You are on page 1of 16

Construction and Building Materials 404 (2023) 133158

Contents lists available at ScienceDirect

Construction and Building Materials


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

Development of engineered cementitious composite with moderate tensile


properties using polyester fibers
Subodh Kumar , Durgesh C. Rai *
Department of Civil Engineering, Indian Institute of Technology Kanpur, Kanpur, UP 208016, India

A R T I C L E I N F O A B S T R A C T

Keywords: In recent years, the use of reinforced cementitious composites has increased manifold for structural rehabilitation
Engineered Cementitious Composites purposes. The present study concerns the development of a special variant of ECC, which is low-cost but has
Polyester moderate tensile strength and ductility. Widely available polyester fiber was chosen, which was about ten times
Fibers
cheaper but about four times weaker than the preferred but expensive PVA fibers. Micromechanics-based
Micro-mechanics
Multiple Cracking
mathematical model was used for fiber optimization to attain the multiple and strain hardening phenomena
of the ECC. Different ECC mixes were prepared to investigate the effect of the matrix toughness on the tensile
properties of ECC. ECC, with a strong matrix, failed to attain multiple cracking after 28-days of tensile testing
because the fracture toughness exceeded the fiber bridging energy. The matrix was made weaker by increasing
the amount of fly ash, and the desired strain hardening phenomena was achieved after 28-days of tensile testing.
A low-cost ECC was developed, which used 3% polyester fiber by volume; the tensile strength and strain capacity
of ECC was 2.17 MPa and 1.88 %, respectively.

1. Introduction cracking. Since its development in the 1990 s by Victor Li and others,
ECC has been extensively researched [14,15], and ECC is shown to
In recent years, the use of innovative materials such as fiber- exhibit a tensile strain capacity from 3% to 5%, which is 300–500 times
reinforced polymer (FRP), textile-reinforced mortar (TRM), fabric- the ordinary concrete (about 0.01%) [16–19]. During multiple micro­
reinforced cementitious composites (FRCM), and engineered cementi­ crack development, the tensile strength increases, and the localized
tious composites (ECC) has been increased by manifold in the retrofit­ catastrophic brittle failure is suppressed, resulting in strain-hardening
ting of unreinforced masonry structures[1–9]. A good composite action behavior. Because of its high tensile ductility and narrow crack width,
between the retrofitting material and the substrate material is desired ECC overcomes many of the problems associated with normal concrete
for the better performance of the structure under different loading. materials, such as brittleness and cracking damage. ECC can be gainfully
Many existing retrofitting materials involve a tedious manufacturing employed to improve the structural and durability performance of the
process, and their application is time-consuming. To address these infrastructure [20,21]. Furthermore, the strain hardening and the mul­
inherent shortcomings of existing strengthening techniques, researchers tiple cracking properties make the ECC a suitable material for retrofit­
have proposed new innovative materials for retrofitting URM walls, such ting and strengthening deficient structures.
as ECC [7–9]. In recent years, the field application of the ECC as retro­ The strain hardening and the multiple cracking properties of ECC are
fitting material has increased, for example, the use of ECC for dam tailored using mathematical models based on micromechanics and
repairing, coupling beams in high-rise buildings, and bridge deck link fracture mechanics theories. ECC was initially developed with poly­
systems [10,11]. Under direct tension or flexural load, normal concrete ethylene (PE) fibers that exhibited significant strain-hardening behavior
experiences sudden loss in its load-carrying capacity after the formation [15,22]. However, the PE fiber-based ECC could not be used in large-
of the first crack, and brittle failure is observed [12], whereas Fiber- scale construction due to the high cost associated with PE fiber. An
reinforced concrete (FRC) experiences stress softening after first alternative fiber, Polyvinyl alcohol (PVA) fiber, was later used to pre­
cracking [13]. ECC is a type of fiber-reinforced cementitious composite pare ECC as it was eight times cheaper than PE fibers, which reduced the
(FRCC) that has ductile and strain-hardening properties due to multiple cost of ECC material to a greater extent [23]. PE fibers are hydrophobic

* Corresponding author.
E-mail addresses: subodh@iitk.ac.in (S. Kumar), dcrai@iitk.ac.in (D.C. Rai).

https://doi.org/10.1016/j.conbuildmat.2023.133158
Received 12 January 2023; Received in revised form 24 August 2023; Accepted 26 August 2023
Available online 13 September 2023
0950-0618/© 2023 Elsevier Ltd. All rights reserved.
S. Kumar and D.C. Rai Construction and Building Materials 404 (2023) 133158

in nature, whereas PVA fibers are hydrophilic in nature due to the 2. Experimental studies
presence of the hydroxyl (–OH) group, which results in high adhesion to
the cementitious matrix. Excessive adhesion between the fiber and the 2.1. Raw materials
cementitious matrix can lead to fiber rupture and loss of fiber-bridging
capacity, which can result in the reduction in strain capacity of ECC The raw material used in preparing the ECC includes Portland
below 1% and a huge variation of mechanical properties. The strain Pozzolana cement (PPC), fly ash (FA), river sand, water, low-cost
capacity of up to 5% can be achieved by deliberate surface coating of the polyester fibers, and polycarboxylate-based high-range water-reducing
PVA fibers using an oiling agent (typically 1.2% oil coating by fiber admixture (HRWRA). The low-cost polyester used in the present study
weight) [24]. The oil coating on the surface of the PVA fibers weakens was available in the open market under the ‘Recron 3S’ brand name
the bond between the fiber and the cementitious matrix, which facili­ (Manufactured by Reliance Industries Pvt. Ltd). The ECC mixture was
tates the pull-out of the fiber that yields the desired tensile behavior of also prepared for comparison purposes using a good quality PVA fiber
ECC. supplied by Kuraray Co. Ltd. (Japan), which was approximately ten
Though PVA fibers are cheaper than PE fibers, PVA at 2% of the times costlier than the polyester fiber in the local market. Most of the
volume fraction accounts for approximately 60–80% of the total cost of research available on ECC materials uses Ordinary Portland cement
the material. The material cost of the ECC may even go higher in (OPC), which is not easily available in the local market for sale; hence,
countries where these fibers are to be imported; for example, in India, Portland Pozzolana cement (PPC) has been selected for the current
these fibers cost about 44 US $ per kg, which is very expensive compared study. Locally available river sand was used as fine aggregate, and the
to 6.6$ per kg in USA [23]. In order to reduce the material cost, poly­ sand was sieved with a sieve size of 1.18 mm. Finer particles of fly ash
propylene (PP) fibers with lower elastic modulus have been investigated were used to densify the matrix, which also helped in enhancing the
as an alternative to the PE/PVA fibers, and it was shown that the tensile aggregate-matrix interface. The particle size distributions and the me­
strength of the PP-ECC was relatively low (2 MPa) [21]. The recycled dian diameter (D50) of the binder materials (cement, fly ash) and sand
polyethylene terephthalate (PET) fibers have been used with PVA fibers are presented in Fig. 1.
with 50% replacement to prepare PET-PVA hybrid ECC that showed The X-ray diffraction (XRD) analysis was performed to investigate
tensile strength and tensile strain of 3.6 MPa and 2.2%, respectively the main reactive minerals present in the PPC and fly ash used in this
[18]. The steel fibers can improve ductility and post-cracking strength; study (Fig. 2). The main crystalline phase identified in the fly ash was
however, the required dosage of the steel fibers cannot be added due to Quartz (SiO2), Mullite (3Al2O3.2SiO2), and Hematite (Fe2O3). Table 1
workability issues [25,26]. Apart from the workability problem, the presents the chemical compositions of the PPC and the fly ash measured
steel fibers are more prone to corrosion, which may compromise the by X-ray fluorescence (XRF) test. The fly ash was confirmed to be class F
durability of the ECC prepared using steel fibers. There are very few fly ash as per ASTM C618 [29]. The physical and mechanical properties
studies conducted on the development of ECC using polyester fibers. of the PPC cement, fly ash, and sand are presented in Table 2. The PPC
Polyester or polyethylene terephthalate (PET) fibers are easily available used in the study was confirmed from IS:1489 (Part 1) [30], and its
in India at a much lower cost, about one-tenth of the PVA fiber cost (US fundamental properties are presented in Table 2.
$4.8/ kg), and are widely used for crack control in concrete structures. The mixed proportions of the ECC using polyester and PVA fibers are
Polyester (PET) fibers have very low tensile strength and elastic presented in Table 3. Three different ECC mixtures were adopted in the
modulus. However, they are a cost-effective alternative to PVA-based study with three levels of the Polyester fiber dosage to investigate the
ECC, which has been extensively studied and used in real-life applica­ feasibility of the multiple cracking and strain hardening phenomena.
tions. A cost-effective PET-based ECC can be easily prepared using The nomenclature of the ECC matrix was adopted based on the dosage of
locally available materials and used for various structural applications fiber, fiber type, and fly ash. For example, the matrix M1.2R1.5 refers to
such as retrofitting masonry structures [27]. A study by Fenwick and the matrix with a fly ash-to-cement ratio of 1.2, and R represents the
Dhakal (2007) examined the test results of beams, columns, and walls polyester fibers of Recron 3S make with 1.5% volume fraction. Similarly,
and found that for nominally ductile beams, the limiting value of the M0.5 K2 represents mix with fly ash to cement ratio 0.5, and K refers to
compressive strain in the concrete and the tensile strain in the rein­ the PVA fibers with Kuraray make. The dosage of HRWRA was adjusted
forcing bars were 0.004 and 0.016, respectively [28]. Materials with a to get a workable mix, as the workability of the mix was reduced
tensile strain capacity of up to 2% can be used for satisfactory structural significantly after the addition of polyester fibers.
performance. The ECC mix M1.2R1.5 was prepared as a trial mix to investigate the
The main objective of the present study is to develop a low-cost ECC micromechanical parameters of the polyester fiber of Recron 3S make.
using polyester (PET) fibers of moderate tensile strength and moderate The elastic modulus and tensile strength of the polyester fiber were
tensile capacity for easy and economic structural applications such as marginally lower than the PVA fibers. The polyester fibers were found to
the strengthening of unreinforced masonry elements. The existing be soft after a physical inspection, whereas the PVA fibers were stiffer
micromechanical model-based design methods have been used in (Fig. 3). Most synthetic fibers have circular cross-sections, whereas the
developing ECC using polyester (PET) fibers. As the interfacial proper­
ties of the polyester fibers were not known, this study will investigate the
interfacial parameters of the ECC with polyester fibers using experi­
ments. The present study also investigates the optimum dosage of the
polyester fiber for multiple cracking and strain hardening phenomena,
considering the workability of the ECC mix. The role of matrix toughness
in the tensile ductility behavior of the ECC will be investigated using
different proportions of the raw materials. The behavior of the different
ECC mixes with different fiber volume fractions has been studied under
uniaxial tensile and compressive loads. The present study intends to
propose a feasible process for preparing ECC, which can be easily
applied in the field without any laboratory-like sophisticated facility.

Fig. 1. Particle size distribution of the materials.

2
S. Kumar and D.C. Rai Construction and Building Materials 404 (2023) 133158

Fig. 2. XRD characterization of PPC cement and Fly ash.

Table 1
Chemical Composition (wt%) of PPC and Fly Ash measured by XRF.
Materials SiO2 Al2O3 Fe2O3 CaO MgO SO3 TiO2 P2O5 Na2O K2 O NiO MnO LOI

PPC 14.04 3.73 5.02 65.5 3.27 2.27 0.70 0.07 0.28 0.66 – 0.04 4.80
Fly Ash 55.04 31.84 4.65 1.18 0.30 – 2.65 0.45 0.04 1.15 0.01 0.04 1.20

mixer of 5-Litre capacity, which can operate at two different mixing


Table 2
speeds. The mixing sequence in the preparation of the ECC mixture can
Physical and mechanical properties of cement, Fly Ash, and sand.
be found in Fig. 4a. The dry powder of the cement, sand, and fly ash was
Raw Material Properties Value first mixed for 2 min at low speed (140 ± 5 rev/min). Thereafter, 90 %
Cement Specific gravity (g/cm3) 2.90 of water and HRWR admixture were added and mixed for 2 min at low
Initial setting time (min) 135 speed, and the remaining water was added and mixed for another 2 min
Final setting time (min) 335
at low speed. The mixer was rotated at high speed (285 ± 10 rev/min) to
Volume expansion (mm) 1.0
3-day compressive strength (MPa) 28.5 get a homogeneous and flowable mix. The fiber was added to the
7-day compressive strength (MPa) 32.7 mixture in two stages, with 50 % in each stage (Fig. 4b) within two
28-day compressive strength (MPa) 44.5 minutes, and the mixing speed was kept low at each stage. The mixer
was finally rotated at high speed for 5 min to ensure the proper
Fly Ash Specific gravity (g/cm3) 2.10 dispersion of the fibers within the cementitious matrix. The flow table
Sand Specific gravity (g/cm3) 2.60 test was performed in accordance with IS: 4031 (part7) [31] to deter­
mine the workability of the prepared ECC mixture, as shown in Fig. 4(c).
The moulds were prepared for the tensile coupons in accordance with
polyester fibers used in the present study have triangular cross-sections.
the JSCE (2008) [32]. The wet ECC mixture was then filled in these
The mechanical properties of the fiber used in the study are listed in
moulds, as shown in Fig. 4(d).
Table 4.
The top surface of the specimen was covered with a jute bag, which
can restrict the loss of water. The specimen should have enough strength
2.2. Specimen preparation at the time of demoulding so that it can resist the force applied to the
specimen during the demoulding process. The specimen was demoulded
A mixing procedure was adopted to ensure easy mixing and obtain a after 3–4 days of casting (Fig. 4e) as the rate of hydration of ECC was
workable ECC mix. The mixing was done using a countertop mortar

Table 3
Mixture proportion of ECC (by weight).
Mix Cement (PPC) Sand Fly ash Water HRWRA % Fibre Fiber Fiber manufacturer
(by volume)

M1.2R1.5 1 0.8 1.2 0.4 0.02 1.5% Polyester Recron 3S (Reliance)


M0.5R2 1 0.5 0.5 0.4 0.02 2.0 % Polyester
M0.5R2.5 1 0.5 0.5 0.4 0.02 2.5 % Polyester
M0.2R3 1 0.5 0.2 0.4 0.02 3.0 % Polyester
M1.5R2 1 0.6 1.5 0.65 0.02 2.0 % Polyester
M1.5R3 1 0.6 1.5 0.65 0.02 3.0 % Polyester

M0.5K2 1 0.5 0.5 0.4 0.01 2.0 % PVA Kuraray


M1.5K2 1 0.6 1.5 0.65 0.02 2.0 % PVA

3
S. Kumar and D.C. Rai Construction and Building Materials 404 (2023) 133158

Fig. 3. Different fibers used for preparing ECC (a) low-cost polyester fibers (b) PVA fibers.

Table 4
Properties of polyester fiber and PVA fiber provided by the manufacturer.
Fibers Density (g/cm3) Tensile strength (MPa) Elastic modulus (GPa) Elongation (%) Length (mm) Diameter (μm)

Polyester (Recron) 1.36 460 5 20 12 30


PVA (Kuraray) 1.40 1600 22 6 12 40

Fig. 4. Preparation of ECC mix: (a) mixing sequence adopted in the study, (b) addition of fiber, (c) flow measurement of ECC mix, (d) Casting of ECC coupons, (e)
tensile coupons of ECC.

slow due to the high dosage of Fly ash and HRWRA. The specimen was 2.3. Test program
cured for 28 days under water prior to the testing.
The compression testing was conducted on a set of at least three
specimens of each ECC mixture of dimension 70.6 × 70.6 × 70.6 mm in
accordance with IS 4031 [31]. Uniaxial tensile test was performed on

4
S. Kumar and D.C. Rai Construction and Building Materials 404 (2023) 133158

the dog bone tensile coupons prepared as per the JSCE recommendation mm × 12 mm were cut from the tested tensile coupons.
for the design and construction of high fiber reinforced cement com­
posites (HPFRCC) with multiple fine cracks (JSCE, 2008) [32] to 3. Experimental results
investigate the performance of ECC under direct tension. The detailed
dimension of the tensile coupon and the test setup configuration can be 3.1. Fresh state tests
found in Fig. 5(a). The tensile coupons were tested under quasi-static
loading conditions at a displacement rate of 0.5 mm/min. A hydraulic The ECC with PVA fibers yielded a workable mix. On the contrary,
Universal Testing Machine (UTM) of 100 kN capacity was used to the workability of the polyester fiber-based ECC mixture was reduced
perform the tensile test. The displacement data was recorded using an significantly after fiber addition. It has been reported that polyester fi­
extensometer attached to the specimen, as shown in Fig. 5(a). bers do not behave as hydrophobic or hydrophilic but present them­
The fracture toughness of the ECC matrix was obtained by per­ selves as a mix of these two distinct properties [36]. The ECC mix
forming a three-point bending test on the notched specimen (Fig. 5b). M1.2R1.5 had the highest spread flow value (180 mm) as the dosage of
The three-point bending test on the notched specimen was conducted in the polyester fiber was 1.5% (Fig. 7a). The spread flow values of the
accordance with ASTM E399 [33] to measure the matrix fracture polyester fiber-based ECC mixes were between 158 mm and 140 mm.
toughness of the ECC. Three specimens of dimension 305 mm × 76 mm The workability of the ECC mixes was reduced as the dosage of the
× 38 mm were prepared for each matrix composition. The distance polyester fiber was increased (Fig. 7a).
between the support was kept at 250 mm, and the notch depth was kept The average flow diameters of the PVA fiber-based ECC mixes were
at 0.4 times the beam height, as shown in Fig. 5 (b). The specimens were between 160 mm and 158 mm. The measured values of the flow di­
then subjected to a three-point bending test (Fig. 5b) using a servo- ameters of the PVA fiber-based ECC mixes were higher than those of the
hydraulic universal testing machine (UTM) at a slow rate of 0.5 mm/ polyester fiber-based ECC mixes (Fig. 7a).
min. The crack mouth opening displacement (CMOD) was obtained
using the extensometer at the notched portion of the specimen (Fig. 5b). 3.2. Compressive strength
The crack mouth opening was also monitored using the digital image
correlation (DIC) technique using MATLAB-based open-source software The compressive strength of the ECC mix with PVA fibers demon­
Ncorr [34]. Random dots were provided to the specimen using a per­ strated higher compressive strength compared to the ECC mix with
manent marker, and the deformation of the specimen was recorded polyester fiber, which may be attributed to the fiber dispersion in the
using a high-speed camera at a frame rate of 30 fps. cementitious matrix (Fig. 7b). The compressive strengths of PVA-based
The single-crack test was performed on the notched specimen to ECC mixes were between 44 and 45 MPa. The maximum compressive
obtain the crack bridging stress vs. crack opening relationship experi­ strength of 55 MPa was obtained for ECC mix M1.2R1.5 as the water-to-
mentally. The ECC mix M1.2R1.5 was chosen to prepare the notched binder ratio (0.18) was lower than the water-to-binder ratio (0.26) of
specimen, and the dimension of the notched specimen was kept as 76 × other mixes. The compressive strengths of the polyester fiber-based ECC
90 × 13 mm (Fig. 6a), which was adopted from the experimental study mixes were found to be between 44.84 MPa and 26.13 MPa. The
conducted by Yang et al. (2008) [35]. The deep notches of 6 mm were compressive strength of M0.2R3 was 35.42 MPa, and the compressive
provided on two laterals, and shallow notches of 2 mm were provided on strengths of the Mixes M0.5R2.5 and M0.5R2.0 were 41.45 MPa and
the front and back surfaces before testing (Fig. 6a). The specimens were 44.84 MPa, respectively. The ECC mixes M0.5R2.0 and M0.5R2.5 had
subjected to uniaxial tensile loading under displacement control the same cementitious matrix, but the compressive strength of the mix
(Fig. 6b). M0.5R2.5 was lower than that of the mix M0.5R2.0 primarily due to
The Scanning Electron Microscope at an accelerated voltage of 10 kV increased volume of fiber. It was also found that the increase in fiber
was used to characterize the microscopic morphology of the tested percentage had an adverse effect on the strength of the ECC (Fig. 7b).
specimens of polyester fiber-based ECC. Samples of size 10 mm × 10 Though the matrix of the mix M0.2R3 was stronger than the other two

Fig. 5. (a) Uniaxial tensile test setup (all units in mm) (b) Matrix fracture toughness setup and specimen dimension.

5
S. Kumar and D.C. Rai Construction and Building Materials 404 (2023) 133158

Fig. 6. Determination of crack bridging relation (a) specimen dimension (b) experimental setup.

Fig. 7. (a) The flow values and (b) compressive strength of various ECC mixes.

mixes, M0.5R2 and M0.5R2.5, it contained 3% volume fraction that the fiber bridging plot [σ (δ)] as shown in Fig. 8. The crack tip toughness
might lead to lower compressive strength. The compressive strength of (Jtip ) can be approximately taken as K2m /Em , where Km and Em are the
the ECC mixture was decreased with the fiber percentage increase, fracture toughness and Young’s modulus of the matrix.
primarily attributed to the reduced workability and, thereby, non-
homogeneous distribution of the fiber [37]. The workability loss σ 0 > σ fc andJtip ≤ J′b (1)
might lead to increased micro-crack density, resulting in reduced ∫ δ0
compressive strength [38]. Similar behavior was observed from an Km2
≤ σ0 δ0 − σ (δ)dδ ≡ J′b (2)
experimental study on the development of Basalt fiber ECC, and it was Em 0

observed that the compressive strength of the ECC was reduced as the There can be two potential pseudo strain hardening (PSH) indicators
fiber percentage was increased [39]. It was also observed that the based on the strength and energy criteria: (1) PSH energy index (PSHE =
polyester fiber-based ECC mixes with lower fly ash dosage (M0.5R2,
J′b /Jtip ) and (2) PSH strength index (PSHs = σ 0 /σfc ) [24]. The higher the
M0.5R2.5, and M0.2R3) demonstrated higher compressive strength as
PSH indices values, the larger the margin for satisfying the strain
compared to the ECC mixes with higher fly ash dosage (M1.5R2 and
M1.5R3). Clearly, it implies the higher dosage of fly ash resulted in a
weaker cementitious matrix.

3.3. Design of ECC based on micromechanical modeling

ECC design is governed by micromechanics and fracture mechanics


of the composites. The multiple cracking and strain hardening behavior
of the ECC depends on the properties of the matrix, the type of fiber, the
tensile strength of the fiber, the aspect ratio of the fiber, the volume
fraction of the fiber, interface properties between the fiber and the
matrix, etc. The micromechanics-based mathematical model can be
useful in optimizing the fiber dosage in designing the ECC.
Two conditions (Energy and strength-based criteria) must be satis­
fied for strain-hardening and multiple cracking behaviors [14]. As per
the strength criterion, the bridging capacity (σ o ) must be greater than
the first cracking strength (σ fc ) of the matrix [Eq. (1)]. According to the
energy criterion, the complementary energy of the fiber bridging (J′b )
must exceed the crack tip toughness (Jtip ) of the matrix [Eq. (2)]. The Fig. 8. The typical σ-δ curve for strain hardening composites (modified
complementary energy of the fiber bridging (J′b ) can be obtained from from [40]).

6
S. Kumar and D.C. Rai Construction and Building Materials 404 (2023) 133158

hardening criteria and, thus, the higher the tensile ductility of ECC [40]. Table 5
For saturated multiple cracking, the energy parameter (PSHE) greater Measurement of the fracture toughness of the ECC mixes.
than 3 and the strength parameter (PSHS) greater than 2 are desirable Mix M0.2R3 M0.5R2/K2/ M1.2R1.5 M1.5R2/K2/
[41]. R2.5 R3
√̅̅̅̅
Km (MPa m ) 0.82 (10.04) 0.75 (16.78) 0.69 (13.40) 0.63 (16.27)
3.3.1. Flexural toughness test on the notched specimen (
Jtip J/m2
)
31.22 26.90 (32.65) 22.50 18.00 (19.87)
A higher margin in the fiber bridging energy and fracture toughness (20.85) (27.46)
is desired for the robust multiple cracking behaviors. However, the Values in brackets present the coefficient of variation (COV.)
fracture toughness of the matrix cannot be too small as it may reduce the
compressive strength of the composite. The load vs. CMOD plot is pre­
conducted by Rathod (2014) [44]; however, the interfacial properties of
sented in Fig. 9(a) for the ECC mixes M1.2R2/K2/R3. As can be seen
the polyester fibers were considered to be the same, and it was assumed
from Fig. 9(b), the load vs. CMOD plot obtained from the experiment
that the interfacial properties were not largely altered by the variation in
compares well with the one obtained by DIC. There was a very fine crack
the matrix properties. The interfacial parameters: chemical bond
that started from the tip of the notch and traveled in the upward di­
strength (Gd), frictional bond strength (τ0 ), slip hardening coefficient
rection, which was well captured in the strain contour obtained from
(β), snubbing coefficient (f), and strength reduction factor (f ′) can be
DIC (Fig. 9c). The average peak load carried by the notched specimen of
deduced from the single fiber pull-out test and the mathematical model
the ECC mix M1.5R2 was 0.75 kN, and the CMOD at the peak load was
proposed by Lin et al. (1999) [43]. The mathematical model curve was
0.05 mm, which was evident from the experiment and the DIC both
calibrated with the experimental curve to obtain the interfacial pa­
(Fig. 9a and b). The fracture toughness values of the matrices were
rameters, as shown in Fig. 10(b). The embedded fiber length (Le = Lf /2)
estimated using the empirical relation as per ASTM E399 [33] and are
tabulated in Table 5. The fracture toughness of the matrix M0.2R3 was was assumed to be half of the fiber length. Many studies [19,35,40,43]
highest as the water-to-binder ratio was lowest (0.18). report on the interfacial properties of PVA fibers. The interfacial pa­
The amount of fly ash used in matrices M0.5R2 and M0.2R3 was rameters of the PVA fiber for the present study were deduced from the
lower, which resulted in a stronger matrix (Table 5). The weaker tensile stress–strain plot of the ECC with 2% PVA fibers (Table 6). It is
matrices were obtained for the matrices M1.2R1.5 and M1.5R2 by widely reported [21,44] that polypropylene and polyester fibers possess
increasing the dosage of fly ash. It was evident that the fracture a hydrophobic nature, and it does not make any chemical bond; hence,
toughness of the matrices decreased as the fly ash dosage was increased the chemical bond strength (Gd) is negligible. From the mathematical
(Table 5), which was in compliance with the other experimental studies model, the chemical bond strength of the polyester fiber was also found
available in the literature [19,42]. to be zero for the present case (Table 6).
The frictional bond strength (τ0 ) and the slip hardening coefficient
3.3.2. Determination of micromechanical parameters (β) of the polyester fiber obtained from the load–displacement plots
The complementary energy is determined from the fiber bridging were 0.75 MPa and 0.01, respectively (Table 6). The values of the
stress vs. crack opening (σ − δ) plot, which can be either determined snubbing co-efficient (f) and strength reduction factor (f ′) of the poly­
through the experiments or through the mathematical model. Based on ester fiber deduced from the single fiber pull-out test results were 0.5
the fiber properties, matrix properties, and interfacial parameters, re­ and 0.3, respectively.
searchers [35,43] have developed a mathematical model that can pre­
dict the fiber bridging stress vs. crack opening behavior (refer to 3.3.3. Single-crack test and illustration of composite optimization
Appendix A). The interfacial properties can be either obtained from the The interfacial parameters of the polyester deduced from the single
single fiber pull-out test or can be deduced from the direct tensile testing fiber pull-out test were validated through the experimental results of the
on the notched specimens. The single fiber pull-out test was not con­ uniaxial tensile test on the notched specimens. The crack bridging stress
ducted in the present study due to the experimental difficulties associ­ vs. crack opening curve was obtained experimentally and compared
ated with the 12 mm fiber. The fiber pull-out test with the longer fiber with the analytical curve by Lin et al. (1999) [43] model. A MATLAB
could have been easier, but the fibers with the longer length were not code was developed for computing the fiber bridging [σ(δ)] relationship,
available in the open market. However, Rathod (2014) [44] conducted a and random 2-D distribution of fibers was assumed within the cemen­
single fiber pull-out test on the polyester fiber of the same manufacturer titious matrix. The original model assumed circular fibers, but the
as used in the preparation of the ECC in the present study. The matrix polyester fibers used in the present study were triangular. The
with a C:S ratio of 1:0.5 was used in the single fiber pull-out test, and the micromechanics-based models used in the study were not revised;
experimental test setup is presented in Fig. 10(a). The matrix properties instead, the equivalent diameter of the triangular shape fibers was used.
of the present study were different from the experimental study The mathematical model using an equivalent diameter for the
triangular-shaped fiber accurately predicted the ultimate cracking

Fig. 9. Flexural toughness test of mix M1.5R2 (a) Load vs. CMOD obtained from the experiment (b) comparison of load vs. CMOD plot from experiment and DIC (c)
strain contour measures using DIC technique.

7
S. Kumar and D.C. Rai Construction and Building Materials 404 (2023) 133158

Fig. 10. Single fiber pull-out test of polyester fiber (a) test setup [44] (b) comparison of load vs. displacement plot from experiment and Lin et al. [43] model.

Table 6
Micromechanical parameters of ECC.
Fiber type Fiber parameters Calculated Interfacial parameters Matrix parameters

df (μm) lf (mm) Ef (GPa) σfu (MPa) τ0 (MPa) Gd (J/m2 ) β f f′ Em (GPa) σm (MPa)

Polyester 30 12 5 460 0.75 0 0.01 0.5 0.3 15 4.5


PVA 40 12 22* 1060* 1.3 5 0.028 0.4 0.2 15 4.5

*In-situ values of elastic modulus and tensile strength of PVA fibers adapted from reference[23].

strength, which was evident from the direct tensile test. The embedded the tensile properties of the fiber, increasing the volume fraction of the
fiber length (Le = Lf⁄2) was assumed to be half of the fiber length. It was fiber, improving the interfacial properties, or lowering the fracture
observed that the fiber bridging curve obtained from the mathematical toughness of the matrix. The polyester fiber had very low tensile
model based on the interfacial parameters deduced from the single fiber strength and elastic modulus (Table 6), and the high fracture toughness
pull-out test was closely matched with the experimental result of the matrix might not be desirable for strain-hardening behavior. A
(Fig. 11a). Thus, these interfacial parameters can be further used for high-volume fraction of fiber may lead to poor workability of the ECC
optimizing the fiber percentage to satisfy the strain hardening and mix; hence, the fiber volume fraction and the matrix toughness shall be
multiple cracking. The complementary energy was determined from the handled in such a way that desirable tensile ductility is obtained without
fiber bridging stress vs. crack opening (σ − δ) relationship. compromising the workability of the ECC mix. The complementary en­
The peak value of the stress was 5.16 MPa from fiber bridging [σ(δ)] ergy was calculated using the Lin et al. [43] mathematical model for
relationship of ECC mix M0.5 K2 (Fig. 11b). As can be seen in Fig. 11(b) different volume fractions of the polyester fiber (Fig. 11a). As there was
that, there was a marginal difference in the bridging stress (σ0 ) of the no surface treatment done to the fibers, the interfacial properties of fi­
ECC with 2% PVA and 2% polyester fibers. This difference in the bers were assumed to be the same in estimating the complementary
bridging stresses can be attributed to the higher tensile strength and energies (Table 7). The variation of the interfacial parameters due to
elastic modulus of the PVA fibers (Table 6). Both the PSH indicators for changes in the matrix properties was ignored. The values of the first
( )
the ECC mix M1.2R1.5 were found to be less than unity, and it can be crack strength σfc were obtained from the uniaxial tensile testing, and
assumed from the PSH indicator that the occurrence of multiple cracking the maximum bridging stress was estimated from the mathematical
is unlikely (Table 7). However, both the PSH indicators of the ECC mix model. The ECC mix M0.5R6 was not prepared, but the PSH indices were
M0.5 K2 were greater than unity, implying that multiple cracking should calculated for these matrices to investigate the effect of the increase in
occur (Table 7). the fiber volume fraction on its PSH indices.
The higher values of the PSH indicator can be obtained by enhancing The matrices M0.5R2.0 and M0.5R6.0 were not physically prepared,

Fig. 11. Fiber bridging relationship of ECC mixes (a) fiber bridging relationship of mix M1.2R1.5 from experiment and various ECC mixes using a mathematical
model (b) comparison of ECC mixes M0.5R2 and M0.5 K2.

8
S. Kumar and D.C. Rai Construction and Building Materials 404 (2023) 133158

Table 7
Calculation of energy and strength criterion using Lin et al.(1999) [43] model for different volume fraction.
( )
Mix Fibre % Fibre Jtip (J/m2 ) J′b (J/m2 ) PSHE J′b /Jtip σfc (MPa) σ0 (MPa) PSHS
( )
(by volume) σ0 /σfc
M1.2R1.5 RF 1.5 % 22.50 19.67 0.87 2.86 1.45 0.51
M0.5 K2 PVA 2.0 % 26.90 67.17 2.49 4.56 5.16 1.13
M0.5R2.0 RF 2.0 % 26.90 26.18 0.97 3.00* 1.94 0.65
M0.5R2.5 RF 2.5 % 26.90 32.67 1.21 3.06 2.42 0.79
M0.5R6.0 RF 6.0 % 26.90 77.4 2.87 3.00* 5.82 1.94
M0.2R3.0 RF 3.0 % 31.20 39.14 1.25 2.45 2.90 1.18
M1.5R2.0 RF 2.0 % 18.00 26.18 1.45 2.57 1.94 0.75
M1.5R3.0 RF 3.0 % 18.00 39.14 2.17 1.81 2.90 1.59
M1.5 K2.0 PVA 2.0 % 18.00 67.17 3.72 3.38 5.16 1.53

*Assumed first crack strength of the ECC.

Fig. 12. The tensile stress–strain curves of ECC mixtures after 14 days and 28 days of curing.

9
S. Kumar and D.C. Rai Construction and Building Materials 404 (2023) 133158

so they were considered only for the parametric study. The matrix PSHE for the ECC mix M0.5R2 was 0.97 (Table 7), the occurrence of
M0.5R2.5 was physically prepared, and the first cracking of the matrix multiple cracking and strain hardening phenomena was unlikely. Hence,
was 3.06 MPa; hence, the first cracking strength of the ECC M0.5R2 and the tensile testing of the ECC mix M0.5R2 was not performed after 28
M0.5R6 mixes was assumed to be the same as that of M0.5R2.5 days of curing. Since the PSHE for the ECC mix M0.5R2.5 was 1.21,
(Table 7). The volume fraction varied between 1.5% and 6%, and the slightly greater than 1, the ECC might demonstrate multiple cracking.
complementary energy was calculated for each case. It was found that The 14-day tensile test showed the strain hardening behavior with the
the PSH energy parameter was greater than 1 for the ECC mixes first cracking strength (0.96 MPa) almost equal to the ultimate tensile
M0.5R2.5, M0.5R3, M0.2R3 but the PSHS for these mixes were less than strength (0.86 MPa), as shown in Fig. 12(c). The multiple cracking
1 (Table 7). The PSH indices of the matrix M0.5R2.0 were less than 1, phenomena of the ECC mix M0.5R2.5 vanished after 28 days of testing,
which implies that multiple cracking would not occur. It was observed and there was a significant drop in the tensile strength of the mix, as
that the stronger matrix required a higher value of complementary en­ shown in Fig. 12(d). The 14 days of testing of the ECC mix M0.2R3
ergy to satisfy the maximum PSHE criteria, which resulted in higher fiber showed consistent multiple cracking and strain hardening behaviors
demand. The PSH indices of the mix M0.5R6.0 were calculated using the (Fig. 12e). The tensile strength and ductility of the mix M0.2R3 after 14
mathematical model, and the values of the PSHE and PSHS for the matrix days were 2.71 MPa, and 1.03%, respectively. As shown in Fig. 12(f), the
M0.5R6.0 were 2.87 and 1.84, respectively. It was also observed that a tensile ductility of the mix M0.2R3 was reduced to 0.42% after 28 days
volume fraction of 6% of polyester fiber was required to obtain the of testing. However, the average value of the post-cracking tensile
saturated multiple cracking, but the fiber mixing could be difficult. From strength was 2.38 after 28 days of testing, but the strain-hardening
practical experience, it was observed that mixing polyester fiber up to behavior of the mix was not consistent. It is evident from the direct
3% volume fraction was easy, and the workability was good. Hence, the tensile testing and the micromechanical modeling that the fiber volume
volume fraction of the polyester fiber in developing the ECC was kept fraction of 3% demonstrates good tensile ductility because the PSHE
less than or equal to 3%. The values of the energy and the strength in­ index was greater than 1. However, the tensile ductility was either
dicators (PSHE and PSHs) increased with an increase in the volume reduced or vanished after 28 days of curing in all the ECC mixes with a
fraction of the polyester fiber (Table 7). The weaker matrices with stronger matrix. This can be primarily attributed to the higher fracture
polyester fiber (M1.5R2 and M1.5R3) had PSHE values greater than 1. energy of the cementitious matrix.
However, the PSHs values were greater than 1 only for the case of The proportion of fly ash was increased to reduce the fracture
polyester-based ECC mixes M0.2R3 and M1.5R3. The tendency of toughness of the matrix. The fly ash-to-cement ratio was increased from
saturated cracking was most likely for the ECC mix M1.5R3, which was 0.5 to 1.5, and the water-to-cement ratio was increased from 0.4 to 0.65
also verified through the uniaxial tensile testing. to obtain a weaker matrix. The 14 days uniaxial tensile test of ECC mix
M1.5R2 demonstrated multiple cracking behaviors (Fig. 13a). The ten­
3.4. Tensile performance of ECC mixes sile strength and ductility of ECC mix M1.5R2 after 14 days of the test
were 1.75 MPa and 0.77%, respectively (Table 8). However, the tensile
The stress–strain plot of the ECC mixtures with a strong matrix ob­ ductility of the ECC mix M1.5R2 was reduced to 0.29% after 28 days
tained from the uniaxial tensile test is presented in Fig. 12 (a)-(f). The (Fig. 13b), which can be attributed to the strength gain of the cemen­
first cracking strength, ultimate tensile strength, and tensile strain ca­ titious matrix. The 14 days tensile test of the ECC mix M1.5R3 demon­
pacity of each ECC mixture were measured from the tensile stress–strain strated the strain hardening phenomena (Fig. 13c). The average values
plot and are presented in Table 8. of tensile strength and ductility of ECC mix M1.5R3 were 1.21 MPa and
The tensile tests were performed first on trial mixes M1.2R1.5 and 2.01%, respectively, after 14 days. It was observed the mix M1.5R3
M0.5 K2 after 28 days of curing. It can be clearly seen in Fig. 12 (a) that demonstrated a consistent multiple cracking and pseudo strain hard­
the ECC mix with PVA fiber demonstrated multiple cracking with an ening behavior even after 28 days of testing (Fig. 13d). The average
average tensile strength capacity of 4.5 MPa. The matrix with M1.2R1.5 tensile strength and ductility of ECC mix M1.5R3 after 28 days were
demonstrated brittle failure (Fig. 12a), as there was no strength gain 2.17 MPa, and 1.88%, respectively. The weaker matrix resulted in
after the first crack, and the ultimate tensile strength (1.36 MPa) was higher tensile ductility, as shown in Fig. 13(d).
marginally lower than the first crack strength (2.86 MPa). The PVA fiber-based ECC mix M1.5 K2 demonstrated a higher tensile
It was already known from the micromechanical model that multiple strength and tensile ductility after 14 days of testing due to the presence
cracking cannot be obtained with a stronger matrix and with the weaker of stronger fibers (Fig. 13e). However, after 28 days of testing, the
polyester fiber with a 1.5 % volume fraction. The matrix properties were tensile strength of the ECC was increased, but the tensile ductility was
changed, and the tensile testing was performed 14 days and 28 days marginally reduced (Fig. 13f). The PVA fibers used in the present study
after the curing period; also, the fiber dosage varied between 2% and were not subjected to any surface treatment. The reduction in the tensile
3%. The 14 days of tensile testing of ECC mix M0.5R2 demonstrated ductility after 28 days of testing can be attributed to the stronger
multiple cracking because the matrix did not attain its full strength chemical bond between the fiber and the cementitious matrix [21,24].
(Fig. 12b). It was found from the micromechanical modeling that the In contrast, the polyester fiber does not possess any chemical bond with

Table 8
Measured compressive strength and tensile properties of ECC mixtures after 14 days and 28 days.
Mix Compressive strength (MPa) First cracking strength Ultimate Tensile strength σ0 (MPa) Tensile strain capacity
σfc (MPa) (%)

14 days 28 days 14 days 28 days 14 days 28 days

M1.2R1.5 55.65 (12.0) – 2.86 – 1.36 – –


M0.5 K2.0 44.38 (17.1) – 4.56 (6.3) – 4.53 (11.2) – 0.36 (14.6)
M0.5R2.0 44.84 (7.3) 1.23 (22.4) – 1.52 (23.5) – 1.41 (27.4) –
M0.5R2.5 41.45 (15.5) 0.96 (16.7) 3.06 (26.8) 0.86 (20.9) 1.58 (23.2) 0.86 (36.0) –
M0.2R3.0 35.42 (2.8) 2.35 (5.3) 2.45 (17.4) 2.71 (7.7) 2.38 (16.9) 1.03 (18.0) 0.42 (40.6)
M1.5R2.0 26.81 (3.0) 1.69 (2.2) 2.57 (5.8) 1.75 (13.9) 1.94 (11.0) 0.77 (35.6) 0.29 (19.7)
M1.5R3.0 24.09 (2.7) 0.99 (19.3) 1.81 (3.9) 1.21(9.3) 2.17 (4.3) 2.01 (23.5) 1.88 (12.0)
M1.5 K2.0 43.65 (7.6) 2.51 (9.2) 3.38 (12.9) 3.00 (12.7) 3.53 (16.9) 1.75 (28.7) 0.45 (24.3)

Values in brackets present the coefficient of variation (COV.).

10
S. Kumar and D.C. Rai Construction and Building Materials 404 (2023) 133158

Fig. 13. The tensile stress–strain curves of ECC mixtures with weaker matrix after 14 days and 28 days of the curing period.

the surrounding cementitious matrix, which helped the ECC mix manner.
M1.5R3 to achieve a higher tensile ductility. The actual failure pattern of Fig. 14(b) depicts the multiple cracking phenomena of the ECC mix
the fibers of the ECC prepared with polyester fiber can be further veri­ M0.2R3 during the tensile testing conducted after 14 days of the curing
fied using Scanning Electron Microscopy (SEM), which is discussed in period. More than three cracks were visible with the naked eye
later sections. distributed over the gauge length (Fig. 14b). Multiple cracking of the
ECC mix M1.5R3 under uniaxial tensile load after 28 days of testing is
clearly visible in Fig. 14(c). Several tight cracks occurred after the first
3.5. Crack pattern cracking, and the ECC sustained the tensile load till its ultimate point.
There was a continuous decrease in the tensile load due to the widening
The compression testing of the ECC cubes of the mix M1.5R3 showed of the bigger cracks (Fig. 14c and Fig. 13d).
ductile failure, as shown in Fig. 14 (a). Several cracks were observed in
the compression cubes, which can be primarily attributed to the
bridging effect of the fibers. The failure pattern of the ECC mix was 3.6. Scanning Electron Microscope (SEM) study
completely different from the normal mortar/concrete. The ECC cubes
maintained their shape even after the failure, and unlike normal con­ The SEM study can be helpful in understanding the micro-level
crete, ECC cubes were not broken into small fragments in a brittle behavior of the composite that led to a global response. It was

11
S. Kumar and D.C. Rai Construction and Building Materials 404 (2023) 133158

Fig. 14. (a) failure of ECC mix M1.5R3 under compression, (b) multiple cracking observed to ECC mix M0.2R3, and (c) multiple cracking of ECC mix M1.5R3 during
uniaxial tensile test.

intended to investigate the fiber pull-out or fiber rupture phenomena the circular fibers used in designing the ECC and optimizing the fiber
under direct tension, which could be an indicator of the global behavior dosage. The equivalent diameter of the triangular polyester fiber was
of the ECC. SEM testing was conducted on the fractured surface of the assumed, which gave a good approximation of the ultimate strength
tensile specimen of the ECC mix M1.5R3. The debonding of the polyester which was evident from the experiments. The fiber tunnel observed from
fiber was confirmed through the visible path of the fibers (Fig. 15a). The the SEM (Fig. 16a) indicates that the fiber was completely pulled out
triangular shape of the polyester fiber was also confirmed from the SEM because the polyester fibers do not possess any chemical bond with the
images (Fig. 15b). The measured thickness of the fiber was found to be cementous matrix. The fiber was bonded with the cementitious matrix
27.5 µm. Fig. 15(a) shows the embedded fibers in the cementitious largely by the frictional bond, and at some locations, it was seen that
matrix, the fiber–matrix interface, and the round-shaped fly ash parti­ surface abrasion and the stretching of the fiber occurred (Fig. 16b).
cles. The accumulation of loose hydrated particles on the debonded However, the interfacial bond was not strong enough to cause the fiber
polyester fibers can be seen in Fig. 15(b), which implies that there was a rupture. As manifested by the pull-out test of the polyester fibers, the
good bond between the fiber and the matrix, which was beneficial for harder surrounding matrix causes abrasion damage by delaminating
the better load-carrying capacity of the ECC. layers through the fiber thickness [44]. The complete pull-out does not
take place in the ECC with a stronger matrix as the fibers undergo pull-
4. Discussion out with delamination of the outer layer with the reduction in the cross-
section and, eventually, fiber rupture. In the case of the ECC with a
The polyester fibers were originally made from the polymerization of weaker matrix (M1.5R3), the fiber did not rupture due to the softer
pure terephthalic acid and Mono Ethylene Glycol using a catalyst. These surroundings, and a complete pull-out of the fibers took place, which
fibers were first made in the form of a single filament yarn, which was was evident from the fiber tunnel observed from SEM images (Fig. 16).
later cut into small pieces. The cross-section area of the polyester fibers Due to weaker interfacial bond of the polyester fibers, the fibers got
was mostly triangular, which was evident from the SEM test. Unlike PVA pulled out, and ECC with lower volume fraction (M1.2R1.5) ended with
fiber, the hydrophobic nature of the polyester fiber avoided strong brittle failure with single cracking. On the other hand, it is reported in
chemical bonds, and the triangular cross-section was helpful in better the literature that the PVA forms a stronger bond with the matrix that
anchoring with the other ingredients. The ECC developed in the past was causes the fiber rupture during the failure, which reduces the ductility of
mostly based on the high strength fibers (PE or PVA), which had circular the ECC. However, due to stronger interfacial bonds, ECC with PVA
cross-sections. The micromechanical models were originally based on (M0.5 K2) possessed multiple cracking, but the tensile strain capacity

Fig. 15. SEM images of ECC mix M1.5R3 (a) images of fiber, interface, and path of debonded fibers (b) image of the single fiber pulled out.

12
S. Kumar and D.C. Rai Construction and Building Materials 404 (2023) 133158

Fig. 16. SEM images of ECC mix M1.5R3 (a) fiber tunnel (b) stretching of the fiber.

was less. The interfacial bond of the PVA fibers can be reduced by the material cost was due to the fiber cost. The cost of individual compo­
application of the oiling agent, which enhances the performance of the nents of the ECC is presented in Table 9. The presented individual ma­
ECC in terms of ductility. The increased tensile strength and a higher terial cost was based on the rates obtained from the supplier in the local
frictional bond strength can be an efficient method for improving the market. The cement bags are available in a package of 50 kg; the cost per
ductility of the ECC. The polyester fibers used in the present study had a kg was obtained from the cost of one bag. Similarly, fly ash and sand are
lower tensile strength and weaker frictional bond strength. generally available in 1 ton for sale; the per kg cost was accordingly
The fiber bridging energy and the fracture toughness were monitored estimated. The cost of the fiber was based on the quotation obtained
with the help of micromechanical modeling in such a way that a higher from the fiber manufacturer. As the PVA fibers were supplied directly
margin between the two was obtained. The lower tensile strength and from Japan, the per kg cost of the Kuraray make PVA fiber also included
elastic modulus of polyester fiber required a large volume fraction for the transportation cost. The polyester fibers can be directly purchased
the higher value of the PSHE. The weaker matrix was desirable for the from the local shops in a packet of 1 kg. The per kg cost of the PVA fiber
ECC for the pseudo strain hardening. The matrix was made weaker by was approximately 10 times higher than that of the polyester fiber
increasing the proportion of the fly ash, and as can be seen in Fig. 17(a), (Table 9).
the fracture energy decreased as the fly ash increased. The values of the The weight of each component in 1 cubic meter volume was calcu­
strength parameter PSHS were lower than 1 for the ECC mixtures lated, and the individual cost was multiplied by its weight to obtain the
M1.2R1.5, M0.5R2, M0.5R2.5, and M1.5R2 (Fig. 17b). The PSH indices material cost per cubic meter, as shown in Table 10. It can be clearly
for the ECC mix M1.5R2 were higher than that of the M0.5R2 which seen that the material cost of ECC mixtures with polyester fiber (USD
indicated that the increase in the fly ash resulted in the increased values 262.5) was significantly lower than that of the ECC mixes with PVA fi­
of the PSH indices. Additionally, the fiber percentage increase resulted bers (USD 1348.5). The fiber cost is also presented in order to estimate
in increased PSH indices values, thus increasing tensile strength the share of the fiber in the total material cost. It was observed that the
(Fig. 18a). No multiple cracking was observed for the ECC mixes fiber contribution varied between 45% and 66% of the total material
M1.2R1.5, M0.5R2, and M0.5R2.5 as the PSH indices were near about 1 cost in the polyester fiber-based ECC mixtures. The contribution of the
(Fig. 18b). The ECC mixtures with stronger PVA fibers resulted in higher fiber in the PVA-based ECC mixtures was about 90% of the total material
values of PSH indices and higher ultimate tensile strength values. The cost. It was clear from Table 10 that the PVA-based ECC mixtures were
ECC mix M1.5R3 had the highest ultimate tensile strain values as the approximately 5 times costlier than that of the polyester-based ECC.
weaker matrix had lower fracture energy and higher fiber bridging en­
ergy. If the fiber volume fraction was kept at 3% for the ECC mixes 6. Conclusion
M0.2R3 and M1.5R3, the ultimate tensile strain capacity of the ECC
M0.2R3 and M1.5R3 was 0.42% and 1.88%, respectively. The higher The present study achieved its prime objective of developing a low-
tensile ductility of the ECC mix 1.5R3 was attributed to a higher value of cost ECC with polyester fiber that possessed moderate tensile strength
PSHE of 2.17 as compared to 1.25 PSHE of the ECC mix M0.2R3. and ductility, which is sufficient for various structural applications. The
existing mathematical models based on micromechanics were used to
5. Cost assessment develop a systematic design procedure for the ECC with the desired
mechanical properties. The ECC developed using low-cost polyester
As explained in the previous sections, approximately 80% of the fiber and readily available PPC cement was able to demonstrate the

Fig. 17. (a) Fracture energy of the matrices with varying fly ash dosage (b) Pseudo strain hardening indices of various ECC Mixes.

13
S. Kumar and D.C. Rai Construction and Building Materials 404 (2023) 133158

Fig. 18. Relationship between the PSH indices and (a) ultimate tensile strength (b) ultimate tensile strain.

Table 9
Cost of each component of ECC.
Material Cement (PPC) Sand Fly ash HRWRA PVA Fibre Polyester Fibre Water
(per kg) (per kg) (per kg) (per kg) (per kg) (per kg) (per m3)

Cost (INR) 9.00 3.00 1.80 100.00 3700.00 380.00 4.90


Cost (USD) 0.11 0.04 0.02 1.20 44.58 4.58 0.06

Table 10
Cost comparison of different mixtures.
Mix Cement Sand Fly ash Water HRWRA Fibre Fiber cost Material (Matrix + Fiber) cost (INR/ Average Material cost of ECC (INR/
(PPC) kg/ kg/m3 kg/ kg/m3 kg/ (INR/m3) m3) m3)
Kg/m3 m3 m3 m3

M1.2R1.5 599 719 479 239 12 20 7637 16,964 21,529


M0.5R2.0 820 410 410 328 16 27 10,133 21,126 (USD 262.5)
M0.5R2.5 816 408 408 326 16 33 12,605 23,544 (Polyester-based ECC)
M0.2R3.0 922 184 461 369 18 40 15,052 26,918
M1.5R2.0 500 750 300 325 10 27 10,133 17,888
M1.5R3.0 495 743 297 321 10 40 15,052 22,732

M0.5 K2.0 827 413 413 331 8 27 101,569 111,827 110,575


M1.5 K2.0 500 750 300 325 10 27 101,569 109,324 (USD 1348.5)
(PVA-based ECC)

multiple cracking and strain hardening behavior. The tensile properties cracking and strain hardening behaviors after 14 days of testing.
and the ductility of polyester fiber-based low-cost ECC were comparable The tensile strength of the ECC mixes varied between 0.96 MPa
with PVA fiber-based ECC. This low-cost ECC can be further used in and 2.73 MPa, and the tensile ductility varied between 0.86 and
structural applications such as strengthening the existing structures, 1.4%. The multiple cracking phenomena could not be seen after
preparation of precast panels, etc., at a reasonable cost. The major 28 days of tensile testing primarily due to the high toughness of
conclusions that can be drawn from this study are as follows: the matrix and lower elastic modulus and tensile strength of the
polyester fibers.
(1) The existing mathematical model based on micromechanics was
helpful in optimizing the volume fraction of the fiber and the CRediT authorship contribution statement
toughness of the matrix. The interfacial parameters for the
polyester fibers were obtained from the necessary experiments. It Subodh Kumar: Writing – original draft, Methodology, Data cura­
was observed that as the volume fraction was increased, the en­ tion, Conceptualization. Durgesh C. Rai: Writing – review & editing,
ergy indices were increased. The optimum dosage of 3% polyester Supervision, Funding acquisition.
fiber was recommended for multiple cracking and strain hard­
ening behavior without any workability issues.
(2) The desired low-cost ECC was developed with a relatively weaker Declaration of Competing Interest
matrix and a 3% volume fraction of polyester fiber (M1.5R3). The
multiple cracking and strength hardening behaviors were The authors declare that they have no known competing financial
observed after 14 days and 28 days of tensile testing. The tensile interests or personal relationships that could have appeared to influence
strength of the developed low-cost polyester fiber-based ECC was the work reported in this paper.
2.17 MPa, and the tensile ductility was 1.88%. The ECC mix
(M1.5R2) prepared using 2% polyester fiber demonstrated a Data availability
tensile strength of 2.5 MPa and tensile ductility of 0.3% after 28
days of testing. The ECC with 2% PVA fiber demonstrated higher No data was used for the research described in the article.
tensile strength (4.5–5 MPa), and the tensile ductility was be­
tween 0.3% and 0.45% after 28 days of testing. Acknowledgment
(3) The ECC prepared with polyester fibers and a stronger matrix
(M0.5R2, M0.5R2.5, and M0.2R3) demonstrated multiple The authors sincerely acknowledge the financial assistance provided
by the Ministry of Education, Government of India, New Delhi. The

14
S. Kumar and D.C. Rai Construction and Building Materials 404 (2023) 133158

authors would also like to thank M/S Kuraray Co. Ltd. (Japan), M/S Solutions (India) for providing the material for the research.
Reliance Industries (Mumbai, India), and M/S Fosroc Constructive

Appendix A. . Mathematical model for determination of the interfacial properties

The single fiber pull-out mathematical model was developed based on the simple stress analysis and energy balancing principle. The
load–displacement plot obtained using Lin et al. (1999) [43] model primarily involves three distinct stages: initial elastic stretching of the fiber-free
length followed by the debonding stage, and finally, the fiber pull-out stage. The theoretical pull-out load (P) versus pull-out displacement (δ)
relationship for the debonding stage and the pull-out stage can be given by equations (3) and (4), respectively.
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
π2 τ0 Ef d3f (1 + η) π 2 τ0 Ef d3f
P= δ+ , 0 ≤ δ ≤ δ0 (3)
2 2
( )
P = πdf τ0 1 + β(δ − δ0 )/df (Le − δ + δ0 ), δ0 ≤ δ ≤ Le (4)

where Le is the fiber embedment length and δ0 corresponds to the displacement at which full debonding occurs, which is given by Eq (5).
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
2τ0 Le2 (1 + η) 8Gd L2e (1 + η)
δ0 = + (5)
Ef df Ef df

where, η = Vf Ef /Vm Em , Vf and Vm are volume fractions, and Ef and Em are moduli of elasticity of fiber and matrix, respectively. The interfacial pa­
rameters: chemical bond strength (Gd), frictional bond strength (τ0 ) and slip hardening coefficient (β) can be deduced from the single fiber pull-out
test.
As the short fibers are randomly oriented in the ECC mix, the alignment of the fiber affects the pull-out load and in-situ fiber strength. The effect of
fiber alignment on the fiber pull-out load, known as the snubbing effect, was investigated by Li (1992) [45], and the empirical relation can be given by
Eq (6).
P(ϕ) = P(0)ef ϕ (6)

where f is the snubbing coefficient, a similar exponential expression was found by (Kanda and Li, 1998) [41] to account for the reduction in the in-situ
fiber strength and the empirical relation can be given by Eq (7).

σ fu (ϕ) = σ fu (0)e− f ′ϕ
(7)

where σfu is the in-situ fiber strength and f ′ is the strength reduction factor. Based on the single fiber pull-out relationship, the crack bridging stress vs.
crack opening relation (σ B − δ) can be obtained by averaging the contributions of only those fibers that cross the matrix cracking plane, as given in Eq
(8).
∫ ∫
4V π/2 (Lf /2)cosϕ
σ B (δ) = 2f P(δ)ef ϕ p(ϕ)p(z)dzdϕ (8)
π df ϕ=0 z=0

where p(ϕ) and are p(z) are the probability density function of the orientation angle ϕ and the centroidal distance of a fiber from the crack plane z,
respectively.

References [7] M.A. Kyriakides, S.L. Billington, Behavior of unreinforced masonry prisms and
beams retrofitted with engineered cementitious composites, Mater Struct Constr 47
(2014) 1573–1587, https://doi.org/10.1617/s11527-013-0138-x.
[1] C.G. Papanicolaou, T.C. Triantafillou, K. Karlos, M. Papathanasiou, Textile-
[8] A. Dehghani, G. Fischer, A.F. Nateghi, Strengthening masonry infill panels using
reinforced mortar (TRM) versus FRP as strengthening material of URM walls: in-
engineered ceme-ntitious composites, Mater Struct Constr 48 (2015) 185–204,
plane cyclic loading, Mater. Struct. 40 (10) (2007) 1081–1097.
https://doi.org/10.1617/s11527-013-0176-4.
[2] V. Turco, S. Secondin, A. Morbin, M.R. Valluzzi, C. Modena, Flexural and shear
[9] M. Deng, S. Yang, Cyclic testing of unreinforced masonry walls retrofitted with
strengthening of un-reinforced masonry with FRP bars, Compos. Sci. Technol. 66
engineered cementitious composites, Constr. Build. Mater. 177 (2018) 395–408,
(2) (2006) 289–296.
https://doi.org/10.1016/j.conbuildmat.2018.05.132.
[3] A. Prota, G. Marcari, G. Fabbrocino, G. Manfredi, C. Aldea, Experimental in-plane
[10] Y.Y. Kim, G. Fischer, V.C. Li, Performance of bridge deck link slabs designed with
behavior of tuff mas-onry strengthened with cementitious matrix-grid composites,
ductile engineered cementitious composite, ACI Struct. J. 101 (2004).
J. Compos. Constr. 10 (3) (2006) 223–233.
[11] G. Fischer, V.C. Li, Deformation behavior of fiber-reinforced polymer reinforced
[4] S. Babaeidarabad, C.F. De, A. Nanni, Out-of-plane behavior of URM walls
engineered cementitious composite (ECC) flexural members under reversed cyclic
strengthened with fabric-reinforced cementitious matrix composite, J. Compos.
loading conditions, ACI Struct. J. 100 (2003).
Constr. 18 (2014), https://doi.org/10.1061/(asce)cc.1943-5614.0000457.
[12] V.C. Li, Engineered cementitious composites (ECC): bendable concrete for
[5] S.L. Sagar, V. Singhal, D.C. Rai, In-plane and out-of-plane behavior of masonry-
sustainable and resilient infrastructure, Eng Cem Compos Bendable Concr Sustain
infilled RC frames strengthened with fabric-reinforced cementitious matrix,
Resilient Infrastruct (2019) 1–419, https://doi.org/10.1007/978-3-662-58438-5/
J. Compos. Constr. 23 (2019), https://doi.org/10.1061/(asce)cc.1943-
COVER.
5614.0000905.
[13] M. Şahmaran, V.C. Li, Durability of mechanically loaded engineered cementitious
[6] I. Koutromanos, M. Kyriakides, A. Stavridis, S. Billington, P.B. Shing, Shake-
composites under highly alkaline environments, Cem. Concr. Compos. 30 (2)
Table Tests of a 3-Story Masonry-Infilled RC Frame Retrofitted with Composite
(2008) 72–81.
Materials, J. Struct. Eng. 139 (2013) 1340–1351, https://doi.org/10.1061/(asce)
[14] V.C. Li, C.K.Y. Leung, Steady-state and multiple cracking of short random fiber
st.1943-541x.0000689.
composites, J. Eng. Mech. 118 (1992) 2246–2264, https://doi.org/10.1061/(asce)
0733-9399(1992)118:11(2246).

15
S. Kumar and D.C. Rai Construction and Building Materials 404 (2023) 133158

[15] V.C. Li, From micromechanics to structural engineering -the design of cementitious [31] BIS: 4031(Part 7), Bureau of Indian Standards, IS:1443-1972, Kisan M, Sangathan
composites for civil engineering applications, Struct Eng Eng 10 (1994). S, et al. Determination of Compressive Strength of Masonry Cement. Bur Indian
[16] V.C. Li, S. Wang, C. Wu, Tensile strain-hardening behavior or polyvinyl alcohol Stand Dehli 1988;1980:New Delhi,India.
engineered cementitious composite (PVA-ECC), ACI Mater. J. 98 (483–92) (2001). [32] Japan Society of Civil Engineers. Recommendations for Design and Construction of
[17] B. Nematollahi, J. Sanjayan, F.U.A. Shaikh, Matrix design of strain hardening fiber High Performance Fiber Reinforced Cement Composites with Multiple Fine Cracks
reinforced engineered geopolymer composite, Compos. B Eng. 89 (2016) 253–265. (HPFRCC). Concr Eng Ser 2008;82.
[18] K. Yu, Y. Wang, J. Yu, S. Xu, A strain-hardening cementitious composites with the [33] ASTM, E399 Standard Test Method for Linear-Elastic Plane-Strain Fracture
tensile capacity up to 8%, Constr. Build. Mater. 137 (2017) 410–419, https://doi. Toughness KIc of Metallic Materials, ASTM B Stand (2013).
org/10.1016/j.conbuildmat.2017.01.060. [34] J. Blaber, B. Adair, A. Antoniou, Ncorr: Open-Source 2D Digital Image Correlation
[19] H. Ma, Z. Zhang, B. Ding, X. Tu, Investigation on the adhesive characteristics of Matlab Software, Exp. Mech. 55 (2015) 1105–1122, https://doi.org/10.1007/
engineered cementitious composites (ECC) to steel bridge deck, Constr. Build. s11340-015-0009-1.
Mater. 191 (2018) 679–691, https://doi.org/10.1016/j.conbuildmat.2018.10.056. [35] E.H. Yang, S. Wang, Y. Yang, V.C. Li, Fiber-bridging constitutive law of engineered
[20] S.Z. Qian, V.C. Li, H. Zhang, G.A. Keoleian, Life cycle analysis of pavement overlays cementitious compo-sites, J. Adv. Concr. Technol. 6 (2008) 181–193, https://doi.
made with engineered cementitious composites, Cem. Concr. Compos. 35 (1) org/10.3151/jact.6.181.
(2013) 78–88. [36] J.D. Rathod, S.C. Patodi, Interface tailoring of polyester-type fiber in engineered
[21] E.H. Yang, V.C. Li, Strain-hardening fiber cement optimization and component cementitious compo-site matrix against pull-out, Mater J 107 (114–22) (2010).
tailoring by means of a micromechanical model, Constr. Build. Mater. 24 (2010) [37] C.D. Atiş, O. Karahan, Properties of steel fiber reinforced fly ash concrete, Constr.
130–139, https://doi.org/10.1016/j.conbuildmat.2007.05.014. Build. Mater. 23 (2009) 392–399, https://doi.org/10.1016/j.
[22] Z. Lin, V.C. Li, Crack bridging in fiber reinforced cementitious composites with slip- conbuildmat.2007.11.002.
hardening interfaces, J. Mech. Phys. Solids 45 (5) (1997) 763–787. [38] V.C. Li, A simplified micromechanical model of compressive strength of fiber-
[23] Z. Zhang, Q. Zhang, Matrix tailoring of Engineered Cementitious Composites (ECC) reinforced cementiti-ous composites, Cem. Concr. Compos. 14 (2) (1992) 131–141.
with non-oil-coated, low tensile strength PVA fiber, Constr. Build. Mater. 161 [39] M. Xu, S. Song, L. Feng, J. Zhou, H. Li, V.C. Li, Development of basalt fiber
(2018) 420–431, https://doi.org/10.1016/j.conbuildmat.2017.11.072. engineered cementitious composites and its mechanical properties, Constr. Build.
[24] V.C. Li, C. Wu, S. Wang, A. Ogawa, T. Saito, Interface tailoring for strain-hardening Mater. 266 (2021), 121173, https://doi.org/10.1016/j.conbuildmat.2020.121173.
polyvinyl alcohol-engineered cementitious composite (PVA-ECC), ACI Mater. J. 99 [40] K. Tosun-Felekoǧlu, B. Felekoǧlu, R. Ranade, B.Y. Lee, V.C. Li, The role of flaw size
(463–72) (2002). and fiber distribution on tensile ductility of PVA-ECC, Compos. B Eng. 56 (2014)
[25] J. Michels, D. Waldmann, S. Maas, A. Zürbes, Steel fibers as only reinforcement for 536–545, https://doi.org/10.1016/j.compo-sitesb.2013.08.089.
flat slab construction – experimental investigation and design, Constr. Build. [41] T. Kanda, C.L. Victor, Multiple cracking sequence and saturation in fiber reinforced
Mater. 26 (1) (2012) 145–155. cementitious composites, Concr Res Technol 9 (1998) 19–33, https://doi.org/
[26] A. Caggiano, M. Cremona, C. Faella, C. Lima, E. Martinelli, Fracture behavior of 10.3151/crt1990.9.2_19.
concrete beams reinfo-rced with mixed long/short steel fibers, Constr. Build. [42] V.C. Li, Tailoring ECC for special attributes: a review, Int J Concr Struct Mater 6 (3)
Mater. 37 (2012) 832–840. (2012) 135–144.
[27] S. Kumar, D.C. Rai, Enhancing flexural strength of unreinforced masonry members [43] Lin Z, Kanda T, Li VC. On Interface Property Characterization and Performance of
using cementitious matrix-based, Composites (2022) 125–142, https://doi.org/ Fiber Reinforced Cementitious Composites 1999.
10.1007/978-981-19-2424-8_6. [44] J.D. Rathod, Effect of single fiber pull out test result on flexural performance of
[28] R.F.P. Dhakal, Material Strain Limits for Seismic Design of Concrete Structures, ECC, J Civ Environ Eng 04 (2014) 2–7, https://doi.org/10.4172/2165-
SESOC J NZ Struct Eng Soc 20 (1) (2007) 14–28. 784x.1000140.
[29] Astm. Standard Specification for Coal Fly Ash and Raw or Calcined Natural [45] V.C. Li, Postcrack scaling relations for fiber reinforced cementitious composites,
Pozzolan for Use. Annu B ASTM Stand 2010:3–6. 10.1520/C0618-22.2. ASCE J. Mater. Civil Eng 4 (1) (1992) 41–57.
[30] BIS:1489 (Part 1). Portland-pozzolana cement-specification. Bur Indian Stand
1991:New Delhi,India.

16

You might also like