with LVDTs. In some test series
longitudinal forces were added which were generated by hydraulic cylinders and kept constant during the tests.

According to the tests, the AR glass reinforcement achieved an average tensile strength of about 410 MPa (59.5 ksi)
nearly independent of the reinforcement ratio, whereas the calculated ultimate strength of the carbon fabric
increased with raising reinforcement ratio. The strength amounted to about 1100 MPa (159.5 ksi) for a beam
reinforced with one layer (reinforcement ratio of 1%) and reached about 1400 MPa (203 ksi) for a three layer
reinforced beam (reinforcement ratio 1.9%). This may be explained as follows: In general, a higher reinforcement
ratio leads to lower crack spacing and more cracks, respectively. This comes along with larger deflections and
increasing curvature of the beam. The latter leads to bigger transversal stresses acting on the roving and, hence, to a
better bond performance of the filaments. In case of the carbon fibers this effect may superpose the damage due to
the transversal action at the crack edges. Furthermore, a higher number of cracks leads to a higher strain of the
tensile zone of the beam before the failure occurs. The higher strains of the inner filaments cause a higher
effectiveness. This effect results in higher ultimate strength of the carbon fabric under bending loading than in a
centric tensile test as described in (Hegger and Voss 2005). In case of AR glass reinforcement the damaging effects
are pronounced so that the load carrying capacity amounts to only about 90% of the strength under centric tensile
loading. Because of the increasing transversal action on the filaments the ratio between the ultimate strength under
bending and tensile loading decreases with increasing reinforcement ratio.

For the calculation of the bending strength the effects due to the transversal action on the reinforcement caused by
the curvature of the beam are considered by the factor kfl,ρ. Besides the reinforcement ratio this factor is affected by
several properties of the reinforcement like the binding, the diameter of the roving or an additional impregnation of
the fabric. Therefore, in this paper only the results for the fabrics described in Table 1 are given. These may be
different for fabrics with other properties.

According to steel reinforced concrete the bending strength can be calculated with the knowledge of the tensile
strength of the reinforcement Fctu and the inner lever arm z:

M u = k fl ,ρ ⋅ Fctu ⋅ z (3)

where kfl,ρ factor for bending loading


AR glass (MAG-07-03): k fl , ρ = 0.9
At
carbon (MAG-04-03, MAG-05-03): k fl ,ρ = 1.0 + 40 ⋅
Ac
Fctu calculated from Equation (1)
z inner lever arm

SHEAR CAPACITY

The same beams as for the bending tests were used for the shear tests (Fig. 1). Only the reinforcement of the web
varied along the length of the beam in order to carry out two shear tests with each specimen. First, a four-point
bending test was carried out in which the part of the web with lower reinforcement ratio failed. After that the
remaining part of the beam with higher shear reinforcement ratio was tested in a three-point bending test.

The test program included tests with different shear reinforcement ratios, different fiber materials and tests with
additional longitudinal force. In addition, in one test series the effect of the horizontal rovings in the web was
investigated on beams with and without horizontal rovings.

An important issue was the precise measurement of the deformation of the shear area with a photogrammetry
measurement system (Fig. 2). Especially the relative movement of the shear crack edges was of major interest. With
these data the effect of the angular change of the shear reinforcement α and the effect of friction forces at the shear
crack edges were evaluated. Based on the global deformations dx and dy which are provided by the photogrammetry
Textile-Reinforced Concrete 73
measurement system the shear crack opening w, the parallel crack shift v and the angular change of the shear
reinforcement α (Fig. 3) have been calculated.

The first well-known model for the description of the bearing capacity of shear reinforced beams is the truss model
derived by Mörsch consisting of 45°-inclined compressive struts and vertical tensile elements simulating the shear
reinforcement. The bearing capacity of the truss depends on the amount and strength of the shear reinforcement and
is limited to the strength of the compression struts. From the comparison with experimental results it is well-known
that this model underestimates the shear capacity particularly of beams with low shear reinforcement ratios.
Therefore, in contemporary enhanced models the shear strength of reinforced beams is composed of a truss
contribution and a concrete contribution, which is effected by the bearing capacity of the compressive zone. Other
models take a variation of the truss inclination into account, which is explained by friction forces at the shear crack
edges and aggregate interlock. Actual investigations demonstrated that the influence of the crack friction forces on
the shear capacity is low because at failure loads the appearing shear crack width does not allow for the transfer of
shear stresses across the cracks. Therefore, a model has been derived consisting of a truss contribution and a
concrete contribution which is in very good agreement with test results (Görtz 2004; Hegger et al. 2005a).

An important factor for the shear strength of reinforced beams is the mechanical reinforcement ratio ωct, calculated
from Equation (4):

Atw ⋅ σ max
ωct = (4)
bw ⋅ f ct

where Atw is the cross-sectional area of the shear reinforcement, σmax is the tensile strength of the reinforcement in
the composite, bw is the thickness of the web and fct is the tensile strength of the concrete. In investigations on
stirrup-reinforced beams a significant influence of ωct on the shear crack angle βr,exp has been detected especially for
low reinforcement ratios (ωct < 1.5%), where an increasing ωct leads to steeper shear cracks (Görtz 2004). However,
due to the relatively high shear reinforcement ratio in the tests on TRC-beams no dependency was detectable. Here
the average value of the shear cracks amounted to about 45°.

In order to estimate the effect of shear crack friction the crack width w and the relative displacement of the crack
edges v were calculated from the photogrammetry measurement data as explained above. The average shear crack
width at failure amounted about to 0.2 mm (0.008 in.), so that due to the small grain size of 0.6 mm (0.024 in.) no
shear friction can be expected. This means that the shear crack angle βr,exp corresponds to the compression strut
inclination.

Besides the knowledge of the strut inclination the tensile strength of the shear reinforcement is required to calculate
the bearing capacity of the truss. Therefore, the angular change α of the shear reinforcement was introduced to
calculate the effective tensile strength of the rovings in the web. The tests showed that the values for α varied
between about 12° - 25° depending on the reinforcement ratio and the fiber material. Higher shear reinforcement
ratios lead to decreasing angles α. Considering the results of the tensile tests with sloped reinforcement (Hegger et
al. 2005) leads to the conclusion that the loss of strength due to the direction change is significant as the exemplary
calculation of the reduction factor k0,α for α = 25° shows:

k0 ,α = 1 − 25° 90° = 0.72 .

The tests revealed that the horizontal reinforcement of the web has to be taken into account for the calculation of the
truss capacity. Since the angular change of the longitudinal reinforcement is much higher (for α = 25° the angular
change of the horizontal rovings is 65°) the resulting tensile strength amounts to only about 15 - 30% of the axial
strength.

Based on the available data for the shear crack angle and deformations of the shear area the tensile strength of the
truss can be calculated:

VF ,exp = (atw ,v ⋅ f t ,res ,v + atw ,h ⋅ f t ,res ,h )⋅ z ⋅ cot β r ⋅ cos α (5)


74 Hegger et al.
where atw,v and atw,h are the cross-sectional area of the vertical and horizontal shear reinforcement, ft,res,v and ft,res,h are
the resulting strength of the reinforcement and z is the inner lever arm. Comparing the failure loads of the tests with
the calculated bearing capacity of the truss VF,exp clarifies that for low reinforcement ratios ωct beside the truss
contribution an additional load-bearing mechanism is existing. This is the concrete contribution which can be
explained by the shear capacity of the compression zone (Görtz 2004). The decreasing influence of the concrete
contribution with increasing mechanical reinforcement ratio ωct is explained by the weakness of the concrete
contribution compared to the increasing stiffness of the truss.

Further investigations concerning the bearing capacity of the concrete compression strut demonstrated that the
strength of the concrete has to be reduced compared to ordinary stirrup-reinforced concrete. The reason is the
development of transversal tensile stresses due to the weakness of the fiber material. Therefore, only about 40% of
the shear capacity calculated with the model described in (Görtz 2004) has been achieved in the tests with a failure
of the compression strut. Based on the results of the tests the shear capacity of TRC beams can be modelled with a
truss contribution VF and a concrete contribution Vc,f, which is mainly influenced by the load-bearing capacity of the
compression zone (Fig. 5).

Based on the described results of the investigations a design model has been derived. According to the model
described in (Görtz 2004, Hegger et al. 2005a) the shear capacity V consists of a truss contribution VF and a
concrete contribution Vc,f. In order to consider the effect of the mechanical reinforcement ratio ωct the factor κf is
established.

V = VF + κ f ⋅ Vc , f (6)

The truss contribution is subdivided into the shear resistance of the shear reinforcement and the concrete
compressive strut, in which the smaller value is decisive.

⎧(a fw,0° ⋅ f t ,res + a fw,90° ⋅ f t ,res ) ⋅ z ⋅ cot β r ⋅ cos α


VF = min ⎨ (7)
⎩0.3 ⋅ f cm ⋅ bw ⋅ z /(cot β r + tan β r )

The shear crack angle βr depends mainly on the mechanical shear reinforcement ratio ωct and is influenced by
additional normal stresses σx:

cotβr = 1+ 0.15 / ωw,ct – 0.18 ⋅ σx / fctm ≤ 2.15 (8)

The concrete contribution can be calculated as the shear strength of the unreinforced element according to the
German standard DIN 1045-1.
13
Vc , f = 1.5 ⋅ 0.1 ⋅ η1 ⋅ κ ⋅ (100 ⋅ ρ l ⋅ f cm ) ⋅ bw ⋅ d (9)

η1 = 1.0 , for normal concrete


200
κ = 1+ , where d is the distance between reinforcement and edge of the element
d
A fl
ρl = ≤ 0.02 , longitudinal reinforcement ratio
bw ⋅ d

As described above the part of the concrete contribution depends on the mechanical reinforcement ratio ωct Equation
(4). This effect is considered with the combination factor κf which is defined in Equation (10):

κf = 1 - ωw,ct / 3 ≥ 0 (10)
Textile-Reinforced Concrete 75
APPLICATIONS

Concrete is commonly applied for load-transmission purposes, especially in multi-storey construction. Perhaps the
most important property of concrete is its three-dimensional mouldability being scarcely exploited to achieve
intelligent load-bearing forms that minimize the use of materials. This was not always the case, however. Especially
from the 1940s to the 1960s, numerous load-bearing roof structures were created in which the economical use of
materials led to buildings of great structural intelligence and design elegance. The leading exponents of this type of
construction included architects and engineers like Pier Luigi Nervi, Felix Candela and Angelo Mangiarotti. One
thing their structures have in common is their extreme slenderness with thickness of 40 to 60 mm (1.6 to 2.4 in.) –
something that was no longer practicable in view of the minimum concrete covering now required to protect the
steel reinforcement against corrosion. The efforts of developing appropriate applications in TRC must aim to reach
this former intelligence and elegance.

The use of TRC for the skins of buildings is conceivable for rear ventilated façade systems as well as for load-
bearing and non-load-bearing sandwich elements. Indeed, this material has already been applied in the façade of the
laboratory building of the Institute of Structural Concrete, Aachen University (Hegger et al. 2004). On the
longitudinal side of the hall, curtain wall panels are used instead of natural stone, which would have been the typical
choice. The upper part of the façade is cladded with 2.685 m (8.8 ft) x 325 mm (12.8 in.) x 25 mm (0.98 in.) panels
(Fig. 6). The panels were produced lying horizontally. First, a 4 mm (0.16 in.) thick layer of self-compacting
concrete was poured in the formwork. Then the first textile reinforcement layer was placed, followed by the pouring
of the next 17 mm (0.67 in.) thick concrete layer and the placing of the second layer of the textile. Finally the last 4
mm (0.16 in.) thick concrete layer was casted. The difficulties in the exact positioning of the textile reinforcement
layer are one disadvantage of this very simple, however, labor intensive production process. The filigree panels are
an efficient and economic alternative for reinforced concrete or natural stone façades and open up new fields for the
application of concrete as building material. All solutions have in common that the beneficial properties of textiles
and concrete must be investigated in order to obtain durable components economical to manufacture, to install and
to maintain.

A different application for TRC is the use as non-ventilated façade elements. The manufacture of thin sandwich
units, consisting of two TRC layers with a high-quality surface finish and an intermediate layer of polystyrene rigid-
foam insulation, would increase the overall rigidity of the units. The use of such elements for load-bearing walls,
floors and roofs in one- and two-story buildings would thus be conceivable. Further research is needed in this area to
investigate aspects like impermeability, vapor diffusion and behavior under conditions of fire. A so-called system
house is currently being developed based on this work (Fig. 7). By using smaller cross-sectional dimensions, the aim
is to achieve a high functional efficiency in terms of building physics, combined with high-quality exposed concrete
surface and a short assembly time.

A further filed of application in which a reduction of weight can be effectively achieved by slender elements are
load-bearing systems. The use of diamond-lattice framing for arched forms is based on an efficient structural
principle that has been applied in hall construction for over one century by now. The diamond-shaped grid is a
geometrical expression of relatively closely spaced arch sheaves that intersect on the diagonal. As a structural
principle, it has its origins in forms developed in 1905 by Friedrich Zollinger for arched halls, where individual,
intersecting timber strips are used.

The efficiency of systems of this kind is based on the work with small, slender, individual components from which
entire structures can be assembled. The diagonal arrangement of the members automatically achieves a stiffening
effect in the longitudinal direction of the building. The often complex detailing at the points of intersection of the
stave members may be regarded as a disadvantage, however. In the constructional system developed by Zollinger,
for example, three linear elements meet at the nodes; in other diamond-lattice grid systems, there are usually four;
and in forms of construction based on a hexagonal geometry, there may be six elements that have to be jointed at
these points.

The diamond-lattice grid principle described above provides an opportunity to prefabricate slender, lightweight,
diamond-shaped components. In this way, the number of elements that have to be connected at the points of
intersection can be reduced to two. Furthermore, in regard to the thickness of the members (25 mm (0.98 in.)), the
use of simple bolted node connections is possible. This allows smaller arched structures, with spans from 8 to 15
76 Hegger et al.
meters (26.3 to 49.2 ft.), to be realized in a very simple manner. The extremely slender elements in textile-reinforced
concrete also result in a finespun appearance that was hitherto not associated with concrete (Fig. 8).

In context of the Collaborative Research Center 532 a diamond-lattice arch has been produced and erected in
February 2005 (Fig. 9) at RWTH Aachen University. The construction consisting of 36 rhombic elements has
spanned 10 meters (32.8 ft.). The elements measuring 1000 by 600 mm (39.4 by 23.6 in.) have been prefabricated in
a cnc-milled formwork.

The production of the formwork, providing the form with textile fabrics, casting the self-consolidating, fine-grained
matrix and assembling the prefabricated elements demonstrated that it is possible to produce accurate elements made
of TRC in a object-related serial production. It could be taken advantage of the plastic mouldability of the concrete
to design formatively fastidious junctures. The bolted joints have constituted an effective connecting device.

In the middle of the 20th century, a wide range of large-span, lightweight prefabricated shell-plate structures were
created for commercial buildings. It was possible to manufacture beams as single elements that spanned distances of
up to 24 m (78.7 ft.). Serial methods of fabrication on a largely mechanical basis facilitated extremely economical
forms of construction. Ribbed and T-beam slabs, folded section beams, channel-section folded beams, V-section or
wave-like shell beams and hyperbolic paraboloid shell roofs were commonly used.

To understand the nature of TRC and to develop appropriate applications for this new material, simple roof
structures are being developed at the Chair of Construction and Design. All forms of construction described are
distinguished by the extremely slender dimensions they allow. The load-bearing behavior is being optimized by
folding or curving the surfaces or by creating wave-like undulations and compression arch forms in cross section.

Even in reinforced concrete work, simple channel-section folded beams belong to the most economical types of
construction. With a material thickness of 25 mm (0.98 in.) in textile-reinforced concrete, provisional dimensions
can be calculated with a span of 9 m (29.5 ft) and a structural depth of 350 mm (13.8 in.).

The shell effect of thin concrete structures can be very effective in the case of barrel-shell roofs (Fig. 10). The shell
is rigid in both the longitudinal and lateral directions. TRC structures – with a thickness of 25 mm (0.98 in.) – are,
therefore, extremely lightweight. With a structural depth of about 500 mm (19.7 in.) and spans of up to 8 m (26.2 ft),
interesting applications for this type can be found in smaller and medium-sized hall structures.

Future examination will broaden the range of roof structures. On the one hand, for example an interesting object will
be the development of double-curved shell structures. On the other hand producing prototypes of selected
construction systems will show the requirements for further research. The load-bearing behavior, production,
transport and assembly can be optimized by applying semi finished parts to compose roof structures as well as
whole buildings. This way it becomes possible to work with relatively small components of a simple geometry (Fig.
11). Carbon slats or steel cables can be applied to brace the elements and to bear tension forces. Thereby a selection
of adequate applications in TRC can be optimized and be prepared for accomplishment.

SUMMARY AND CONCLUSIONS

The load-bearing behavior of TRC is influenced by material, amount and orientation of the textile reinforcement in
the composite. In particular, the bond behavior of the textiles is important for the load-bearing characteristics of
TRC. Models known from steel reinforced concrete cannot be directly applied to TRC without additional
considerations. This is because of the inhomogeneous cross sections of the rovings that lead to basically different
bond behavior compared to the reinforcing bars in steel reinforced concrete. However, design models for the tensile,
bending, and shear strengths of TRC elements have been derived by analogy with known models for steel reinforced
concrete. Therefore, extensive test series have been carried out. Significant results and conclusions from this
investigation are as follows:
• The fiber material as well as the binding and the cross section of the roving have significant influence on
the bond performance of the reinforcement.
• The cross-sectional area of the rovings cannot be fully activated due to the different bond performance of
the inner and the outer filaments.
• Biaxial tensile loading leads to decreasing strength of the reinforced element because of the formation of
Textile-Reinforced Concrete 77
cracks in loading direction. These cracks cause worse bond performance and damaging of the roving.
• Different directions of load and rovings cause loss of strength compared to axial loading mainly initiated by
the actions on the filaments at the crack edges. Presuming that the loss of strength increases linearly with
raising slope between loading and roving direction results in good accordance with test results.
• Bending loading causes transversal action on the reinforcement. In case of AR glass fabrics this leads to a
loss of strength. In contrast the strength of carbon fabrics in bending tests is higher compared to axial
tensile loading and increases with raising reinforcement ratio.
• The bending capacity of TRC-elements can be calculated by analogy with steel reinforced concrete with an
additional factor considering the effects of the curvature of the beam.
• The shear capacity of beams made of TRC cannot be described by a pure truss model. An additional
concrete contribution exists which is the capacity of the concrete compression zone. For the calculation of
the truss contribution the angular change of the shear reinforcement has to be considered. In the tests this
angular change amounted to 12 - 25° for the vertical rovings. In contrast to steel reinforced concrete the
horizontal reinforcement of the web has to be taken into account for the calculation of the shear capacity.
The proposed design model is in good accordance with the shear capacities measured in the tests.

Further investigations are required to confirm the presented results and to provide a basis for a general design
concept for TRC-elements. In addition, the influence of cyclic loading on the load-bearing behavior needs to be
investigated. These aspects are part of the current investigations at the Collaborative Research Center 532 at RWTH
Aachen University.

ACKNOWLEDGMENTS

The authors gratefully acknowledge the financial support of the Deutsche Forschungsgemeinschaft (DFG) in context
of the Collaborative Research Center (SFB) 532 “Textile Reinforced Concrete - Development of a new technology”.

REFERENCES

Brameshuber, W.; Brockmann, T.; Banholzer, B.: “Material and Bonding Characteristics for Dimensioning and
Modelling Textile Reinforced Elements, 2005,” Materials and Structures - accepted for publication

DIN-1045-1, 2001, Tragwerke aus Beton, Stahlbeton und Spannbeton, Teil 1: Bemessung und Konstruktion.

Görtz, S.: “Shear cracking behavior of prestressed and nonprestressed beams of normal and high performance
concrete”, Ph.D. Thesis, RWTH Aachen University, Germany, 2004.

Hegger, J.; Sherif, A.; Görtz, S., 2005a, “Shear capacity of beams made of high performance concrete”. Seventh
International Symposium on Utilization of High-Strength/ High-Performance Concrete, Washington D.C, USA,
2005. ACI SP-228-46.

Hegger; J. and Voss, S., 2005, “Design of Textile Reinforced Concrete Structures”, ConMat`05, Vancouver,
Canada, 22 - 24. August 2005, p. 299 and CD-ROM.

Hegger, J. and Voss, S.: “Textile Reinforced Concrete under Biaxial Loading”, 6th RILEM Symposium on Fiber
Reinforced Concrete (FRC), BEFIB 2004, Varenna (Italy), 20 - 22 September 2004, pp. 1463 - 1472.

Hegger, J.; Will, N.; Bruckermann, O.; Voss, S.: “Load-bearing behaviour of Textile Reinforced Concrete”, 2005,
Materials and Structures, in review.

Hegger, J.; Will, N.; Voss, S., 2004a, “Textile Reinforced Concrete Façades”. Concrete Structures: the challenge of
creativity, fib - Symposium, Avignon (France), 26 - 28. April 2004, S. 168 - 169 and on CD-ROM.

Jesse, F.: “Load Bearing Behaviour of Filament Yarns in a Cementicious Matrix”, Ph.D. Thesis, TU Dresden,
Germany, 2005.
78 Hegger et al.
Molter, M.: “Load-bearing behavior of Textile Reinforced Concrete”. Ph.D. Thesis, RWTH Aachen University,
Germany, 2005.

Ohno; S. and Hannant; D. J.: “Modelling the Stress-Strain Response of Continuous Fiber Reinforced Cement
Composites”, 1994, ACI Materials Journal, Vol. 91, No. 3, pp. 306-312.

Roye, A and Gries, Th.: “Tensile behavior of rovings, textiles and concrete elements - possible to compare
directly?” Proc. of Third International Conference Composites in Construction (CCC2005), Lyon, France, 2005;
1147-1154.

Vorechovsky, M; Chudoba, R; Konrad, M.: “Stochastic modeling of multi-filament yarns, part I, random properties
within the cross-section and size-effect.” International Journal of Solid and Structures, 2005, available online:
www. sciencedirect.com.

NOTATION

ARov = cross section of one roving


At = cross section area of the textile reinforcement
βr = shear crack inclination
fcm = concrete compressive strength
Fctu = tensile strength of the component
fct = tensile strength of the concrete
ft = tensile strength of the fiber from test on roving
k1 = factor accounting for the bond behavior
k0,α = factor accounting for the fiber orientation
kfl,ρ = factor accounting for bending
k2 = factor for biaxial loading
Mu = bending capacity
σmax = axial tensile strength of the fiber material in the composite
V = shear capacity
Vc,f = concrete contribution
VF = truss contribution
ωct = mechanical shear reinforcement ratio
z = internal moment arm
α = angular change of reinforcement
bw = width of the web of a beam
κf = combination factor for concrete contribution
κ = factor for seize effect

CONVERSION FACTORS
1 tex = 1 g/km
1 in. = 25.4 mm
1 ft = 0.3048 m
1 kip = 4.448 kN
1 ft-kip = 1.356 kN-m
1 psi = 6.89x10-3 MPa
Textile-Reinforced Concrete 79
Table 1 — Fabrics used for the investigations
Name Material Filaments Filament Mesh Cross-sectional
number diameter size area
(mm) (mm) (mm²/m )
(0.04 in.) (0.04 in.) (0.00047 in²/ft.)
MAG-07-03 AR glass 1300 0.029 8x8 107
MAG-04-03 Carbon 24000 0.007 8x8 108
MAG-05-03 Carbon 24000 0.007 8x8 54

Table 2 — Concrete properties


Name Compression strength Flexural strength Young’s Modulus
(MPa) (0.145 ksi) (MPa) (0.145 ksi) (MPa) (0.145 ksi)
PZ-0899-01 85 7.6 33400

Table 3 — Mix of fine-grained concrete


Cement Fly Silica Plasticizer Sand Quartz Water w/c
Ash Fume 0-2 mm flour
(kg/m3) (0.062 lb/ft)
490 175 35 10.5 714 499 245 0.4

load cell section A-A


110 (4.3)

(0.5) (3.8) (0.5)


12
A

120 (4.7)
18

96
strain gauge (0.7)
textile

12
A
110 (4.3)
lvdt
[mm]
300 (11.8) 300 (11.8) 300 (11.8) [(in.)]
[mm]
[(in.)]

Fig. 1 — Test setup of four-point bending test and profile geometry

Fig. 2 — Photogrammetry measurement system


80 Hegger et al.
concrete
e
e dg
ck
cra
roving
βr
dx
α
y
w v dy
x

Fig. 3 — Direction change of roving at the shear crack edges

c σc
Fc a
xd

N zf,flange zf,web
Mu xz —f,web Ff,web
f,web

f,flange
—max,bending Ff,flange

Fig. 4 — Model for calculation of bending capacity

Fig. 5 — Shear-carrying mechanisms


Textile-Reinforced Concrete 81

Fig. 6 — Façade made of TRC

Fig. 7 — System housing


82 Hegger et al.

Fig. 8 — Diamond-lattice grid

Fig. 9 — Prototype of diamond-lattice grid


Textile-Reinforced Concrete 83

Fig. 10 — Barrel shells

Fig. 11 — Segmented wave-form shells


84 Hegger et al.

You might also like