You are on page 1of 11

Early Age Cracking in Cementitious Systems 1

CHAPTER 1
INTRODUCTION: OVERVIEW OF EARLY AGE CRACKING
Arnon Bentur
National Building Research Institute-Faculty of Civil Engineering
Technion, Israel Institute of Technology

1.1 Introduction

Early age cracking sensitivity is a well known phenomenon, which is associated with drying
shrinkage. In special structures and concretes this can be aggravated by thermal effects. The
cracking sensitivity induced by drying can occur before or at the onset of setting (plastic shrinkage
cracking) or at the period of few days after beginning of setting and hardening. The traditional
methods to deal with this problem include the application of proper curing procedures (sealing or
water ponding) and the preparation of contraction joints in slabs and pavements.

The advent of modern concrete technology renewed the interest in this issue from several points of
view: application of new means to control cracking (e.g. fiber reinforcement and shrinkage
reducing admixtures) and the development of a variety of new concretes of superior performance
in the fresh and hardened state, that may be more sensitive to early age cracking (e.g. high strength
concretes of low water/binder, w/b, ratio matrix). This is reflected in a large number of
conferences which have been held in recent years on this topic [1- 7].

In order to optimize the use and design with such concretes and admixtures, there is a need for a
comprehensive understanding and treatment of early age cracking, which is much more complex
than the simplistic approach where cracking sensitivity is quantified in terms of free shrinkage
only. It is essential to evaluate, test and model the systems under restrained conditions, to take into
account the stresses developed and assess the risk of cracking by considering fracture criteria.

With this in mind, the RILEM technical committee TC 181-EAS was established to develop a
state of the art report which will provide a comprehensive treatment, to cover basic mechanisms,
engineering and design considerations as well as testing procedures. The present section highlights
in a concise manner the issues covered in this report and it serves as an introductory overview to
this topic.

1.2 Scope of the Early Age Cracking Treatment

Two types of early age cracking can be considered. The first is the one which may occur during
the stage where the concrete can be considered as a fluid, roughly before setting. This is the well
known plastic shrinkage cracking, to which some modern concretes, such as high strength/high
2 RILEM TC 181-EAS: Final Report - July 2002

performance concrete (HSC/HPC) may be more sensitive as they contain more fines than normal
concrete and therefore bleed less. The measures to overcome this sensitivity are similar to those of
conventional concrete, namely prevention of evaporation by sealing, or water ponding of the
surface. The use of small volume (approximately 0.1% by volume) of low modulus polymer fibers
can also be used to control the cracking at this stage.

The other type of early age cracking is the one which occurs as the concrete is turning from a fluid
into a visco-elastic solid, roughly at the setting time. At this stage conventional curing practices,
which are considered as adequate for normal concrete, such as sealing of the concrete surface,
may not be effective for the HSC. This type of cracking sensitivity is the topic to be covered by
the RILEM committee.

This early age cracking sensitivity in modern concretes is to a large extent the result of the low
w/b ratio typical to HSC/HPC which can lead to autogenous shrinkage, as well as the consequence
of a higher than usual cement content which is characteristic of HSC/HPC and concretes of other
compositions, such as shotcrete, which may result in “conventional” early age shrinkage as well as
thermal deformation. The superposition of shrinkage and thermal deformation should be
considered with respect to early age cracking sensitivity. An understanding of these mechanisms is
of particular significance, since the cracking seems sometime to be a random and irreproducible
event. Resolving the nature of these mechanisms and their quantification can provide the tools for
rational design to prevent, or at least mitigate, this kind of cracking.

In order to deal with early age cracking one must realize that it should be addressed at three levels
of materials and structure: (i) microscopic behavior of the cementitious matrix, (ii) the
macroscopic behavior of the concrete taking into consideration also the size of the component,
and (iii) the overall restrain induced in the structure.

The driving forces for the early age cracking are associated with mechanisms that are related to
the microstructural changes and chemical reactions taking place during the first few days after
mixing the binder with water. However, whether these driving forces will eventually develop into
sufficiently large tensile stresses that will lead to cracking depends on macro-structural
characteristics, such as the size of the component and restrain imposed internally (e.g. reinforcing
bars) or externally (e.g. supports) during the first few days. In an analysis of this kind one must
take into consideration that the shrinkage and thermal deformation are not necessarily uniform
within the cross-section of the concrete (i.e. temperature and moisture gradients). Also, the
concrete at the early age of its life may be more visco-elastic in nature, and stress relaxation may
occur under the sustained restraining load, resulting in stress which is lower than the one
calculated simply on the basis of free deformation and modulus of elasticity. Additional factor that
should be considered during the first few days is the “dynamic” nature of the system, where the
properties are changing with time. It is thus essential in a treatment of this kind to evaluate and
model the development of the engineering properties of the concrete at early age.

The internal driving forces can generate a tensile stress early on which could be quite high. This
implies that even if cracking does not seem to occur, there is considerable residual stress within
the concrete which could be easily higher than 50% of the tensile strength. Thus, any additional
unexpected load, superimposed on the internal residual stress, such as one induced by temperature
changes or movement of the formwork, can tilt the balance and result in an overall stress greater
than the tensile strength, and lead to cracking. This is perhaps the basis for the apparently
Early Age Cracking in Cementitious Systems 3

misleading site observations that the cracking seems sometimes as a random and an irreproducible
event. However, if one resolves the driving forces and evaluates the stresses that they generate, a
rational approach to design of the material and structure, to overcome this shortcoming, can be
developed. This can also be useful in the development of meaningful test methods, which can
provide a better indication of the actual performance of the material with respect to early age
cracking sensitivity and provide data of physical significance which could be used for design of
crack control.

A comprehensive document dealing with this problem should therefore include a treatment at
different levels as outline below:
(a) Definitions
Defining the scope of the problem, in particular with respect to the time of onset
of stress development (Chapter 2).
(b) Microstructural aspects
Effect of hydration on mechanical and thermal properties, moisture and
temperature distribution and the resulting deformation (Chapter 3).
(c) Engineering properties
Elastic response, visco-elastic relaxation, fracture properties, thermal properties,
shrinkage strains, activation energy (Chapter 4).
(d) Modeling
Microstructural models and calculations of autogenous shrinkage (Chapter 5).
(e) Testing
Initiation time, free shrinkage, thermal deformation, visco-elastic properties,
Internal relative humidity, stress development in restrained conditions and onset
of cracking (Chapter 6).
(f) Early age cracking of special concretes
HSC/HPC, fiber reinforced concretes, shotcrete and lightweight aggregate
concrete (Chapter 7).

1.3 Microstructural Characteristics and Driving Forces

The kinetics and nature of the hydration reactions as well as the physical characteristics of the
binder (w/b ratio and particle size distribution) are the fundamental processes and parameters
controlling directly the early age deformation (shrinkage and thermal effects). This deformation is
the driving force leading to development of stresses and possible cracking.

The hydration process of the cementitious system involves two basic inherent characteristics
which generate the effects leading to deformation and internal stress development. These are the
chemical shrinkage associated with the hydration reaction and its exothermic nature. Chemical
shrinkage is of particular concern in HSC/HPC concretes whereas thermal deformation is also an
issue in compositions rich in cement. Several proceedings of conferences dealing with these issues
were published recently [1,2,3,4]. Shrinkage stresses and cracking induced by external drying at
early age will not be considered here since this is a well established field, and engineering
practices are available to deal with it.

1.3.1 Chemical shrinkage and autogenous deformation


Chemical shrinkage is the result of the reduction in the volume of the hydrated products
compared with the reacting constituents (cement, reactive (pozzolanic) fillers and water). There is
4 RILEM TC 181-EAS: Final Report - July 2002

a distinct difference between the consequences of this chemical shrinkage at the stage where the
concrete can be considered as a fluid, and after it starts to behave as a visco-elastic solid. The
dividing time between the two is roughly the setting time.

At the stage where the concrete is a fluid like mass, the chemical shrinkage will result in reduction
of the volume of the mass, by a value which is equal to the chemical shrinkage [8,9,10]. The linear
dimensional changes are not necessarily isotropic and in a horizontal beam the volume changes
will show up as vertical contraction without any linear horizontal deformation. This is the result of
the fluid nature of the material which is free to dilate only in the vertical direction [9].

At the stage where the concrete is starting to behave like a visco-elastic mass with a rigid self-
supporting skeleton, and when the surface is sealed, the deformation of the specimen due to
chemical contraction will become more isotropic and linear contraction may also be observed
[9,10,11]. At this period, in a sealed concrete, self-desiccation due to the formation of gas filled
pores will occur. This is a process of self-desiccation, will be augmented as hydration progresses,
as long as the sealing conditions are kept [12].

Within the partially saturated pores, menisci are formed and capillary stresses develop leading to
self desiccation shrinkage [9,11,13,14].

On the basis of these concepts several models have been developed (e.g. reference 14) to predict
the self desiccation and self desiccation shrinkage as a function of the hydration kinetics, the water
binder ratio and the composition of the binder, in particular with respect to the presence of fillers
such as silica fume and fly ash. In the modeling of the self desiccation shrinkage the properties of
the solid skeleton and its visco-elastic nature have been considered, since there is contribution of
creep of the solid under the sustained capillary stress.

The extent of self desiccation is greater in lower w/b ratios systems, where the chemical
contraction makes up a bigger portion of the water present initially, and as a result autogenous
shrinkage becomes bigger (e.g. references 11,15). Binder of a finer particle size is expected to
develop a solid with finer pore structure, and therefore, for the same self desiccation, the capillary
stresses generated will be higher, leading to greater self desiccation shrinkage, e.g. systems with
silica fume, where the autogenous shrinkage under isothermal conditions are higher than a
Portland cement system of similar w/c ratio (11,16). Similar effects were predicted analytically
and demonstrated experimentally for cements of fine particle size distribution [11,17]. It was
suggested that in a system with silica fume additional autogenous shrinkage occurs once the
pozzolanic reactions begin, at about 70 hours [13,18].

Although self desiccation shrinkage induced by capillary stresses is considered by many as the
main mechanism for autogenous shrinkage, there is growing bulk of evidence that other processes
may contribute to autogenous shrinkage, in particular when temperature effects are considered:
changes in surface tension [18] and redistribution of water within the hydrated material [8,19].
This recognition came around as attempts to predict the magnitude of autogenous shrinkage in a
“real” system in which temperature changes occur over time, using the maturity concept, were not
successful. Thus, it was suggested to use the term autogenous deformation to include all these
phenomena, rather than autogenous shrinkage which is more limiting [8].
Early Age Cracking in Cementitious Systems 5

1.3.2 Thermal deformation


An additional driving force leading to linear and volume changes as well as stress development is
associated with the exothermic nature of the hydration reactions, which results in the development
of heat and thermal gradients within the concrete mass. This effect, with significant contribution at
early age, occurs simultaneously with autogenous shrinkage. Thus, the deformation measured at
this stage, under sealed conditions, is not necessarily due to autogenous deformation only, and a
contribution of thermal deformation should be considered. One of the strategies which is often
taken is to measure the temperature and the deformation simultaneously, and determine the
contribution of the thermal deformation by calculating it from the temperature change, assuming a
constant coefficient of thermal expansion. The slight expansion often observed just after setting,
between 6 to 12 hours, can be related to this thermal effect [20]. This kind of approach is often
used for modeling these combined influences, with the implied assumption that the coefficient of
thermal expansion is constant [21].

It was demonstrated that such an approach, assuming a simple additive effect of thermal and
autogenous deformation is too simplistic, as the coefficient of thermal deformation is not a
constant value: it decreases before setting and increases afterwards due to the self desiccation
process [8,19].

1.4 Engineering Behavior

1.4.1 Bleeding
Frequently, some small expansion is observed around the setting time, before autogenous
shrinkage is registered. Such expansion can be the result of temperature build up, as discussed
above, but it has also been observed in tests which can be considered as isothermal. It was
suggested that this may be due to re-absorption of bleed water which accumulated on the surface
of the beam specimen, at the time of setting [8,19].

The influence of early transient expansion due to bleeding is obviously associated with the mix
design of the concrete. It was demonstrated that it reduces at very low w/b ratio mixes where
bleeding is largely eliminated [8]. This phenomenon, although not a basic one, has practical
implications with regards to the effects which take place in the actual concrete, as well as the
definition of characteristic points in the testing and evaluation of properties.

1.4.2 Size effects


There are other influences in concrete which are related to the size of the members. They can be
classified into two:

(i) The most obvious effect is the temperature build up which will be higher for
specimens or components with larger cross section. This will compensate partially
for the autogenous shrinkage at the first day or two. To accurately model the
behavior and effect of size with respect to the temperature build up, it is essential to
consider the temperature gradient within the concrete as well as the gradients created
on cooling [22].
(ii) In sealed concrete, the autogenous shrinkage is expected to be uniform across the
whole cross-section. It is generally assumed that underwater curing will eliminate the
autogenous shrinkage, as water will readily penetrate into the partially dried pores.
This effect may not occur in low w/b ratio systems, where sealing of the surface
6 RILEM TC 181-EAS: Final Report - July 2002

layer due to the development of discontinuous pore structure may occur at a low
degree of hydration. As a result, the core of the concrete may become effectively
sealed from the water available on the surface, and self desiccation will take place at
the core even in submerged concrete.

1.4.3 Admixtures
The use of admixtures to reduce the autogenous shrinkage effects was evaluated in several studies.
Shrinkage reducing admixtures (SRA) were found to be effective in reducing the shrinkage of
HSC as well as the stress development and risk for cracking [11,23,24]. The mechanisms by
which these admixtures influence shrinkage is the reduction of the surface tension of the pore
water, which leads to reduction in the capillary stresses developed during drying. This mechanism
may be effective in reducing autogenous shrinkage, since this phenomenon is to a large extent
driven by self desiccation which induces capillary stresses [11]. It should be noted that the SRA
admixture reduces the autogenous shrinkage, but it also reduces the creep relaxation capacity [25].
These two influences are opposing, but the data suggest that the first influence, of reducing
shrinkage, is the overriding one and thus, the ‘net’ effect in restrained shrinkage conditions is a
reduction in stress development and mitigation of the risk for cracking.

1.4.4 Curing
In sufficiently small cross section the autogenous shrinkage can be eliminated if ponding is
applied immediately after casting [26]. It has been claimed that such a procedure should be
required for HSC. However, even under continuous water curing, one should not rule out the
occurrence of self desiccation if the concrete cross-section is sufficiently big, as discussed in
section 1.4.3 with respect to size effect.
A different approach to solve the curing problem has been suggested in several reports
[27,28,29,30], using the concept of internal curing by means of saturated lightweight aggregates
which serve as internal reservoirs of water.

1.4.5 Development of mechanical properties of concrete


Autogenous deformation is the driving force for the development of early age stresses and
cracking. However, the magnitude of the stresses and the likelihood of cracking will depend to a
large extent on the development of the mechanical properties at this stage. Particularly important
is the time at which significant mechanical property develops, since this is the time when stress
build up will start to occur. From an engineering point of view this time has been suggested to be
defined as the “zero” time for modeling.

Early age mechanical properties were characterized in several studies and served as a basis for
empirical modeling [31,32,33]. It was demonstrated that for 0.40 w/b ratio concrete the time to,
where significant mechanical properties start to develop, is about 10 hours [33]. For 0.50 w/b ratio
concrete this time corresponds to a degree of hydration of about 25% [31]. The trends in both
studies suggest that the rise in modulus of elasticity is quicker than that of tensile strength [31].
This is a manifestation of the cracking sensitivity at this age since the rise in modulus of elasticity
will result in build up of tensile stress.

1.5 Structural Effects: Behavior in Restrained Conditions

The practical consequences of the autogenous deformations that should be considered are the
stresses and cracking which may develop in the structure. These are dependent not only on the
Early Age Cracking in Cementitious Systems 7

magnitude of the autogenous and thermal deformation, but also on the restraining conditions in the
structure, the mechanical properties of the concrete which develop at this early age (in particular
modulus of elasticity and tensile strength) as well as the visco-elastic properties which may lead to
relaxation of the stresses. Thus, although the autogenous deformation is the driving force for the
early age distress, the magnitude of the distress is dependent on other processes and properties
which develop in the concrete during this time period. Therefore, evaluation of the autogenous
shrinkage only is not sufficient to estimate the risk for distress or to model it.

In order to study the process of the development of stresses and the occurrence of cracking, the
overall behavior should be investigated experimentally and modeled in restrained configurations
(see section 1.6 here and chapter 6 in the report).

Studies of early age stress development clearly emphasize the significant impact of the visco-
elastic behavior, which can result in considerable stress relaxation [34], amounting to more than
50% of the stress which would have been produced considering shrinkage without relaxation [34].

Thus, in assessing the influence of the composition of the concrete and the binder the influences
on shrinkage and stress relaxation should be considered simultaneously. For example, silica fume
causes much larger autogenous shrinkage, yet the stress development in a system with this mineral
admixture is smaller than that expected on the basis of the autogenous shrinkage. This trend was
shown to be the result of the higher visco-elastic response of the silica fume concrete, which at
early ages shows greater creep than a similar w/b ratio concrete consisting of Portland cement
binder only [34,36].

The restrain in longitudinal close-loop rigs (see section 6.3) might be considered as “full restrain”
since the concrete is kept at a constant deformation and is not allowed to shrink. In longitudinal
rigs with external restraining rods (e.g. reference 37), the level of restrain is a function of the ratio
between the rigidity of the rods and the concrete. This rigidity is changing over time due to the
hardening of the concrete at the early age stages. Since the overall behavior is dependent on the
restraining conditions (e.g. the visco-elastic component which induces stress relaxation) this
parameter should be considered if tests are to be developed to simulate the actual conditions. An
estimate of the restrain existing in an typical beam component is provided in reference [24] and is
outlined below.

Degree of restrain in terms of the ratio of the stiffness of the restraining member, Kr, and the
shrinking element, Ks. When the two are equal the degree of restrain is defined as 50%; in such
conditions the shrinkage of the restrained shrinking component is 50% of its free shrinkage. The
degree of restrain in % values is defined as:

% restrain = [Kr/(Ks+Kr)]*100

Calculation of the restrain of a typical bridge deck, taking into account the reinforcing steel bars,
leads to a level of 33% [24]. For an early age concrete, where the modulus of elasticity of the
concrete is one or two orders of magnitude smaller, the degree of restrain for this typical element
will be 83% and 98% for modulus of elasticity which is 10% and 1% of that of mature concrete,
respectively.
8 RILEM TC 181-EAS: Final Report - July 2002

1.6 Testing

The restrained shrinkage tests can be classified into three categories: ring tests with a restraining
core, panel tests in which the restrain is at the circumference of the panel and longitudinal tests in
which the restrain is by external rigid frame, or by instrumentation which enables to apply a load
on one of the grips (the other one is fixed) to prevent contraction of the specimen. A review of
testing techniques to evaluate early age cracking is provided in reference [38].

1.6.1 Ring tests


The ring test is perhaps the most common test that has been used to evaluate plastic shrinkage
cracking as well as the shrinkage cracking of hardened concrete. In this test the concrete ring is
cast around a restraining core (usually steel) and is allowed to shrink against it. Tensile stresses
are generated and cracking is monitored.

In view of the non-uniformity of the stress field in such tests, their common use is for comparative
purposes, to evaluate for example the efficiency of different means applied to reduce cracking
sensitivity (e.g. fiber addition and shrinkage reducing admixtures). However, it was suggested that
for certain geometries the difference between the tangential radial stresses at the inner and outer
surface of the concrete ring are small enough (10 to 20%) to consider the stress field as uniform
[39].

In the application of this test for routine evaluation, attention was given to acceleration of the test.
One such approach was based on a core of the ring which is an "active" material that expands and
induces additional tensile stress, on top of the one induced by the drying of the concrete ring [40].
This was achieved by having Perspex as the core material, and heating it up by 10oC, from 20 to
30 oC. Tests to evaluate this concept showed that for a "conventional" ring test with steel core and
concrete ring diameter ratio of 0.77, no cracks appeared during one week of drying, whereas for
the heated Perspex core of similar dimensions cracks occurred within less than 30 minutes.

1.6.2 Plate tests


The plate test was developed in several laboratories, with the intent of using it to evaluate the
performance of low volume content polymer fibers, to reduce plastic shrinkage cracking (e.g.
references 41,42). It is based on casting concrete in a slab mold with dimensions of about 1*1m,
with restrain being achieved by reinforcement at the edges of the plate. The surface is exposed to
drying which is usually achieved by a fan, simulating windy conditions. The performance of the
concrete is expressed by quantification of the extent of cracking.

1.6.3 Longitudinal tests


The tests in this category can be divided into three:
(i) The restrain along or at the end of the longitudinal geometry is used to generate cracks
(referred to as longitudinal-qualitative). The test results are quantified in terms of the size,
length and time of crack occurrence [43-46].
(ii) The rig is partially instrumented to determine the restraining forces and stresses. The test
is passive in nature, since the restrain is achieved by longitudinal bars on which strain gauges
are mounted (referred to as longitudinal-passive). The test results are quantified in terms of
the stress which is generated as well as the size, length and time of crack occurrence [37, 47].
(iii) Fully instrumented test, in which a closed loop system is operating, to adjust the position
of the grips to keep the specimen at zero deformation or close to it. This test may be viewed
Early Age Cracking in Cementitious Systems 9

as an active one, since there is a continuous need for external intervention to keep the
specimen at zero strain (referred to as longitudinal-active). The test set up enables to
characterize the stress development, stress relaxation as well as characterization of the cracks
[16, 48, 49, 50].

1.7 Conclusions

The review presented in this paper indicates the complexity of the issue of early age cracking
sensitivity. In order to advance a rational approach to deal with this problem and develop models
which could be used for design, there is a need to deal with the issue simultaneously at the three
levels identified above: basic micro-analysis of the driving forces for the phenomena of
deformation, engineering properties of the concrete and the nature of the structural system with
respect to the restrain it induces. All of these should be combined to calculate the stresses
developed and the subsequent risk of cracking.

In order to quantify these models with the appropriate parameters there is a need to resort to tests
which will provide information which is of physical relevance to the basic characteristics of the
concrete. This requires more sophisticated test methods.

From a practical point of view, the approach to reduce or eliminate autogenous deformation may
require the development of innovative curing methods. Porous lightweight aggregates is a
demonstration for such an approach.

1.8 References

1. Tazawa, E., (editor), AUTOSHRINK’98, Proc. Int. Workshop on autogenous shrinkage of concrete, Japan
Concrete Institute, Japan (1998).
2. Persson, P. and Fagerlund, G., (editors), ‘Self Desiccation and Its Importance in Concrete Technology’,
Proc. Int. Res. Seminar, Lund, Sweden, (1997).
3. Persson, B. and Fagerlund, G., (editors), ‘Self Desiccation and Its Importance in Concrete Technology’,
Proc. Int. Res. Seminar, Lund, Sweden, (1999).
4. Holand, I. and Sellevold, E.J., (editors), ‘Utilization of High Strength/High Performance Concrete’,
Proc. 5th Int. Symp., Sandefjord, Norway, (1999).
5. Mihashi, H., (editor), ‘International Workshop on Control of Cracking in Early Age Concrete’, Tohoku
University, Sendai, August 2000, Japan.
6. Baroghel-Bouny, V. and Aitcin, P.-C., ‘Shrinkage of Concrete – Shrinkage 2000’, Proc. RILEM
Workshop, Paris, October 2000, Proceedings PRO 17, (RILEM, 2000).
7. Kovler, K. and Bentur, A., ‘Early Age Cracking in Cementitious Systems’, Proc. International RILEM
Conference, Haifa, March (2001).
8. Bjøntegaard, Ø., ‘Thermal Dilation and Self Desiccation as Driving Forces to Self-Induced Stresses in
High Performance’, Ph.D thesis, Division of Structural Engineering, The Norwegian University of
Science and Technology, Norway, (1999).
9. Barcelo, L., Boivin, S., Rigaud, S., Acker, P., Clavaud B.and Boulay, C., ‘Linear vs. Volumetric
Autogenous Shrinkage Measurements: Material Behavior or Experimental Artifact’, ref. 3, 109-126.
10. Justens, H., Van Gemart, A., Verboven F. and Sellevold, E.J., ‘Total and External Chemical Shrinkage of
Low W/C Ratio Cement Pastes’, Advances in Cement Research, 8 (31) (1996) 121-126.
11. Tazawa E. and Miyazawa, S., ‘Influence of Constituents and Composition on Autogenous Shrinkage of
Cementitious Materials’, Magazine of Concrete Research, 49 (178) (1997) 15-22.
12. Persson, B., ‘Self-desiccation and its Importance in Concrete Technology’, Materials and Structures, 30
(1997) 293-305.
10 RILEM TC 181-EAS: Final Report - July 2002

13. Jensen, O.M., and Hansen, P.F., ‘Autogenous Deformation and Change of the Relative Humidity in Silica
Fume Modified Cement Paste’, ACI Materials Journal, 93 (6) (1996) 539-543.
14. Hua, C., Acker P. and Ehrlacher, A., ‘Analysis and Models of the Autogenous Shrinkage of Hardening
Cement Paste I. Modeling at Macroscopic Scale’, Cement and Concrete Research, 25 (7) (1995) 1457-
1468.
15. Tazawa E. and Miyazawa, S., ‘Experimental Study on Mechanisms of Autogenous Shrinkage of
Concrete’, Cement and Concrete Research, 25 (8) (1995) 1633-1638.
16. Bloom, R. and Bentur, A., ‘Free and Restrained Shrinkage of High Strength Concrete with and without
Silica Fume’, Amer. Concr. Inst. Maters. J., 92 (2) (1995) 211-217.
17. Bentz, D.P. and Haecker, C.J., ‘An Argument for Using Coarse Cements in High-Performance
Concretes’, Cement and Concrete Research, 29 (1999) 615-618.
18. Jensen, O.M. and Hansen, P.F., ‘Influence of Temperature on Autogenous Deformation and Relative
Humidity Change in Hardening Cement Paste’, Cement and Concrete Research 29 (1999) 567-575.
19. Bjøntegaard, Ø. and Sellevold, E.J., ‘Thermal Dilation-Autogenous Shrinkage: How to Separate?’ ref. 1,
233-244.
20. Lepage, S.. Baalbaki, M., Dallaire E. and Aitcin, P.-C., ‘Early Shrinkage Development in a High
Performance Concrete’, Cement, Concrete and Aggregates, 21 (2) (1999) 31-35.
21. Technical Committee on Autogenous Shrinkage of Concrete, , Committee Report, Japan Concrete
Institute, ref. 1, 5-66.
22. Jonassen, J.-E., Groth P. and Hedlund, H., ‘Modeling of Temperature and Moisture Field in Concrete to
Study Early Age Movements as a Basis for Stress Analysis’, in ‘Thermal Cracking at Early Ages’, R.
Springenschmid (editor), E&FN SPON (1994) 45-52.
23. Folliard, K.J. and Berke, N.S., ‘Properties of High Performance Concrete Containing Shrinkage Reducing
Additives’, Cement and Concrete Research, 27 (9) (1997) 1357-1364.
24. Bentur, A., Berke, N.S., Dallaire M.P. and Durning, T., ‘Crack Mitigation Effects of Shrinkage Reducing
Admixtures’, ‘Design and Construction Practices to Mitigate Cracking’, E.G.Nawy, F.G.Barth and
R.J.Frosch, editors, ACI SP-204, (The American Concrete Institute, 2001) 155-170.
25. Brooks, J.J. and Jiang, X., ‘The Influence Of Chemical Admixture on Restrained Drying Shrinkage of
Concrete’, in Superplasticizers and Other Chemical Admixtures in Concrete, Proc. 5th CANMET/ACI
Conference, M. Malhotra (editor), ACI SP-173 (1997) 249-265.
26. Bentur, A., Igarashi S. and Kovler, K ‘Control of Autogenous Shrinkage Stresses and Cracking in High
Strength Concretes’, ref. 4, 1017-1026.
27. Bentur, A., Igarashi S. and Kovler, K., ‘Internal Curing of High Strength Concrete to Prevent Autogenous
Shrinkage and Internal Stresses by Use of Wet Lightweight Aggregates’, Cement and Concrete
Research 31 (2001) 1587-1591.
28. Kohno, K., Okamoto, T., Ishikawa, Y., Sibata T. and Mori, H., ‘Effects of Artificial Lightweight
Aggregate on Autogenous Shrinkage of Concrete’, Cement and Concrete Research, 29 (1999) 611-614.
29. Van Breugel, K. and de Vries, H. ‘Mix Optimization of HPC in View of Autogenous Shrinkage’, ref. 4,
1041-1050.
30. Bentz, D.P. and Snyder, K.A., ‘Protected Paste Volume in Concrete, Extension to Internal Curing Using
Saturated Lightweight Fine Aggregate’, Cement and Concrete Research, 29 (1999) 1863-1867.
31. De Schutter, G. and Taerwe, L., ‘Degree of Hydration-Based Description of Mechanical Properties of
Early Age Concrete’, Materials and Structures, 29 (1996) 335-344.
32. De Schutter, G., and Taerwe, L., ‘Fracture Energy of Concrete at Early Ages’, Materials and Structures,
30 (1997) 67-71.
33. Bjøntegaard, Ø., Kanstad, T. Sellevold E. and Hammer, T.A. ‘Stress-Inducing Deformation and
Mechanical Properties of Concrete at Very Early Ages’, ref. 4, 1027-1040.
34. Igarashi, S., Bentur A. and Kovler, K., ‘Stresses and Creep Relaxation Induced in Restrained Autogenous
Shrinkage of High-Strength Pastes and Concretes’, Advances in Cement Research, 11 (4) (1999) 169-
177.
35. Igarashi, S., Bentur A. and Kovler, K., ‘Autogenous Shrinkage and Induced Restraining Stresses in High
Strength Concretes’, Cement and Concrete Research, 30 (11) (2000) 1701-1707.
Early Age Cracking in Cementitious Systems 11

36. Kovler, K., Igarashi S. and Bentur A., ‘Tensile Creep Behavior of High Strength Concretes at Early
Ages’, Materials and Structures, 32 (1999) 383-387.
37. RILEM TC 119-TCE, ‘Avoidance of Thermal Cracking in Concrete at Early Ages-Recommendations’,
Materials and Structures, 30 (1997) 451-461.
38. Bentur, A. and Kovler, K., ‘Early Age Shrinkage Stresses and Cracking in Cementitious Systems’, in
Construction Materials-Theory and Application, Hans-Wolf Reinhardt zum 60. Geburstag, H.von
R.Eligehausen (editor), Germany (1999) 45-56.
39. Grysbowski, M. and Shah, S.P., ‘Shrinkage Cracking of Fiber Reinforced Concrete’, American Concrete
Inst. Materials J., 87 (2) (1990) 395-404.
40. Kovler, K., Sikuler, J. and Bentur, A., ‘Restrained Shrinkage Tests of Fibre-Reinforced Concrete Ring
Specimens: Effect of Core Thermal Expansion’, Materials and Structures, 26 (1993) 231 - 237.
41. Kraii, P.P., ‘Proposed Test to Determine the Cracking Potential Due to Drying Shrinkage of Concrete’,
Concrete Construction, 30 (1985), 775-778.
42. Yokoyama, K., Hiraishi, S., Kasai, Y. and Kishitani, K., ‘Shrinkage and Cracking of High Strength
Concrete and Flowing Concrete at Early Ages’, in ACI SP-148, American Concrete Institute (1994)
243-258.
43. Berke, N.S. and Dallaire, M.P., ‘The Effect of Low Addition Rates of Polypropylene Fibers on Plastic
Shrinkage Cracking and Mechanical Properties of Concrete’, in ‘Fiber Reinforced Concrete:
Developments and Innovations’, J.J.Daniels and S.P.Shah (editors), ACI SP-142-2, American Concrete
Institute (1994) 19-42.
44. Banthia, N., Azzabi, M. and Pigeon, M., ‘Restrained Shrinkage Cracking in Fibre-Reinforced
Cementitious Composites’, Materials and Structures, 26 (1993) 405-413.
45. Banthia, N., Azzabi, M. and Pigeon, M., ‘Restrained Shrinkage Tests on Fiber Reinforced Cementitious
Composites’, in ACI-SP 155-7, American Concrete Institute (1995) 137-151.
46. Banthia, N., Yan, C. and Mindess, S., ‘Restrained Shrinkage Cracking in Fiber Reinforced Concrete: A
Novel Technique’, Cement and Concrete Research, 26 (1) (1996) 9-14.
47. Breitenbucher, R., ‘Investigation of Thermal Cracking with the Cracking Frame’, Materials and
Structures, 23 (1990) 72-177.
48. Springenschmid, R., Breitenbucher R. and Mangoid, M., ‘Development of the Cracking Frame and
Temperature Testing Machine’ , in Thermal Cracking at Early Ages, R.Springenschmid (editor), E&FN
SPON (1994) 137-144.
49. Van Breugel, K., de Vries, J. and Takada, K., ‘Mixture Optimization of Low Water/Cement Ratio High
Strength Concretes in View of Reduction of Autogenous Shrinkage’, Paper presented at the Int. Conf.
on High Performance Concretes, Sherbrooke, Canada, August, 1998.
50. Kovler, K., ‘Testing System for Determining the Mechanical Behavior of Early Age Concrete Under
Restrained and Uniaxial Shrinkage’, Materials and Structures, 27 (1994) 324-330.

You might also like