You are on page 1of 19

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/273796261

A MODELLING STUDY OF DRYING SHRINKAGE DAMAGE IN CONCRETE


REPAIR SYSTEMS

Conference Paper · July 2014

CITATIONS READS

14 617

5 authors, including:

Mladena Luković Branko Šavija


Delft University of Technology Delft University of Technology
78 PUBLICATIONS 1,871 CITATIONS 209 PUBLICATIONS 5,006 CITATIONS

SEE PROFILE SEE PROFILE

Erik Schlangen Guang Ye


Delft University of Technology Delft University of Technology
458 PUBLICATIONS 17,360 CITATIONS 427 PUBLICATIONS 14,439 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Mladena Luković on 21 March 2015.

The user has requested enhancement of the downloaded file.


A MODELLING STUDY OF DRYING SHRINKAGE DAMAGE IN CONCRETE
REPAIR SYSTEMS
Mladena Luković1, Branko Šavija1, Erik Schlangen1, Guang Ye 1,2, Klaas van Breugel1
1-Delft University of Technology 2- Ghent University
Faculty of Civil Engineering and Geosciences Magnel Laboratory for Concrete Research,
Stevinweg 1, 2628 CN Technologiepark-Zwijnaarde 904 B-9052,
Delft, Ghent (Zwijnaarde),
The Netherlands Belgium

KEYWORDS: Concrete repair, shrinkage, strain-hardening cementitious composite (SHCC), lattice


model

ABSTRACT
Differential shrinkage between repair material and concrete substrate is considered to be the main cause
of premature failure of repair systems (Martinola, Sadouki et al. 2001, Beushausen and Alexander 2007).
Magnitude of induced stresses depends on many factors, for example the amount of restraint, moisture
gradients caused by different curing and drying conditions, type of repair material, etc. Once stresses
exceed the strength, two damage mechanisms may take place: debonding (curling) of the repair material
or/and cracking inside the repair material or concrete substrate.

Numerical simulations combined with experimental observations can be of great use when determining
the influence of governing parameters and predicting the performance of a repair system. In this work, a
lattice type model is used first to simulate moisture distribution inside a repair system in time, and then to
model cracking caused by resulting shrinkage. The influence of substrate surface preparation, bond
strength between the two materials, thickness of the repair material, and different types of repair
materials is investigated. One of the promising materials which was developed recently and showed to
have promising properties for application in a concrete repair is a strain hardening cementitious
composite - SHCC. Benefits of applying either SHCC or concrete as a repair material over non-reinforced
repair mortar are discussed.

INTRODUCTION
Concrete repair implies integration of the new (repair) material with old concrete substrate in order to
form a composite system capable of enduring exposure to mechanical loads and varying environmental
conditions. What makes this goal difficult is mismatch in age, properties, and performance of two
materials which often leads to premature failure of the multi-layered system (Emberson and Mays , Tilly
and Jacobs 2007).

After casting, the repair material is exposed to ambient temperature and relative humidity. Due to the
environmental drying, ongoing hydration, and moisture absorption by the concrete substrate, repair
material loses water and tends to shrink. However, its deformation is restrained by the concrete substrate.
As a consequence, stresses build up in the repair system and cracking and/or debonding follows. Cracking
or debonding of the overlay reduces the load-carrying capacity of the system and allows water and other
hazardous substances to penetrate into concrete and further speed up the deterioration process.

In order to make some practical design recommendations, a number of analytical models for bonded
overlays subjected to differential shrinkage have been developed (Birkeland 1960, Silfwerbrand 1997,
Denarié, Granju et al. 2004, Beushausen and Alexander 2007, Zhou, Ye et al. 2008, Denarié,
Silfwerbrand et al. 2011). Also, a number of two dimensional continuum models were used (Martinola
and Wittmann 1995, Asad, Baluch et al. 1997, Granger, Torrenti et al. 1997, Sadouki and Van Mier 1997,
Martinola, Sadouki et al. 2001, Habel, Denarié et al. 2006). A way of coupling moisture transport and
fracture simulations, while taking into account influence of drying on cracking mechanism is introduced
also through discrete lattice type modelling (Bolander and Berton 2004). Benefits of using lattice models
for fracture simulations are that they can mimic physical structure and processes, when realistic crack
patterns can be achieved (Schlangen 1993, Landis and Bolander 2009). Besides modelling fracture of
brittle materials (i.e. concrete, mortar or cement paste), a ductile behaviour such as that of fibre reinforced
composites can be simulated (Li, Perez Lara et al. 2006, Schlangen and Qian 2009). Material
micro/mesostructure in these fracture models can be explicitly represented, while randomness of the
lattice mesh enables that material aspect are captured in a probabilistic sense. The same concept can be
applied for lattice based transport processes modelling. The corresponding transport properties are
assigned to the lattice elements which are idealized as conductive “pipes”. Material structure can be
captured thorough assigning different diffusivity (conductivity) properties to lattice elements that
represent a certain material phase (Sadouki and Van Mier 1997, Šavija, Pacheco et al. 2013).

The aim of this paper is to couple moisture and fracture processes during drying in concrete repair
systems. A 3D mesoscale lattice type model is used for simulating moisture distribution and resulting
damage development in the repair system. Transport and fracture phenomena are studied based on
incorporated mesostructure. The influence of different parameters, for example substrate surface
preparation (substrate roughness and moisture preconditioning), bond strength between two materials,
thickness of the repair material and different types of repair material on the damage caused by differential
shrinkage in concrete repair is investigated. The damage in the repair system can be reduced if repair
material is reinforced by fibres, as in case of strain-hardening cementitious composite-SHCC (Li, Horii et
al. 2000). Fibres bridge crack faces and prevent that large crack widths are reached. An alternative way of
improving performance of concrete repair is through reducing differential shrinkage in the system. This
can be achieved by including coarse aggregates in the repair material mixture. Aggregates act as stiff,
non-shrinking inclusions in the repair material which restrain shrinkage deformation and enable more
stable crack propagation in the repair system. How damage progress changes and subsequent benefits
when either fibres or coarse aggregates are included in the repair material mixture are studied.

LATTICE MODEL
Lattice models have been widely used to simulate fracture, moisture transport and chloride diffusion in
cement-based materials (Schlangen 1993; Bolander and Berton 2004; Šavija, Pacheco et al. 2012). In the
transport lattice approach, concrete is treated as an assembly of one-dimensional “pipes”, through which
the flow takes place. In the mechanical lattice approach, concrete is discretized as a set of truss or beam
elements which transfer forces. Output from the moisture model is used as an input for simulating
fracture. This is a one-way coupling process– moisture transport does effect mechanical fracture, but
there is no influence of the (micro)cracking on the moisture transport. This is a reasonable assumption as
drying shrinkage microcracks should not have significant effect on drying rate of mortars and concrete
(Bisschop and van Mier 2008).

Spatial discretization
The approach proposed here uses the same lattice network for both fracture and moisture simulations. For
the spatial discretization of the specimen in three dimensions, the basis is the prismatic domain (Šavija,
Pacheco et al. 2013). Discretization of the domain is performed according to the following procedure:
• A cubical grid is chosen (square for 2D lattice) and the domain is divided into a number of cubic
cells. In all simulations presented herein, linear dimension of the cubical grid for mesh generation
is 1 mm.
• In each cell (square for 2D, cube for 3D lattice), a random location for a lattice node is selected.
First the nodes are randomly positioned inside a sub cell of size s in a regular grid with size A
(Figure 1a). The ratio s/A is defined as randomness of a lattice and here it is set to be 0.5. This
means that some disorder is built into the lattice mesh itself.
• Voronoi tessellation of the prismatic domain with respect to the specified set of nodes is
performed. Nodes with adjacent Voronoi cells are connected by lattice elements (Figure 1). Since
Voronoi diagrams are dual with Delaunay tessellation, this approach is equivalent to performing a
Delaunay tessellation of the set of nodes (Šavija, Pacheco et al. 2013).
• In order to take material heterogeneity into account, either a computer-generated material
structure, or a material structure obtained by micro-CT scanning (Schlangen 2008, Landis and
Bolander 2009) can be used. Here, the concrete mesostructure was simulated using the Anm
material model originally developed by Garboczi (Garboczi 2002) and implemented in a 3D
packing algorithm by Qian (Qian 2012). It is based on placing multiple irregular shape particles
separated into several sieve ranges into a predefined empty container. Aggregates smaller than 4
mm are not explicitly modelled and together with matrix are considered as a mortar.
• Material overlay procedure (schematically shown in Figure 1a and 1b) is employed: the beams
which belong to each phase are identified by overlapping material mesostructure (i.e. substrate/
repair material mortar and aggregates) on top of the lattice. Interface elements are generated
between substrate nodes and repair material nodes (Interface MS/RM, Figure 1a) while aggregate-
paste interface (ITZ) elements are generated between mortar nodes and aggregates (Figure 1b). In
this manner different transport and mechanical properties are assigned to different phases (Table 1
and Table 2 respectively). Interfaces, as used in the present model, do not exactly coincide with
the size of real interfaces. In reality, interface thickness is in a range of tens of micrometres, while
interface elements in the present model take up also a piece of aggregate (or repair material) and a
piece of mortar (Figure 1). Therefore, the actual size of the interface in the model depends on the
characteristic element size and, in presented simulations, is around 1mm.

a) b)
Figure 1: Two-dimensional overlay procedure for generation of the lattice model in interface zone
a) between substrate and repair material b) between aggregate and substrate mortar (ITZ)

• For the fracture simulations, fibre elements are added in the repair material according to a design
volume content (2%), fibre length (8mm) and diameter (80 microns). The location of the first node
of each fibre is chosen randomly in the specified volume and a random direction is defined which
determines the position of the second node. If the second node is outside of the mesh boundary,
then the fibre is automatically cut off and accounted for in order to ensure preservation of
prescribed volume content.
• Extra nodes inside the fibres are generated at each location where the fibre crosses the square (in
3D cubical) grid.
• Fibre/matrix interface elements are generated between fibre nodes and the matrix nodes in the
neighbouring cell. Also, the end nodes of the fibres are connected with an interface element to the
matrix node in the cell where the fibre end is located (figure 1).
• Both aggregates and fibres are simulated with periodic boundary conditions. This means that one
side of the specimen is connected to the other end and that the properties are periodically
repeating. The example for the periodic boundary conditions of the fibres is given in Figure 2a and
for aggregates in Figure 2b. In following simulations, as the third dimension of the specimen is
small (5mm), periodic boundary conditions are simulated only in one direction.
• Fibre elements and fibre/matrix interface elements do not take part in the moisture transport and
therefore are not modelled in moisture simulations.
a) b)
Figure 2: Periodic boundary conditions for a) fibres in two directions b) aggregates in three
directions (Qian 2012)

LATTICE MOISTURE MODEL


A random lattice (s/A=0.5) is used first to model moisture transport caused by water exchange in repair
system. The proposed model treats concrete as an assembly of one-dimensional linear, “pipe” elements,
through which moisture transport takes place. A governing partial differential equation for moisture
transport is:
  
  (1)
  
where H is the relative humidity and D(H) is a humidity dependent diffusion coefficient which can be
given as:

  (2)
Here β and γ are parameters that can be determined by calibration with experimental results (Ayano and
Wittmann 2002). In the model presented herein, although transport phenomena is in reality a multi-scale
problem, no distinction is made between the different flow mechanisms that take place at the micro-scale
(Quenard and Sallee 1992).

If equation (1) is discretized using the standard Galerkin method (Lewis, Nithiarasu et al.), the following
set of linear equations arises (in matrix form):
     (3)
where M = the element mass matrix; K = the element diffusion matrix and f = forcing vector. Vector H is
the vector of unknown quantities (in this case relative humidity) and dot over H indicates the time
derivative. M and K, have the following forms:
  2 1   1 1
           (4)
1 2  1 1
where lij is the length of the element between nodes i and j, Aij is its cross sectional area, and D(H) its
diffusion coefficient. Cross sectional areas of individual elements are assigned using the so-called
Voronoi scaling method (Yip, Mohle et al. 2005). All matrices are equivalent to those of regular one-
dimensional linear elements (Lewis, Nithiarasu et al.), except the correction parameter ω in the mass
matrix (equation 4). This parameter is used to convert the volume of all lattice elements to the volume of
the specimen, due to overlap of volume of adjacent lattice elements (figure 3). Therefore, ω corresponds
to the ration between the total area of Voronoi facets through which moisture transport takes place and
lattice represented volume, and can be determined as (Nakamura, Srisoros et al. 2006):
∑"  
 #$%   (5)
where m is the total number of elements, Ai and li cross sectional area and length of each lattice element, k
element number, and V the total volume of the specimen.

When flux, qs, occurs between the material boundary and the atmosphere, it is necessary to account for
convective boundary conditions:
q '  () *  + (6)
Figure 3: Definition of overlap area for determination of parameter ω

Here, Cf is the film coefficient, qs is the moisture flux across the boundary, Hs and Ha are the relative
humidities at the material surface and surrounding atmosphere, respectively. In the lattice model,
evaporation rate is implemented through force vector in the element e:
,-  1 01    +
  .     .   *,3  (7)
0 0
where Hs,i is the relative humidity of the surface node (node at the surface exposed to drying) and ϑ is the
correction factor which is determined as
∑5 
4  #$ -
(8)

Here, n is the total number of elements corresponding to the surface nodes, As is the area of Voronoi
facets corresponding to these elements and A is the area of the surface exposed to drying. The concept is
similar as for determination of the correction factor ω.

Using the Crank-Nicholson procedure (Lewis, Nithiarasu et al.), the system of linear equations is then
discretized in time, and the following equation results:
  0.5∆9 :;<  :    0.5∆9 :;<  :;<  ∆9 (9)
This equation is then solved for each discrete time step (∆t) and moisture contents are obtained. Since
parameter D and therefore, K is dependent on relative humidity H, iterative procedure is avoided by
calculating relative humidity in each step (n) based on values of hydraulic diffusivity determined from the
previous step (n-1). Although this implies a certain amount of error, it significantly shortens the
simulation time and, the error is small for small time step ∆t. Benefits of using lattice type moisture
model is that transport and fracture simulations can be further coupled while the same material
mesostructure and mesh are used.

VERIFICATION OF MOISTURE TRANSPORT MODEL-OVERLAY SYSTEM


For the analysis of the time dependent drying shrinkage, overlay system is simulated. The same problem
was studied numerically by Wittmann and Martinola (Martinola and Wittmann 1995), Sadouki and van
Mier (Sadouki and Van Mier 1997) and Bolander and Berton (Bolander and Berton 2004). Repair
material thickness is 40 mm while substrate thickness is 200 mm. System is exposed to environment with
50% relative humidity (Ha=0.5) and film coefficient of the surface is 0.7mm/day. Concrete substrate was
pre-saturated, which means that top layer of 20 mm has the same initial relative humidity as repair
material (H=1). The third dimension of the simulated sample is 5 mm.

To check the accuracy of the lattice model, non-linear solution procedure, and implementation of
convective boundary conditions, simulation results are compared with those from commercial finite
element model FEMMASSE MLS (Femmasse 1996). The same material parameters and boundary
conditions are used for both simulations. In following simulations, the terms in the diffusivity equation
(2), β and γ, were set to obtain humidity profiles similar to those proposed by Martinola and Wittmann
(Martinola and Wittmann 1995, Martinola, Sadouki et al. 2001). As in the continuum model concrete
microstructure is not take into consideration, concrete substrate was first simulated as homogenous
material, with the same diffusivity properties of all elements (Mortar substrate from Table 1). Moisture
profiles at 1 day, 10 days and 110 days, obtained with FEMMASSE and lattice model are presented in the
Figure 4. Both simulations show the same tendency. Due to the drying, repair system loses water from the
top. Also moisture exchange inside the substrate occurs as unsaturated part tends to get water from
initially pre-saturated substrate layer.

For different w/c ratio and aggregate content of the materials, diffusion parameters (Table 1) will change.
In order to obtain moisture diffusivity, concrete substrate and repair mortar should be exposed to different
environmental conditions, and an inverse numerical technique should be applied on processing the
experimental data (Martinola and Wittmann 1995). Therefore, for different repair materials and
substrates, diffusivity profiles have to be measured and subsequently included as input in simulation.

Table 1: Diffusivity parameters for lattice moisture model and FEMMASSE


D Repair Interface Mortar ITZ Aggregates Evaporation
[mm2/day] material Substrate Cf [mm/day]
β 0.022 0.022 0.022 0.066 0.00022
0.7
γ 7.5 7 4 4 0

Figure 4: Moisture profiles at certain time steps from lattice model and FEMMASSE model

Once the model is verified, concrete substrate mesostructure is included. The volume percentage of the
crushed stone in the simulated concrete substrate is 30%. 67% of aggregates are in the sieve range [8,16]
mm and 33% are in the sieve [4,8] mm. Aggregates are considered impermeable and are ascribed very
low diffusivity properties (Table 1). The surface nodes of the aggregates, which are in contact with
surrounding matrix, have the same initial relative humidity as the surrounding matrix. As a zone with
higher porosity, ITZ has higher diffusivity properties than bulk matrix. In the simulations, 3 times higher
diffusivity compared to the bulk matrix diffusivity is ascribed to ITZ which is in line with ratio obtained
by Šavija (Šavija, Pacheco et al. 2013). In their study, effective diffusion rate for chloride penetration in
concrete was calculated, and 2.5-7 times higher diffusivity properties for ITZ elements compared to bulk
matrix were obtained. Since both diffusivity of chloride penetration and water movement are directly
related to the porosity, the same ratio is assumed to be valid in this study.

From the moisture profiles (Figure 5), it can be observed that ITZ and material heterogeneity have
significant influence of the drying process and moisture exchange. In the continuum model and
homogenous lattice mesh, obtained profiles are uniform (Figure 5a and 7b). However, in the system with
explicitly modelled material mesostructure, local conditions around aggregates and higher diffusion rate
in the ITZ enables that the moisture profile is not uniform but strongly dependant on aggregate
distribution and connectivity of locally more porous ITZ zone. Therefore, although at the location of
aggregates no moisture transport occurs, drying around the aggregates is faster compared to drying front
in the substrate mortar (Figure 5c). Faster drying in the ITZ zone may cause local stress concentrations
around the aggregates, which can trigger microcracking in this locally weaker zone. Microcracking leads
to the reduction of the local constrains in the microstructure and further determines fracture propagation
during the ongoing drying. With the included material mesostructure and local moisture gradients, time
dependent shrinkage cracking in heterogeneous and multi-layered materials are further studied.

a) b) c)
Figure 5: Moisture distribution at the 110 days when used a) FEMMASSE model b) lattice model without
aggregates c) lattice model with material meso-structure included

LATTICE FRACTURE MODEL


In lattice fracture models the material is discretized as a network of truss or beam elements connected at
the ends. All the single elements have linear elastic behaviour. In each loading step, an element that
exceeds limit stress or strain capacity is removed from the mesh. The analysis procedure is then repeated
until a pre-determined failure criterion is achieved. In this way realistic crack patterns can be obtained.
Further on, without introducing material softening (elements have locally brittle behaviour) structural
softening and ductile global behaviour can be simulated (Schlangen 1993).

For the fracture criterion, only axial forces are taken into account to determine the stress in the beams. An
element in lattice fracture model can fail either in tension or in compression, when the stress exceeds its
strength.

Coupling lattice moisture model with lattice fracture model


For coupling moisture and fracture analysis, it might be assumed that the moisture distribution produces a
shrinkage field according to equation (Martinola, Sadouki et al. 2001):
=*>  ?*>  ∆ (11)
where εsh is the unrestrained shrinkage strain, ∆H is the moisture gradient and αsh(H) hygral coefficient
of shrinkage which can be measured from drying tests at different relative humidities. Hygral coefficients
are taken directly from the experimental measurements of Martinola and Wittmann (Martinola and
Wittmann 1995). Therefore, for repair material αsh,RM=0.0048, for interface between repair material and
substrate αsh,INT=0.0028 and for substrate αsh, SUB=0.0013. As well as diffusivity parameters (Table 1), for
different mixtures of repair material and concrete substrate, hygral coefficients will also change and
should be measured experimentally.

If material is completely free to deform, shrinkage deformation due to hygral gradients results only in
volume change of material and no stress occurs. However, internal restrains in material (i.e. aggregates,
stiff inclusions) and external restrains (i.e. bond with substrate) result in eigenstresses and local stress
concentrations which might exceed material strength. Once material strength is exceeded, cracking in the
material occurs. This behaviour is mimicked by the lattice mesh. Due to hygral gradients, every element
tends to change its length (Figure 6). However, the deformation is restrained by element’s connectivity
with other elements in the mesh and this results in generation and building up of tensile stresses. Each of
the elements, according to its cross section A, its local E modulus, and the resulting hygral gradient is
loaded axially with following force:
@*>  =*> A B (12)
Once the element breaks, it is removed from the mesh, stresses in that element are released and other
elements connected to the broken element are less restrained. During the further analysis, new restrain
levels and moisture gradients in elements lead to stress generation and redistribution among the surviving
beams until the strength of the next element is reached. This is analogue to material shrinkage-once the
material cracks the magnitude of restraints in the material is reduced and the stresses are partially
released. Lower restraint levels lead to higher deformability of the system and therefore lower stresses.

Figure 6: Coupling lattice moisture and lattice fracture model in time (H-relative humidity)

Mechanical properties of the elements in the lattice mesh


In fracture simulations, different mechanical properties are ascribed to the elements that characterize
certain phase (Table 2). Three different mixtures of repair material are simulated. First, repair material is
simulated as a repair mortar, with homogenous local properties and without coarse aggregates. Then,
fibres that enable crack bridging properties and strain hardening behaviour are implemented in repair
mortar (SHCC). Finally, aggregates acting as a stiff, non-shrinking inclusions are added in the repair
material mixture (repair concrete) and simulated in the concrete repair system. Volume fraction of the
simulated crushed aggregate in the repair material is 30%. Due to the limitation in thickness and
dimensions of patch repairs in practice, all the aggregates in repair material are somewhat smaller than
aggregates in substrate and are in the sieve range [4,8] mm.
Values for the mechanical properties of interface elements which connect repair material and mortar
substrate are assumed values. As the interface is usually the weakest zone in the system, lower properties
are ascribed to elements which characterize this zone. Influence of interface properties of a repair system
in a direct tension test and drying shrinkage test is further studied.
Table 2 :Input values for lattice fracture model
Element E [GPa] ft [MPa] fc [MPa]
Matrix (repair mortar-RM) 20 3.5 -10 ft
Fibre [14] 40 7380 - ft
Interface (Matrix/Fibre) 20 90 - 10 ft
Mortar substrate (MS) 20 3.5 -10 ft
Mortar substrate (MS with meso- 25 4 -10 ft
structure)
5
Aggregate 70 8 -10 ft
ITZ 15 2.5 -10 ft
3.5
Interface (MS/RM) 15 1 -10 ft
3

Bond performance of the repair system in direct tension test


Direct tension test is performed to characterize the influence of interface properties on bond behaviour
and mechanical performance of the repair system. Surface roughness is considered to have the greatest
influence on the bond strength (Denarié, Silfwerbrand et al. 2011). Sandblasted surface is usually advised
for surface pre-treatment of the substrate (Julio, Branco et al. 2004, Bissonnette, Courard et al. 2011).
Surface laitance is removed without providing a high degree of roughness and aggregates are exposed of
less than a half of the diameter of the maximum size aggregate (Garbacz, Górka et al.). In following
simulations, sandblasted (rough) surface of the concrete substrate is imitated.

Load displacement curves for smooth and rough surface with low interface properties (1 MPa) and
smooth surface with high interface properties (3 MPa) are shown in figure 7a. The final fracture patterns
are presented in Figure 7b.

b1 2 3
Figure 7: a) Stress-strain diagrams of direct tension test applied to repair system and b) final damage
patterns of: 1) rough surface and 2) smooth surface with low interface properties and 3) smooth surface
with high interface properties

In both smooth and rough surface with low properties (1MPa) specimens fail at the interface (Figure 7b-1
and Figure 7b-2). When high strength properties are ascribed to interface elements, specimen fails in the
substrate (Figure 7b-3). As the weakest link in the repair system is the original concrete, the bond strength
between two materials is higher than the measured strength of the composite system.

For the low interface properties, increased roughness does not affect substantially the bond strength itself,
but it enables more stable fracture and more ductile performance under uniaxial tension test. Larger
fracture energy is needed for failure, and failure itself is not as brittle as in case of smooth surface. As
failure propagates also around the aggregate, there is more interlocking and crack bridging, which enables
that more energy is spent during crack formation. This leads to more ductile behaviour of the bond and
more stable fracture propagation. These results are in line with previous results about the influence of
substrate roughness on different bond strength tests (Lukovic, Schlangen et al. 2013).

DRYING SHRINKAGE DAMAGE IN AN OVERLAY SYSTEM AND DISSCUSION


Performance of the repair system exposed to environmental drying conditions is further simulated.
Typical sequence of damage development in a repair system with repair mortar is shown in figure 8.
Smooth surface with low interface properties (1MPa) is simulated. Due to the moisture gradient and
drying from the top surface, first crack develops at the surface of the repair material (Figure 8a). Soon
after forming this initial crack, another crack develops (Figure 8b). Homogeneity and brittleness of a
repair mortar enables that first crack propagate straight and fast until it reaches the interface. Depending
on the interface properties, it either continues into substrate and/or kinks in the interface leading to
debonding. Since the interface properties are low and smooth surface does not enable crack faces friction
or crack interlocking to stop the crack propagation, fast debonding and wide crack opening at the very
early stage in the repair system are observed (Figure 8c and 8d).

Figure 8: Damage development in a repair system with low bond strength (interface properties 1 MPa)
and 40 mm thickness of repair material after 2, 3, 17 and 110 days (Deformation scaled 300x)

In order to improve this performance, influence of different parameters for the time dependent damage
development due to drying shrinkage is further studied.

Influence of surface roughness


Moisture profiles at 110 days, for both smooth and rough surface are presented in the Figure 9. Top 20
mm of the concrete substrate was pre-saturated. The same input parameters and initial conditions are used
as in model verification (Table 1). Surface roughness does not have significant influence on the resulting
moisture profile (Figure 9b). Rough profile has slightly more uneven drying front compared to smooth
surface (Figure 9a) which is probably caused by geometry and more exposed aggregates on the surface of
the substrate. Locally more porous ITZ and higher conductivity properties in the surrounding of
aggregates disturbs the uniform drying front and causes some local moisture gradients.

a b c
Figure 9: Moisture distribution at the 110 days when in the system with a) smooth surface, 40mm repair
thickness b) rough surface, 40mm repair thickness c) smooth surface, 20mm repair thickness

If final crack pattern is compared for smooth (Figure 10a) and rough (“sandblasted”) surface (Figure 10b)
with the same interface properties (1 MPa), there is a substantial difference in the performance of the
system. In Figure 10 the final damage pattern and crack widths at the age of 110 days are presented.
Cracks wider than 75 microns are represented in red colour. Rough surface enables high mechanical
interlocking and therefore more constraint at the contact zone compared to the smooth surface. This
results in more cracking in the repair material, smaller length of debonding, but also more microcracking
in the concrete substrate as the crack will propagate directly from the repair material. Similar is observed
when higher interface properties (3 MPa) are ascribed to the smooth surface (Figure 10c). There is less
debonding, but more cracking in the repair material due to higher restrain levels at the interface. Rough
surface with lower bond strength enables similar performance as smooth surface with high bond strength.
This explains the significant role of the amount of energy that is lost during crack propagation. Although
having lower peak bond strength (Figure 7), rough surface enables high fracture energy and more area
which will spend energy during crack formation. Significance of post peak behaviour and achieved
ductility, over peak strength, is especially important for imposed deformation type of loading such as in
case of restrained shrinkage. Repair systems are not crack free, and therefore, even more important that
the peak bond strength itself is the way that fracture propagates and the amount of energy that is lost
during its propagation. Therefore, depending on the repair system application, small amount of debonding
may be beneficial, as long as there is something that will either stop this crack or deviate it in other
direction (inside the repair material or concrete substrate). More heterogeneity in the interface provided
by the rough surface enables that the specific surface area of the crack is increased. Consequently, more
energy is spent on crack formation and fracture propagation in the system is less brittle.

a b c
Figure 10: Fracture pattern and crack width in a repair system after 110 days, 40 mm thickness of repair
material with a) smooth surface (interface properties 1 MPa) b) rough surface (interface properties 1
MPa) and c) smooth surface (interface properties 3 MPa) (Deformation scaled 100x)

Maximum crack width, delamination height (debonding) and crack width in the substrate for different
types of surface preparation are presented in Figure 15. Improved bonding properties are beneficial for
reducing debonding but lead to higher crack widths inside the concrete substrate. Once the crack reaches
the interface, it does not branch on debonding but propagates straight to the concrete substrate. Therefore,
care should be taken when original concrete is reinforced as the larger crack widths will enable fast
propagation route for chloride penetration and corrosion of the reinforcing steel.

Type of repair material


For repair applications, usually repair mortars (without coarse aggregates) are designed. This is
convenient for good workability properties, low viscosity, applications of repair mortars in thin layers or
small patch repairs. Fast setting time of the repair material, which is preferred in order to have minimal
interruption of operations during repair work, leads to low w/c ratio and application of accelerating
admixtures which further increase the brittleness of the system. Without coarse aggregates, the mixture is
almost homogenous at the mesoscale, with high amount of cement and high global shrinkage
deformations. Without heterogeneities (aggregates or other inclusions), as soon as the crack is initiated, it
will tend to propagate extremely fast, in more or less straight manner until is stopped by other pre-
existing crack or defect (Colina and Roux 2000). Due to the inherent brittleness of the cement based
material and continued drying, this crack will further open and reach large crack widths. Heterogeneities,
however, will favour the nucleation of stable microcracks and more “controlled” development of
shrinkage cracking (Colina and Roux 2000). Therefore, the influence and benefits of including coarse
aggregates in the repair material are also simulated. Volume fraction of the simulated crushed aggregate
in the repair material is 30%. All the aggregates are in the sieve range [4,8] mm.

In the repair mortar, once the crack is opened it propagates straight until it reaches the interface (figure 8).
No aggregates are present to stop this single crack or to enable branching from the straight path. Due to
the brittleness of homogenous material, as drying continues, this crack widens and becomes susceptible to
fast chloride and water penetration. With aggregates, locally porous ITZ zone enables faster moisture loss
from the repair material and uneven moisture profile. Further on, crack propagate around the aggregates
and enables more tortuous cracks paths for the aggressive substance ingress. As it enables that more
cracking occurs in the ITZ zone, locally lower restrain level exist. Also, by including the non-shrinking
aggregate, the global shrinkage is reduced and the final crack width is smaller than in case with
homogenous repair material (repair mortar). More crack trapping and crack branching, caused by
aggregates acting as a stiff inclusions, is observed. This is in agreement with the statement that
heterogeneities favour the nucleation of stable micro cracks and give rise to a controlled development of
cracking compared to fast growing, straight cracks in brittle homogenous materials (Colina and Roux
2000).

Microcracking in the ITZ zone appears before the main crack initiates (Figure 11a). Cracks that propagate
through the matrix bridge several aggregate particles (Figure 11b-d), similar as observed by Wong
(Wong, Zobel et al. 2009). With concrete as a repair material, final crack widths are lower compared to
crack widths when repair mortar is used (Figure 15). Observations from the simulations and obtained
crack patterns are in line with the literature data. Influence of size and amount of aggregates on the
shrinkage induced cracking was studied both experimentally (Wong, Zobel et al. 2009) and numerically
(Grassl, Wong et al. 2010). It was observed that crack increased with decreasing aggregate volume
fraction. Therefore, when less aggregates are present, larger crack widths in the system are measured.

Figure 11: Damage development of a 40 mm thick repair concrete in a repair system with low bond
strength (interface properties 1 MPa) after 1, 3, 14 and 110 days (Deformation scaled 300x)

Although more controlled shrinkage cracking with lower crack widths is achieved (Figure 15), aggregates
do not prevent the main crack from further widening. Aggregates enable more tortuosity of cracking path,
more distributed microcracking and lower global shrinkage. As higher energy is consumed during
cracking, some of the cracks are locked by stiff aggregate inclusions. However, concrete is still a quasi-
brittle material, which means that once a large crack opens it will grow further until crack faces
completely separate. There is nothing except crack faces friction that will transfer stresses and prevent
this crack from further widening. Aggregates therefore, enable more stable crack propagation and reduce
global shrinkage, but do not limit the final crack widths.

Crack widths in the repair system can be limited and controlled if a repair material is reinforced by
discrete fibres, such as in the case of strain-hardening cementitious composite-SHCC (Li, Horii et al.
2000, Li 2009). Fine PVA fibres introduced in this material ensure that once the crack is open, fibres
bridging that crack take the stresses and prevent the crack from further widening. As a result, new crack
at a different location will open. Consequently, SHCC will result in high ductility which makes it very
suitable for application in concrete repairs.

When PVA fibres are added, high shrinkage strains of repair mortar are still measured (Li and Li 2006,
Zhou 2010). PVA fibres are not stiff enough nor there are coarse aggregates that will restrain and reduce
shrinkage deformation. In addition, once the microcracking starts, it propagates straight through the
material, similar as in non-reinforced repair mortar. However, once the fibres are activated, they arrest
this crack and prevent it from further widening.

Figure 12: Damage development of a 40mm thick SHCC in a repair system with low bond strength, rough
surface (interface properties 1 MPa) after 1, 10, 37 and 110 days (Deformation scaled 300x)

Fracture propagation if fibre reinforced repair material is exposed to drying shrinkage is presented in
Figure 12. Instead of one crack with large maximum crack widths (~140 microns) in repair mortar (Figure
13b) or repair concrete (~55 microns, Figure 13a), 4 cracks with small crack widths (~35-40 microns) in
SHCC are achieved (Figure 13c). Rough surface in SHCC repair system further enables more cracking
with smaller crack widths (5 cracks, max crack width ~30microns) compared to smooth surface (4 cracks,
max crack with ~38 microns). As previously addressed, rough surface enables more interlocking and
higher constraints level at the interface and therefore higher probability of cracking inside the repair
material. Similar was observed in drying shrinkage test in SHCC/concrete and concrete/concrete repair
systems (Li and Li 2006). Rougher surface profiles in SHCC/concrete repair system provided more
cracking with smaller crack widths and less delamination height between two layers. It was concluded
that when material ductility satisfied, free drying shrinkage of repair material becomes less important for
performance of the system. Therefore, although SHCC has significantly higher drying shrinkage
compared to concrete, measured crack widths in SHCC/concrete repair system were in range of 10-60
microns while in concrete/concrete repair system they were 120-360microns. In addition, provided that
roughness profile and bond strength are high enough in SHCC/concrete repair system, delamination
length and height are also comparable to concrete/concrete repair system.

a b c
Figure 13: Fracture pattern and crack width in a repair system with smooth surface and 40 mm thickness
of repair material, low bond strength and a) with aggregates b) without aggregates-homogenous
mesostructure c) with fibres (SHCC) (deformation scaled 100x)

In structural, three point bending tests and reflective cracking tests, it was shown that smooth surface
enables more debonding and lower amount of restraint (Kamada and Li 2000, Lukovic, Schlangen et al.
2013). If SHCC is used as a repair material, this might be beneficial as debonding results in more cracks
in the repair material, higher energy dissipation and more ductility of the repair system. However, in these
simulations and laboratory experiments, drying shrinkage of repair material was not taken into account.
Therefore, care should be taken, as smooth surface might be beneficial only in certain applications when
pre-damage through the optimal curing is limited.
Repair material thickness
Another important factor is the thickness of the repair material. It is expected that in the samples with
smaller thickness there is a higher probability for cracking compared to the thicker overlays (Zhou, Ye et
al. 2008). This is also dependant on the bond strength of the repair material as this determines the amount
of constraint at the interface. Therefore, different thickness of the repair material with different bond
strengths are simulated. Again, top surface of the substrate (20 mm) was saturated. In Figure 9, drying
profiles at the age of 110 days for two samples with repair thickness of 20mm and 40mm are compared. If
the thickness of the overlay is reduced, faster drying occurs. In both specimens, concrete substrate with
smooth surface is simulated. Again, the benefits of using SHCC as a repair material on fracture
propagation in the repair system are investigated.

a b c

d e f
Figure 14: Fracture pattern and crack widths in a smooth surface repair system with interface properties
of: a) 1 MPa, 20mm repair mortar b) 3 MPa, 20 mm repair mortar c) 3 MPa and 40 mm repair mortar, d)
1 MPa and 20 mm SHCC, e) 3 MPa and 20 mm SHCC and f) 3MPa and 40 mm SHCC

When high bond strength between the concrete substrate and repair material is simulated, there is higher
constrain level at the interface and therefore, more cracks develop (Figure 14b). High constrain levels and
moisture gradients results in radial cracks localized between two main cracks. When the interface has
lower properties, due to the debonding, main crack opens very wide, and the system completely
delaminates (Figure 14a). It is also interesting to notice that for the high bond strength, in a system with
20 mm repair material thickness there are two large cracks (Figure 14b) while in the case of 40 mm repair
material (Figure 14c), only one crack is opened. This is in accordance with the results obtained by Colina
(Colina and Roux 2000). In their study, in the system consisting of a layer of a paste made of clay, sand
and water, deposited on a rigid substrate and exposed to drying, final crack spacing seemed to be simply
proportional to the sample thickness. This means that the mean crack spacing is primarily controlled by
the thickness of the drying sample. The same was also observed in (Groisman and Kaplan 1994). In both
of these experiments bottom friction with the substrate critically influence the final crack pattern. When
compared to repair system performance, there is an analogy between friction with the substrate and the
bond strength within the repair system.
The importance of good bond when SHCC used as a repair material is also observed from damage
patterns in Figure 14d and Figure 14e. With low bond strength, although fibres are used (Figure 14d),
system delaminates and only small number of cracks open. These cracks reach higher crack widths
(~58microns) and it seems that there are no substantial benefits from fibre reinforcement (Figure 15).
However, in specimen with high bond strength, high constraint levels leads to more cracks with limited
crack widths (8 cracks, ~28microns, Figure 14d). When 40 mm SHCC thickness simulated, final fracture
pattern for smooth surface with high bond strength (Figure 14f) resembles the failure pattern in rough
surface with lower bond strength (Figure 12). This is in accordance with results obtained for non-
reinforced repair mortar.

Comparison of the maximum crack width in the repair, substrate and maximum debonding for all
simulated specimens is shown in the Figure 15 and Figure 16. With increase of the interface properties in
smooth surface, debonding is reduced, but the crack width in the substrate is enlarged. Roughness profile,
which gives lower peak bond strength but high fracture energy (Figure 7), results in the same damage
pattern as the smooth surface with high bond strength. This gives indication that bond strength is not the
main parameter when the repair system is exposed to imposed deformations. More meaningful for the
damage development and the final crack widths becomes the fracture energy that is lost during crack
propagation.

Once the crack is opened, interface properties will determine how wide this crack will further open. If the
interface is weak, warping of the overlay results in large surface cracks (Figure 16).

Figure 15: Maximum crack widths in the repair material, interface zone and concrete substrate for 40 mm
repair thickness with different types of surface preparation, repair material and bond strength

Figure 16: Maximum crack widths in the repair material, interface zone and concrete substrate for 20 mm
repair thickness with different types of repair material and bond strength
With the aggregates in the repair material mixture, global shrinkage is reduced, more stable and tortuous
cracking is achieved, and due to the distributed microcracking in the ITZ lower restrain levels exist in the
repair mortar. This leads to lower crack widths and more distributed microcracking in the repair system.
Although providing lower shrinkage, once the cracks are opened, aggregates do not prevent these cracks
from further widening. This can be achieved by implementation of fibres. More cracks with limited crack
widths are achieved. However, in order to exhaust full capacity of SHCC, bond strength should be high
enough in order to prevent complete interface delamination. If the repair material is delaminated, there is
no benefit from the fibres.

CONCLUSIONS
In the present work, a lattice based model is proposed to simulate moisture transport and drying shrinkage
induced cracking in the repair systems. Based on above discussion, following conclusions can be drawn:

• In order to enable good prediction of the performance of the repair system, it seems that material
mesostructure needs to be included. Existence of aggregates and ITZ as locally more porous and
weak area influence both the moisture distribution and the damage development in the system.
• For the behaviour of the system under imposed deformation, the way that fracture develops
(whether system exhibits brittle or ductile behaviour) and energy that is lost on crack formation
might have as high influence as peak strength values. From the modelling study, higher roughness
profiles with even lower bond strength enable more stable crack propagation and less brittle
behaviour of the interface.
• When SHCC is used as a repair material, although in flexural and reflective cracking laboratory
tests higher fracture energy is measured with smooth surface substrate, in practical applications
smooth surface is not advised. Due to the moisture gradients system exhibits pre-damage. Drying
shrinkage might cause large and uncontrolled debonding and once the system is debonded, there
are no benefits from fibres anymore. Therefore, in order to exhaust full capacity of SHCC under
restrained deformation, it seems that adequate bonding should be achieved.
• With adequate bonding, SHCC performed best in the simulated drying shrinkage tests. It shows
small crack widths which are beneficial for durability properties of repair systems. Repair material
should have strain-hardening behaviour in order to enable multiple crack formation, stable
propagation and to prevent wide opening of the crack.
• Depending on the repair application, use of coarse aggregates for the repair material is advised.
Aggregates reduce total shrinkage and enable more stable fracture propagation during the drying
and less brittle behaviour of the system. From the model, smaller crack widths are obtained
compared to the more homogenous materials at the same exposure conditions.
• With the same bond strength and the same drying conditions, more cracks (with lower crack
spacing) are observed in thinner overlays. This is in accordance to experimental observations in
drying shrinkage tests of different brittle and quasi brittle materials.

Although 3D simulation was done, due to the limited size of specimen thickness (5mm), periodic
boundary conditions were applied only in one direction. In real applications, however, material is
restrained in two directions and this will probably lead to faster fracture propagation. This will be part of
the future study.

In the model, edge boundary conditions were not included. However, it was considered that, once the
crack is opened, boundary conditions around this crack represent edge effect in an overlay system.

Model result shows that general performance of the repair system due to drying shrinkage can be well
imitated. However, a number of assumptions for the input parameters were made. Diffusivity parameters
and hygral coefficients that are used as input are different for every repair material and concrete substrate.
Therefore, for every mixture they should be independently measured as they will influence the state of
stresses in the repair system. Due to difficulties in experimental quantification of the diffusivity and
fracture properties of interfaces (interface RM/CS, ITZ), values used in the model are also only
reasonable assumptions.

The influence of time dependent phenomena, such as development of material properties in repair system
is not considered but is important for estimating the performance of the repair system. Due to relaxation
properties of repair material, the amount of stresses due to internal and external restrains will be reduced,
and therefore, crack initiation will be postponed. As in the case of repair system, relaxation has beneficial
influence, neglecting the influence of this parameter is a conservative approach to determining the
performance of the repair system. However, in order to enable more thorough estimation and
quantification of the performance of the repair system, these aspects should also be considered.

ACKNOWLEDGEMENTS
Financial support by the Dutch Technology Foundation (STW) for the project 10981-“Durable Repair and
Radical Protection of Concrete Structures in View of Sustainable Construction” is gratefully
acknowledged.

REFERENCE
Asad, M., M. Baluch and A. AI-Gadhib (1997). "Drying shrinkage stresses in concrete patch repair systems."
Magazine of Concrete Research 49(181): 283-293.
Ayano, T. and F. Wittmann (2002). "Drying, moisture distribution, and shrinkage of cement-based materials."
Materials and Structures 35(3): 134-140.
Beushausen, H. and M. Alexander (2007). "Localised strain and stress in bonded concrete overlays subjected to
differential shrinkage." Materials and structures 40(2): 189-199.
Birkeland, H. W. (1960). Differential shrinkage in composite beams. ACI Journal Proceedings, ACI.
Bisschop, J. and J. G. van Mier (2008). "Effect of aggregates and microcracks on the drying rate of cementitious
composites." Cement and Concrete Research 38(10): 1190-1196.
Bissonnette, B., L. Courard, D. Fowler and J. L. Granju (2011). Bonded Cement-based Material Overlays for the
Repair, the Lining Or the Strengthening of Slabs Or Pavements: State-of-the-art Report of the RILEM Technical
Committee 193-RLS, Springer Verlag.
Bolander, J. E. and S. Berton (2004). "Simulation of shrinkage induced cracking in cement composite overlays."
Cement and Concrete Composites 26(7): 861-871.
Colina, H. and S. Roux (2000). "Experimental model of cracking induced by drying shrinkage." The European
Physical Journal E 1(2-3): 189-194.
Denarié, E., J. Granju and J. Silfwerbrand (2004). Structural behaviour of bonded concrete overlays. International
RILEM TC 193-RLS Workshop on Bonded Concrete Overlays, RILEM Publications SARL.
Denarié, E., J. Silfwerbrand and H. Beushausen (2011). Bonded Cement-Based Material Overlays for the Repair,
the Lining or the Strengthening of Slabs or Pavements.
Emberson, N. and G. Mays "Significance of property mismatch in the patch repair of structural concrete Part 1:
Properties of repair systems." Magazine of Concrete Research 42(152): 147-160.
Femmasse, B. (1996). Physical and mechanical Models implemented in the modules HEAT–MLS–MES/2.5 D of
Femmasse.
Garbacz, A., M. Górka and L. Courard "Effect of concrete surface treatment on adhesion in repair systems."
Magazine of Concrete Research 57(1): 49-60.
Garboczi, E. J. (2002). "Three-dimensional mathematical analysis of particle shape using X-ray tomography and
spherical harmonics: Application to aggregates used in concrete." Cement and Concrete Research 32(10): 1621-
1638.
Granger, L., J.-M. Torrenti and P. Acker (1997). "Thoughts about drying shrinkage: experimental results and
quantification of structural drying creep." Materials and structures 30(10): 588-598.
Grassl, P., H. S. Wong and N. R. Buenfeld (2010). "Influence of aggregate size and volume fraction on shrinkage
induced micro-cracking of concrete and mortar." Cement and concrete research 40(1): 85-93.
Groisman, A. and E. Kaplan (1994). "An experimental study of cracking induced by desiccation." EPL
(Europhysics Letters) 25(6): 415.
Habel, K., E. Denarié and E. Brühwiler (2006). "Time dependent behavior of elements combining ultra-high
performance fiber reinforced concretes (UHPFRC) and reinforced concrete." Materials and structures 39(5): 557-
569.
Julio, E. N., F. A. Branco and V. t. D. Silva (2004). "Concrete-to-concrete bond strength. Influence of the
roughness of the substrate surface." Construction and Building Materials 18(9): 675-681.
Kamada, T. and V. C. Li (2000). "The effects of surface preparation on the fracture behavior of ECC/concrete
repair system." Cement and Concrete Composites 22(6): 423-431.
Landis, E. and J. Bolander (2009). "Explicit representation of physical processes in concrete fracture." Journal of
Physics D: Applied Physics 42(21): 214002.
Lewis, R., P. Nithiarasu and K. Seetharamu Fundamentals of the finite element method for heat and fluid flow.
2004, John Wiley & Sons.
Li, M. (2009). Multi-scale design for durable repair of concrete structures, PhD Thesis, The University of
Michigan.
Li, M. and V. C. Li (2006). "Behavior of ECC/concrete layer repair system under drying shrinkage conditions."
Restoration of Buildings and Monument 12(2): 143-160.
Li, V. C., H. Horii, P. Kabele, T. Kanda and Y. M. Lim (2000). "Repair and retrofit with engineered cementitious
composites." Engineering Fracture Mechanics 65(2-3): 317-334.
Li, Z., M. A. Perez Lara and J. Bolander (2006). "Restraining effects of fibers during non-uniform drying of cement
composites." Cement and concrete research 36(9): 1643-1652.
Lukovic, M., E. Schlangen, G. Ye and B. Šavija (2013). Impact of surface roughness on the debonding mechanism
in concrete repairs. Proceedings of the 8th International Conference on Fracture Mechanics of Concrete and
Concrete Structures (FraMCoS-8), Toledo, Spain.
Martinola, G., H. Sadouki and F. Wittmann (2001). "Numerical Model for Minimizing the Risk of Damage in a
Repair System." Journal of materials in civil engineering 13(2): 121-129.
Martinola, G. and F. Wittmann (1995). "Application of fracture mechanics to optimize repair mortar systems."
Fracture Mechanics of Concrete Structures’, Proceedings of FRAMCOS-2, FH Wittmann Ed., Aedificatio: 1481-
1492.
Nakamura, H., W. Srisoros, R. Yashiro and M. Kunieda (2006). "Time-dependent structural analysis considering
mass transfer to evaluate deterioration process of RC structures." Journal of Advanced Concrete Technology 4(1):
147-158.
Qian, Z. (2012). Multiscale Modeling of Fractrue Processes in Cementitious Material, PhD thesis, PhD thesis, Delft
University of Technology.
Quenard, D. and H. Sallee (1992). "Water vapour adsorption and transfer in cement-based materials: a network
simulation." Materials and Structures 25(9): 515-522.
Sadouki, H. and J. Van Mier (1997). "Meso-level analysis of moisture flow in cement composites using a lattice-
type approach." Materials and Structures 30(10): 579-587.
Sadouki, H. and J. Van Mier (1997). "Simulation of hygral crack growth in concrete repair systems." Materials and
structures 30(9): 518-526.
Šavija, B., J. Pacheco and E. Schlangen (2013). "Lattice modeling of chloride diffusion in sound and cracked
concrete." Cement and Concrete Composites 42(Complete): 30-40.
Schlangen, E. (1993). Experimental and numerical analysis of fracture processes in concrete
PhD thesis, Delft University of Technology.
Schlangen, E. (2008). "Crack development in concrete, Part 1: Fracture experiments and CT-scan observations."
Key Engineering Materials 385: 69-72.
Schlangen, E. and Z. Qian (2009). "3D modeling of fracture in cement-based materials." Journal of Multiscale
Modelling 1(2): 245-261.
Silfwerbrand, J. (1997). "Stresses and strains in composite concrete beams subjected to differential shrinkage." ACI
Structural Journal 94(4).
Tilly, G. P. and J. Jacobs (2007). Concrete repairs: Observations on performance in service and current practice,
CONREPNET Project Report, IHS BRE Press, Watford, UK.
Wong, H., M. Zobel, N. Buenfeld and R. Zimmerman (2009). "Influence of the interfacial transition zone and
microcracking on the diffusivity, permeability and sorptivity of cement-based materials after drying." Magazine of
Concrete Research 61(8): 571-589.
Zhou, J. (2010). Performance of engineered cementitious composites for concrete repairs, PhD Thesis, Delft
University of Technology.
Zhou, J., G. Ye, E. Schlangen and K. Van Breugel (2008). "Modelling of stresses and strains in bonded concrete
overlays subjected to differential volume changes." Theoretical and Applied Fracture Mechanics 49(2): 199-205.

View publication stats

You might also like