Vous êtes sur la page 1sur 8

High Performance Fiber Reinforced Cement Composites (HPFRCC7),

Stuttgart, Germany – June 1-3, 2015

DEVELOPMENT OF ENGINEERED CEMENTITIOUS COMPOSITES


WITH LOCAL MATERIAL INGREDIENTS
H. Ma (1), S. Qian (2), V.C Li (1, 3)

(1) Institute of Highway and Railway Engineering, School of Transportation, Southeast


University, Nanjing, China
(2) School of Civil and Environmental Engineering, Nanyang Technological University,
Singapore, Singapore
(3) Dept. of Civil and Environmental Engineering, University of Michigan, Ann Arbor, USA

Abstract
Polyvinyl Alcohol (PVA) fibre reinforced engineered cementitious composites (ECC) is a
kind of composites reinforced with moderate fibre volume fraction, typically 2% by volume.
Of special interest is the capability of ECC material to deform to high tensile strains,
commonly over 4%, while maintaining very tight crack width. The standard ECC mix
proportion is widely available in the literature; nevertheless direct adoption of such mixture
cannot guarantee high tensile ductility due to wide material variations geographically.
Furthermore, to reduce the material cost, domestic PVA fibres must be utilized, which has
different physical/mechanical properties compared with commonly used PVA fibres from
Japan. In this paper, Chinese domestic PVA fibres were adopted to develop highly ductile
ECC material, along with other local ingredients, such as fly ash, cement, sand and crumb
rubber. Preliminary results from tensile test suggest that it is feasible to use local ingredients,
especially Chinese domestic PVA fibres to produce highly ductile ECC material in order to
reduce material cost, promoting the use of this material in civil engineering fields in China.

1 INTRODUCTION
Engineering Cementitious Composites (ECC), a special type of High Performance Fiber
Reinforced Cementitious Composites (HPFRCC), has been researched widely since it was
developed by Li [1] in 1990s. ECC has an extreme tensile ductility, in the 3~5% range (~ 300-
500 times that of concrete or FRC) [2][3]. After first cracking, the crack width of ECC
stabilizes at about 60μm while the number of cracks increases [4]. Typical ECC material
consists of cement, sand, fly ash, water, additives, and short, randomly oriented polymeric
fibers (e.g. Polyethylene, Polyvinyl Alcohol). Fiber volume fractions are minimized (Vf
=1.5~2%) for execution and economic feasibility in infrastructure applications [5][6]. At
present, almost all of the PVA fiber used in ECC is produced by Kuraray Co. Ltd., Japan; and
the price of this PVA fiber is 220,000yuan/t in Chinese market. The resulting price of ECC is

259
High Performance Fiber Reinforced Cement Composites (HPFRCC7),
Stuttgart, Germany – June 1-3, 2015

about 6700yuan/m3, while that of traditional concrete is 360yuan/m3. The intimidating high
price impedes the popularization and large-scale application of ECC. Therefore, development
of ECC with local material, especially PVA fiber, is significant to push ECC from laboratory
into practical applications. Domestic ECC using local material have been developed by many
researchers, including Mechtcherine [7] in Germany and Margareth da Silva Magalhães [8] in
Brazilian. In China, a preliminary research focusing on only flexural properties of ECC with
domestic PVA fibers was conducted by Zhang and Qian [9] in 2011.
This study focuses on developing domestic ECC with Chinese raw materials, including
cement, silica sand, fly ash, crumb rubber and most importantly PVA fiber. According to
micromechanics design theory of ECC, the strain hardening behavior of ECC was largely
governed by proper tailoring of matrix fracture toughness, fiber bridging behavior and initial
flaw size. In this study, matrix fracture toughness tailoring with fly ash and crumb rubber
were used to develop domestic ECC. The mechanical properties were assessed through
compressive and uniaxial tensile test.

2 EXPERIMENTAL PROGRAM

2.1 Materials and mixture design


All materials used in this study were produced from local factories. Silicate cement PII
42.5R, type I fly ash(FA) (Chinese standards [ 10 ][ 11 ]) and fine silica sand with size
distribution from 106 to 212 μm and a mean size of 150 μm are all produced in Jiangsu
province, China. The chemical composition of cement and fly ash are presented in Table 1. It
is well-know that fiber plays a important role in ECC development. There are three kinds of
domestic Polyvinyl Alcohol (PVA) fibre were used in this study, one is produced from
Wanwei High-tech Co. Ltd (WW) in Anhui province; and other two are produced from Bao
Hualin Co. Ltd (BHL) in Fujian province. The REC-15 fiber produced from Kuraray Co. Lt
was used to for comparison. The physical/mechanical properties of PVA fibers are listed in
Table 2.
ECC is designed based on a mutilscale framework, whereby macroscopic composite tensile
strain hardening behaviour is governed by mesoscopic fiber bridge properties. The fiber
bridging capacity is further determined by the microscopic characteristics of fiber, matrix and
fiber/matrix interface. The fiber characteristics are difficult to tailoring due to involvement of
fiber manufacturing process. Therefore, the matrix toughness and interface were tailored
through adding fly ash and crumb rubber (0.2 mm sieve size) to design domestic ECC in this
study. The mix proportions are listed in Table 3. The fly ash to cement ratio changed from 2.2
to 4.0. For the second series (D4-D6), fly ash to cement ratio was kept constant at 2.2,
whereas crumb rubber was added at15%, 25% and 35% volume replacement of the sand. The
contents of PVA fibers were fixed at 2% by total volume.

Table 1: Chemical composition of cement and fly ash


Material SiO2 Al2O3 Fe2O3 CaO SO3 P2O5 Na2O K2O TiO2 MgO
Cement 21.26 7.67 2.88 57.82 4.04 5.26 0 0.78 0.21 --
Fly ash 54.21 22.64 7.17 8.55 1.14 0.5 1.75 1.42 1.52 1.00

260
High Performance Fiber Reinforced Cement Composites (HPFRCC7),
Stuttgart, Germany – June 1-3, 2015

Table 2: Physical properties of PVA fibers


Diameter Length Elongation Density Elastic Modulus Tenacity
Fiber type MFG
(μm) (mm) (%) (g/cm3) (GPa) (MPa)
WW 35 12 7.3 1.3 31.3 1287 Wanwei
Bao
BHL 39 9 7 1.3 22 1250
Hualin
REC-15 39 12 7 1.3 42.8 1620 Kuraray

Table 3: Mix design of domestic ECCs (weight ratios)

Crumb
Sand- Water-
Mixture Fly rubber- HRW PVA fiber (by
Cement binder binder
ID ash Sand ratio RA vol)
ratio ratio
(by vol)
D0 1 0.36 2.2 -- 0.25 0.02 2%(REC-15)
D1(w/b) 1 0.36 2.2 -- 0.25 0.02 2%(WW/BHL)
D2(w/b) 1 0.36 3.0 -- 0.25 0.02 2%(WW/BHL)
D3(w/b) 1 0.36 4.0 -- 0.25 0.02 2%(WW/BHL)
D4(w/b) 1 0.31 2.2 15% 0.25 0.02 2%(WW/BHL)
D5(w/b) 1 0.27 2.2 25% 0.25 0.02 2%(WW/BHL)
D6(w/b) 1 0.24 2.2 35% 0.25 0.02 2%(WW/BHL)

2.2 Specimen and test


All ECC mixtures in this study were mixed using planetary mixer with 10L capacity. The
fresh mixtures were poured into steel modes and demoded after 24 hours; then the specimens
were cured under 95 ± 5% RH and 20 ± 2℃ until 28 days. The compressive and tensile tests
of domestic ECC were conducted in this study. Three cube specimens with the size of
75*75*75 mm were prepared for compressive test. The uniaxial tensile test was conducted by
the setup shown in Figure 1(a) under displacement control with a loading rate of 0.5 mm/min
as recommended by the Japan Society of Civil Engineers (JSCE) [12]. For each ECC mixture,
four dogbone specimens were prepared; and the dimensions of dogbone specimen are shown
in Figure 2(b). Two external LVDTs were attached on the each side of dogbone specimen to
measure the tensile extension of the narrow section.

261
High Performance Fiber Reinforced Cement Composites (HPFRCC7),
Stuttgart, Germany – June 1-3, 2015

(a) (b)
Figure 1: (a) The uniaxial tensile test setup and (b) dogbone specimen dimensions

3 RESULTS AND DISCUSSION


Figure 2 shows the typical uniaxial tensile stress-strain curves of ECCs with local
ingredients and different PVA fibers. The ECC specimens were mixed with the proportions
D0 and D1 listed in Table 3. As can be seen from this figure, ECC with REC-15 fiber reveals
an excellent tensile strain-hardening behaviour with tensile strain capacity of 5.5% and
ultimate tensile stress of 5 Mpa. By contrast, ECCs with domestic PVA fibers have a similar
moderate tensile strain capacity of about 1.8%, significantly lower compared with that of
REC-15 fiber. Their tensile strengths, however, are very different for three domestic fibers,
ranging from 3.2 to 5.5 MPa. It suggested that ECC with BHL1 fiber has a higher fiber
bridging capacity than other two fibers. Nevertheless all domestic fibers reveal relatively low
tensile ductility, which requires further improvement via micromechanical tailoring.

Figure 2: Typical uniaxial tensile stress-strain curves of ECCs with local ingredients and
different fiber (REC-15, BHL and WW)

262
High Performance Fiber Reinforced Cement Composites (HPFRCC7),
Stuttgart, Germany – June 1-3, 2015

In this study, fly ash and crumb rubber contents were adjusted to tailor the matrix
toughness of domestic ECC. Compressive strength of ECCs with different fly ash to cement
ratio are shown in Figure 3(a), where compressive strength reduces gradually with increasing
fly ash content. Fly ash reaction mainly takes place at later stage which results in the
reduction of compressive strength when the fly ash content increases. Overall, it can be seen
from this figure, there are no significant difference in compressive strength for three PVA
fibers. It suggests that the three kinds of domestic fiber have no explicit influence on ECC
compressive strength.
The typical tensile stress-strain curves of ECC with different fly ash content for three PVA
fibers are shown in Figure 4. It can be seen that all ECC specimens exhibit strain hardening
behaviour under direct tensile load. The tensile ductility increases with the increase of fly ash
to cement ratio. ECCs with different domestic PVA fiber exhibit an increase of tensile strain
capacity from 1.8% to 4.5% when the fly ash to cement ratio varies from 2.2 to 4.0.It suggests
that increasing fly ash content is beneficial to the performance of domestic ECC in tensile
ductility. Although the ultimate tensile strength reduces gradually with increasing fly ash
content, it was still maintained at 4.5 MPa for the ECCs with BHL fibers, as shown in Figure
4(a) and (b). In terms of ECC with WW fiber, however, the tensile strength is much lower
than that of BHL fiber. The ultimate tensile strength is only 2.5 MPa when the fly ash to
cement ratio is 4.0. As mentioned in the last section, it indicates that the BHL fiber in ECC
provides a higher fiber bridging capacity than WW fiber.
Another method adopted in this study to tailoring domestic ECC is via partial volume
replacement of silica sand by crumb rubber, as shown in Table 3 (D4-D6). Figure 3(b) shows
the change of compressive strength with different replacement ratio of crumb rubber. Similar
to the addition of fly ash, compressive strength decreases from 32 MPa to 18 MPa, with
increasing crumb rubber content. The crumb rubber has a lower elastic modulus than that of
silica sand, which may result in the reduction of compressive strength. Furthermore, the
reduction may also be attributed to weak interfacial bond between hydration products and
hydrophobic crumb rubber, resulting in higher porosity of ECC mixture.
Figure 5 shows the influence of crumb rubber on the tensile strain hardening behaviour of
ECCs with different domestic PVA fibers. It can be seen that the first cracking strength and
ultimate tensile strength in general reduce with the addition of crumb rubber. The first
cracking strength decrease potentially due to weak interfacial bond mentioned before. The
ultimate tensile strength reduction can be explained by the declining fiber bridging capacity
likely due to weak bond between crumb rubber and fiber. In terms of tensile ductility, the
tensile strain capacity of ECCs with BHL fiber have almost no change with the crumb rubber
content increasing, as shown in Figure 5(a) and (b). However, for ECCs with WW fiber, the
tensile ductility shows a significant increasing with the addition of crumb rubber. The tensile
strain capacity could increase to 7% when 35% volume of silica sand was replaced by crumb
rubber.

263
High Performance Fiber Reinforced Cement Composites (HPFRCC7),
Stuttgart, Germany – June 1-3, 2015

(a) (b)
Figure 3: Compressive strength of ECCs: (a) with fly ash adjustment; (b) with crumb rubber
adjustment

(a) (b)

(c)
Figure 4: Typical uniaxial tensile stress-strain curves of ECCs with different fly ash content:
(a) BHL1 fiber; (b) BHL2 fiber; (c) WW fiber

264
High Performance Fiber Reinforced Cement Composites (HPFRCC7),
Stuttgart, Germany – June 1-3, 2015

(a) (b) (c)


Figure 5: Typical uniaxial tensile stress-strain curves of ECCs with different crumb rubber
content: (a) BHL1 fiber; (b) BHL2 fiber; (c) WW fiber

4 CONCLUSIONS
This study focuses on the development of low-cost domestic ECC using Chinese
ingredients and PVA fibers. Through the addition of fly ash and partial replacement of silica
sand with crumb rubber, domestic ECC with reasonable strain hardening behavior can be
achieved. The compressive and tensile properties were investigated in this study. Following
conclusions can be drawn based on this study:
1. This preliminary study indicated that it was possible to achieve strain capacity of 3%
using local ingredients and domestic PVA fiber.
2. The use of high quantity of fly ash is conducive to producing domestic ECC with a
reasonable tensile strain capacity. The ECC mixture with BHL fiber and fly ash to cement
ratio of 4 revealed a tensile strain capacity of above 4% and ultimate tensile strength of 4.5
MPa. However, for WW fiber, the tensile strength is not below 3 MPa.
3. Replacing silica sand partially with crumb rubber in ECC composition is another
approach to enhancing tensile ductility. The results demonstrate that crumb rubber
replacement together with WW fiber can increase the tensile strain capacity significantly,
even though the ultimate tensile strength still needs to be improved. For BHL fiber, however,
the tensile ductility has no improvement by replacing sand with crumb rubber. Detailed
micro-scale investigations are needed to clarify the underlying mechanisms.

ACKNOWLEDGMENTS
The authors would like to graciously thank the National Natural Science Foundation of
China (No.51008071, 51278097), Nanyang Technological University start-up grant
(M4081208) and Southeast University for one thousand people short-term program.

265
High Performance Fiber Reinforced Cement Composites (HPFRCC7),
Stuttgart, Germany – June 1-3, 2015

REFERENCES

[1] Li, V.C. “From Micromechanics to Structural Engineering – The Design of Cementitious
Composites for Civil Engineering Application”, JSCE J. of Structural Mech. and Earthquake
Eng., 10(2). pp. 37-48, 1993.
[2] Li, V.C., “ECC – tailored composites through micromechanical modeling.” Fiber Reinforced
Concrete: Present and the Future edited by Banthia et al, CSCE, Montreal, 64-97. 1998.
[3] Li, V.C., “On Engineered Cementitious Composites (ECC) – A review of the material and its
applications.” J. Advanced Concrete Technology, 1(3): 215-230. 2003.
[4] Weimann, M. B., and V. C. Li, "Drying Shrinkage and Crack Width of ECC," Brittle Matrix
Composites-7, Warsaw, Poland, pp.37-46, Oct. 2003.
[5] Wang, S. and V. C. Li, "Polyvinyl Alcohol Fiber Reinforced Engineered Cementitious
Composites: Material Design and Performances", Proceedings of Int'l workshop on HPFRCC in
structural applications, Honolulu, Hawaii, USA, May 23-26, 2005.
[6] Li, V.C., “High Performance Fiber Reinforced Cementitious Composites as Durable Material for
Concrete Structure Repair,” Int’l J. for Restoration of Buildings and Monuments, Vol. 10, No 2,
163–180 (2004).
[7] Mechtcherine V and Schulze J. Testing the behavior of strain hardening cementitious composites
in tension. In: Proceedings of the International RILEM Workshop on High Performance Fiber
Reinforced Cementitious Composites in Structural Applications; 2006. 2006. p. 37-46.
[8] Margareth da Silva Magalhães, Romildo Dias Toledo Filho, Eduardo de Moraes Rego Fairbairn,
Influence of Local Raw Materials on The Mechanical Behaviour and Fracture Process of PVA-
SHCC, Materials Research. 2014, 17(1): 146-156.
[9] Zhang Z.G. and Qian S.Z. The feasibility of engineered cementitious composites with local
compositions. In: Proceedings of the 2nd International RILEM Conference; 2011. pp. 323-328.
[10] GB175-2007, Common Portland cement [S]. 2007. (in Chinese)
[11] GB1596-88, Fly ash Used in Cement and Concrete [S]. 1988. (in Chinese)
[12] JSCE. Recommendations for design and construction of high performance fiber reinforced
cement composites with multiple fine cracks. Tokyo: Japan Soc. Of Civil Engineers; 2008.

266

Vous aimerez peut-être aussi