Вы находитесь на странице: 1из 7

Second International Conference on Science, Engineering & Environment,

Osaka City, Japan, Nov.21-23, 2016, ISBN: 978-4-9905958-7-6 C3051

STEEL FIBRE REINFORCED CONCRETE:


FROM X-RAY IMAGING OBSERVATION TO MODELLING
Trevor Htut1, Tian Sing Ng1 and Stephen Foster2
Department of Civil Engineering, Curtin University, Australia;
2
School of Civil & Environmental Engineering, University of New South Wales, Australia
1

ABSTRACT
In developing the constitutive model for steel fibre reinforced concrete, insight of fibre behaviour during the
pullout is vital. In a freely and randomly orientated multi-fibre composite, not all fibres are aligned in the direction
of the applied load; instead, fibres lie at various angles to the loading direction. This paper reports on the results
of X-ray imaging of steel fibre reinforced concrete in tension. In this study, uniaxial tension tests with multi-fibres
randomly distributed and oriented in dog bone shaped specimens were conducted. It is observed that crack paths
find ways of minimum resistance, through a section with poor fibre dispersion and divert around fibre ends, where
possible. In large scale structural members where crack locations are not fixed by either geometry or load state,
variability in fibre dispersion and fibre-matrix interaction should be considered at the model level, rather than
incorporated in materials or member safety factors.
Keywords: Steel Fibre, Tension, Dispersion, Distribution, X-Ray, Fibre-Matrix Interaction
agents [12] and fibre type and fibre orientation [2, 13].
In spite of a belief sometimes expressed [2] that no
correlation exists between the behaviour of a single
fibre pullout test and the behaviour of bulk fibres in a
real composite matrix, the effectiveness of a fibres as
a medium of stress transfer is often assessed using
fibre pullout tests where slip between the fibres and
the matrix is monitored as a function of the applied
load.

INTRODUCTION
In the early 1960's, pioneering research into fibre
reinforced concrete was undertaken by Romualdi and
Batson [1] where it was demonstrated that the tensile
strength and crack resistance of concrete can be
improved by providing suitably arranged, closely
spaced, wire reinforcement. After 50 years of
research in the development and placement of fibres
in reinforced concrete, the concept has matured to the
stage where it is finding increasing use in practice.
The materials used for fibres have also seen
significant advancements including stainless steels
and complex polymers. As early as 1994, Banthia and
Trottier [2] recorded that fibres are used as a form of
shear reinforcement in reinforced concrete (RC)
structural elements, for blast resistance in structures,
as shotcrete in tunnel linings, for use in slope
stabilisation works and to limit early age shrinkage
cracking in large concrete pavements.
By adding fibres to a concrete mix the objective is
to bridge discrete cracks providing for some control
to the fracture process and increase the fracture
energy. Since the early work, the pullout mechanism
of discontinuous fibres embedded in a variety of
cementitious materials has been studied by a number
of researchers [3-7].
The current understanding of the behaviour of
fibre-matrix interfacial mechanics is based on a
number of pullout studies using single or multiple
fibres where steel fibres are embedded within a
cementitious matrix. The experimental parameters
investigated include the rate of loading [8-9], curing
and environmental temperatures [2, 10], the quantity
and quality of the matrix [11], addition of adhesive

EXPERIMENTAL PROGRAM
The study consists of nineteen 30 mm thick dogbone shaped specimens with randomly distributed
25 mm long by 0.3 mm diameter end hooked fibres
with nominal yield strength of 2300 MPa and volume
of between 0.5% and 2% were cast and eleven
specimens were X-ray imaged prior to tensile testing.
The mortar mix used in the specimens was
composed of kiln dried Sydney sand and general
purpose Portland cement mixed with a sand and
cement ratio of 3:1 and water-cement ratio of 0.4. No
other additives were added in the mix design.
The testing arrangement is shown in Fig. 1. The
displacements in the direction of movement of the
loading jacks were measured using two linear
variable differential transducers (LVDTs), one placed
on each side of the specimen to allow for the X-ray
film to be attached. The displacement plotted is taken
as the average of the LVDT readings. Loading was
conducted using displacement control at a rate of
0.12 mm per minute up to the crack opening
displacement of 2 mm. The rate was then increased to
a minimum of 0.1 mm per minute, with further rate
increases introduced as necessary during the test.
1

SEE-Osaka, Japan, Nov. 21-23, 2016

Each test specimen was loaded until the fibres were


pullout completely from the section. Load and
displacement readings were recorded at 1 second
intervals.

presented in Fig 5.

3.0

200

Epoxy

0.5%

2.0
1.5

Tensile Stress (MPa)

2.5

Tensile Stress (MPa)

35

3.0

UT-Hom-0.5% (3)
UT-Hom-1.0% (2)
UT-Hom-1.5% (3)
UT-Hom-2.0% (3)

2.5
2.0

1.5%

1.5

0.5%

1.0

1.0%

0.5

2.0%

0.0

Note: Number of specimens averaged in ( )


0.0

1.0

0.5
1.0
1.5
Crack Opening Displacement (mm)

2.0

1.0%
0.5

1.5%

R1
45

2.0%

0.0
0.00

chemical
anchors

1.25

2.50

3.75

5.00

6.25

7.50

Crack Opening Displacement (mm)

Figure 2: Plot of average tensile stress versus crack


opening response for unixail tension test

35

LVDT
Epoxy

P
200
Fig. 1

Fig. 3

Test arrangement for uniaxial tension test

TEST RESULTS & ANALYSIS


In this study, a total of 19 tests were conducted of
which 7 were conducted under x-ray observations.
The average load versus displacement results are
presented in Fig. 2 and it shows that for the 25 mm
long high yield strength fibres used in this study, there
was no increase in performance from a 0.5% fibre
volume to a 2.0% fibre volume.
The images taken prior to testing were then
analysed for fibre concentration over various regions
(Fig. 3); to avoid the wall effect in the analyses, the
region near the side walls was excluded.
The digital image analysis was undertaken using
IMAQTM Vision Builder software. Each sample
image was filtered to distinguish the fibres from the
background image (Fig. 4). An image analysis was
then undertaken to determine the area of fibres in the
image (white area in Fig. 4b) with the fibre
dispersion/distribution factor (Ffd) defined as the ratio
of white area to the total sample area. Only a limited
improvement in result of the analysis (approximately
2%) was observed when the sampling size was
reduced much below the length of the fibres and a
sampling size of 25 mm 25 mm was adopted for
further investigation. The dispersion results are

Sampling locations for fibre dispersion


analysis: 12.5 mm 12.5 mm and 25 mm
25 mm.

(a)

(b)
Figure 4: Digital X-ray image along the crack path:
(a) before the image analysis and (b) after the
image analysis.
For a sample of statistical significance, the median
value of Ffd for samples taken within one dog-bone
specimen represents the average fibre volume
fraction, f. This data is plotted in Figure 6 for the
dog-bone shaped specimens for the different, known,
fibre volumetric ratios (f from 0.5% to 2.0%). With
the relationship Ffd and f established, for the given
fibre type and specimen thickness, (Fig. 6), the
volume fraction as a function of Ffd is determined as:

SEE-Osaka, Japan, Nov. 21-23, 2016

= (126 136 )

Relative Frequency (%)

35

(fibre dispersion factors of approximately 0.375 to


0.677). Beyond the limits of the data, extrapolation is
needed and the results are less certain.

(1)

f = 0.005

30

0.025

20

0.020

15
10

0.015

5
0
0.000

0.010

0.002

0.004

0.006

0.008

0.010

Variation of f

0.005

(a)
Relative Frequency (%)

20

f = 0.010

0.000
0.0

15

0.2

0.3

0.4

0.5

0.6

0.7

Fibre Dispersion Factor, Ffd

Figure 6: Fibre volume concentration ratio versus


average fibre dispersion along the crack
path.

0.006

0.009

0.012

0.015

0.018

From the fibre dispersion data, the standard


deviation for the test series was = 0.27f, and,
considering the fibres to be normally distributed, the
75th and 90th percentiles are 75 = 0.82f and 90 =
0.65f , respectively.
It is important to note that, in this study, the
specimen thickness was 30 mm or 100 times the
diameter of a single fibre. Fibres are more likely to
overlap one another in the digital image analysis as
the specimen thickness increases. Consequently, in
the digital image analysis, specimen thickness is a
significant component in determining of the fibre
dispersion factor. For the hypothetical case of a fibre
diameter thick specimen, there is no possibility of
fibre being overlapped and the fibre dispersion factor
is linearly proportional to the fibre volume
concentration. Probability of fibre overlapping is a
function of the specimen thickness, and, consequently,
a nonlinear relationship is observed.

0.021

Variation of f

(b)
20

Relative Frequency (%)

0.1

10

0
0.003

f = 0.015

15

10

0
0.004

0.008

0.012

0.016

0.020

0.024

0.028

Variation of f

(c)
15

Relative Frequency (%)

fd
Mean Fibre Dispersion Factor, Eq. (1) f
126 136 F fd
th
90 percentile characteristic value
th
75 percentile characteristic value

Fibre Volume Concentration , f

25

f = 0.020

10

Influence of Fibre Dispersion


0
0.005

0.010

0.015

0.020

0.025

Variation of f

0.030

It was observed that the crack initiation is likely


to commence at the point of minimum resistance and,
consequently, cracks initiate from the location at
which least number of fibres are located; that is,
cracks initiate from areas with poor fibre dispersion
rather than at the narrowest part of the dog-bone
specimen. Thus, fibre dispersion plays a significant
role in crack initiation and, consequently, on the
tensile strength.
The dispersion of fibres in the matrix and
deformation of the end-hooks significantly influences
the tensile behaviour of a SFRC composite. In the
development of material models for design, this

0.035

(d)
Figure 5: Distribution of fibres for fibre volume
percentage (f) of 0.5, 1, 1.5 and 2.
With the fitting Eq. (1) established, variations in the
fibre volume fraction within one specimen can then
be determined. Note that the best predictions from Eq.
1 are when the results are interpolated, that is between
fibre volume fractions of between 0.005 and 0.02
3

SEE-Osaka, Japan, Nov. 21-23, 2016

observation needs consideration.


The digitised X-ray images taken during the
tension test of the dog-bone specimens show the
importance of fibre dispersion on the crack
formation/initiation and propagation processes. To
validate the findings presented above, further digital
image analysis was undertaken to determine the fibre
dispersion along the crack path of the specimens
containing fibre volume concentrations of 0.5%,
1.0% and 1.5%. A typical X-ray image around the
crack path is shown in Fig 4a. Digital image analysis
was undertaken on a sample size of 12.5 mm square
(Fig 4b) and the fibre volume ratio versus fibre
dispersion ratio is plotted in Fig 6.
The results show that the cracks are likely to form
or propagate along the path of least resistance. The
fibre volume concentration along the crack path was
found to average through the 75th percentile
characteristic value (Fig 6). This confirms the
conclusion that fibre dispersion contributes
significantly to the fracture process in uniaxial
tension and this observation needs to be taken into
consideration during the development of behavioural
models.
Based on clear evidence from his X-ray imaging,
Htut [14] noted that the dominant crack forms where
the local fibre concentration is at its lowest across a
section and where near the end of a fibre, deflects
around it. When specimens have a dominant notch,
however, the crack path forms at the notch and the
impact of fibre dispersion is negated [16].
Consequently, fibres have a higher possibility of
being fully deformed in the notched section tests and,
thus, higher tensile strengths and ductility are
observed. On the other hand, if the notch is not
dominant, failure may occur away from the notch at a
section where low fibre numbers, through dispersion
variations, occur. As a part of this testing programme,
uniaxil tension testing was conducted on three of his
dog-bone shaped specimen that had a pair of 5 mm
notches at the critical section, reducing the crosssectional area by 9%. In each case, it was observed
that the dominant crack formed well away from the
notched section. From the observations with regards
to the influence of fibre dispersion of the previous
section, it may be postulated that a cross-sectional
area reduction of the order of 20% would be needed
for a reasonable probability that cracking would occur
through the notched section.

adverse effect on the tensile performance of a fibre


reinforced composite. In this study and Markovic et
al. [16] observed that, the crack path follows the
easiest propagation route and is often near the end of
fibres or around them (Fig 7). Consequently, many of
the end-hooked fibres fail to engage and do not
deform during the fracture process.

Fig. 7 Crack propagation during a uniaxial tension


test
Figures 8 shows the final fracture (or failure)
plane of the dog-bone shaped specimens with various
fibre volume concentrations. Based on visual
observations, a higher proportion of fibres were
deformed (i.e. straightened) in the case of specimens
with fibre volume concentration of 0.5%. It was
observed that an increase in fibre volume
concentration lead to a decreased in proportion of
end-hooked fibres straightening through the endhook (Fig. 8). Rather, failure of the matrix
surrounding the hook significantly influenced the
behaviour. Similarly, it is observed that the shape and
contour of the failure plane is dependent on the fibre
volume concentration. Figure 8(a) shows that
relatively flat and even fracture plane was observed
in the case of specimen with 0.5% fibre volume
concentration. An increase in fibre volume
concentration, however, resulted in a rotation of the
fracture plane (Fig. 8b) and it was observed to be a
function of fibre volume concentration (Fig. 8c & d).
In order to investigate the phenomena discussed
above, the end-hooked fibres across the cracking
planes were counted after the tests were conducted
and the numbers of fibres that had, and had not,
straightened through the end hook were noted. A total
of eight dog-bone shaped specimens were studied in
this process and the results are presented in Fig 9. The
results support the observation that the proportion of
straightened end-hooked fibres is inversely
proportional to the fibre volume ratio.
Figure 9 shows that for the specimen containing
0.5% of fibre volume concentration, approximately
45% of fibres were deformed during the uniaxial
tension tests, which is in agreement with the findings
of Markovic et al. [17]. As the fibre volume ratio
increases, more energy is required to fully deform the
end-hooks sufficiently to pull through the fibre tunnel.
It was observed that due to the higher energy required
to fully deform the end-hooked fibres, cracks
propagate around or near the end of the end-hooked
fibres, which results in an uneven crack surface (Fig.

Influence of Fibre and Matrix Relationship on


fracture process
It is a common assumption that the improvement
in tensile performance of a fibre composite can be
achieved by incorporating a high fibre volume
fraction, especially in the case of steel fibres [17].
In this study, however, it was observed that
increases in fibre volume concentration can have an
4

SEE-Osaka, Japan, Nov. 21-23, 2016

8b to c). The additional energy in the matrix cracks


going around the fibre ends is more than offset by the
loss of efficiency of the fibres being pulled through
the fibre tunnel.

concrete.
The pullout load, Pf, of an individual fibre can be
determined from

Pf b la d f

(2)

For multi-fibres, the strength assuming failure is


by fibre pullout is

P b d f

(a)

la
i 1

where N is the number of fibres crossing the cracking


plane, b is the average fibre bond stress, la is initial
fibre embedment length and df is fibre diameter.
On the other hand, the matrix tensile capacity may
be calculated as

(b)

Pc k0 fct bD

(d)
Figure 8: Fracture plane of dog-boned shaped
specimens for fibre volume concentration of
(a) 0.5%, (b) 1.0%, (c) 1.5% and (d) 2.0%
Fully and/or Partially Deformed
End-Hooked Fibres (%)

b d f

End-Hooked Fibres
lf = 25 mm
df = 0.3 mm

50

fy = 2300 MPa
Cement Mortar
fcm.ave = 41 MPa

40

(4)

where k0 is a factor to account for the energy required


to deflect the crack around the fibre ends. For
argument we shall conservatively assume k0 as unity.
It is hypothesised that fibre pullout will occur
when the matrix strength, Pc is greater than the fibre
pullout load, Pf. If the tensile resistance provided by
the fibre is greater than that provided in a path through
the concrete, failure will occur through the concrete,
which is undesirable.
Based on the hypothesis presented, fibres will pull
out from the matrix without cracks being propagating
around the ends of fibres provided that:

(c)

60

(3)

la k0 fct Db

(5)

i 1

For the case of randomly distributed fibres in a


cementitious matrix, the average length of initial
embedment across a crack is given by Marti et al. [18]
as

30
20

la.ave

10

lf
4

(6)

0
0.0

0.5
1.0
1.5
Fibre Volume Concentration (%)

2.0

Aveston and Kelly [19] showed that for a crack


plane with a cross-sectional area of Ac, the total
number of randomly distributed fibres in the threedimensional space crossing the plane is:
A n fi f Ac
N c

(7)
2 i 1 A fi 2 A f

Figure 9: Plot of fibre volume concentration versus


surveyed fully and/or partially deformed
end-hooked fibres after the tests

The importance of the inter-relationship between


the mechanical properties of fibres and the
mechanical properties of the matrix can be used to
optimize the performance of steel fibre reinforced

Substituting Eqs. (6) and (7) into Eq. (5) gives the

SEE-Osaka, Japan, Nov. 21-23, 2016

minimum tensile strength of the matrix, fct, to ensure


fibre pullout action as

fct

b f f
2k0

REFERENCES
[1] Romualdi JP and Batson GB, Behaviour of
reinforced concrete beams with closely spaced
reinforcement, ACI Journal, Vol. 60, June 1963,
pp. 775-789.
[2] Banthia N and Trottier JF, Concrete reinforced
with deformed steel fibres, Part I: Bond-slip
mechanisms, ACI Materials Journal, Vol. 91,
Sep. 1994, pp. 435-446.
[3] Gray RJ, Analysis of effect of embedded fibre
length on fibre debonding and pull-out from an
elastic matrix. Part 2 - application to steel fibrecementitious matrix composite system, Journal
of Materials Science, Vol. 19, May 1984, pp.
1680-1691.
[4] Gopalaratnam VS and Shah SP, Tensile failure
of steel fibre-reinforced mortar, Journal of
Engineering Mechanics, Vol. 113, Jan. 1987, pp.
635-652.
[5] Mandel JA, Wei S and Said S, Studies of the
properties of the fibre-matrix interfaced in steel
fibre reinforced mortar, ACI Materials Journal,
Vol. 84, Mar. 1987, pp. 101-109.
[6] Namur GG and Naaman AE, A bond stress
model for fiber reinforced concrete based on
bond stress slip relationship, ACI Materials
Journal, Vol. 86, Jan. 1989, pp. 45-57.
[7] Wang Y, Li CV and Backer S, Experimental
determination of tensile behavior of fiber
reinforced concrete, ACI Materials Journal,
Vol. 87, Sep. 1990, pp. 461-468.
[8] Hughes BP and Fattuhi NI, Fiber bond strength
in cement and concrete, Magazine of Concrete
Research, Vol. 27, Sep. 1975, pp. 161-166.
[9] Banthia N and Trottier JF, Deformed steel fibercementitious matrix bond under impact, Cement
and Concrete Research, Vol. 21, Jan. 1991, pp.
158-168.
[10] Banthia N and Trottier JF, Effects of curing
temperature and early freezing on the pullout
behaviour of steel fibres, Cement and Concrete
Research, Vol. 19, May 1989, pp. 400-410.
[11] Gray RJ and Johnston CD, Effect of matrix
composition on fibre/matrix interfacial bond
shear strength in fibre-reinforced mortar,
Cement and Concrete Research, Vol. 14, Mar.
1984, pp. 285-296.
[12] Guerrero P and Namman AE, Effect of mortar
fineness and adhesive agents on pullout response
of steel fibres, ACI Materials Journal, Vol. 97,
Jan. 2000, pp. 12-20.
[13] Naaman AE and Shah SP, Pullout mechanism
in steel fibre-reinforced concrete, ASCE Journal
of the Structural Division, Vol. 102, Iss. 8, 1976,
pp. 1537-1548.
[14] Htut TNS, Fracture Processes in Steel Fibre
Reinforced Concrete, PhD Thesis, School of

(8)

and, similarly, an upper limit of fibre volume (or


optimum fibre volume) to ensure that a failure plane
passes through the fibres, and not around the fibre
ends can be determined as

f .optimum

2k0 fct

b f

(9)

where f is the fibre volume friction and f is the fibre


aspect ratio (lf / df).
The average bond stress (b) is calculated from
discrete fibre pullout tests as

Pf .max

d f la

(10)

where Pf.max is the maximum force in the fibre and la


is the embedded length of the fibre at the point of
engagement. In the development of Unified Variable
Engagement model by Htut [14], the point of
engagement is taken as the crack (or separation plane)
opening displacement corresponding to a force in the
fibre of 0.5Pf.max. Thus, in Eq. (10), la = lf/2-we , where
we is the separation plane displacement at the point of
engagement.
CONCLUSION
While the tensile and fracture behaviour of steel
fibre reinforced concrete have been researched for
nearly five decades, their use in structures has been
limited by a lack of design models and
standardization. With design rules for SFRC
introduced in some national concrete structures
standards, it could be expected that more use will be
made of this higher performance material in building
and bridge structures for the carrying of tensile
stresses. Ideally, design models should be based on a
physical understanding of fracture processes and
physical-mechanical models adopted as the model
basis, rather than empirically based modelling
approaches. Through X-ray imaging of tensile tests,
pre-loading and under load, the effect of fibre
distribution, dispersion and influence of fibre-matrix
relationship on the fracture processes is quantified.
Crack diversion around fibre ends is also observed
and needs consideration in the development of
rationally based, physical, design models.

SEE-Osaka, Japan, Nov. 21-23, 2016

Civil & Environmental Engineering, The


University of New South Wales, 2010, pp. 389.
[15] di Prisco M, Plizzari G and Vandewalle L, Fibre
reinforced concrete: new design perspectives,
Materials and Structures. Vol. 42, Nov. 2009,
pp.1261-1281.
[16] Markovic I, Walraven JC and Van Mier JGM,
Proceeding of the 5th International Conference
on Fracture Mechanic of Concrete & Concrete
Structure, 2006, pp 1113-1120.
[17] Li Z, Li F, Chang TYP and Mai YW, "Uniaxial
tensile behaviour of concrete reinforced with
randomly distributed short fibres", ACI

Materials Journal, Vol. 95, Sep. 1998, pp. 564574.


[18] Marti P, Pfyl T, Sigrist V and Vlaga T,
Harmonized test procedures for steel fibrereinforced concrete, ACI Materials Journal,
Vol. 96, Nov. 1999, pp. 676-685.
[19] Aveston J and Kelly A, "Theory of multiple
fracture of fibrous composites", Journal of
Materials Science, Vol. 8, Mar. 1973, pp. 352362.
[20] Ng UVEM

Вам также может понравиться