Sie sind auf Seite 1von 58

Elsevier Editorial System(tm) for Engineering Structures

Manuscript Draft
Manuscript Number:
Title: Application of air cooled pipes for reduction of early age cracking risk in a massive RC wall
Article Type: Research Paper
Keywords: Cement hydration; service life conditions; differential shrinkage; cracking; numerical
simulation
Corresponding Author: Dr. Miguel Azenha, PhD
Corresponding Author's Institution: University of Minho
First Author: Miguel Azenha, PhD
Order of Authors: Miguel Azenha, PhD; Rodrigo Lameiras, MSc; Christoph de Sousa, MSc; Joaquim
Barros, PhD
Abstract: The construction of massive concrete structures is often conditioned by the necessity of
phasing casting operations in order to avoid excessive heat accumulation due to cement hydration. To
accelerate construction and allow larger casting stages (usually increasing lift height), it is usual to
adopt internal cooling strategies based on embedding water pipes into concrete, through which water
is circulated to minimize temperature development. The present paper reports the use of horizontally
placed ventilated prestressing ducts embedded in a massive concrete wall for the same purpose, in line
with a preliminary Swedish proposal made in the 1990's. The application herein reported is a holistic
approach to the problem under study, encompassing extensive laboratory characterization of the
materials (including a technique developed for continuous monitoring of concrete E-modulus since
casting), in-situ monitoring of temperatures and strains, and 3D thermo-mechanical simulation using
the finite element method. Based on the monitored/simulated results, it is concluded that the aircooling system is feasible and can effectively reduce early cracking risk of concrete, provided adequate
planning measures are taken.
Suggested Reviewers: Stephanie Staquet PhD
Associate Professor, Universit Libre de Bruxelles
sstaquet@ulb.ac.be
Stphanie Staquet is an expert in material/structural behaviour of concrete since early ages.
Eduardo Fairbairn PhD
Associate Professor, Federal University of Rio de Janeiro
eduardo@coc.ufrj.br
Eduardo Fairbairn is a known expert in the simulation of large concrete structures under the effect of
heat of hydration at early ages.

Cover Letter
Click here to download Cover Letter: cover letter.doc

Campus de Azurm
4800-058 Guimares

ESCOLA DE ENGENHARIA
Departamento de Engenharia Civil

Miguel Azenha
School of Engineering University of Minho
Civil Engineering Department
Azurm Campus, 4800-058 Guimares
PORTUGAL

May 1, 2013

Engineering Structures
Elsevier B.V.

To whom it may concern:


Please find enclosed (electronic submission) the manuscript Application of air cooled pipes
for reduction of early age cracking risk in a massive RC wall by Miguel Azenha, Rodrigo
Lameiras, Christoph de Sousa and Jos Granja, that we intend to submit to Engineering
Structures.
Should you need to contact me, please use the above address or call me at phone number
(+351)-93 840 4554. You may also contact me by fax at number (+351)-253.510.217 or via
email at miguel.azenha@civil.uminho.pt.
Sincerely,

Miguel Azenha (the corresponding author)


(Assistant Professor, PhD)

*Highlights (for review)


Click here to download Highlights (for review): Highlights.docx

Highlights

Use of horizontally placed ventilated prestressing ducts embedded in a massive concrete


wall to avoid excessive heat accumulation due to cement hydration;

Extensive laboratory characterization of the materials;

In-situ monitoring of temperatures and strains;

3D thermo-mechanical simulation using the finite element method

Good coherence between field monitoring and numerical simulation both in terms of
temperatures and strains;

Air cooling system proved to be effective, reducing cracking risk.

*Abstract
Click here to download Abstract: Abstract.docx

Abstract
The construction of massive concrete structures is often conditioned by the necessity of phasing
casting operations in order to avoid excessive heat accumulation due to cement hydration. To
accelerate construction and allow larger casting stages (usually increasing lift height), it is usual
to adopt internal cooling strategies based on embedding water pipes into concrete, through which
water is circulated to minimize temperature development. The present paper reports the use of
horizontally placed ventilated prestressing ducts embedded in a massive concrete wall for the
same purpose, in line with a preliminary Swedish proposal made in the 1990s. The application
herein reported is a holistic approach to the problem under study, encompassing extensive
laboratory characterization of the materials (including a technique developed for continuous
monitoring of concrete E-modulus since casting), in-situ monitoring of temperatures and strains,
and 3D thermo-mechanical simulation using the finite element method. Based on the
monitored/simulated results, it is concluded that the air-cooling system is feasible and can
effectively reduce early cracking risk of concrete, provided adequate planning measures are
taken.

*Manuscript
Click here to download Manuscript: Manuscript.docx

Click here to view linked References

Application of air cooled pipes for reduction of early age cracking risk in
a massive RC wall
By Miguel Azenha1, Rodrigo Lameiras2, Christoph de Sousa3 and Joaquim Barros4
Abstract: The construction of massive concrete structures is often conditioned by the necessity
of phasing casting operations in order to avoid excessive heat accumulation due to cement
hydration. To accelerate construction and allow larger casting stages (usually increasing lift
height), it is usual to adopt internal cooling strategies based on embedding water pipes into
concrete, through which water is circulated to minimize temperature development. The present
paper reports the use of horizontally placed ventilated prestressing ducts embedded in a massive
concrete wall for the same purpose, in line with a preliminary Swedish proposal made in the
1990s. The application herein reported is a holistic approach to the problem under study,
encompassing extensive laboratory characterization of the materials (including a technique
developed for continuous monitoring of concrete E-modulus since casting), in-situ monitoring of
temperatures and strains, and 3D thermo-mechanical simulation using the finite element method.
Based on the monitored/simulated results, it is concluded that the air-cooling system is feasible
and can effectively reduce early cracking risk of concrete, provided adequate planning measures
are taken.
Keywords: Cement hydration, service life conditions, differential shrinkage, cracking, numerical
simulation
1

INTRODUCTION
The combined effect of the exothermic nature of cement hydration reactions and the

relatively low thermal diffusivity of concrete leads concrete structures to endure temperature
1

Assistant Professor, ISISE Institute for Sustainability and Innovation in Structural Engineering, University of
Minho, School of Engineering, Civil Engineering Dept., Azurm Campus, 4800-058 Guimares, Portugal, Phone:
(+351)938404554, Fax: (+351) 253 510 217, E-mail: miguel.azenha@civil.uminho.pt.
2
PhD Student, ISISE Institute for Sustainability and Innovation in Structural Engineering, University of Minho,
School of Engineering, Civil Engineering Dept., Azurm Campus, 4800-058 Guimares, Portugal, Phone: (+351)
938928308, Fax: (+351) 253 510 217, E-mail: rmlameiras@civil.uminho.pt.
3
PhD Student, ISISE Institute for Sustainability and Innovation in Structural Engineering, University of Minho,
School of Engineering, Civil Engineering Dept., Azurm Campus, 4800-058 Guimares, Portugal, Phone: (+351)
938928308, Fax: (+351) 253 510 217, E-mail: christoph@civil.uminho.pt.
4
Full Professor, ISISE Institute for Sustainability and Innovation in Structural Engineering, University of Minho,
School of Engineering, Civil Engineering Dept., Azurm Campus, 4800-058 Guimares, Portugal, Phone:
(+351)93.840.4554, Fax: (+351) 253 510 210, E-mail: barros@civil.uminho.pt.

-1-

increases at early ages, and eventually return to thermal equilibrium with the surrounding
environment. These early temperature variations induce volumetric changes in concrete that are
partially restrained by adjoining previously cast members, or even due to non-uniform
temperature distributions within a concrete member itself. Such restraint to deformation may
induce stresses that can be relevant enough to induce early age thermal cracking in concrete,
which is usually unacceptable in view of aesthetics, durability and even structural performance.
Contractors usually attempt to avoid this thermal cracking by adopting concrete compositions
and construction schedules that maintain temperature gradients in concrete below prescribed
limits, both along time and space [1]. It has however been recognized that such approach leads to
erroneous conclusions, as several important issues are disregarded [1], such as the degree of
restraint to deformation and the actual mechanical properties of concrete. In view of the
limitations of temperature-based criteria for crack risk assessment, it has been widely
acknowledged [2] that more realistic crack risk assessment can be made through multi-physics
approaches that encompass numerical simulation of temperatures and corresponding stresses in
concrete: thermo-mechanical analyses. The use of thermo-mechanical simulation models allows
the evaluation of alternative construction scenarios (for casting procedures, concrete mixes,
environmental conditions), and thus permits the optimization of construction without
compromising the safety in regard to thermal cracking. The numerical studies for the assessment
of the optimum construction strategy frequently involve diminishing the temperature rise in
concrete at early ages. In fact, if the thermal variation is diminished, the corresponding
volumetric changes also decrease, as well as the developed stresses. The diminishment of early
temperature rises is usually achieved by partial replacement of cement by additions as fly ash [3,
4], or by cooling water/aggregates before mixing operations [5-7], or even by introducing
internal cooling pipes in concrete with cooling fluids (usually water) [8-12]. An attempt to use
air as the cooling fluid in cooling pipes has been made in the 1990s by Hedlund and Groth [8,
9], who have shown the feasibility of such technique in thick columns. Nonetheless, no further

-2-

application of such technique was found in the literature, except for an in-situ use of air for
localized cooling reported by Ishikawa et al. [12].
Even though thermo-mechanical simulation approaches have been applied to concrete
structures for decades [13-17], they involve several complexities and problematic issues
particularly in view of the assessment of material properties and model parameters (thermal and
mechanical), which frequently demand specific laboratory characterization: heat of hydration,
thermal boundary coefficients, creep of concrete, evolution of E-modulus and tensile stress,
among others. Even though several scientific works have been done either on thermal simulation
[18], thermo-mechanical simulation [19] or monitoring the concrete behaviour at early ages [2023], some combine the numerical simulation with experimental data obtained in laboratory [2426] or with temperature monitoring for partial validation [27-32], whereas others go further and
additionally include in-situ strain monitoring for validation [33]. Nonetheless, in the scope of
internal cooling of concrete through the use of embedded pipes, no works were found to adopt
holistic approaches that simultaneously include material characterization, in-situ monitoring of
temperature/strain and thermo-mechanical simulation. The works found to focus on concrete
cooling with embedded pipes are mostly limited to thermal [34], or thermo-mechanical analyses
only [35-38], and thermal [39-41] or thermo-mechanical analyses together with partial validation
through in situ monitoring [12, 42].
The present paper pertains to a case study of the thermal stresses in the central wall of the
entrance organ of a dam spillway. Such wall is 27.5m long, with a maximum width of 2.8m and
height of 15.0m, and attention is given to the most unfavourable construction phase in which a
2.0m tall batch had to be made (total 150m3 of concrete). Due to the materials and equipment
available at the construction site, it was decided to attempt internal cooling of concrete with aircooled prestressing ducts placed longitudinally along the wall.
The present paper regards to the in-depth study of the early age performance of concrete in
the wall, encompassing laboratory thermal and mechanical characterization of concrete, as well

-3-

as in-situ monitoring of temperatures/strains and the corresponding thermo-mechanical


simulation with the finite element method.
The extensive laboratory characterization encompassed quantification of the heat of
hydration

evolution

(isothermal

and

semi-adiabatic

calorimetry),

evaluation

of

compressive/tensile strength and E-modulus through cube/cylinder testing, creep testing at


several ages and assessment of mechanical activation energy. In particular regard to E-modulus
testing, a methodology that allows continuous measurement of concrete E-modulus since casting
(EMM-ARM [43]) was applied. This is a pioneering use of this methodology for the purpose of
stress simulation on concrete since its very early ages, with important advantages in regard to
previous approaches that tend to extrapolate values of E-modulus at very early ages.
A relatively complete monitoring program has been carried out in-situ, involving the use of
20 temperature sensors and 7 vibrating wire strain gauges embedded in concrete. Particular
attention was given to the evaluation of the effectiveness of the cooling system, with
temperatures being measured at several points along the prestressing ducts, and with air velocity
measurements taken with handheld anemometers.
Bearing in mind the information gathered with the laboratory characterization and in-situ
monitoring, a thermo-mechanical simulation was carried out with recourse to a three dimensional
finite element model. Such simulation model included the explicit modelling of the cooling
ducts, as well as the phased construction of the wall. The simulation model was made with
DIANA software [44].

THERMO-MECHANICAL MODEL
The thermo-mechanical simulation approach presented here has strong similarities with that

described in a previous work [30]. Nonetheless, some particularities are distinct, namely: (i)
solar radiation is explicitly considered according to a model based on the incidence angle of the

-4-

sun beams; (ii) the effect of internal cooling pipes is taken into account (iii) the evolution of
mechanical properties is simulated according to the equivalent age concept (instead of the
degree of hydration concept). The following sub-sections pertain to a general description of the
modelling strategy with specific emphasis on topics (i) to (iii) mentioned above.
2.1

Thermal model
The calculation of temperature fields in concrete is based on the heat balance equation,

whose solution is made through the finite element method [45]:

k T Q cT

(1)

where k is the thermal conductivity, c is the volumetric specific heat and T is the temperature.
Q is the volumetric heat generation rate due to cement hydration, formulated as an Arrhenius

type law [46]:

Q A f () e

Ea
RT

(2)

where A is a rate constant, Ea is the apparent activation energy, is the degree of heat
development (ratio between the heat Q released up to time t and the total heat Qfinal released upon
completion of cement hydration), R = 8.314 Jmol-1K-1 is the Boltzmanns constant and f () is a
normalized function for heat.
Thermal boundary conditions are applied through a prescribed flux per unit area qT
formulated as [47]:

qT hcr Tb Te

(3)

where hcr is a mixed convection-radiation boundary transfer coefficient, Tb is the boundary


surface temperature and Te is the environmental temperature.
The simulation of thermal inputs associated to solar radiation in concrete structures can be
made with significant accuracy through the adoption of models that are readily used in
meteorological sciences [31, 48]. Such models can take into account the effects of the spatial

-5-

relationship between the earth and sun at a given time of the day/year and thus predict the solar
radiation that reaches a certain point on earth at sea level (i.e. low atmosphere). It is further
possible to compute the angle between the sunbeam and any arbitrarily oriented/inclined surface,
and evaluate the intake of energy throughout the day of such surface.
The calculation of the solar energy that reaches earth at sea level, qm, is based on the solar
constant, q0, which represents the total radiation energy received from the sun at a distance
corresponding to 1 Astronomical Unit. Even though q0 varies slightly throughout the year by
~7%, it is usually acceptable to consider q0=1367 Wm-2. The estimation of qm can be done
through the following empirical equation [48, 49]:

q m q0 e

Tl
0.9 9.4sin(h )

(4)

where Tl is the Linke turbidity factor that summarizes the turbidity of the atmosphere
(attenuation of the direct beam solar radiation) and h is the solar elevation that corresponds to the
angle between the direction of the sunbeam and the idealized horizon. Tl is known to usually
vary between 3 and 7, whereas h can be calculated by taking into account latitude, date and time.
Further geometrical considerations allow the calculation of the angle between an incident
sunbeam and the vector orthogonal to an arbitrarily oriented/inclined surface, termed as i (see
detailed description of models to calculate h and i in [45, 48]).
Based on the knowledge of qm and i at a given instant, and considering the absorvity of the
material of the target surface (S), it is possible to calculate the radiation energy qs that is
actually absorbed:

qS S qm cos(i)

(5)

Another particularity of the present application in regard to previous works [30] is the use of
prestressing ducts acting as cooling pipes. A formulation is thus necessary to describe the added
internal heat fluxes that are caused by the presence of an embedded cooling pipe, which can be

-6-

expressed in the following energy balance equation, to be applied throughout the length z along
the pipe [50]:

m cp

dTc
hc P Tc Tw
dz

(6)

where m is the mass flow rate of the coolant (air in this case), cp is the specific heat of the
coolant, hc is the boundary transfer coefficient between the coolant and the surrounding concrete,
P is the perimeter of the cooling pipe, Tc is the temperature in the cooling pipe and Tw is the bulk
temperature of concrete around the cooling pipe. The mass flow rate of the coolant m can be
estimated through the product of the cooling fluid density () by the fluid mean velocity (m) and
by the cross sectional area of the cooling pipe (Ac). The implementation of equation (6) into a
finite element software [44] brings further nonlinearities due to the interaction between the
cooling fluid and the surrounding concrete, which results in progressive heating of the cooling
fluid along the pipe.
For updating age-dependent properties along time in the mechanical model, the equivalent
age of concrete teq is adopted. Its formulation is based on an Arrhenius type equation established
for a reference temperature Tref (usually 20C) [51]:
t

teq e

Ea
R

1
1

T
Tref

(7)

2.2

Mechanical model
The mechanical model is relatively similar to those adopted for time-dependent mechanical

analysis of hardened concrete, except for some particularities associated to the facts that: (i) it is
being preceded by a thermal analysis (de-coupled), with imposition of strains that are calculated
with basis on the thermal dilation coefficient of concrete (T) and the previously calculated
temperature field; (ii) there is a strong evolution of mechanical properties throughout the
analysis, dully taken into account through the equivalent age concept; (iii) the strong viscoelastic
behaviour of concrete at early ages makes it necessary to use creep formulations that can provide

-7-

adequate estimates within such time span. In specific regard to the last point (iii), basic creep of
concrete was accounted for through the use of the Double Power Law (DPL), which has a
reasonably good performance on both early age and long term time spans [52]:
J (t , t , )

1
E0 (t , )

1
E0 (t , )

(t , ) m (t t , ) n

(8)

,
where J (t , t , ) is the compliance function at time t for a load applied at instant t , , E0 (t ) is the
asymptotic elastic modulus, and 1, m and n are material parameters. Since drying creep is
negligible for an application that only envisages early age behaviour, it was disregarded [53].
Bearing in mind that the aim of the thermo-mechanical simulations is to assess the risk of
cracking, the post-cracking behaviour is not considered relevant and it is thus not simulated.
Thus, linear elastic behaviour (with creep) is considered for concrete both in compression and in
tension.

THE PARADELA DAM SPILLWAY: DESCRIPTION AND MONITORING

3.1

Overview
The Paradela dam, located in the North of Portugal, is a rockfill gravity dam built in the

1950s, with 540m longitudinal development and maximum height of 112m from foundation.
Due to recent hydraulic problems in one of the dams spillways, it was necessary to build a new
complementary ski-jump spillway on the right margin of the river [54]. The case study reported
in this paper concerns the cooling measures and assessment of cracking risk in the construction
of the central wall of the spillway entrance.
3.2

Description of the spillway entrance

3.2.1

Geometry and construction phasing

The spillway functions in free surface conditions, and has two main entrances at the top level
of the dam, each with 5.5m width, being separated by a hydro-dynamic shaped wall with

-8-

maximum width of 2.8m and 17.4m height. A three dimensional representation of the entrance
region of the dam spillway is shown in Error! Reference source not found.a, whereas its
corresponding plan view at approximately mid-height of the wall is depicted in Error!
Reference source not found.b. The reinforcement of the middle wall can be generally
characterized by 16//200mm vertically and 12//200mm horizontally near each surface with a
concrete cover of 60mm.
The construction of the wall was generally performed with 1.2m tall construction phases,
with empirically defined waiting periods being defined by a target temperature in the core
regions during the cooling period (approximately 27C, which corresponded to 17C above
average daily temperature during construction). In order to minimize such waiting periods, an
air-cooling system based on ventilated prestressing ducts placed horizontally was implemented,
allowing lower peak temperatures and faster return to temperature equilibrium with the
surrounding environment. The main scope of the present paper is the study of a specific
construction phase that corresponds to the zone of embedment of the fixation parts of the sluice
gates. Such fixation parts were approximately 2.5m tall, and it was thus desirable to perform a
2.5m tall construction phase, labelled as 9th phase in Error! Reference source not found.a. Due
to its larger thickness, this construction phase is the critical one in terms of peak temperatures
and cracking risk, being therefore the object of analysis.
3.2.2

Materials

The wall of the spillway entrance was generally cast using concrete of class C30/37 [55]
with the composition labelled as S1-D32 in Error! Reference source not found.. In the 9th
construction phase, due to increased complexity of reinforcement near the downstream extremity
of the wall (related to the salient concrete blocks), two slightly different compositions with
higher fluidity were used in the vicinity of such region, as shown in Error! Reference source
not found. (S3-D32 and S3-D16). In spite of this, the areas of most interest to this study
(thickest regions of the wall and monitored sections) correspond to the upstream region.
-9-

Therefore, and also taking into account the fact that the compositions have similarities, all the
characterizations and modelling in the scope of this work pertain to mix S1-D32. Steel
reinforcement was S400C [55], with characteristic yield stress of 400MPa.
3.2.3 Cooling system
In view of the work reported by Hedlund and Groth [8, 9], which proposed the possibility of
using ventilated prestressing ducts for concrete cooling, and bearing in mind the easy availability
of the corresponding necessary equipment in the construction site of the wall, it was decided to
test the feasibility of this kind of cooling technique. However, in view of practical limitations
posed by contractor/owner, this pilot application of air-cooling system was slightly different
from that of Hedlund and Groth [8]. Instead of placing the tubes vertically along the wall, they
were placed horizontally, even though this implied more limited cooling capacity as the length of
tube along freshly cast concrete is longer. In the particular case of the 9th construction phase, a
total of 6 prestressing ducts of 90mm diameter have been used, with their air intake being made
horizontally at the downstream extremity of the wall, and the outtake made near the upstream
extremity, on the top surface of the casting phase, in order to avoid a direct upstreamdownstream potential leakage channel after construction. The overall path of the ducts is shown
in the schemes of Error! Reference source not found., with the ducts labelled from T1 to T6.
For ventilation, a 0.60m diameter fan was used, with 1200m3/h ventilation capacity, that
collected air from the environment and blowed it into the ducts at an internal air speed of
approximately 8.6m/s (measured with anemometer at the outtake of the ducts). The fan had to be
placed in the downstream extremity of the wall due to practical constraints of the contractor. The
cooling system was only started at the age of 14h after casting to avoid introducing undesirable
vibrations to the freshly cast concrete. The ventilation system was disconnected 8.6 days after
casting in view of the similarity of temperature between the walls core and the surrounding
environment. After that, the prestressing ducts were filled with mortar using standard procedures
[56].

- 10 -

It is remarked that this cooling system had been previously tested in the 8 th phase of casting,
with three prestressing ducts placed at mid-height. Details on this test can be found elsewhere
[57].
3.3

Monitoring and material characterization

3.3.1

General remarks

In order to better understand the effectiveness of the cooling system, its influence on the
cracking risk, and assess the capabilities of the adopted numerical simulation strategy, an
extensive set of actions has been carried out, comprising in-situ internal monitoring of
temperaures/strains of concrete, in-situ validation tests, as well as laboratory material
characterization.
3.3.2

Temperature monitoring

Temperatures inside the 9th construction phase have been monitored with K-type
thermocouples, aiming particularly at assessing temperature profiles in a region near the
maximum width of the wall (i.e. at approximately 5.3m from the upstream extremity: section AA as identified in Error! Reference source not found.a). The placement of temperature
sensors in section A-A is depicted in the scheme of Error! Reference source not found.a,
where thermocouples are identified by the prefix TC. Temperatures in the locations labelled as
VW in Error! Reference source not found.a have also been monitored with resistive
temperature sensors, as these are the locations of vibrating wire strain gauges, which contain
internal temperature sensors for strain compensation. The internal air temperature of ducts T1-T3
has been monitored both in section A-A and in neighbouring areas, as shown in Error!
Reference source not found.b. Environmental temperature (dry-bulb) has been assessed with a
thermocouple.
Monitoring was carried out since the instant of casting during a period of 10 days, and the
measurement frequency was set to 1 reading per each 30 minutes. The internally monitored

- 11 -

temperatures in concrete are shown in Error! Reference source not found. for a vertical and a
horizontal alignment of sensors that passes through sensor VW3. From this figure it can be seen
that the initial temperature of concrete was ~15C, and the peak temperature was approximately
42C in the core regions (VW3 and VW5). Furthermore, it can be observed that the ascending
branch of temperature development is clearly affected at the age of 14 hours, when the cooling
system is activated. In specific regard to the vertical profile of temperatures shown in Error!
Reference source not found.a, the expectable behaviour was captured: the core region has the
highest peak temperatures (VW3, VW5), whereas a decrease trend is seen towards the top
surface. In fact, sensors TC6 and TC7 exhibit maximum temperatures of ~36C, while VW6
(near the top surface) has the lowest peak temperature (circa 27C). Near the bottom surface of
this construction phase, sensor VW1 highlights the importance of the heat storage effect caused
by the previously cast concrete: in fact, even though the temperature peak is lower than that of
the core regions, it occurs later and the heat loss rate observed afterwards is lower than in other
regions. It should also be remarked that all sensors are almost in equilibrium with environmental
temperature by the age of 8 days.
In regard to the temperature development in the sensors located along a horizontal alignment,
shown in Error! Reference source not found.b, it can be noticed that the sensors located in the
vicinity of vertical boundaries exhibit lower temperature variations (VW4 and TC5), with
temperature peaks under 35C. It is interesting to observe that TC5 is placed nearer the surface
(15cm apart) than the case of VW4 (20cm apart), and consequently the temperature peak of
VW4 is slightly higher. By looking at the temperature evolution after the age of 8 days (heat of
hydration has been dissipated), it can be seen that temperatures in VW4 remain higher than those
of TC5 due to a combination of two main reasons: VW4 is located in the vicinity of a surface
oriented to southeast, which is bound to receive more energy through solar radiation than the
surface near TC5, which is oriented to northwest; VW4 is 5cm deeper than TC5, thus being
slightly nearer the inner core (higher thermal inertia);.

- 12 -

The temperature evolution along the air inside duct T3 measured by sensors TC3, TC9 and
TC10 (Error! Reference source not found.b) is shown in Error! Reference source not
found., where the environmental temperature and the temperature in VW5 (hottest region in
concrete) are also represented for comparative purposes. The information provided by such
figure allows the clear identification of the instant at which air ventilation began (14 hours), as
the rate of temperature rise is clearly disrupted inside the duct. Furthermore, the rising
temperature tendency along the ducts length is identifiable, as the temperature is consistently
higher in TC3 in regard to TC10, and in TC9 in regard to TC3. Taking as example the
temperatures recorded at the instant of peak temperature (1.88 days), TC9 measured a
temperature of 32.2C, whereas TC10 indicated a temperature of 29.4C. This represents a shift
in temperature of approximately 3C in 3.5m length of duct. At three instants of this study,
temperatures were measured also at the entrance of T3 (z=0) and at z=1m through the use of a
handheld temperature probe (PT1000). By joining such data with the results of TC3, TC9 and
TC10, it was possible to plot a temperature profile along the duct for t=3.84d, t=4.56d and
t=6.58d see Error! Reference source not found.b. It can be observed that the temperature at
the inlet of the tube matches the environmental temperature, and that the heating of the air along
the tube is strongly dependent on the combination of environmental temperature and internal
temperature of concrete. In fact, in the most unfavourable situation shown by Error! Reference
source not found.b, air was heated from ~11C at the air inlet to ~26C at a point located 25m
away from the air inlet (t=4.56d). This increase of air temperature is bound to reduce its cooling
capacity. However, in spite of such diminishment of cooling capacity, the temperature of the air
in the hotter regions of the duct remained at least 10C below that of concrete during the periods
at which temperature in concrete was near its peak (see t=1d to t=3 d in Error! Reference
source not found.a), showing that the heat removal potential was not negligible at all.
The observed diminishment of cooling capacity along the length of the duct highlights the
fact that the adopted configuration for the tubes does not maximize cooling capacity, which

- 13 -

would be conversely maximized if the length of tube inside concrete had been minimized. Such
goal could have been achieved by providing a vertical arrangement for the tubes and introducing
more individual smaller tubes.
3.3.3

Strain monitoring

Strain measurement was carried out with vibrating wire strain gauges of metallic casing with
14cm reference length (TES/5.5/T Gage Technique). Past laboratory tests and in-situ
applications [45, 58, 59] have shown that this kind of sensor is robust and adequate for strain
measurement in concrete at early ages. The strain gauges were placed at the locations identified
in Error! Reference source not found.a (VW1 to VW6), dully positioned in order to measure
strains in the longitudinal direction of the wall. VW7 has distinct intents and shall be specifically
addressed later. Measurements were taken with the same datalogger and at the same sampling
rate as it was the case for temperature sensors.
One important issue to tackle is the zeroing of the measured strains. In fact, before concrete
sets and has enough stiffness to drive the sensor into the same deformation state, the
measurements taken by the sensor do not have any relevant physical meaning. It is thus
necessary to assess the instant of solidarization (i.e. the full bond) between concrete and the
sensor. In a previous work [58], the zeroing operation has been made by assessing the instant at
which two sensors with different casing (plastic and metallic), placed under the same conditions
inside concrete, started yielding the same results. This means that both sensors are solidarized (as
the plastic sensor is bound to solidarize earlier due to its smaller stiffness). Since the plastic
cased sensor was not available, an alternative methodology for zeroing the data was
implemented. By interpretation of the findings reported in [58] it can be considered that the
solidarization instant coincides with a progressive change in the derivate of strain variation
detected by the sensor, that can be obtained by geometrical intersection of tangents of measured
strains, as shown in Error! Reference source not found.. It was decided to use such zeroing
criterion in the scope of this research work. As a result of the application of such rule, the

- 14 -

solidarization instant of each sensor VW1 to VW5 was, respectively: 0.17, 0.19, 0.19, 0.20 and
0.17 days (due to malfunctioning of the VW6, the strain results from this sensor are not
available). The solidarization instants seem coherent, since they have a trend to increase with the
distance of the sensor from the bottom surface of the casting block, due to the natural delay in its
involvement by concrete during the casting process.
The measured strains in sensors VW1 to VW5 are shown in Error! Reference source not
found.. Even though the strain output is dependent on several factors that interact with each
other (thermal deformation, restraint, creep), it is possible to find a set of common points and
reasoning. Overall, all deformations are strongly commanded by the temperature variation,
following the same kinetics. Sensors VW2, VW3 and VW5, which are located in regions near
the core of the walls cross section and had similar temperature development histories, also have
similar strain developments. This is bound to be caused by similar thermal deformations and
restraints for these locations of measurement. The smallest deformations are recorded in VW1,
which is located in the bottom of the casting phase (i.e. near the existing concrete) and thus
having less temperature rise (thus less expansion), while being more restrained by the existing
concrete below at lower temperatures. Finally VW4, which is near the surface and thus has lower
temperature rises (maximum temperature of ~33C), also has a smaller deformation variation
when compared to VW2, VW3 and VW5.
In order to assess free deformations of the concrete used in the construction (associated to
unrestrained autogenous shrinkage and thermal deformations), strain was measured in a concrete
cylinder placed in specially devised conditions. The concrete cylinder (150mm diameter and
300mm tall), cast simultaneously with the studied construction phase and by using the same
concrete, was cast into a special mould internally coated with a soft membrane (with lids also
coated with such material). A strain gauge was placed inside the mould to measure longitudinal
strains see photo of the open mould in Error! Reference source not found.a. After casting
concrete into such mould, it was placed horizontally inside the studied construction phase

- 15 -

(during its casting procedures) at the location that is identified as VW7 in Error! Reference
source not found.a. This kind of procedure, here termed as use of a no-stress specimen, has
been reported in Choi et al. [60], and it allows to measure free deformations of concrete, which
can in turn be used to assess the thermal dilation coefficient, TDC, (if the temperature inside the
concrete cylinder is assumed to be relatively uniform, and autogeneous shrinkage deformations
are known). Unfortunately, due to undetermined causes, the output of the sensor could not be
read during the first 0.8 days, and thus the reported data only starts at such age, as shown in
Error! Reference source not found.b.
3.3.4 Heat generation and activation energy
In an extensive experimental program on the characterization of the cements marketed in
Portugal, Azenha [45] has reported a library of heat generation obtained through isothermal
calorimetry under several temperatures for plain cement pastes with w/c=0.5. The cement used in
the construction concerned in this paper was also characterized (same supplier and
manufacturing plant), and the resulting information for calorimetry tests under 20C, 30C, 40C
and 50C is shown in Error! Reference source not found.. A reasonable estimate of the heat
generated by concrete can be obtained by multiplying the heat generation reported in Error!
Reference source not found. by the volumetric content of cement, which is of 224kg/m3. By
using the speed method algorithm [45, 61], the necessary data for the numerical simulation of
heat generation according to equation 2 was obtained: Ea = 37.31 kJ/mol, A = 4.989109W/m3,
Qpot = 8.295107J/m3, and function f() characterized by the following set of data
[; f()] = [0.00; 0.00], [0.05; 0.58], [0.10; 0.85], [0.15; 0.98], [0.20; 1.00], [0.30; 0.94], [0.40;
0.69], [0.50; 0.41], [0.60; 0.22], [0.70; 0.13], [0.80; 0.07], [0.90; 0.02], [1.00; 0.00]. Even though
this data pertains to CEM I 42.5R of the same company that supplied the cement to this
construction site, there may be deviations caused by inevitable variations in the characteristics of
the cement. Also, the extrapolation procedure mentioned above did not take into account the
presence of fly ash in the mix (96kg/m3), which may have non-negligible effects on the heat
- 16 -

generation potential and hydration kinetics. Therefore, in order to assess the potential importance
of such deviations, a semi-adiabatic test was conducted in-situ (simultaneously with the casting
operations) in a 30cm edge concrete cube, duly isolated by 2.1cm plywood and 12cm of
polystyrene. The results of such semi-adiabatic calorimetry test shall be addressed in section 4,
upon the simulation of its temperature development through the finite element method.
3.3.5

Complementary laboratory characterization

Compressive strength evolution was assessed with concrete cubes (150mm edge) at the ages
of 1, 3, 7 and 29 days, whereas tensile strength was measured with splitting tests on cylinders
(150mm diameter and 300mm tall) and at the same ages. The evolution of both tensile and
compressive strength for concrete cured at 20C (saturated conditions) is shown in Error!
Reference source not found.a (average results of three specimens at each age). In order to
assess the activation energy suitable for compressive strength maturity estimations, a set of
concrete cubes was cured at 40C with the compressive strength measured at the same ages. The
corresponding results are also shown in Error! Reference source not found.a. By applying the
equivalent age concept [62], together with the superposition method [61], it was possible to
asses that the activation energy based on mechanical testing has the value of 37 kJ/mol, which is
rather consistent with the activation energy obtained through isothermal calorimetry for the same
cement (yet without fly ash), 37.31 kJ/mol [45]. Such coincidence in activation energy for
thermal and mechanical phenomena had already been reported by Ulm [63].
Basic creep was assessed in creep rigs on prismatic specimens (sealed) with dimensions
15cm15cm60cm, loaded at 30%~40% of the concrete compressive strength and internally
monitored with vibrating wire strain gages. Such creep tests were conducted at the ages of 1, 3
and 7 days, and the corresponding specific creep curves are shown in Error! Reference source
not found.b.
The experimental program also included a single specimen for measurement of autogenous
shrinkage. Such specimen was a 150mm diameter and 300mm long cylinder, internally

- 17 -

instrumented with a vibrating wire strain gage, which was kept in its formwork during the
experiment and sealed with a plastic film on the top surface. Unfortunately, two factors
contributed to render the results of this specimen unusable for this research: on one hand, the
monitoring only could be started at the age of 2 days in the laboratory due to unavailability of
datalogging system; on the other hand, the measurements of autogenous taken since t=2 days
were disturbed by an inefficient sealing, which promoted undesired drying of the specimen. This
was not considered a critical problem in view of the low values of autogenous shrinkage that are
usually expectable in concretes of low cement content and high w/c ratio [64, 65].
3.3.6 Continuous monitoring of concrete stiffness
The evolution of elasticity modulus along time was measured through compressive cyclic
testing in concrete cylinders (150mm diameter and 300m tall) at the ages of 1, 3, 7, 15 and 29
days. Concomitantly, E-modulus of concrete was continuously assessed through a methodology
termed as EMM-ARM (Elasticity Modulus Measurement through Ambient Response Method).
This methodology has been developed by Azenha et al. [43] and consists in a variant to classic
resonant frequencies that allows the quantification of E-modulus continuously since the instant
of casting of the specimen inside the testing mould. The basic principle of EMM-ARM is the
following: the specimen is cast inside the testing mould, which is in turn placed in simply
supported conditions and continuously subject to modal identification (using accelerometers)
without any explicit excitation of the beam, as ambient vibration suffices. As concrete hardens,
the first resonant frequency of the composite beam evolves, and the stiffness of concrete can be
inferred by applying the equations of motion. Details about the testing setup and procedure
applied for the concrete of this spillway application can be found in [57, 66, 67], as it
corresponds to an improved version of the originally devised test (a steel mould is used). The
collected results with EMM-ARM and cyclic compression tests on cylinders are shown in
Error! Reference source not found., where the feasibility of EMM-ARM is confirmed in view

- 18 -

of the resemblance of results. Also, the richness of information that can be obtained through
EMM-ARM represents an added value for the numerical simulation.

NUMERICAL MODELLING

4.1

Geometry, mesh, materials, initial/boundary conditions and time integration

4.1.1

Geometry of the model and finite element mesh

A cross-sectional scheme of the model for simulation is shown in Error! Reference source
not found.a, where the construction stages considered in the analysis can be observed. The first
stage of the model encompasses all concrete until the 7th phase of concreting (inclusive),
considered as hardened concrete, together with the 8th phase of concrete evaluated as freshly cast
concrete. The second and third stages correspond to the 9th and 10th phases of concreting
respectively. This strategy diminishes the computational cost of the model without relevant
effect on the accuracy of results. For similar reasons, the underlying subgrade is not modelled, as
it is far from the construction phases of interest. Due to the geometrical symmetry of the wall, a
longitudinal plane of symmetry is considered.
The simulation was made with a 3D finite element model comprising rectangular brick FE of
8 nodes (222 integration scheme) for concrete in the thermal model, and coincident 20 nodes
brick FE (333 integration scheme) in the mechanical analysis. Convective boundaries were
modelled with 4 node planar elements (22 integration scheme), and the cooling ducts were
considered with linear elements of 2 nodes (2 point integration scheme) [44]. The schematic
representation of geometry, casting phases and cooling duct location for section A-A is shown
in Error! Reference source not found.b. The reader is reminded that phase 8 also had cooling
ducts, according to the description of section 3.2.3. The longitudinal layout of all the ducts was
considered straight on an horizontal plane, in correspondence to simplifications in the vicinity of
the extremities of the wall.

- 19 -

It should be remarked that due to the phased analysis, the mesh evolved along time with
some elements/boundaries being activated/de-activated (e.g. the convective top boundary of a
given phase is de-activated upon the beginning of the next casting phase).
The mesh adopted for this simulation is shown in Error! Reference source not found.c, with a
total of 18738 elements and 66037 nodes. As consequence of the symmetry simplification, it was
considered that the diameter of the tubes located in the symmetry axis was equal to half the
actual diameter. As solar radiation does not represent a symmetrical energy input to the structure,
the symmetry simplification adopted is not truly valid. However, as solar radiation has most of
its effect near the surface, it was decided to keep the symmetry simplification by considering the
southeast half of the wall, which is most subject to solar radiation effects.
4.1.2 Materials (thermal and mechanical properties)
The thermal conductivity and specific heat of concrete were estimated with basis on the
pondered average of the corresponding thermal properties of the constituent materials of the mix
[45, 68]. The adopted values for k and c for concrete were, respectively 2.40 W/mK and
2.400E+06 J/m3K. Even though it is known that these thermal properties suffer variations during
early ages [69-75], the adopted modelling approach considers them constant in view of the
conclusions of the parametric analyses reported by Azenha [45], where a relatively small impact
of considering evolving k and c was found on computed temperatures in hardening concrete.
In regard to the data for heat of hydration, the parameters mentioned in section 3.3.4 were used.
The adequacy of these parameters was evaluated through the semi-adiabatic calorimeter
described in the same section, whose behaviour was simulated through a FE simulation model
that explicitly considered the extruded polystyrene (XPS) and wood walls of the calorimeter. The
material modelling parameters for concrete coincide with those herein described, whereas
additional information and mesh representation are shown in Error! Reference source not
found.a. The results of the simulation of the calorimeter were quite coherent with those collected
experimentally, as seen in Error! Reference source not found.b, leading to the confirmation
- 20 -

that the strategy described in section 3.3.4 to determine the heat generation and activation energy
was adequate for defining the heat of hydration modelling parameters adopted in the simulations.
The thermal dilation coefficient (TDC) of concrete was assessed with basis on the no-stress
specimen described in section 3.3.3. However, in order to obtain the thermal deformation and
calculate the TDC, it was necessary to subtract the autogenous shrinkage deformation from the
total deformation. As data on autogenous shrinkage was not available, an estimate of the
autogenous shrinkage evolution based on Eurocode 2 [55] was used. Another issue to take into
account is the fact that the TDC of concrete is not constant during the first hours of age. In fact,
several authors have dealt with this subject, and it generally agreed that the initial TDC tends to
be larger than that of hardened concrete, and tends to decrease sharply within the first 12 to 24
hours of age, reaching then the plateau level corresponding to hardened concrete [69, 73, 76].
The no-stress specimen cast in the scope of this research cannot be used to estimate TDC at
concrete early ages, due to the absence of data in the first 8 hours reported in Error! Reference
source not found.b. Nonetheless, as the peak temperature occurred later than 24 hours age, and
full data is available for temperatures and strains occurred in such period, calculations could be
made under the assumption that TDC was already at its plateau value. The TDC was estimated
between instants t=2.0d (peak temperature) and t=8.0d (local minimum) as shown in Error!
Reference source not found.b, and the autogenous shrinkage strain in such period was
estimated to be of 8.45, (considering fcm=42.3MPa in Eurocode 2 [55]). The estimated
constant TDC to be used in the numerical simulation was 11.07C. Nonetheless, since it is
known that TDC varies during the first 24 hours of age, an alternative formulation for the
evaluation of the TDC was considered, based on experimental evidence reported by Laplante and
Boulay [77]. Therefore, this alternative formulation considers that during the first 16h the TDC
varies according to TDC (t ) 0.16t 2 4.88t 48.93 (t in hours), and remains constant after such
age.

- 21 -

The E-modulus evolution of concrete for the simulation model was directly extracted from
EMM-ARM data reported in Error! Reference source not found., whereas creep modelling
was made through the adjustment of DPL parameters to the creep data experiments. The best-fit
creep parameters and their adjustment to the experimental data are shown in Error! Reference
source not found.b. Prestressing ducts were modelled with consideration of their inner
perimeter of 282.74mm (90mm inner diameter). Steel reinforcement was disregarded in
temperature calculations due to its low interference in temperature development [45, 53].
Regarding mechanical field simulations, steel was not considered because post-cracking
behaviour was not sought. In the non-cracked stage, the similitude in TDC of steel in regard to
that of hardened concrete strongly minimizes the restraint to concrete thermal deformation, thus
rendering the effect of reinforcement negligible for the computation of thermal stresses at early
ages [53].
4.1.3 Initial/boundary conditions and construction phasing
In what concerns the boundary conditions in the thermal problem, it was assumed that an
average wind speed of 2.5m/s occurred (confirmed during 3 days with an anemometer) and the
resulting convection/radiation coefficient for concrete surfaces in contact with the environment
was estimated to be 15.25 W m2 K in accordance to the predictive formula of [78]. For the
particular case of formworks surrounding concrete, an equivalent boundary convection
coefficient was adopted according to an electrical analogy [47]. Bearing in mind that the
formworks were made of wood (kwood=0.175W/m2K) and their thickness was 18mm, the
resulting equivalent boundary coefficient was 6W/m2K. Formwork was applied to the vertical
surfaces of each casting stage during the first 7 days of age, and removed afterwards. All
convective boundaries were subjected throughout the analysis to the environmental temperature
that was monitored in-situ see Error! Reference source not found.. The symmetry plane of
the model was considered as an adiabatic boundary.

- 22 -

In what regards to solar radiation intake, the several surface directions of the model were
taken into account, and the absorbed radiation was calculated according to the model described
in section 2.1. Absorvity of concrete was considered as 0.6 [79] , the latitude was 41.77 N, and
the casting date was 28/03/2011. The Linke turbidity factor was considered as 2.5, and its
feasibility was confirmed by comparing computed solar radiation on horizontal surfaces with
solar radiation data from a nearby weather station (Rio Torto Station) in conditions of clear
skies. The effect of cloudiness was taken into account by normalizing the predictions of the
adopted solar radiation model according to information obtained from the piranometer of the
neighbouring weather station. The diminishment of solar absorption caused by shadows cast by
neighbouring objects was considered negligible throughout the entire day, as the wall was one of
the tallest elements in the landscape.
Bearing in mind the construction phases taken into account in this calculation model
(identified in Error! Reference source not found.a), the initial temperatures were considered as
follows. For the existing concrete at the beginning of analysis, it was assumed that concrete was
already in thermal equilibrium with the environment, and so, the average daily temperature of
the preceding week (14.5C) was considered for the existing concrete. For the subsequent stages
of construction, the initial temperature of concrete was obtained from in-situ monitoring (average
value), which was also of approximately 14.5C.
In regard to the cooling ducts, due to a limitation of the adopted software, the inlet
temperature had to be considered constant, equal to the average environmental temperature
during the time in which the tubes were active. The following temperatures were considered for
each phase: 8th: 7.13C and 9th: 13.91C. The internal convection coefficient in the ducts was
obtained with basis on their internal air speed of 8.6m/s, which according to the studies of
Hedlund and Groth [8] should correspond to a convection coefficient of 30.0W/m2K. The
cooling duct elements were activated at each construction phase when the surrounding concrete
had age of 14 hours.

- 23 -

Taking into account the directions of the axes of the coordinate system presented in Error!
Reference source not found.a, the mechanical boundary conditions consisted in placing Z
direction supports on the bottom surface of the wall and Y supports in all the elements of the
symmetry plane.
The time step strategy adopted for the model consisted in considering the initial time
coincident with the casting instant of 8th phase. Casting of the 9th and 10th phases were
considered at the relative instants t=24 days and t=41 days in accordance to the actual
construction. All analyses were conducted with a constant time step of 1h duration, even though
some localized adjustments were necessary in view of construction phasing and ventilation
activation. Nonetheless, all adjustments were carefully made to assure that the duration of all
time steps remained under 1h.
4.2

Results and discussion


The presentation and discussion of results is centred in the 9th construction phase, with

particular emphasis for comparisons between monitored and simulated temperatures/strains. The
temperature simulations in concrete were quite coherent with the monitored ones, as it can be
confirmed for a set of representative locations (VW1, VW2, VW4 and VW5), whose results are
shown in Error! Reference source not found.. In fact the largest deviations regarding the
monitored temperatures during the entire calculation always remain under 4C, thus providing
confirmation of the feasibility of the modelling strategy, particularly in regard to the cooling
capacity of the ventilated ducts. Based on the confidence gained on the thermal simulation
model, a further numerical simulation was made, in which the effect of the cooling ducts was
disregarded. The corresponding results for the hottest region of both models are shown in Error!
Reference source not found.b. It can be confirmed that the inclusion of the duct had a twofold
effect: not only was the peak temperature diminished by 5C with benefits for cracking safety,

- 24 -

but also the return of the internal temperature to thermal equilibrium with the environment was
accelerated, with advantages for construction phasing.
The fact that the calculated temperatures matched well the monitored ones is a solid starting
point for the analysis of results of the mechanical simulation, as any detected deviations are
bound to be solely attributed to issues in the mechanical simulation itself. The calculated and
measured strains for the same set of sensors that has just been discussed for temperature
development are shown in Error! Reference source not found.. The experimentally measured
strains in this figure are represented by their value according to the zeroing procedure mentioned
in Section 3.3.3 (Error! Reference source not found.), but also with a lower and upper bound
related to possible uncertainties in the instant for zeroing of the sensors output of 2 hours. It can
be seen that all the computed strains with consideration of constant TDC underestimate the peak
strain at 1.96 day, but the post-peak kinetics seems to have been well captured. As the constant
TDC assumption may lead to underestimations of early strain development [58], a further
calculation was made using a plausible TDC evolution during the first 24 hours, as discussed in
4.1.2. The corresponding simulation results are shown in Error! Reference source not found.,
where a better overall fit is seen between experimental and calculation data (particularly for core
regions). Even though the variable TDC was not based on experimental evidence obtained in the
scope of this research, it is feasible to assume that a significant part of the strain deviations
regarding experimental results can be explained by the variable TDC at early ages. It has to be
kept into consideration that another possible source of deviation of results may be related to the
instant at which measured strains were zeroed, which can be debatable. Nonetheless, the
adequate prediction of strains that was attained is a good indication of the feasibility of the
computed stresses which are to be analysed.
The discussion of cracking risk is now addressed by comparing the computed principal
tensile stress at the most unfavourable part of the model (located in the core region: x=8.5m,
y=0.36m, z=11.7m), as shown in Error! Reference source not found.. This figure contains the

- 25 -

equivalent age-adjusted measured evolution of the tensile strength, and the calculation results for
the cases of constant and variable TDC, as well as the case of inexistent cooling ducts and
constant TDC. The first comment that can be made from Error! Reference source not found.,
is that the consideration of constant or variable TDC had very marginal effect on the results. The
reason for the very small difference is bound to be related with the very low stress level that is
induced during the first 24 hours, as the E-modulus of concrete is still very small and
creep/relaxation is very high. Regardless of the comparison between these two models, it can be
observed from Error! Reference source not found. that the ratio between the tensile stress and
the tensile strength of concrete at the most unfavourable instant is of approximately 0.9, which
corresponds to a significant cracking risk. Nonetheless, even though the cracking risk was higher
than the desirable one, the structure did not present thermal cracks neither later through-cracks
(evaluations made until 2 years after casting). It should be remarked that this point of stress
analysis was the most unfavourable one within the structures, and significantly lower cracking
risks were calculated for distinct regions, resulting in a global scenario of much more cracking
safety (compared to a single-point analysis).
Interestingly, a simulation of the same construction situation without consideration of
cooling ducts would yield to higher cracking risk at the same location (as seen in Error!
Reference source not found.), with the ratio between the tensile stress and the tensile strength
of concrete reaching 1.2. Therefore, if the calculations made here are considered trustworthy, the
use of the cooling ducts may have been the differentiating factor that avoided a cracking scenario
in this concrete lift.

CONCLUSIONS
A case study regarding the assessment of the cracking risk of a thick wall in the entrance of a

dam spillway, internally cooled with air-filled prestressing ducts, was presented in this paper.
The use of ventilated prestressing ducts is considered more straightforward than water ducts due

- 26 -

to the often easy availability of ducts and fans in construction works associated to massive
concrete structures.
The case study has involved a comprehensive experimental part, including laboratory
characterization of materials and in-situ monitoring for temperatures and strains. A 3D thermomechanical simulation of the construction phasing was shown, with input data duly based in the
laboratory characterization program.
In regard to previous works reported in the literature, the work presented in this paper has its
main original contributions in the following fields: (i) the EMM-ARM methodology for
continuous monitoring of concrete E-modulus since casting was applied for the first time as a
characterization tool for stress simulation in concrete at early ages, thus enhancing the quality of
input data; (ii) this is the first reported application of horizontally placed air-cooling pipes, with
its efficiency being assessed and numerically simulated; (iii) the work reported here is relatively
unprecedented in view of its holistic approach, with the authors being involved in all tasks of
laboratory characterization (allowing sustainable estimates of material properties and modelling
strategy for numerical simulations), field monitoring and numerical simulation with thermomechanical analysis.
The numerical simulation results were compared to those collected by the in-situ monitoring
and good coherences were observed both in terms of temperature and strain, providing good
prospects in regard to the simulation capabilities of the models and the soundness of the
experimentally obtained data. The monitoring/simulation results allowed concluding that the
effectiveness of the air cooling system with horizontally placed pipes is limited in view of the
significant heating that air suffers along the first meters of tube, thus diminishing its capacity of
cooling further regions of concrete. Also, the duct efficiency ends up being quite dependent of
the environmental temperature, which cannot be easily anticipated during planning. Nonetheless,
in spite of the acknowledged limitations of air cooling, it may prove quite feasible in relatively
cool climates and small lengths of embedment (e.g.: 10m or less).

- 27 -

Furthermore, the risk of cracking on the studied construction phase has resulted acceptable in
most of its regions, even though a non-negligible risk of internal cracking was observed in some
regions. The fact that no surface or through cracking was observable in the construction
corroborates the cracking risk evaluation. It was also concluded that the same construction
phasing without the use of the cooling ducts had a significantly higher cracking probability, thus
confirming the usefulness of the cooling system.

ACKNOWLEDGEMENTS
Funding provided by the Portuguese Foundation for Science and Technology to the Research

Unit ISISE, as well as to the second author through the PhD grant SFRH/BD/64415/2009, and to
the research project PTDC/ECM/099250/2008 is gratefully acknowledged. The kind assistance
of the contractor (Teixeira Duarte S.A.) and the owner (EDP Eletricidade de Portugal) are also
deeply appreciated. The contribution of ngelo Costa to the experimental work here reported is
also gratefully acknowledged.

REFERENCES

[1] Bernander S. Practical Measures to Avoiding Early Age Thermal Cracking in concrete
structures. In: Springenschmid R, editor. Prevention of Thermal Cracking in Concrete at Early
Ages : State-of-the-art Report Prepared By RILEM Technical Committee 119 (RILEM Report
15): E & FN Spon. 1998; 255-314.
[2] Springenschmid R. Thermal Cracking in Concrete At Early Ages RILEM Proceedings
(RILEM Report 25): Taylor & Francis Routledge. 1995.
[3] Du C. Dam construction - concrete temperature control using fly ash. Concrete International.
1996; 18:34-6.
[4] Utsi S, Jonasson J-E. Estimation of the risk for early thermal cracking for SCC containing fly
ash. Materials and Structures. 2012; 45:153-69.
[5] Kurita M, Goto S, Minegishi K, Negami Y, Kuwahara T. Precooling concrete using frozen
sand. Concrete International. 1990; 12:60-5.

- 28 -

[6] Takeuchi H, Tsuji Y, Nanni A. Concrete precooling method by means of dry ice. Concrete
International. 1993; 15:52-6.
[7] St. John J. Construction of the Hoover Dam Bypass. Concrete International. 2011; 33:30-5.
[8] Hedlund H, Groth P. Air cooling of concrete by means of embedded cooling pipes-Part I:
Laboratory tests of heat transfer coefficients. Materials and Structures. 1998; 31:329-34.
[9] Groth P, Hedlund H. Air cooling of concrete by means of embedded cooling pipesPart II:
Application in design. Materials and Structures. 1998; 31:387-92.
[10] Roush K, O'Leary. Cooling concrete with embedded pipes. Concrete International. 2005;
27:30-2.
[11] Maggenti R. From passive to active thermal control. Concrete International. 2007;29:24-30.
[12] Ishikawa S, Matsukawa K, Nakanishi S, Kawai H. Air pipe cooling system. Concrete
International. 2007; 29:45-9.
[13] Nobuhiro M, Kazuo U. Nonlinear thermal stress analysis of a massive concrete structure.
Computers & Structures. 1987; 26:287-96.
[14] Emborg M. Thermal Stresses in Concrete Structures at Early Age: Lule University of
Technology; 1989.
[15] Ishikawa M. Thermal stress analysis of a concrete dam. Computers & Structures. 1991;
40:347-52.
[16] de Borst R, van den Boogaard AH. Finite-Element Modeling of Deformation and Cracking
in Early-Age Concrete. Journal of Engineering Mechanics. 1994; 120:2519-34.
[17] Torrenti JM, De Larrard F, Guerrier F, Acker F, Grenier F. Numerical simation of
temperatures and stresses in concrete at early ages: the French experience. In: R.Springenschmid,
editor. Thermal Cracking in Concrete at Early Ages (RILEM Proceedings 25). Paris: E & FN
Spon. 1994; 281-8.
[18] Fairbairn EMR, Ferreira IA, Cordeiro GC, Silvoso MM, Filho RDT, Ribeiro FLB.
Numerical simulation of dam construction using low-CO2-emission concrete. Materials and
Structures/Materiaux et Constructions. 2010; 43:1061-74.
[19] De Schutter G, Vuylsteke M. Minimisation of early age thermal cracking in a J-shaped nonreinforced massive concrete quay wall. Engineering Structures. 2004; 26:801-8.
[20] Dolmatov AP, Neidlin SZ. Temperature control of the massive concrete of the Krasnoyarsk
hydroelectric station dam. Hydrotechnical Construction. 1968; 2:956-60.
[21] Blinov IF, Shaikin YP, Gal'perin IR. Field studies of the temperature regime and stress state
of the concrete in arch dams subjected to various methods of cooling. Hydrotechnical
Construction. 1980; 14:19-23.

- 29 -

[22] Torrenti J-M, Buffo-Lacarrire L, Barr F. CEOS.FR experiments for crack control of
concrete at early age. RILEM-JCI International Workshop on Crack Control of Mass Concrete
and Related Issues concerning Early-Age of Concrete Structures (ConCrack 3). Paris. 2012; 310.
[23] Cussigh F. Experience in limiting early age concrete temperature for DEF prevention.
RILEM-JCI International Workshop on Crack Control of Mass Concrete and Related Issues
concerning Early-Age of Concrete Structures (ConCrack 3). Paris. 2012; 79-88.
[24] Benboudjema F, Torrenti JM. Early-age behaviour of concrete nuclear containments.
Nuclear Engineering and Design. 2008; 238:2495-506.
[25] Craeye B, De Schutter G, Van Humbeeck H, Van Cotthem A. Early age behaviour of
concrete supercontainers for radioactive waste disposal. Nuclear Engineering and Design. 2009;
239:23-35.
[26] Briffaut M, Benboudjema F, Torrenti JM, Nahas G. Numerical analysis of the thermal
active restrained shrinkage ring test to study the early age behavior of massive concrete
structures. Engineering Structures. 2011; 33:1390-401.
a L, Cussigh F, Lecrux S. Using the maturity method in concrete cracking
control at early ages. Cement and Concrete Composites. 2004; 26:589-99.
[28] Xiang Y, Zhang Z, He S, Dong G. Thermalmechanical analysis of a newly cast concrete
wall of a subway structure. Tunnelling and Underground Space Technology. 2005; 20:442-51.
[29] Faria R, M. A, J.A. F. Modelling of concrete at early ages: application to an externally
restrained slab. Cement & Concrete Composites. 2006; 28(6):572-85.
[30] Azenha M, Faria R. Temperatures and stresses due to cement hydration on the R/C
foundation of a wind tower-A case study. Engineering Structures. 2008; 30:2392-400.
[31] Boutillon L, Linger L, Kolani B, Meyer E. Effects of sun irradiation on the temperature and
early age stress distribution in external concrete structure. In: Toutlemonde F, Torrenti J-M,
editors. RILEM-JCI International Workshop on: Crack control of mass concrete and related
issues concerning early-age of concrete structures ConCrack 3. Paris: RILEM. 2012; 181-92.
[32] Li J. Predicting early-age thermal behaviour of mass concrete for bridge foundations [MSc
Thesis]. Ames, Iowa, United States of America: Iowa State University. 2012.
[33] Zreiki J, Bouchelaghem F, Chaouche M. Early-age behaviour of concrete in massive
structures, experimentation and modelling. Nuclear Engineering and Design. 2010; 240:2643-54.
[34] Zhu Y-M, Liu Y-Z, Xiao Z-Q, He J-R, Lin Z-H, Ma Y-F. Analysis of pipe-cooling system
in mass concrete. Nanjing, China: College of Water Conservancy & Hydropower Eng., Hohai
Univ. 2004.

- 30 -

[35] James RJ, Dollar DA. Thermal Engineering for the construction of large concrete arch
dams. The 6th ASME-JSME Thermal Engineering Joint Conference. Hapena Beach, Hawaii.
2003.
[36] Xie H, Chen Y. Influence of the different pipe cooling scheme on temperature distribution
in RCC arch dams. Communications in Numerical Methods in Engineering. 2005; 21:769-78.
[37] Myers TG, Fowkes ND, Ballim Y. Modeling the cooling of concrete by piped water. ASCE
Journal of Engineering Mechanics. 2009; 1375-83.
[38] Yamamoto T, Ohtomo T. Pratices for crack control of concrete in Japan. In: Toutlemonde F,
Torrenti JM, editors. RILEM-JCI International Workshop on Crack Control of Mass Concrete
and Related Issues Concerning Early-Age of Concrete Structures (ConCrack 3). Paris: RILEM
Publications. 2012; 193-200.
[39] Tanabe T-a, Yamakawa H, Watanabe A. Determination of convection coefficient at cooling
pipe surface and analysis of cooling effect. Proceedings of JSCE. 1984; 34:171-9.
[40] Yang J-K, Lee Y, Kim J-K. Heat Transfer Coefficient in Flow Convection of Pipe-Cooling
System in Massive Concrete. Journal of Advanced Concrete Technology. 2011; 9(1):103-14.
[41] Yang J, Hu Y, Zuo Z, Jin F, Li Q. Thermal analysis of mass concrete embedded with
double-layer staggered heterogeneous cooling water pipes. Applied Thermal Engineering. 2012;
35:145-56.
[42] Kim JK, Kim KH, Yang JK. Thermal analysis of hydration heat in concrete structures with
pipe-cooling system. Computers & Structures. 2001; 79:163-71.
[43] Azenha M, Magalhes F, Faria R, Cunha . Measurement of concrete E-modulus evolution
since casting: A novel method based on ambient vibration. Cement and Concrete Research.
2010; 40:1096-105.
[44] TNO-DIANA-BV. Diana Users Manual - Release 9.4.4. Delft, The Netherlands. 2012.
[45] Azenha M. Numerical Simulation of the Structural Behaviour of Concrete Since its Early
Ages. Porto, Portugal: University of Porto. 2009.
[46] Reinhardt H, Blaauwendraad J, Jongedijk J. Temperature development in concrete
structures taking account of state dependent properties. Int Conf Concrete at Early Ages. Paris,
France. 1982.
[47] Faria R, Azenha M, Figueiras JA. Modelling of concrete at early ages: Application to an
externally restrained slab. Cement and Concrete Composites. 2006; 28:572-85.
[48] Breugel K, Koenders EAB. IPACS Report BE96-3843/2001:31-1. Solar radiation: effect on
solar radiation on the risk of cracking in young concrete. 2001.

- 31 -

[49] Kasten F. A simple parameterization of the pyrheliometric formula for determining the
Linke turbidity factor. Meteorl Rdsch. 1980; 33:124-7.
[50] Incropera FP, DeWitt DP, L. BT, Lavine AS. Fundamentals of Heat and Mass Transfer:
Wiley. 2006.
[51] Freiesleben Hansen P, Pedersen EJ. Maturity computer for controlled curing and hardening
of concrete. Nordisk Betong. 1977; 1:19-34.
[52] Baant ZP, L'Hermite R. Mathematical modeling of creep and shrinkage of concrete. John
Wiley & Sons. 1988; 484 pages.
[53] Japan Concrete Institute. JCI Guidelines for Control of Cracking of Mass Concrete. JCI.
2011. p. 153.
[54] Couto LT, Oliveira MS, Dias da Silva J, Magalhes AP. New Paradela dams spillway:
Design and testing in reduced hydraulic model. 9 Congresso Nacional da gua. Estoril,
Portugal: Associao Portuguesa de Recursos Hdricos. 2008 (in Portuguese).
[55] CEN. EN 1992-1 European Standard Eurocode 2: Design of concrete structures Part 1: general rules and rules for buildings. 2004.
[56] CEN. EN 446:2007 Grout for prestressing tendons: Grouting procedures. 2007. p. 20.
[57] Costa MV. Thermo-mechanical analysis of self-induced stresses in concrete associated to
heat of hydration: a case study of the spillway of Paradela dam. Guimares: University of Minho.
2011 (in Portuguese).
[58] Azenha M, Faria R, Ferreira D. Identification of early-age concrete temperatures and
strains: Monitoring and numerical simulation. Cement and Concrete Composites. 2009; 31:36978.
[59] Yeon JH, Choi S, Won MC. In situ measurement of coefficient of thermal expansion in
hardening concrete and its effect on thermal stress development. Construction and Building
Materials. 2013; 38:306-15.
[60] Choi S, Won M. Thermal strain and drying shrinkage of concrete structures in the field. ACI
Materials Journal. 2010; 107:498-507.
[61] D'Aloia L, Chanvillard G. Determining the apparent activation energy of concrete: Ea
numerical simulations of the heat of hydration of cement. Cement and Concrete Research. 2002;
32:1277-89.
[62] Carino NJ, Lew H. The Maturity Method: Theory and Application. Structures Congress &
Exposition. Washington (DC): American Society of Civil Engineers. 2001; 15 pages.
[63] Ulm F-J, Coussy O. Strength Growth as Chemo-Plastic Hardening in Early Age Concrete.
Journal of Engineering Mechanics. 1996; 122:1123-32.

- 32 -

[64] Aitcin P-C, Neville A, Acker P. Integrated view of shrinkage deformation. Cement
International. 1997; 19:35-41.
[65] fdration internationale du bton. fib Model Code 2010 (final draft). Lausanne,
Switzerland. 2011; 653 pages.
[66] Azenha M, Faria R, Ferreira D. Monitoring and numerical simulation of the construction
process of a concrete gravity dam. BE2008 - Encontro Nacional Beto Estrutural 2008.
University of Minho, Guimares. 2008 (in Portuguese).
[67] Azenha M, Ramos LF, Aguilar R, Granja JL. Continuous monitoring of concrete E-modulus
since casting based on modal identification: A case study for in situ application. Cement and
Concrete Composites. 2012; 34:881-90.
[68] Breugel K. Prediction of temperature development in hardening concrete. In:
Springenschmid R, editor. Prevention of thermal cracking in concrete at early ages Report 15,
RILEM: E & FN SPON; 1998.
[69] Boulay C, Patis C. Strain measurements on concrete at early age (Mesures des
dformations du bton au jeune ge). Mat Struct. 1993; 26:307-11 (in French).
[70] Bastian G, Khelidj A. The thermophysical properties of a freshly cast concrete. Propriets
thermophysiques d'un bton frachement coul. Bulletin de liaison des LPC. 1995; 200:25-35.
[71] De Schutter G, Taerwe L. Specific heat and thermal diffusivity of hardening concrete.
Magazine of Concrete Research. 1995; 47:203-8.
[72] Kim KH, Jeon SE, Kim JK, Yang S. An experimental study on thermal conductivity of
concrete. Cement and Concrete Research. 2003; 33:363-71.
[73] Bjontegaard O, Sellevold E. Effects of Silica Fume and Temperature on Autogenous
Deformation of High Performance Concrete. In: Jensen O, Bentz D, Lura P, editors. ACI SP 220,
Autogenous Deformation of Concrete: American Concrete Institute, Farmington Hulls, MI.
2004.
[74] Bentz DP. Transient plane source measurements of the thermal properties of hydrating
cement pastes. Materials and Structures/Materiaux et Constructions. 2007; 40:1073-80.
[75] Choktaweekarn P, Saengsoy W, Tangtermsirikul S. A model for predicting thermal
conductivity of concrete. Magazine of Concrete Research. 2009; 61:271-80.
[76] Viviani M, Glisic B, Smith IFC. Separation of thermal and autogenous deformation at
varying temperatures using optical fiber sensors. Cement and Concrete Composites. 2007;
29:435-47.
[77] Laplante P, Boulay C. Evolution du coefficient de dilatation thermique du bton en fonction
de sa maturit aux tout premiers ges. Materials and Structures. 1994; 27:596-605.

- 33 -

[78] Branco F, Mendes P, Mirambell E. Heat of hydration effects in concrete structures. ACI
Materials Journal. 1992; 89:139-45.
[79] Ruiz J, Schindler A, Rasmussen R, Kim P, Chang G. Concrete temperature modeling and
strength prediction using maturity concepts in the FHWA HIPERPAV software. 7th
International Conference on Concrete Pavements. Orlando, Florida, USA. 2001.

- 34 -

Figure Captions

Error! Reference source not found.


Error! Reference source not found.
Error! Reference source not found.
Error! Reference source not found.
Error! Reference source not found.
Error! Reference source not found.
Error! Reference source not found.
Error! Reference source not found.
Error! Reference source not found.
Error! Reference source not found.
Error! Reference source not found.
Error! Reference source not found.
Error! Reference source not found.
Error! Reference source not found.
Error! Reference source not found.
Error! Reference source not found.
Error!

Reference

source

- 35 -

not

found.

Table Captions
Error! Reference source not found.

- 36 -

Table
Click here to download Table: Table01.docx

Table 1 Mix proportions of concrete in the 9th phase of the spillway wall
C30/37
S1-D32
Components

C30/37 C30/37
S3S3D32
D16

(kg/m3) (kg/m3) (kg/m3)


gravel
14-32 mm
gravel
10-16 mm
gravel
4-8 mm

449

400

438

400

490

306

316

387

sand
CEM I
42.5R
fly ash

621

646

763

224

238

280

96

102

120

plasticizer

2.2

3.4

water

170

180

200

Concrete pour plan

Figure01
Click here to download Figure: Figure01.docx

a)

b)

Figure 1 Entrance of the dam spillway: a) three dimensional representation; b) plan view of the
central wall.

Figure02
Click here to download Figure: Figure02.docx

a)
b)
Figure 2 a) Construction phasing of the central wall (elevation); b) Overall photo during the
construction of the 9th phase.

Figure03
Click here to download Figure: Figure03.docx

a)

b)

c)
Figure 3 Cooling system at the 9th construction phase: a) plan view; b) longitudinal
section; b) cross-section A-A.

Figure04
Click here to download Figure: Figure04.docx

a)

b)

Figure 4 Temperature and strain measurement devices: a) Location of sensors


within section A-A b) Location of sensors within the tubes T1, T2 and T3. Units: [m]

Figure05
Click here to download Figure: Figure05.docx

a)

b)
Figure 5 Monitored temperatures in the a) vertical and b) horizontal alignment that
encompasses sensor VW3.

Figure06
Click here to download Figure: Figure06.docx

a)

b)

Figure 6 Monitored temperatures in duct T3: a) Global results for TC3, TC9 and TC10; b)
temperature profile along the duct at three selected instants. Units: [m]

Figure07
Click here to download Figure: Figure07.docx

Figure 7 Validation of methodology adopted for determination of the instant of solidarization


of VW sensor to concrete.

Figure08
Click here to download Figure: Figure08.docx

Figure 8 Strains measured by sensors VW1 to VW5.

Figure09
Click here to download Figure: Figure09.docx

a)

b)

Figure 9 - In situ determination of TDC: a) overview of installed no-stress specimen; b)


evolution of strain and temperature evolutions in VW7.

Figure10
Click here to download Figure: Figure10.docx

Figure 10 Heat generation rate of a cement paste with CEM I 42.5R and w/c=0.5.

Figure11
Click here to download Figure: Figure11.docx

a)

b)

Figure 11 a) Evolution of compressive and tensile strength of concrete; b) Basic creep of


concrete: specific creep for loading at the ages of 1, 3 and 7 days.

Figure12
Click here to download Figure: Figure12.docx

Figure 12 E-modulus of concrete assessed by compressive cyclic testing and EMM-ARM.

Figure13
Click here to download Figure: Figure13.docx

a)

b)

c)
Figure 13 a) Scheme of the simulation model and phasing; b) Section A-A of the model; c)
Finite element mesh

Figure14
Click here to download Figure: Figure14.docx

a)

b)

Figure 14 a) FE model for the semi-adiabatic calorimeter (geometry and general data / units in
millimetres); b) Calculated and recorded temperature in the geometrical centre of the calorimeter

Figure15
Click here to download Figure: Figure15.docx

a)

b)

c)

Figure 15 a) Measured and simulated temperatures for VW1, VW2, VW4 and VW5;
b) Temperature distribution within the thicker section of the wall for the instants that the
maximum temperatures are attained; b) Calculated temperatures in the points where the highest
temperatures are attained in the thicker section of the wall, for the computed model and for an
alternative model in which the cooling ducts were not considered.

Figure16
Click here to download Figure: Figure16.docx

Figure 16 Measured and simulated strains for VW1, VW2, VW4 and VW5

Figure17
Click here to download Figure: Figure17.docx

Figure 17 Principal tensile stresses at the point corresponding to the maximum tensile stress in
the wall (x=8.5m, y=0.36m, z=11.7m) for three simulation models

Das könnte Ihnen auch gefallen