Sie sind auf Seite 1von 8

Concrete Repair, Rehabilitation and Retrofitting II Alexander et al (eds) 2009 Taylor & Francis Group, London, ISBN 978-0-415-46850-3

Durability of Strain-Hardening Cement Composites (SHCC) An overview


G.P.A.G. van Zijl
Department of Civil Engineering, University of Stellenbosch, South Africa Department of Architecture, Delft University of Technology, Delft, The Netherlands

ABSTRACT: It has become possible to design fibre reinforced cement-based composites to desired mechanical performances. Among the various classes of High Performance Fibre-Reinforced Cement-based Composites (HPFRCC) that have been developed, a particular class of moderate tensile strength (36 MPa), but with pseudo strain-hardening tensile behaviour of ultra ductility, in the range 26%, is of interest here. Fibre-reinforced Strain-Hardening Cement Composites (SHCC) have such superior tensile behaviour. Matrix cracking does not lead to loss of tensile resistance, but increased tensile resistance, albeit at a lower stiffness. This is achieved by effective fibre bridging of the matrix crack, enabling resistance to higher tensile loads without significant crack widening, but rather successive cracks at the next weak points in the matrix. It is the crack control, generally to widths below 100 m despite large tensile deformation, which is of significance for durable repair, rehabilitation and retrofitting strategies. The largest source of damage in cement-based composites like steel Reinforced Concrete (RC) may be attributed to moisture, gas and chlorides ingress, whereby steel reinforcement is subjected to degradation processes. Crack width limitation or control is a well established durability concept in RC design. Design standards and codes for concrete suggest limiting values for crack widths for different environments to assure durability of RC structures. In RC crack width limitation is generally obtained through steel detailing. In SHCC this is an inherent material property. This is shown for SHCC on the material level and on the structural scale, as well as for steel reinforced SHCC elements (R/SHCC).

INTRODUCTION

High performance fibre-reinforced cement-based composites (HPFRCC) are actively researched and developed internationally. This is driven by trends in the construction industry such as increased structural size, increase in the amount of prefabrication of infrastructural elements, replacement or reduction of steel bar reinforcement to reduce labour cost or enable free form, and general better understanding of cementbased composite behaviour, enabling the design of tailor made concrete (Walraven 2007) for specified performance. In addition to structural and aesthetical requirements, improved durability of cement-based construction materials has been a continuous striving. This has lead to the development of various classes of HPFRCC, distinguished by their particular mechanical characteristics.

1.1

Classification of HPFRCC: Strength and ductility

One categorization is based on tensile strength and ductility. Fibre reinforcement was initially introduced

to improve concrete tensile strength and to reduce brittleness. The tensile response could be improved from the brittle response of plain concrete. It has now become possible by careful mix design and/or processing techniques to achieve high performance fibrereinforced cement-based composites (HPFRCC) with tensile responses shown in Figure 1. As shown in the figure, two classes define the extremities of tensile ductility and strength in HPFRCC. Strain-hardening cement composites (SHCC) do not significantly increase the matrix tensile strength, but are designed for increased tensile strength beyond first cracking and large tensile ductility. SHCC contain low to moderate amounts of fibres (1% Vf 3%). In contrast, ultra-high performance fibre-reinforced concretes (UHPFRC) are designed to have high tensile and flexural strength, as well as extremely high compressive strength (180240 MPa), but reach these strengths at moderate strain levels. It has become possible to achieve such strengths through developments such as reactive powder concrete (RPC) technology (Richard & Cheyrezy 1995) and multi-scale cement composite (MSCC) technology (Rossi 1997). Fibrereinforcement is added for ductility, at moderate fibre volumes (3% Vf 6%).

199

UHPFRC
2.1

CRACK CONTROL AS MEASURE OF DURABILITY Crack width limitation

SHCC 2

SHCC 1

Figure 1. Typical tensile behaviour of UHPFRC and SHCC.

Crack width limitation is a well established concept in RC design. Limiting values for crack widths are prescribed by design standards and codes for different environments, or so-called classes of exposure, to assure durability of structures built of RC in these environments. Distribution of steel reinforcing bars for crack width control is reasonably well understood. The pseudo strain-hardening of SHCC shown in Figure 1 involves the formation of multiple cracks. Thereby individual crack widths are arrested and new cracks arise at increased deformation, which is a form of crack control. This may be regarded as an extension of the phenomenon of crack-control in RC by reduction in steel bar diameter, but increase in number of bars to maintain the reinforcement level. However, it should be noted that unlike R/C, the crack width in SHCC does not depend on steel reinforcement, but should be regarded as an intrinsic material property. 2.2 Moisture migration in SHCC

1.2

Durability of SHCC It is generally agreed that resistance to gas and moisture ingress and migration in cement-based composites is increased by matrix densification. However, upon the formation of cracks, permeability is increased in concrete (e.g. Wang et al. 1997), as well as for mortar (Lepech & Li 2005) see Figure 2. A significant increased coefficient of permeability is found beyond a crack width threshold of roughly 0,1 mm. At a crack width of 0,3 mm the permeability is increased by several (approximately 5) orders of magnitude. This trend in crack width dependent permeability and threshold of roughly 0,1 mm is confirmed by research results by Rapoport et al. (2001).

Whereas increased strength and deformability in compression and tension, as well as general stress states (flexure, shear) remain important achievements for improved structural behaviour in various applications, durability considerations are increasingly important HPFRCC materials design criteria. It is generally agreed that gas and moisture permeability is an important measure of concrete durability, as they may be media for ingress of deleterious materials which may lead to degradation of the cementbased material or steel reinforcement. Also, general consensus exists that capillary sorption and moisture diffusion are models describing the most important mechanisms of moisture ingress and migration. In the near surface zone capillary sorption dominates moisture intake (Neithalath 2006) while moisture diffusion governs the longer term migration of water in the material through the micro-pores (Baant & Najjar 1971, Neithalath 2006). By matrix densification the capillary absorption is significantly reduced in UHPFRCC (Denari et al. 2007). In SHCC diffusivity is reduced by inherent crack control (Lepech & Li 2005, Sahmaran et al. 2007). In this paper the case of structural durability through inherent crack control of SHCC is argued. Recent developments in SHCC, in terms of characterization of micro-cracking in tension, shear and flexure of both plain SHCC and SHCC containing steel reinforcement (R/SHCC) are reviewed. The influence of loading path on crack spacing and width is discussed, with examples monotonic tensile load, cyclic tensile loading and sustained tensile load (creep).

Figure 2. SHCC and steel mesh reinforced mortar water permeability normalised by the number of cracks (Lepech & Li 2005).

200

2.3

Chloride diffusion and corrosion

Increased crack width can be related to higher chloride penetration rate in cement composites. Chloride ingress and migration in cement composites is predominantly as solvent in moisture, thus sharing the driving mechanisms of absorption and diffusion. Sahmaran et al. (2007) studied chloride penetration and permeability of SHCC in comparison to mortar. Based on results of immersion tests, chloride penetration depth was found to be reduced in uncracked SHCC specimens compared to uncracked mortar. Based on ponding tests of pre-cracked specimens, the effective chloride diffusion coefficient was found to be strongly dependent on crack width in mortar

(Figure 3a). The diffusion coefficient in SHCC was found to be comparative for equal crack widths in SHCC and mortar. However, the crack width in SHCC was found to be insensitive to the deformation level, which in this case was induced by four point bending. This explains the diversion of chloride diffusivity of mortar specimens from that of SHCC with increased deformation level (Figure 3b). The reduced penetration depth in SHCC versus reinforced mortar/concrete is confirmed by observations of Maalej et al. (2002) and Miyazato & Hiraishi (2005), in comparative SHCC and concrete beams loaded in flexure to the same deflection. In addition to chlorides ingress, Miyazato & Hiraishi (2005) studied corrosion rates in steel reinforced SHCC in comparison to R/mortarFigure 4. A significantly increased corrosion rate occurs at the location of the localized crack in R/mortar. Due to the controlled, distributed multiple cracks, this is not found in R/SHCC. 3 CRACK CONTROL IN SHCC

(a)

To exploit the inherent crack control as fundamental mechanism of structural durability, it is imperative that all modes of mechanical and environmental resistance are studied to ensure that crack control is maintained. The behaviour must be fully understood, modelled and the model parameters characterised to allow sound prediction and design for use in design of durable structures. 3.1 Tensile behaviour of SHCC

(b)
Figure 3. (a) Diffusion vs crack width in mortar. (b) Diffusion vs four point bending deformation level of SHCC and steel mesh reinforced mortar (Sahmaran et al. 2007). Engineered Cementitious Composites (ECC) form a class of SHCC.

Through balanced properties of the cement matrix, fibres and their interfaces, based on consideration of micro-mechanical mechanisms (Li et al. 1995), pseudo strain-hardening tensile behaviour is achieved. The conditions for pseudo strain-hardening are well understood and the data base of experimental results towards confirming micromechanical design models is growing. In Figure 5 the tensile response of a typical SHCC, containing 2% by volume Polyvinyl
4

Corrosion Rate (mm/year)

0.010

Stress (MPa)

Location of Large Precrack Macrocell Microcell (a) Total

(b)

Macrocell Microcell Total

3 2 1 0

0.005

0.000 5 10 15 20 25 30 Steel Location (cm) 355 10 15 20 25 30 Steel Location (cm) 35

4 5 Strain (%)

Figure 4. Corrosion rate along rebar for (a) R/mortar, and (b) R/SHCC (Miyazato & Hiraishi 2005).

Figure 5. Direct tensile stress-strain response of SHCC.

201

Alcohol (PVA) fibres, and proportional (by mass) paste ingredients: CEM I 42.5 : fly-ash : water : aggregate = 1 : 1 : 0, 4 : 0,5. Note that fine sand is used, with maximum particle size <0,2 mm.

In the figure the tensile strain-induced multiple crack formation is shown. Crack widths remain small throughout the pseudo strain-hardening part until crack saturation, after which widening and eventual localisation at a single crack occurs. For similar SHCC compositions crack widths have been measured and reported to develop to a controlled maximum width in the range 4060 m (Li et al. 2001, Weimann & Li 2003). This is within the threshold crack width for low diffusivity shown in Figures 2 and 3. In steel reinforced SHCC (R/SHCC), the pseudo strain-hardening allows compatible tensile strains between the steel bar and the SHCC, avoiding typical debonding failure modes in RCsee Figure 6. 3.2 SHCC flexural behaviour

Crack control is retained in flexure. The response to three point bending of the same SHCC mix as the previous section, is shown Figure 7. Due to the extreme tensile ductility (Figure 5), which is in fact caused by the inherent crack control, large deflections are possible, while retaining crack control. The typical load-deflection response suggests that the multiple cracking phase may be utilized in structural design even in the service condition (up to point II) in Figures 8 and 9. Typically this point marks crack saturation, after which gradual widening follows until
R/SHCC Pu Py II III

Measured responses Suggested response phases

SHCC Pe e I y u Deflection [mm]

Figure 6. Crack control and tensile strain compatibility in R/SHCC as opposed to delamination in RC (Fischer & Li 2004). Note that Engineered Cementitious Composites (ECC) form a class of SHCC.

Force [N]

Measured responses Suggested response phases

Pu III Py Pe I II

e y

Deflection [mm]

Figure 7. Three point bending response of SHCC.

Figure 8. Flexural responses of (top) SHCC and R/SHCC beams of 100 100 mm2 cross section and (bottom) thin SHCC plates. Dashed lines indicate limits of (I) linear-elastic response, (II) crack widening with subsequents localisation.

202

P
Pu
Py
Pe
II
III
II

Plastic hinge

III
Localisation

Multiple cracking

I
e

Figure 9. Schematic representation of (R/)SHCC beam load-deflection phases.

shear test, but which are not strain-hardening (Vf < 2%) as opposed to SHCC specimens (Vf 2%). The dependence of crack orientation on principle stress direction is illustrated in Figure 11. The behaviour of shear dominated beams of R/SHCC, and the resistance to combined bending and shear require further investigation to ascertain crack control under service conditions. The test results on SHCC, R/SHCC and RC beams by Visser (2008) indicate crack width control retention until close to peak flexural-shear response in three point bending of 100 mm 100 mm section beams with a span of 400 mm. Steel bars at a reinforcement level of 1% were included in the R/SHCC and RC specimens. Full details of the SHCC and RC mix are given in Visser (2007). The results are shown in Figure 12 in terms of

Steel load beam

Steel base support

1 '2

'2 1 2 1

h cos

'2

'

2 1 h sin

Figure 11. Illustration of principal stress rotation beyond first cracking in shear (van Zijl 2007).

Vf = 0

Vf = 1%

Vf = 2%

Figure 10. Ioscipescu shear test: (Top) Schematic representation of the test and specimen. (Bottom) mortar, (centre) non strain-hardening and (right) SHCC shear responses (van Zijl 2007).

localization (point III). Note that the same typical load-deformational response is found for R/SHCC, although the relative sizes of the branches III and IIIII may differsee Figure 8. Crack width control is retained beyond stage II, which implies that structural resistance to gas, moisture and chlorides ingress and diffusion is retained in a large range of mechanical resistance. 3.3 Crack control in shear

Left: R/SHCC beam symmetrical half. Crack pattern pre-peak 45 Below: R/SHCC beam Crack pattern post-peak

There is evidence (van Zijl 2007) that the multiple cracking characteristic of SHCC is retained under shear action. It is postulated that the tensile ductility enables micro-crack arrest and apparent crack rotation. This is observed in the experimental crack patterns in Figure 10 for specimens subjected to the Ioscipescu

Figure 12. Load-deflection results of R/SHCC and RC beams with 100 mm 100 mm cross-sections, showing crack patterns at (mid) pre-peak and (bottom) beyond peak of an R/SHCC beam (Visser 2007).

203

load-deflection response and typical crack patterns in an R/SHCC beam in stage III, as well as beyond peak (after stage III). The responses of R/SHCC beams produced by casting, as well as extrusion are shown in the Figure. A significantly lower initial stiffness can be seen for cast R/SHCC, which is due to higher porosity of these specimens (Visser 2007).

such time-dependent widening must considered to ensure sustained durability. This is a current research focus.

CONCLUSIONS

LOADING PATH DEPENDENCE

If structures built of (R/)SHCC are designed to operate in the multiple cracking region under in-service actions, it must be ensured that crack width control is retained under various loading paths, including cyclic loading, loading at varying rates and sustained load, whether it is sustained deformation (relaxation) or sustained force (creep). Figure 13 shows various tensile loading histories schematically. Research results from monotonic tensile tests, as well as cyclic tensile tests of SHCC specimens under deformation control, as well as under force control by Yun & Mechtcherine (2007) indicate that the number of cracks that arise during the multiple cracking phase is relatively insensitive to these different loading histories, although a 2025% lower average number of cracks arose under deformation controlled cyclic loading. The total deformability was virtually unchanged, which indicates that crack widths may be insensitive to the loading path as well. From SHCC tensile creep test results of Boshoff & van Zijl (2007), it appears that a smaller number of cracks arise in specimens which have reached a particular tensile deformation level in creep, compared with tensile specimens loaded monotonically to the same level of tensile deformation. Although new cracks do arise during the sustained load phase, widening of existing cracks appear to contribute to the time-dependent increase in deformation. In applications of SHCC with high sustained loads,

The extension of the durability design of undamaged RC through dense concrete cover of reinforcing steel, to durability in the cracked, in-service state, is addressed in this paper. Evidence of the inherent crack control of SHCC under various loading conditions and histories has been presented. Furthermore, characterised model parameters of degradation driving forces, such as water and chloride diffusivity of SHCC have been presented. SHCC has low water and chloride diffusivity through crack control to small widths (<0.1 mm) in a large deformation range. Through this characteristic, SHCC has significant potential as a repair material and general construction material which affords structural durability in service conditions, which inevitably include cracks. For applications of high sustained loads, SHCC time-dependent crack widening remains to be fully characterised to enable durability design also for this loading path.

ACKNOWLEDGEMENTS The support of the South African Ministry of Trade and Industry through the Technology and Human Resources for Industry Programme (THRIP), as well as the industrial partners of the THRIP project SAPERCS is gratefully acknowledged.

REFERENCES
Mechtcherine, V and Lieboldt, M. 2007. Effect of cracking on . air-permeability and water absorption of SHCC, RILEM PRO 53: HPFRCC5, Mainz, Germany pp. 305312. Banthia, N. and Bhargava, A. 2007. Permeability of Stressed Concrete and Role of Fiber Reinforcement, ACI Materials J. 104(1), pp. 7076. Baant, Z.P., and Najjar, L.J. 1971. Drying of Concrete as a Non-Linear Diffusion Problem, Cement and Concrete Research 1, pp. 461473. Boshoff, W.P. and van Zijl, G.P.A.G. 2007. Tensile creep of SHCC, RILEM PRO 53: HPFRCC5, Mainz, Germany pp. 8795. Jun, P. and Mechtcherine, V 2007. Behaviour of SHCC under . repeated tensile loading, RILEM PRO 53: HPFRCC5, Mainz, Germany, pp. 97104. Kunieda, M., Denari, E., Brhwiler, E. and Nakamura, H. 2007. Challenges for SHCCdeformability versus matrix density Proc. RILEM PRO 53: HPFRCC5, Mainz, Germany, pp. 3138.

Figure 13. Various tensile loading paths.

204

Lepech, M. and Li, V.C. 2005. Water Permeability of Cracked Cementitious Composites, Proceedings of ICF11, Turin, Italy, Mar. 2005, pp. 113130. Li, V .C., Mishra, D.K. and Wu, H.C. 1995. Matrix design for pseudo strain-hardening FRCC, Materials and Structures, 28, pp. 586595. Li, V .C., Wang, S. and Wu, C. 2001. Tensile strain-hardening behaviour of Polyvinyl Alcohol Engineered Cementitious Composites (PVA-ECC). ACI Materials Journal, Nov.Dec. 2001, pp. 483492. Maalej, M., Ahmed, S.F.U., and Paramasivam, P. 2002. Corrosion durability and structural response of functionally-graded concrete beams. J Adv Concr Technol, 1(3):30716. Miyazato, S. and Hiraishi, Y. 2005. Transport Properties and Steel Corrosion in Ductile Fiber Reinforced Cement Composites, Proceedings of the Eleventh International Conference on Fracture, Turin, Italy, Mar. 2025. Neithalath, N. 2006. Analysis of Moisture Transport in Mortars and Concrete using Sorption-Diffusion Approach, ACI Materials J. 103(3), 209217. Rapoport, J., Aldea, C., Shah, S.P., Ankenman, B. and Karr, A.F. 2001. Permeability of Cracked Steel FiberReinforced Concrete. Technical Report Number 115, January, 2001, National Institute of Statistical Sciences.

Richard, P. and Cheyrezy, M. 1995. Composition of reactive powder concretes, Cement and Concrete Research, 25(7), pp. 15011511. Rossi, P. 2000. Ultra-high performance fiber reinforced concrete (UHPFRC): an overview, Proceedings of the 5th International RILEM Symposium on Fiber-Reinforced Concrete (BEFIB 2000), pp. 87100. Sahmaran, M., Li, M. and Li, V 2007. Transport Prop.C. erties of Engineered Cementitious Composites Under Chloride Exposure, ACI Materials J. 104(6), 604611. Van Zijl, G.P.A.G. 2007. Improved mechanical performance: Shear behaviour of strain hardening cement-based composites (SHCC). Cement and Concrete Research, 37(8), pp. 12411247. Visser, C.R. 2007. Mechanical and structural characterisation of extrusion moulded SHCC, MScEng-thesis, University of Stellenbosch. Wang, K., Jansen, D.C., Shah, S. and Karr, A.F. 1997. Permeability study of cracked concrete, Cement and Concrete Research, 27(3) 381393. Weimann, M.B. and Li, V 2003. Hygral Behavior of Engi.C. neered Cementitious Composites (ECC), International Journal for Restoration of Buildings and Monuments 9(5), 513534.

205

Das könnte Ihnen auch gefallen