Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

i An update to this article is included at the end

Composite Structures 185 (2018) 634–645

Contents lists available at ScienceDirect

Composite Structures
journal homepage: www.elsevier.com/locate/compstruct

Basic mechanical properties of ultra-high ductility cementitious composites: T


From 40 MPa to 120 MPa

Yao Dinga, Jiang-tao Yua, Ke-Quan Yua, , Shi-lang Xub
a
Department of Civil Engineering, Tongji University, China
b
Department of Civil Engineering and Architecture, Zhejiang University, China

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

Keywords: A systematic study of the basic mechanical properties of ultra-high ductility cementitous composites (UHDCC),
Ultra-high ductility cementitous composites including the compressive, tensile and shear properties from normal strength to high strength were investigated
Normal strength in present research. The compressive strength of cylinder specimens with diameter of 100 mm and height of
High strength 200 mm were in the range of 43–115 MPa. The compressive strain corresponding to the peak stress, the Young’s
Mechanical properties
modulus and the Poisson’s ratio were calculated. The tensile properties including the peak stress, the strain
capacity and the strain energy were obtained as well as the relationships between the aforementioned para-
meters and the compressive strength. The peak tensile stress and the strain capacity of UHDCC were ranged from
6.2 MPa to 16.5 MPa, and 8.0% to 11.1%, respectively. Additionally, the shear stress-strain curves and the re-
lationship between the shear strength and the corresponding compressive strength were also established. During
the whole loading process of direct tensile and shear tests, the crack patterns were monitored by digital image
correlation method.

1. Introduction of SHCC at the strength of around 60 MPa in term of strain capacity and
Young’s modulus. Zhou et al. [38] studied the compressive behavior of
Strain hardening cementitious composites (SHCC) featuring ex- Polyvinyl alcohol (PVA) fiber-reinforced SHCC with compressive
cellent ductility and multiple instances of micro-cracking with self- strength ranging from 35-60 MPa, including the elastic modulus, the
controlled widths [21,16,2,11,9,23,24] and with the addition of Poly- compressive strain capacity and the Poisson’s ratio. Additionally, Xu
vinyl Alcohol (PVA) fibers at a volume fraction of 2%, can achieve a and Cai [26] proposed a new constitutive model for SHCC under uni-
tensile strain capacity up to 3–5%. The high ductility of SHCC is axial compression. The shear performance of SHCC has also been lim-
achieved by optimizing the microstructure of the composites based on ited explored. Li et al. [13] conducted Ohno shear beam tests on various
micromechanics. The mechanical interactions between the fiber, matrix cement-based composites, including plain concrete, reinforced concrete
and interface are accounted to attain strain-hardening behavior with (RC) and polyethylene (PE) fiber-reinforced SHCC and found that SHCC
the proper amount of fibers and steady state crack propagation. The exhibited superior shear ductility than RC.
high fracture toughness and controlled crack width and space ensured In previous researches, the compressive properties of SHCC with
SHCC perform better than traditional cementitious composites in many high strength (lager than 60 MPa) have been rarely studied and the
infrastructure applications [15]. SHCC has been successfully utilized in relationships between these basic mechanical parameters, i.e., com-
dam repair [15], bridge deck overlays [10] and coupling beams in high- pressive, tensile and shear parameters have not been established. Since
rise buildings [7]. the tensile behaviors of SHCC have been particularly explored, it is
In the last two decades, majority researches of SHCC have been beneficial to establish the relationship between the compressive
focused on its tensile properties [9,11,16,31,34–37,22], while the strength and the tensile properties for easy application of SHCC.
compressive [32,33,26,2,38] and shear [23,13] behavior of SHCC have In this paper, a systematical study on the mechanical properties of
not been widely studied, especially for the SHCC with compressive Polyethylene (PE) fiber-reinforced SHCC with compressive strength
strength higher than 60 MPa. A comprehensive understanding of the ranging from normal strength (43 MPa) to high strength (115 MPa) was
compressive behavior of SHCC is critical for its safe and wide applica- conducted. For compressive properties, the compressive stress-strain
tion in practical engineering. Li [14] studied the compressive properties curves were obtained and from which the compressive parameters, i.e.,


Corresponding author.
E-mail addresses: zjzjykq@163.com, 12yukequan@tongji.edu.cn (K.-Q. Yu).

https://doi.org/10.1016/j.compstruct.2017.11.034
Received 5 July 2017; Received in revised form 31 October 2017; Accepted 14 November 2017
Available online 15 November 2017
0263-8223/ © 2017 Elsevier Ltd. All rights reserved.
Y. Ding et al. Composite Structures 185 (2018) 634–645

100 were utilized to investigate the shear behavior of SHCC. The entire
90 shear stress-strain curves of SHCC specimens with different compressive
80 strengths and the shear responses under different shear loads were
70
elaborated. The relationships between the shear strength and the
Pass percentage (%)

compressive strength, tensile strength were also established.


60
It is noticed that the ductility of all PE fiber-reinforced SHCC from
50 normal to high strength developed in this research were beyond 8%
40 with maximum strain capacity up to 11%, which was far more lager
30 than the conventional SHCC reinforced by PVA fiber. Hereafter, this
LP
material was named as ultra-high ductility cementitious composites
20 GGBFS
OPC (UHDCC) in the following sections.
10
sand
0 2. Experimental program
0.1 1 10 100 1000 10000
Particle size ( m) 2.1. Materials and mix proportions
Fig. 1. Particle size distributions of the component.
The UHDCC binder was prepared using the combination of the or-
dinary Portland cement 52.5R (OPC), ground granulated blast furnace
Table 1
Chemical composition, physical and mechanical properties of OPC, SF and GGBFS.
slag (GGBFS), limestone powder (LP) and silica fume (SF). Pozzolanic
materials like GGBFS and SF were incorporated to promote the for-
OPC SF GGBFS mation of the secondary hydration products (calcium silicate hydrate).
Fine LP was utilized to partially replace cement to reduce the hydration
Chemical composition Properties of cement
SiO2 20.10 92.26 39.66 Specific gravity (g/cm3) 3.13
heat. In order to achieve an acceptable flowability of the UHDCC
CaO 62.92 0.49 34.20 Cement (m2/kg) Blaine 380 binder, a polycarboxylate-based high range water reducer (HRWR) was
Al2O3 5.62 0.89 12.94 Initial setting times (min) 130 used. Fine silica sand with a maximum grain size of 181 μm and a mean
Fe2O3 2.17 1.97 1.58 Final setting times (min) 210 size of 100 μm was utilized as the fine aggregates. Fig. 1 shows that the
MgO 1.14 0.96 6.94 Volume expansion (mm) 1.00
fine particle sizes from small to large successively are SF (0.1–1 μm), LP
Na2O 0.30 0.42 0.20 Compressive strength (MPa)
K2O 0.85 1.31 1.44 7 days 53.2 (1–10 m) and GGBFS (5–100 μm). The chemical compositions and the
SO3 2.92 0.33 0.72 28 days 61.9 critical physical and mechanical properties of the cementitious mate-
L.O.I. 3.84 1.2 Specific surface (m2/kg) rials including OPC, SF and GGBFS are listed in Table 1. Only the
Silica Fume 20000 water/binder ratio (from 0.14 to 0.32) was selected as the parameter to
GGBFS 720
obtain the normal to high strength UHDCC. The mix proportions along
with the weight of each constituent per volume are given in Table 2. As
Table 2
a consequence of rheology control for easy fiber dispersion and matrix
Mix proportions of UHDCC. toughness control for strain-hardening behavior, a high cementitious
material content was adopted in UHDCCs. The matrix fracture tough-
Cement (kg/m3) 700 687 647 605 528 ness had to be limited hence that the multiple cracking could fully occur
Limestone powder (kg/m3) 100 96 91 85 74
before the maximum fiber bridging stress was reached. Large ag-
Silica fume (kg/m3) 150 144 136 127 111
GGBFS(kg/m3) 750 736 693 648 566 gregates were excluded in the mixture.
Silica sand (kg/m3) 500 485 458 446 446 The Polyethylene (PE) fiber with an aspect ratio of 750, aiming to
Water(W)(kg/m3) 230 267 284 320 388 achieve high strength as well as high complementary energy of crack
PE fiber (kg/m3) 20 20 20 20 20
bridging which is favorable for multiple steady-state cracking was se-
HRWR (kg/m3) 25 20 12 8 3
W*/(C + SF + GGBFS) 0.1 0.2 0.2 0.2 0.3
lected as the reinforcement material in all mixtures. The dimensions
and properties of the PE fiber used in this study are listed in Table 3.
Note: W* include the water from HRWR.
2.2. Mixing procedure
Table 3
Properties of PE fibers. The mixing procedure for all mixtures was designed to uniformly
disperse the PE fibers to achieve an optimal material performance. The
PE Fiber
mixtures were prepared in a planetary type, vertical axis, speed ad-
Diameter df, μm 24 justable mixer (two speed grades). Dry powders and aggregates were
Length Lf, mm 18 mixed at the agitating speed of 140 rpm for about 1 min. Water and
Aspect ratio Lf/df 750 HRWR were slowly added and mixed for about another 5 mins at
Experimental strength, MPa 2400
140 rpm and then mixed for another 1 mins at 420 rpm. Once a con-
Nominal Young’s modulus, GPa 100
Elongation at breakage, % 2–3 sistent mixture was reached, the fiber was gradually added and con-
Specific gravity, g/cm3 0.97 tinued mixing at 140 rpm until uniform fiber dispersion was observed.
Melting temperature, °C 150 Then, the agitating speed was turned up to 420 rpm again for 1 mins to
complete the mixing procedure. The final mixture was then poured into
the prepared steel molds and vibrated for 3 mins to ensure the com-
the elastic modulus, the Poisson’s ratio and the strain at peak stress
pactness. All specimens were demolded after 24 h and cured at water
were investigated. The failure modes of the specimens under uniaxial
tank for 28 days.
compression with different strengths were compared. For tensile
properties, the stress-strain curves under uniaxial tension and the re-
2.3. Testing procedure
lationships between the compressive strength and tensile parameters,
including the initial cracking stress, the peak stress, the strain capacity
2.3.1. Tensile test
and the strain energy were discussed. Finally, Iosipescu shear beams
Dogbone-shaped specimens recommended by the Japan Society of

635
Y. Ding et al. Composite Structures 185 (2018) 634–645

average reading of two LVDTs was used to calculate the tensile strain.
The load was measured using a load cell directly attached to the loading
frame. The initial cracking stress σtc, the peak stress σtu, the strain ca-
pacity εtu and the strain energy gse were obtained from the tensile stress-
strain curves. Additionally, the crack number Nc, crack width wc and
crack spacing sc during the loading process and after unloading were
monitored.
Twenty notched prism specimens of dimension 40 × 40 × 160 mm
with a notch-depth ratio of 0.25 were prepared for three-point bending
test to measure the matrix toughness of all strengths. The bridging
stress (σ)-crack opening (δ) relationship (σ-δ curve) of UHDCC (from
43 MPa to 115 MPa) was experimentally determined from twenty
single-crack tests conducted on notched rectangular coupons. The detail
dimensions of single crack specimen are shown in Fig. 4.

2.3.2. Compressive test


The uniaxial compression test was conducted to study the com-
pressive properties of UHDCCs from normal to high strength. Cylinder
specimens with diameter of 100 mm and height of 200 mm were pre-
pared and tested according to ASTM C469 [1]. Both ends of the cylinder
specimens were capped with sulfur compound prior to testing. For each
mix proportion listed in Table 2, at least three specimens were tested to
obtain the stress-deformation curves, and the elastic modulus and
Fig. 2. Dimension of dog bone specimens. Poisson’s ratio were then calculated according to these curves. The
compression tests were conducted using a servo-hydraulic testing ma-
chine with a 4600 kN capacity. To obtain the elastic modulus and the
Civil Engineers [8] were used in this study to measure the full-range
Poisson’s ratio, two LVDTs and two strain gauges centering approxi-
stress-strain behavior of UHDCC under direct uniaxial tension. The
mately at the mid-height of the specimen were settled to measure the
geometry of the test specimens and the test set-up are shown in Fig. 2
longitudinal strain and the transverse deformation, respectively. The
and Fig. 3. At least four specimens were prepared and tested for each
test setup is illustrated in Fig. 5.
mix proportion indicated in Table 2 to study the 28-day tensile prop-
erties of UHDCC from normal strength to high strength. The dogbone
geometry ensured that majority cracks occur in the central gauge re- 2.3.3. Shear test
gion. The dimensions of the cross-section of the specimens within the Iosipescu shear beam was adopted to study the shear performance of
gage length of 80 mm were 30 × 13 mm. UHDCC as shown in Fig. 6. The v-shaped notches precast on both the
The alignment of tensile set-up was carefully checked before loading top and bottom sides of the beam were beneficial to reduce the risk of
to avoid potential eccentricity. The steel grips were hinged at the two bending failure. In addition, an approximately uniform shear stress
ends to allow multi-directional rotations. The tensile tests were con- distribution in the notch section could be generated though choosing
ducted at a constant loading rate of 0.5 mm/min according to JSCE [8]. suitable notch geometry. Van Zijl [23] refined the Iosipescu beam
Two linear variable displacement transducers (LVDTs) fixed at a testing geometry for SHCC based on the finite element analysis to ensure that
frame, which is attached to the specimen in the gauge length, were used the crack initiated in the notch area and pure shear area existed. The
to measure the elongation of the specimens, as shown in Fig. 3. The geometry of Iosipescu beam (as shown in Fig. 7) adopted in this paper
was followed Van Zijl [23] recommendation. A servo-hydraulic testing

Fig. 3. Test setup of direct tensile test.

636
Y. Ding et al. Composite Structures 185 (2018) 634–645

machine with a 4600 kN capacity was employed and operated under


crosshead displacement control at the speed of 0.5 mm/min. The shear
deformation was recorded using two LVDTs placed diagonally at 45°.
One specimen of every mixture was monitored by digital image corre-
lation method (DIC) to record the crack pattern and diagonal dis-
placement.

3. Experimental results

3.1. Compressive properties

3.1.1. Compressive strength and failure mode


The compressive strength of UHDCC cylinder specimens increased
from 43 MPa to 115 MPa when the water/binder ratio decreased from
0.32 to 0.14 as summarized in Table 4. The reduction in water/binder
Fig. 4. Dimensions of single-crack specimen.
ratio resulted in forming a more compact and homogenous micro-
structure leading to a higher compressive strength.
The incorporation of fibers which served as crack arrests or barriers
in the matrix could turn the failure mode of the specimen into a much
more ductile one. As the external loading increased, micro cracks ap-
peared and propagated in the matrix. The bridging effect of fibers was
activated and provided lateral restriction to the specimen as the cracks
extended. Hence, the cracks along the specimen were controlled to
relative small size and numerous tiny cracks were observed near the
main crack (as shown in Fig. 8c). The resulting of crack control led to a
delay of failure and turned the brittle splitting failure to a more ductile
shear failure. The cracking plane of inclined shear crack formed along
the specimens was approximate 70–90° from the horizontal plane, as
shown in Fig. 8a and b.

3.1.2. Compressive stress-strain curve


The typical stress-strain curves of UHDCC cylinders with compres-
sive strength ranging from 43 MPa to 115 MPa under uniaxial com-
pression are shown in Fig. 9. The shadowed areas present the scatter of
experimental results. To describe the complete failure process of
UHDCCs, the stress-strain curve is mainly divided into four distinct
branches in this paper, i.e., a linearly ascending branch, a short non-
linear ascending branch, a descending branch, and a nearly horizontal
branch. The initial slope of the linearly ascending branch reflects that
the elastic modulus of UHDCC increased with the compressive strength.
Due to the crack initiation, a short non-linear branch appeared before
the peak load. During this process, micro-cracks formed due to the
Fig. 5. Test setup of uniaxial compression test.
extended lateral expansion of the specimen and paralleled to the di-
rection of loading. Fibers acted as cracks barriers and hindered cracks
extension. At the descending branch, the load reduced gradually, the
post-peak ductility of the specimen was improved with fiber in-
corporation. The failure mode was changed from brittle failure to a
more ductile mode. When the load decreased to a value around 30% of
the peak stress (i.e., 35 MPa) for high strength UHDCC (i.e.,
fc = 115 MPa) or about 65% of the peak stress (i.e., 28 MPa) for normal
strength UHDCC (i.e., fc = 43 MPa), the stress-strain curve entered a
nearly horizontal platform. In this stage, the load decreased slowly
while the deformation increased fast due to the propagation of the di-
agonal shear crack.

3.1.3. Elastic modulus


The relationship between the elastic modulus and the cylinder
compressive strength of UHDCC is shown in Fig. 10. It is clear that the
elastic modulus of UHDCC increased with the increasing compressive
strength, which was similar with the commonly observation of ce-
mentitious composites by previous researchers [26,38]. The elastic
modulus of normal weight concrete can be predicted using the em-
pirical formulae suggested by codes of practice such as ACI 318 (2005)
and CEB-FIP [3] based on numerous experimental results. The com-
Fig. 6. Test setup of shear test.
parisons of the elastic modulus of UHDCC obtained directly from ex-
periment with those predicted by codes of practice are shown in Fig. 10.

637
Y. Ding et al. Composite Structures 185 (2018) 634–645

Fig. 7. Iosipescu shear beam geometry.


P
Steel load beam

LVDT
Steel load beam

Table 4
Summary of the average tensile parameters of UHDCC specimens.

w/b ratio Compressive strength Initial cracking stress Peak stress σtu Strain capacity εtu Strain energy gse Number of Crack width Crack spacing sc
(MPa) σtc (MPa) (MPa) (%) (KJ·m3) cracks Nc wc (μm) (mm)

0.14 115(3) 9.19(1.62) 16.46(0.90) 8.00(1.00) 1036(1 6 6) 42(5) 150(10) 1.90(0.22)


0.16 94(3) 8.81(0.61) 13.21(0.52) 10.13(1.23) 997(80) 55(1) 147(19) 1.45(0.03)
0.18 77(3) 7.50(0.47) 11.35(1.13) 9.02(1.36) 757(1 2 8) 52(8) 139(11) 1.57(0.27)
0.22 64(2) 7.22(0.78) 9.53(0.84) 11.14(2.12) 816(1 7 1) 52(9) 170(3) 1.57(0.31)
0.32 43(2) 4.94(0.49) 6.17(0.47) 10.64(0.09) 528(35) 49(1) 173(4) 1.64(0.04)

Note: The figure in parentheses is the standard deviation of corresponding parameter.

Fig. 8. Typical failure modes and microcracks of cylinder


specimens under compression.

77 MPa 115 MPa

(a) Failure modes of fc=77MPa (b) Failure modes of fc=115MPa (c) Microcracks

120 It can be seen from Fig. 10 that the elastic modulus of UHDCC was
fc=115MPa
fc=77MPa
overestimated by codes of practice, indicating that the formulae for
fc=43MPa concrete were not suitable for UHDCC material. It is known that the
90
elastic modulus of concrete mainly depends on the type and content of
Stress (MPa)

the coarse aggregate and the elastic modulus of the paste. Hence, it is
60 35MPa reasonable that UHDCC containing no coarse aggregate and small
amount of sand had lower elastic modulus than concrete. As shown in
Fig. 10, the regression equations proposed by Zhou et al. [38] for cy-
30
linder cementitious composites specimens (with the dimensions of
28MPa 100 × 200 mm) with the compressive strength of less than 70 MPa
0 predicted reasonably well of the elastic modulus of the normal strength
0 3000 6000 9000 12000 15000 18000
UHDCC (i.e., 43 MPa and 64 MPa) in this study.
Strain ( m)

Fig. 9. Typical stress-strain curves of UHDCCs under uniaxial compression. 3.1.4. Poisson’s ratio
The Poisson’s ratio plotted against the compressive strength is
shown in Fig. 11. It is clear that the Poisson’s ratio of UHDCC varied
insignificantly with the compressive strength. For simplified purpose,
the average value of 0.237 for Poisson’s ratio was regarded as a

638
Y. Ding et al. Composite Structures 185 (2018) 634–645

60
Experimental results ACI 318
Zhou et al. (2014) CEB-FIP
Elastic modulus (GPa) 50 ( tu
, tu
)

40

Stress (MPa)
( , )
30 tc tc

gse
20

10
40 60 80 100 120
Compressive strength (MPa)
Strain (%)
Fig. 10. Relationship between elastic modulus and prism compressive strength of
Fig. 14. Modeled tensile stress-strain curve and critical parameters.
UHDCC.

0.4 reference value for structural analysis. This value fell in the range
presented in the previous researches of Xu and Cai [26] (i.e., ν = 0.255)
and Zhou et al. [38] (i.e., ν = 0.170) and also closed to the value of
0.3
conventional concrete [3,28]. It can be also seen from Fig. 12 that the
Poisson's ratio

Poisson’s ratio of UHDCC changed marginally with the normalized


0.2
section stress before peak load.

0.1 3.1.5. Strain at peak stress


The strain at peak stress reflects the deformation capacity of spe-
0.0 cimens at the ultimate stress. The relationship between the strain at
40 60 80 100 120 peak stress and the corresponding cylinder compressive strength is
Compressive strength (MPa)
shown in Fig. 13. In general, it can be seen from Fig. 13 that the strain
Fig. 11. Relationship between Poisson’s ratio and compressive strength of UHDCC. at peak stress of UHDCC increased with the compressive strength. In
previous researches [26,38], the strain at peak stress had no obvious
0.30 correlation with the compressive strength, which might be due to the
narrow strength range chose in these researches. Although numerous
fc=64MPa
researches have been conducted to verify whether the strain at peak
fc=115MPa
stress of concrete depends on its compressive strength or not, the con-
0.25
Poisson's ratio

clusions are conflicting. In addition, the strains at peak stress fell in the
range of 3000–4000 με for UHDCC specimens with strength varying
from normal to high strength (i.e., 43–115 MPa) which were lower than
0.20 those obtained by previous researches [26,38] attributing to the higher
sand/binder ratio used in this research.

3.2. Tensile stress-strain curve and tensile properties


0.15
0.2 0.4 0.6 0.8 1.0
Normalized section stress The modeled tensile stress-strain curve of UHDCC is illustrated in
Fig. 14. The critical parameters reflecting the tensile properties of
Fig. 12. Relationship between Poisson’s ratio and normalized stress of UHDCC before
UHDCC, including the initial cracking strength (σtc), the peak stress
peak load.
(σtu), the strain capacity corresponding to the peak stress (εtu) and the
energy absorption capacity (gse) are highlighted in Fig. 14. The stress
value at the kink point from the linear elastic portion to the strain-
4000 hardening portion of the stress-strain curve was defined as the initial
Strain at peak stress ( )

cracking stress (σtc). The strain energy (gse) was calculated using the
area enclosed by the ascending branch of stress-strain curves. The area
3500 underneath the descending branch of the stress-stain curves during the
crack localization stage was not counted in strain energy. In addition,
the number of cracks (Nc), the average crack width (wc) and the average
3000 crack spacing (sc) were also observed after unloading. The number of
cracks (Nc) of specimen was obtained by visually observing and
counting the cracks on both sides of specimen. It was found that most of
2500 the micro-cracks went through entire section of the specimen. The
40 60 80 100 120
average crack width (wc) at the loading stage was calculated using the
Compressive strength (MPa)
elongation divided by the number of cracks within the gauge length
Fig. 13. Relationship between strain at peak stress and compressive strength of UHDCC. (80 mm), and similarly the crack spacing (e.g., total length after de-
formation divided by number of cracks) could be obtained.
The typical experimental tensile stress-strain curves of UHDCCs
with different compressive strengths are presented in Fig. 15. The initial
cracking strength, the peak stress and the strain energy of the UHDCC

639
Y. Ding et al. Composite Structures 185 (2018) 634–645

20 20

16 16
Stress (MPa)

Stress (MPa)
12 12

8 8

4 4

0 0
0 3 6 9 12 15 0 3 6 9 12 15
Strain (%) Strain (%)

(a) Tensile stress-strain curve of fc=115 MPa (b) Tensile stress-strain curve of fc=77 MPa
Fig. 15. Typical stress-strain curves of UHDCCs under uniaxial tension.

Fig. 16. Crack patterns of UHDCCs.

640
Y. Ding et al. Composite Structures 185 (2018) 634–645

20 20 Fig. 17. Typical stress-strain curves and fracture surface of


UHDCCs under shear loading.

16 16

12
Stress (MPa)

12

Stress (MPa)
8 8

4 4

0 0
0 2 4 6 8 10 0 2 4 6 8 10

Strain (%) Strain (%)

(a) Shear stress-strain curve of fc=115 MPa (b) Shear stress-strain curve of fc=77 MPa

20
fc=115MPa
20
fc=94MPa
16 fc=77MPa
16
Shear stress (MPa)

fc=64MPa
Stress (MPa)

12 12 fc=43MPa

8 8

4
4

0
0 0 2 4 6 8 10 12
0 2 4 6 8 10 Diagonal strain (%)
Strain (%)

(c) Shear stress-strain curve of fc=43 MPa (d) Shear stress-strain curve at all mixtures

Fig. 18. Shear responses of UHDCC specimen at compressive strength of 64 MPa.

specimens increased with the compressive strength. On the contrary, distributed (i.e., mostly under 1500 μm and with an average spacing of
the strain capacity corresponding to the peak stress of the specimens around 1000 μm) and the crack widths (i.e., mostly under 100 μm) were
generally decreased with an increase in compressive strength. The in- relatively small. Most of the fibers were pulled out from the matrix (as
crease in the complementary energy due to the raise of the fiber-matrix shown in Fig. 16b) resulting in a ductile failure mode. Regardless of the
interfacial stress was lower than the increase of the fracture toughness strength of the UHDCC specimens, multiple cracking behaviors were all
for higher strength matrices. Thus the decreasing ratio of com- observed and the crack widths of UHDCC specimens were no more than
plementary energy to fracture toughness of matrix led to the decline in 200 μm with the crack spacing less than 2 mm, detailed values can be
strain capacity, which will be shown in Section 4.1.2. referred to Table 4.
The cracking patterns of UHDCC after unloading are shown in The values of the above mentioned critical parameters (i.e., σtc, σtu,
Fig. 16. The microscopic observations of the crack width and crack εtu, gse, Nc, wc and sc) of UHDCCs under uniaxial tension with different
spacing are shown in Fig. 16a, indicating that the cracks were uniformly compressive strengths are summarized in Table 4. When the

641
Y. Ding et al. Composite Structures 185 (2018) 634–645

20 4. Discussion
peak stress
initial cracking stress
16 4.1. Relationship between compressive strength and tensile properties
ACI 318
Peak stress (MPa)
EC2
12 Due to the brittle nature of Portland cement (PC) concrete, the
uniaxial tension test is extremely hard to be conducted due to its high
8 requirement on testing machine stiffness, extremely low loading rate
and the possibility of failure on the end of the samples [18,5,20]. For
4 this reason, numerous researches were conducted to establish the re-
lationship between the tensile properties and the compressive strength
0 of ordinary concrete for its easy application. Hence, the relationship
40 60 80 100 120 between the tensile properties and compressive strength of UHDCCs
Compressive st rength (MPa)
from normal to high strength is illustrated in the following sections. In
Fig. 19. Relationship between stress and compressive strength of UHDCC. addition, the relationship between the compressive strength and the
shear strength of UHDCCs is also stated.
15
4.1.1. Tensile initial cracking stress and peak stress
12 The relationship between the initial cracking stress of UHDCC under
uniaxial tension and the compressive strength is illustrated in Fig. 19. In
Strain capacity (%)

general, the initial cracking stress of UHDCC increased with its corre-
9
sponding cylinder compressive strength. The initial cracking stress in-
creased from 4.94 MPa to 9.19 MPa (i.e., a 86.0% increase) when the
6 compressive strength increased from 43 MPa to 115 MPa. However, the
growth of initial cracking strength decreased as the compressive
3 strength increased. When the compressive strength of UHDCC increased
from 43 MPa to 64 MPa (i.e., 62.5%), the initial cracking strength in-
0 creased from 4.94 MPa to 7.22 MPa (i.e., a 46.1% increase); while this
40 60 80 100 120 improvement reduced to 27.3% (i.e., from 7.22 MPa to 9.19 MPa) when
Compressive strength (MPa) the compressive strength had a 76.9% improvement from 64 MPa to
Fig. 20. Relationship between strain capacity and compressive strength of UHDCC. 115 MPa.
It can be seen from Fig. 19 that the peak stress of UHDCC specimen
under uniaxial tension had a strong correlation with the corresponding
compressive strength of UHDCC specimens increased from 43 MPa to
compressive strength. The peak stress increased almost linearly with the
115 MPa (i.e., a 167.4% increase), the initial cracking stress σtc almost
increase of the compressive strength. The empirical formulae suggested
doubled from 4.94 MPa to 9.19 MPa, the peak stress increased from
by codes of practice such as ACI 318 model (2005) and Eurocode 2
6.17 MPa to 16.46 MPa (i.e., a 166.8% increase), and the strain energy
model (2004) based on numerous experimental results which predict
gse was nearly two times from 528 KJ·m3 to 1036 KJ·m3. However, in
reasonably well of the splitting tensile strength of PC concrete sig-
general, the deformability of the specimens, represented by the strain
nificantly underestimate the peak stress of UHDCC. When the com-
capacity εtu corresponding to the peak stress, reduced from 10.64% to
pressive strength of UHDCC cylinder was 80 MPa, the corresponding
8.00% (i.e., a 24.8% decrease). On the other hand, the number of cracks
peak strength under uniaxial tension was 11.35 MPa; while the splitting
Nc changed not apparently with the improvement of the compressive
tension strength predicted by ACI 318 model and EC2 model were
strength, while the average crack width wc decreased about 30% from
5.00 MPa and 3.76 MPa respectively, which were 55.9% and 66.9%
173 m to 150 m when the compressive strength increased from 43 MPa
lower than the value of UHDCC. The tensile strength of ordinary con-
to 115 MPa.
crete is only a fraction of its compressive strength due to the brittleness
natural of ordinary concrete. However, the tensile stress of UHDCC was
much higher and the failure mode was more ductile than ordinary
3.3. Shear stress-strain curve and shear response
concrete.
It can be also seen that the discrepancies between the initial
The typical shear stress-strain curves of UHDCC with compressive
cracking stress and the peak stress increased with the increase of
strength ranging from 43 MPa to 115 MPa are presented in Fig. 17. It is
compressive strength. When the compressive strength was 43 MPa, the
clear that the initial shear cracking stress and the peak shear stress of
peak strength (i.e., 6.17 MPa) was 24.9% higher than the initial
the UHDCC specimens increased with the compressive strength. How-
cracking stress (i.e., 4.94 MPa). While, as the compressive strength
ever, it seems that the shear deformability of the specimens do not
further increased to 115 MPa, the peak stress (i.e., 16.46 MPa) was
depend on the compressive strength.
nearly two times of the initial cracking stress (i.e., 9.19 MPa). The
The shear responses of the UHDCC Iosipescu beam specimen at the
improvement of initial cracking stress with compressive strength was
compressive strength of 64 MPa corresponding to different shear loads
less significant than that of the peak stress. The increase of initial
based on Digital Image Correlation (DIC) analysis are shown in Fig. 18.
cracking stress with compressive strength was mainly due to the
It indicates that the shear failure was restricted to the notch area. The
strengthening of matrix itself, while the improvement of peak stress was
crack initiated at point “A” indicated in the figure, confirming the
the result of the enhancement in both matrix and fiber-matrix inter-
maximum principal stress was firstly achieved at this point. Subse-
facial stress.
quently, multiple diagonal cracks occurred in the notch zone, indicating
the uniform shear stress distribution in the notch section.
4.1.2. Tensile strain capacity
From Fig. 20 it can be seen that the strain capacity corresponding to
the peak stress under uniaxial tension generally decreased with the
increase of compressive strength. When the compressive strength was
43 MPa, the average strain capacity of the UHDCC specimens was

642
Y. Ding et al. Composite Structures 185 (2018) 634–645

Fig. 21. σ (δ) curve for tensile strain-hardening composite and single crack test setup.

10.64%. However, the strain capacity reduced to 8.00% (i.e., a 33.0% modulus of matrix.
decrease) at the compressive strength of 115 MPa. This results illus- The PSH values decreased slightly with compressive strength from
trates that the capacity of deformation decreased as the compressive 94 at 115 MPa to 80 at 43 MPa, which well explained the decrease of
strength increased, indicating a more brittle matrix with higher com- strain capacity of UHDCCs with strength increase as shown in Fig. 22.
pressive strength. However, it is noticed that the tensile strain capacity The increase of crack tip toughness of matrix was more significant with
of UHDCC is approaching to the value of mild steel of 9% according to compressive strength than the interfacial bridging stress between PE
the code [4], which indicates a potential application of UHDCC in 3D- fiber and matrix. However, the PSH values of all UHDCCs were higher
printing structures without steel bars. enough to ensure the potential of strain capacity compared to the
The pseudo strain-hardening (PSH) values for all UHDCC mixtures previous researches [27,31].
[6], an index indicating the ability of strain capacity, were calculated
and the relationship to compressive strength was established to explain
the tendency of strain capacity with compressive strength. The PSH
performance index has been used to quantitatively evaluate the margin
120
and is defined as follows [6].
Jb′ 100
PSH =
Jtip (1)
80
PSH values

where Jtip is the crack tip toughness of matrix obtained from three-point
bending test and the geometry of test specimens could be referred to 60
authors’ previous research. The calculation of Jtip is shown in Eq. (2), J′b
is the complementary energy, calculated from the bridging stress (σ) 40
versus crack opening (δ) curve of single-crack test, as illustrated in
Fig. 21c and d [12,17]. The values of J'b of UHDCCs from 43 to 115 MPa 20
ranged from 0.82 to 1.44 N/mm, which were relatively smaller than
those of ultra-high performance concrete [29,30] 0
20 40 60 80 100 120
Km2
Jtip = Compressive strength (MPa)
Em (2)
Fig. 22. Relationship of PSH values and compressive strength.
where Km is the fracture toughness of matrix and Em is the Young’s

643
Y. Ding et al. Composite Structures 185 (2018) 634–645

1500 obvious that the shear strength increased almost linearly with the in-
crease of the compressive strength. The shear strength increased from
3
1200
Strain energy (KJ·m ) 8.01 MPa to 16.90 MPa (i.e., a 110.9% increase) when the compressive
strength of UHDCC increased from 43 MPa to 115 MPa. A strong cor-
900 relation is also observed between the shear strength and the tensile
strength of the UHDCC specimens as shown in Fig. 25. It is clear that the
600 shear strength was higher than the tensile strength due to the fiber
inclination in the matrix [19], which increased the fiber-matrix inter-
300 facial stress. The fracture surface of shear specimen is also shown in
Fig. 25.
0
40 60 80 100 120 5. Conclusions
Compressive strength (MPa)
Fig. 23. Relationship between strain energy and compressive strength of UHDCC.
This paper conducted a systematical research on the compressive
performance, the tensile behavior and the shear response of UHDCC,
with compressive strength ranging from 43 MPa to 115 MPa. The re-
20
lationships between the compressive strength and the tensile para-
meters were established as well as the relationships between the shear
16
strength and the compressive/tensile strength. The main conclusions
Shear strength (MPa)

are drawn in the following.


12

(1) With the incorporation of PE fibers, the failure mode of UHDCC


8
under uniaxial compression is ductile shear failure. The elastic
modulus of UHDCC increases with the compressive strength, while
4
the ACI 318 model and CEB-FIP model significantly overestimate
the elastic modulus of UHDCC due to the absence of coarse ag-
0
40 60 80 100 120 gregate. The Poisson’s ratio of UHDCC varies insignificantly with
Compressive st rength (MPa) compressive strength and can be settled at 0.237 for simplified
utilization. The values of Poisson’s ratio keep almost constant to the
Fig. 24. Relationship between shear strength and compressive strength of UHDCC.
peak compressive stress. The strain at peak stress falls in the range
of 3000–4000 με and increases with the increasing compressive
4.1.3. Tensile strain energy strength.
The strain energy of UHDCC specimen plotted against cylinder (2) The initial cracking stress and peak stress of UHDCC under uniaxial
compressive strength is shown in Fig. 23. As expected, the strain energy tension increase with compressive strength, whereas the strain ca-
generally increased with the increase of the compressive strength. The pacity corresponding to the peak stress exhibits opposite trend. The
strain energy doubled from 528 KJ·m3 to 1036 KJ·m3 when the com- ACI 318 model and EC 2 model underestimate the initial cracking
pressive strength increased from 43 MPa to 115 MPa. A higher com- stress and peak stress of UHDCC. As expected, the strain energy of
pressive strength led to a stronger matrix as well as a better bond be- UHDCC under uniaxial tension generally increases with increasing
tween the matrix and the fiber, and resulted in absorbing more energy compressive strength. The pseudo strain-hardening (PSH) also
during crack initiation and crack propagation. The high strain energy slightly decreased with compressive strength, which well explains
would benefit the application of UHDCC in seismic area. Additionally, the tendency of strain capacity. The establishment of the relation-
the values of strain energy of UHDCC were approximately two orders ships between the tensile parameters and the compressive strength
larger than that of the conventional PVA fiber-reinforced strain hard- could help the easy understanding of this material and thus the
ening cementitious composites (PVA-SHCC) [15] and ultra-high per- better application in practical engineering.
formance concrete (UHPC) [25]. (3) The shear strength of UHDCC increases almost linearly with the
corresponding compressive strength. Multiple diagonal cracks are
4.2. Shear strength observed in the notch area. The shear strength of UHDCC is higher
than the tensile strength due to the enhancement of fiber-matrix
The relationship between the shear strength of UHDCC specimens interfacial stress in an inclined angle.
and the corresponding compressive strength is shown in Fig. 24. It is

Fig. 25. Relationship between shear strength and tensile


20 strength of UHDCC.
y=1.09x
16
Shear strength (MPa)

12

4
4 8 12 16 2 0
Tensile strength (MPa)

644
Y. Ding et al. Composite Structures 185 (2018) 634–645

References [21] Shao Y, Shah SP. Mechanical properties of PVA fiber reinforced cement composites
fabricated by extrusion processing. ACI Mater J 1997;94(6):555–64.
[22] Sui L, Luo M, Yu K, et al. Effect of engineered cementitious composite on the bond
[1] ASTM C469-94. Standard test method for static modulus of elasticity and Poisson’s behavior between fiber-reinforced polymer and concrete. Compos Struct
ratio of concrete in compression. 1994. 2017;184:775–88.
[2] Cai X, Xu S. Uniaxial compressive properties of ultra high toughness cementitious [23] Van Zijl GPAG. Improved mechanical performance: shear behaviour of strain-
composite. J Wuhan Univ Technol Mater Sci Ed 2011;26(4):762–9. hardening cement-based composites (SHCC). Cem Concr Res 2007;37(8):1241–7.
[3] Comite Euro-International du Benton (CEB). (1990). CEB model code 90. Bulletin [24] Van Zijl GPAG, Wittmann FH, Oh BH. Durability of strain-hardening cement-based
d’information No. 203, Paris. composites (SHCC). Mater Struct 2012;45(10):1447–63.
[4] GB 50011-2010. Code for seismic design of buildings. Beijing: China Building [25] Xu M, Wille K. Fracture energy of UHP-FRC under direct tensile loading applied at
Industry Press; 2007. low strain rates. Compos B Eng 2015;80:116–25.
[5] Gopalaratnam VS, Shah SP. Softening response of plain concrete in direct tension. [26] Xu SL, Cai XR. Experimental study and theoretical models on compressive prop-
ACI J Proc 1985;82(3):310–23. erties of ultrahigh toughness cementitious composites. J Mater Civ Eng
[6] Kanda T, Li VC. Multiple cracking sequence and saturation in fiber reinforced ce- 2010;22(10):1067–77.
mentitious composites. Concr Res Technol 1998;9(2):19–33. [27] Yang EH, Yang YZ, Li VC. Use of high volumes of fly ash to improve ECC mechanical
[7] Kunieda M, Rokugo K. Recent progress on HPFRCC in Japan. J Adv Concr Technol properties and material greenness. ACI Mater J 2007;104(6):620–8.
2006;4(1):19–33. [28] Yoo DY, Kang ST, Lee JH, et al. Effect of shrinkage reducing admixture on tensile
[8] JSCE. Recommendations for design and construction of high performance fiber and flexural behaviors of UHPFRC considering fiber distribution characteristics.
reinforced cement composites with multiple fine cracks. Tokyo, Japan: Japan Cem Concr Res 2013;54:180–90.
Society of Civil Engineers; 2008. p. 1–16. [29] Yoo DY, Kang ST, Banthia N, et al. Nonlinear finite element analysis of ultra-high-
[9] Jun P, Mechtcherine V. Behaviour of strain-hardening cement-based composites performance fiber-reinforced concrete beams. Int J Damage Mech 2015;78(2):1–23.
(SHCC) under monotonic and cyclic tensile loading: part 1-experimental in- [30] Yoo DY, Kim SW, Park JJ. Comparative flexural behavior of ultra-high-performance
vestigations. Cement Concr Compos 2010;32(10):801–9. concrete reinforced with hybrid straight steel fibers. Constr Build Mater
[10] Lepech MD, Li VC. Application of ECC for bridge deck link slabs. RILEM J Mater 2017;132:219–29.
Struct 2009;42(9):1185–95. [31] Yu JT, Lin JH, Zhang ZG, et al. Mechanical performance of ECC with high-volume
[11] Li VC, Wang S, Wu C. Tensile strain-hardening behavior of polyvinyl alcohol en- fly ash after sub-elevated temperatures. Constr Build Mater 2015;99:82–9.
gineered cementitious composite (PVA-ECC). ACI Mater J 2001;98(6). [32] Yu KQ, Dai JG, Lu ZD, et al. Mechanical properties of engineered cementitious
[12] Li VC. Postcrack scaling relations for fiber reinforced cementitious composites. J composites subjected to elevated temperatures. J Mater Civ Eng
Mater Civ Eng. 1992;4(1):41–57. 2014;27(10):04014268.
[13] Li VC, Mishra DK, Naaman AE, et al. On the shear behavior of engineered ce- [33] Yu KQ, Lu ZD, Yu JT. Residual compressive properties of strain-hardening ce-
mentitious composites. Adv Cem Based Mater 1994;1(3):142–9. mentitious composite with different curing ages exposed to high temperature.
[14] Li VC. Engineered cementitious composites - tailored composites through micro- Constr Build Mater 2015;98:146–55.
mechanical modeling. J Adv Concr Technol 1998;1(3). [34] Yu KQ, Yu JT, Dai JG, et al. Development of ultra-high performance fiber-reinforced
[15] Li VC. On engineered cementitious composites (ECC). J Adv Concr Technol composites using polyethylene fibers. Constr Build Mater 2018;158:217–27.
2003;1(3):215–30. [35] Yu K, Yu J, Lu Z. Mechanical characteristics of ultra high performance strain
[16] Li H, Xu S, Leung CKY. Tensile and flexural properties of ultra high toughness ce- hardening cementitious composites. In: International conference on strain-hard-
montious composite. J Wuhan Univ Technol Mater Sci Ed 2009;24(4):677–83. ening cement-based composites. Springer, Dordrecht, 2017: 230–237.
[17] Marshall DB, Cox BN. A J-integral method for calculating steady-state matrix [36] Yu KQ, Wang YC, Yu JT, et al. A strain-hardening cementitious composites with the
cracking stresses in composites. Mech Mater 1988;7(2):127–33. tensile capacity up to 8%. Constr Build Mater 2017;137:410–9.
[18] Petersson PE. Crack growth and development of fracture zones in plain concrete [37] Zhan K, Yu J, Wang Y, et al. Development of cementitious composites with tensile
and similar materials. Division of Building Materials, Lund Institute of Technology; strain capacity up to 10%. In: International conference on strain-hardening cement-
1981. based composites. Springer, Dordrecht, 2017: 147–153.
[19] Ranade R, Li VC, Stults MD, William FH, et al. Micromechanics of High-Strength. [38] Zhou J, Pan J, Leung CKY. Mechanical behavior of fiber-reinforced engineered
High-Ductility Concrete. ACI Mater J 2013;110(4):375–84. cementitious composites in uniaxial compression. J Mater Civ Eng
[20] Reinhardt HW, Cornelissen HA, Hordijk DA. Tensile tests and failure analysis of 2014;27(1):04014111.
concrete. J Struct Eng 1986;112(11):2462–77.

645
Update
Composite Structures
Volume 224, Issue , 15 September 2019, Page

DOI: https://doi.org/10.1016/j.compstruct.2019.111096
Composite Structures 224 (2019) 111096

Contents lists available at ScienceDirect

Composite Structures
journal homepage: www.elsevier.com/locate/compstruct

Corrigendum

Corrigendum to “Basic mechanical properties of ultra-high ductility T


cementitious composites: From 40 MPa to 120 MPa” [Compos. Struct.
185(2018) 634–645]”
Yao Dinga, Jiang-tao Yub, Ke-Quan Yua, , Shi-lang Xub

a
College of Civil Engineering, Tongji University, China
b
Department of Civil Engineering and Architecture, Zhejiang University, China

The authors regret “The Table 2 in the current version

Table 2
Mix proportions of UHDCC.
Cement (kg/m3) 700 687 647 605 528
Limestone powder (kg/m3) 100 96 91 85 74
Silica fume (kg/m3) 150 144 136 127 111
GGBFS (kg/m3) 750 736 693 648 566
Silica sand (kg/m3) 500 485 458 446 446
Water (W) (kg/m3) 230 267 284 320 388
PE fiber (kg/m3) 20 20 20 20 20
HRWR (kg/m3) 25 20 12 8 3
W*/(C + SF + GGBFS) 0.14 0.16 0.18 0.22 0.32

Note: W* include the water from HRWR.

DOI of original article: https://doi.org/10.1016/j.compstruct.2017.11.034



Corresponding author.
E-mail address: 12yukequan@tongji.edu.cn (K.-Q. Yu).

https://doi.org/10.1016/j.compstruct.2019.111096

Available online 04 June 2019


0263-8223/ © 2019 Elsevier Ltd. All rights reserved.
Y. Ding, et al. Composite Structures 224 (2019) 111096

The revised table has been shown below.


And Fig. 21(c) has been changed to

18
16
14
12

Stress/MPa
10
8
6
4
2
0
0.0 0.5 1.0 1.5 2.0
Crack opening /mm

(c) Stress-crack opening curves of fc=115MPa>.

The authors would like to apologise for any inconvenience caused.

You might also like