revised engstruct d 13 00590r1

Upload: rodrigo-lameiras

Post on 03-Jun-2018

225 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/12/2019 Revised Engstruct d 13 00590r1

    1/63

    Elsevier Editorial System(tm) for Engineering Structures

    Manuscript Draft

    Manuscript Number: ENGSTRUCT-D-13-00590R1

    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; thermal 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 usingthe 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.

  • 8/12/2019 Revised Engstruct d 13 00590r1

    2/63

    Title of Paper:

    Application of air cooled pipes for reduction of early age cracking risk in a massive RC wall

    M. Azenha, R. Lameiras, C. de Sousa, J. Barros

    Ms. Ref. No.:ENGSTRUCT-D-13-00590

    Dear Prof. Herbert Mang:

    We have uploaded a revised version of the manuscript, where several modifications were introduced to

    attend the reviewers comments and criticisms.

    The file Letter_To_Reviewers is dedicated to the reviewers, explaining how their comments were

    addressed in the revised manuscript. The revised version of the manuscript is supplied in the file

    Revised_Manuscript (please note that the main changes associated to the revision are highlighted in

    yellow colour to facilitate the location of changes in regard to the original manuscript).

    I look forward to hearing from you soon.

    Yours sincerely,

    Miguel Azenha (the corresponding author)

    (Research Assistant)

    Miguel ngelo Dias Azenha, PhDSchool of Engineering University of MinhoCivil Engineering DepartmentCampus de Azurm, 4800-058 Guimares, PortugalTel: +351 253510248; E mail address:[email protected]

    University of MinhoSchool of Engineering

    Prof. Herbert A. Mang

    Editor of

    Engineering Structures

    Date: November 29, 2013

    tter To Editor

    ck here to download Cover Letter: Letter_To_Editor.doc

    mailto:[email protected]:[email protected]:[email protected]://ees.elsevier.com/engstruct/download.aspx?id=457898&guid=356417b5-6995-46ad-b077-6123a315f1c8&scheme=1http://ees.elsevier.com/engstruct/download.aspx?id=457898&guid=356417b5-6995-46ad-b077-6123a315f1c8&scheme=1mailto:[email protected]
  • 8/12/2019 Revised Engstruct d 13 00590r1

    3/63

  • 8/12/2019 Revised Engstruct d 13 00590r1

    4/63

    To Reviewer #3

    Concerning the remarks raised in the next comments (italic text), the following changes wereintroduced in the manuscript, or alternatively the following clarification comments are made:

    Page 10, lines 24 - 26: "The cooling system was ... to avoid inducing undesirable vibrations ...concrete." This remark from the authors is not accurate as the concrete have been pokervibrated during pouring.

    Thank you for your comment. Our remark was not accurate indeed. We have corrected thesentence in the revised version of the manuscript:

    The cooling system was only started at the age of 14h after the end of castingoperations to avoid introducing potentially undesirable vibrations to the freshly castconcrete before its structural setting time.

    Page 17, line 9: Comment why the 29-day strength was used instead of 28-day strength according

    to Standard testing (EN 206-1) is missing.There were limitations due to problems that occurred in the laboratory, which did not allowus to test the specimens at 28 days age. However, we tested in the 29 thday of age and,which is likely to provide us very similar results to those that would have been obtained at28 days. Nevertheless, the following clarification has been added to the manuscript:

    It is remarked that testing at the reference age of 28 days was not possible due tolaboratorial constraints.

    Page 18, line 8: Add reference "(Hedlund, H., (2001), Hardening Concrete. Measurements andevaluation of non-elastic deformation and associated restraint stresses. Doctoral Thesis2000:25, ISBN 91-89580-00-1.)" concerning autogenous shrinkage and evaluation ofdeformations. Results used in thermal stress calculations.

    The reference has been added.

    Page 19, line 13: Calculation with suggested geometrically symmetry will not be a good estimation.Authors should explain the consequences of the engineering approach.

    We believe that the reviewer may have misunderstood the drawings, owing to a mistake onour behalf in labelling the axes Z and Y in Figure 13b. We have corrected such mistake, andthe revised version of Figure 13 is shown below. Except for some details quite near the baseslab, the wall is geometrically symmetric along its middle plane. This holds true even withconsideration of the cooling ducts, which are also symmetric in regard to this plane. Interms of solar radiation intake, the wall is not symmetrical indeed. However, such lack ofsymmetry and the corresponding simplified approach was already dully justified at the end

    of section 4.1.1 of the original manuscript.

    As we believe that the reviewers comment may have been induced by a misunderstandingof our text and Figure 13, we have corrected the figure as mentioned above, and tried tomake our text clearer in the revised manuscript:

    Due to the geometrical symmetry of the wall, a longitudinal plane of symmetry isconsidered, identifiable by a ZX plane in Figure 13.

  • 8/12/2019 Revised Engstruct d 13 00590r1

    5/63

    a) b)

    c)

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

    Page 20, line 6: "... was equal to half actual diameter". This is an engineering approach that shouldbe explained better as the correct solution using the full diameter, but correcting the heattransfer coefficient of the tube by an reduction of 0,8.

    The consideration of half diameter for the tubes located in the symmetry plane was not anengineering approach. Please look at the figure below, were we identify the situation at across-sectional level. In fact, the real situation has half a tube in each side of the symmetryplane, which means that the modelling of the tube, which is made through a bar elementcoincident with the symmetry axis should only promote a heat transfer corresponding to

    half a perimeter of the tube. We have considered half diameter, which actually correspondsto a tube with half the perimeter. However, our text had some inaccuracies that may havebeen misleading: (i) we mentioned symmetry axis instead of symmetry plane; (ii) wementioned half a diameter instead of half perimeter for the tube in the symmetry plane.Therefore, in the revised manuscript, the corresponding text has been rephrased to:

    As consequence of the symmetry simplification, it was considered that theperimeter of the tube elements located in the symmetry plane of the model wasequal to half the actual perimeter of the tubes.

  • 8/12/2019 Revised Engstruct d 13 00590r1

    6/63

    All figures indication the time = 0 must be better explained what zero represents.

    All graphs in the manuscript have the time zeroed to coincide with the end of castingoperations of the studied phase of the wall. This clarification has been introduced in therevised manuscript upon the presentation of the first graph into which it is applicable (i.e.Figure 5):

    It is remarked that the results of this figure and all upcoming figures of the paper(either concerning experimental results or numerical simulation results) have theircorresponding time axis zeroed in regard to the instant at which the casting

    operations of the studied phase were finished.

    Figure 17: The authors should comment on the lower tensile strength tested on cylinders comparedwith the evaluated tensile stresses, which is higher than the assumed tensile strength.

    The results of Figure 17 in which the tensile strength is lower than the evaluated tensilestresses correspond to a hypothetical scenario of not having used the cooling ducts. Thecorresponding comment is made in the last paragraph of Section 4 and is replicated below:

    Interestingly, a simulation of the same construction situation without considerationof cooling ducts would yield to higher cracking risk at the same location (as seen inFigure 17), with the ratio between the tensile stress and the tensile strength ofconcrete reaching 1.2. Therefore, if the calculations made here are consideredtrustworthy, the use of the cooling ducts may have been the differentiating factorthat avoided a cracking scenario in this concrete lift.

    Yours sincerely,

    Miguel Azenha (the corresponding author)(University of Minho, PhD)

  • 8/12/2019 Revised Engstruct d 13 00590r1

    7/63

    Highlights

    Internal cooling of a massive concrete wall with ventilated prestressing ducts; 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.

    ghlights

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

    http://ees.elsevier.com/engstruct/download.aspx?id=457876&guid=6c2b2d78-2a58-416b-a813-622c14c54df0&scheme=1http://ees.elsevier.com/engstruct/download.aspx?id=457876&guid=6c2b2d78-2a58-416b-a813-622c14c54df0&scheme=1
  • 8/12/2019 Revised Engstruct d 13 00590r1

    8/63

    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-situmonitoring 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.

    bstract

    ck here to download Abstract: Abstract.docx

    http://ees.elsevier.com/engstruct/download.aspx?id=457877&guid=830e8d23-b16e-4928-8a6f-6b8b873ea714&scheme=1http://ees.elsevier.com/engstruct/download.aspx?id=457877&guid=830e8d23-b16e-4928-8a6f-6b8b873ea714&scheme=1
  • 8/12/2019 Revised Engstruct d 13 00590r1

    9/63

    - 1 -

    Application of air cooled pipes for reduction of early age cracking risk in1

    a massive RC wall2

    By Miguel Azenha1, Rodrigo Lameiras

    2, Christoph de Sousa

    3and Joaquim Barros

    43

    Abstract: The construction of massive concrete structures is often conditioned by the necessity4

    of phasing casting operations in order to avoid excessive heat accumulation due to cement5

    hydration. To accelerate construction and allow larger casting stages (usually increasing lift6

    height), it is usual to adopt internal cooling strategies based on embedding water pipes into7

    concrete, through which water is circulated to minimize temperature development. The present8

    paper reports the use of horizontally placed ventilated prestressing ducts embedded in a massive9

    concrete wall for the same purpose, in line with a preliminary Swedish proposal made in the10

    1990s. The application herein reported is a holistic approach to the problem under study,11

    encompassing extensive laboratory characterization of the materials (including a technique12

    developed for continuous monitoring of concrete E-modulus since casting), in-situmonitoring of13

    temperatures and strains, and 3D thermo-mechanical simulation using the finite element method.14

    Based on the monitored/simulated results, it is concluded that the air-cooling system is feasible15

    and can effectively reduce early cracking risk of concrete, provided adequate planning measures16

    are taken.17

    Keywords: Cement hydration, service life conditions, thermal shrinkage, cracking, numerical18simulation19

    1 INTRODUCTION20The combined effect of the exothermic nature of cement hydration reactions and the21

    relatively low thermal diffusivity of concrete leads concrete structures to endure temperature22

    1Assistant Professor, ISISE Institute for Sustainability and Innovation in Structural Engineering, University ofMinho, School of Engineering, Civil Engineering Dept., Azurm Campus, 4800-058 Guimares, Portugal, Phone:(+351)938404554, Fax: (+351) 253 510 217, E-mail:[email protected].

    2PhD 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:[email protected].

    3PhD 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:[email protected].

    4Full Professor, ISISEInstitute 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:[email protected].

    evised Manuscript

    ck here to download Manuscript: Revised_Manuscript.docx Click here to view linked References

    mailto:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]:[email protected]://ees.elsevier.com/engstruct/download.aspx?id=457915&guid=10e459c7-a690-40e2-8b3e-4f57c02a915f&scheme=1http://ees.elsevier.com/engstruct/viewRCResults.aspx?pdf=1&docID=10631&rev=1&fileID=457915&msid={746304AE-B738-4757-B211-08CB6A86EDFA}http://ees.elsevier.com/engstruct/viewRCResults.aspx?pdf=1&docID=10631&rev=1&fileID=457915&msid={746304AE-B738-4757-B211-08CB6A86EDFA}http://ees.elsevier.com/engstruct/download.aspx?id=457915&guid=10e459c7-a690-40e2-8b3e-4f57c02a915f&scheme=1mailto:[email protected]:[email protected]:[email protected]:[email protected]
  • 8/12/2019 Revised Engstruct d 13 00590r1

    10/63

    - 2 -

    increases at early ages, and eventually return to thermal equilibrium with the surrounding1

    environment. These early temperature variations induce volumetric changes in concrete that are2

    partially restrained by adjoining previously cast members, or even due to non-uniform3

    temperature distributions within a concrete member itself. Such restraint to deformation may4

    induce stresses that can be relevant enough to induce early age thermal cracking in concrete,5

    which is usually unacceptable in view of aesthetics, durability and even structural performance.6

    Contractors usually attempt to avoid this thermal cracking by adopting concrete compositions7

    and construction schedules that maintain temperature gradients in concrete below prescribed8

    limits, both along time and space [1]. It has however been recognized that such approach leads to9

    erroneous conclusions, as several important issues are disregarded [1], such as the degree of10

    restraint to deformation and the actual mechanical properties of concrete. In view of the11

    limitations of these temperature-based criteria, it has been widely acknowledged [2] that more12

    realistic crack risk assessments can be made through multi-physics approaches that encompass13

    numerical simulation of temperatures and corresponding stresses in concrete: thermo-mechanical14

    analyses. The use of thermo-mechanical simulation models allows the evaluation of alternative15

    construction scenarios (for casting procedures, concrete mixes, environmental conditions), and16

    thus permits the optimization of construction without compromising the safety in regard to17

    thermal cracking. The numerical studies for the assessment of the optimum construction strategy18

    frequently involve diminishing the temperature rise in concrete at early ages. In fact, if the19

    thermal variation is diminished, the corresponding volumetric changes also decrease, as well as20

    the developed stresses. The diminishment of early temperature rises is usually achieved by21

    partial replacement of cement by additions as fly ash [3, 4], or by cooling water/aggregates22

    before mixing operations[5-7], or even by introducing internal cooling pipes in concrete with23

    cooling fluids (usually water) [8-12]. An attempt to use air as the cooling fluid in cooling pipes24

    has been made in the 1990s by Hedlund and Groth [8, 9], who have shown the feasibility of25

    such technique in thick columns. Nonetheless, no further application of such technique was26

  • 8/12/2019 Revised Engstruct d 13 00590r1

    11/63

    - 3 -

    found in the literature, except for an in-situuse of air for localized cooling reported by Ishikawa1

    et al.[12].2

    Even though thermo-mechanical simulation approaches have been applied to concrete3

    structures for decades [13-17], they involve several complexities and problematic issues4

    particularly in view of the assessment of material properties and model parameters (thermal and5

    mechanical), which frequently demand specific laboratory characterization: heat of hydration,6

    thermal boundary coefficients, creep of concrete, evolution of E-modulus and tensile stress,7

    among others. Even though several scientific works have been done either on thermal simulation8

    [18], thermo-mechanical simulation [19]or monitoring the concrete behaviour at early ages [20-9

    23], some combine the numerical simulation with experimental data obtained in laboratory [24-10

    26]or with temperature monitoring for partial validation [27-31], whereas others go further and11

    additionally include in-situ strain monitoring for validation [32]. Nonetheless, in the scope of12

    internal cooling of concrete through the use of embedded pipes, no works were found to adopt13

    holistic approaches that simultaneously include material characterization, in-situmonitoring of14

    temperature/strain and thermo-mechanical simulation. The works that focus on concrete cooling15

    with embedded pipes are mostly limited to thermal [33], or thermo-mechanical analyses only16

    [34-37], and thermal [38-40] or thermo-mechanical analyses together with partial validation17

    through in situmonitoring [12,41].18

    The present paper pertains to a case study of the thermal stresses in the central wall of the19

    entrance organ of a dam spillway. Such wall is 27.5m long, with a maximum width of 2.8m and20

    height of 15.0m, with attention being given to the most unfavourable construction phase in21

    which a 2.0m tall batch had to be made (total 150m3 of concrete). Due to the materials and22

    equipment available at the construction site, it was decided to attempt internal cooling of23

    concrete with air-cooled prestressing ducts placed longitudinally along the wall.24

    The present paper regards to the in-depth study of the early age performance of concrete in25

    the wall, encompassing laboratory thermal and mechanical characterization of concrete, as well26

  • 8/12/2019 Revised Engstruct d 13 00590r1

    12/63

    - 4 -

    as in-situ monitoring of temperatures/strains and the corresponding thermo-mechanical1

    simulation with the finite element method.2

    The extensive laboratory characterization included quantification of the heat of hydration3

    evolution (isothermal and semi-adiabatic calorimetry), evaluation of compressive/tensile strength4

    and E-modulus through cube/cylinder testing, creep testing at several ages and assessment of5

    mechanical activation energy. In particular regard to E-modulus testing, a methodology that6

    allows continuous measurement of concrete E-modulus since casting (EMM-ARM [42]) was7

    applied. This is a pioneering use of this methodology for the purpose of supporting stress8

    simulation on concrete since its very early ages, with important advantages in regard to previous9

    approaches that tend to extrapolate values of E-modulus at very early ages.10

    A relatively complete monitoring program has been carried out in-situ, involving the use of11

    20 temperature sensors and 7 vibrating wire strain gauges embedded in concrete. Particular12

    attention was given to the evaluation of the effectiveness of the cooling system, with13

    temperatures being measured at several points along the prestressing ducts, and with air velocity14

    measurements taken with handheld anemometers.15

    Bearing in mind the information gathered with the laboratory characterization and in-situ16

    monitoring, a thermo-mechanical simulation was carried out with recourse to a three dimensional17

    finite element model. Such simulation model included the explicit modelling of the cooling18

    ducts, as well as the phased construction of the wall. The simulation model was made with19

    DIANA software [43].20

    21

    2 THERMO-MECHANICAL MODEL22The thermo-mechanical simulation approach presented here has strong similarities with that23

    described in a previous work [30]. Nonetheless, some particularities are distinct, namely: (i)24

    solar radiation is explicitly considered according to a model based on the incidence angle of the25

  • 8/12/2019 Revised Engstruct d 13 00590r1

    13/63

    - 5 -

    sun beams; (ii) the effect of internal cooling ducts is taken into account (iii) the evolution of1

    mechanical properties is simulated according to the equivalent age concept (instead of the2

    degree of hydration concept). The following sub-sections pertain to a general description of the3

    modelling strategy with specific emphasis on topics (i) to (iii) mentioned above.4

    2.1 Thermal model5The calculation of temperature fields in concrete is based on the heat balance equation,6

    whose solution is made through the finite element method [44]:7

    k T Q cT (1)

    where kis the thermal conductivity, cis the volumetric specific heat and Tis the temperature.8

    Q is the volumetric heat generation rate due to cement hydration, formulated as an Arrhenius9

    type law [45]:10

    RT

    Ea

    efAQ

    )( (2)

    where A is a rate constant, Ea is the apparent activation energy, is the degree of heat11

    development (ratio between the heat Qreleased up to time tand the total heat Qfinalreleased upon12

    completion of cement hydration),R= 8.314 Jmol-1K-1is the Boltzmanns constant and f() is a13

    normalized function for heat.14

    Thermal boundary conditions are applied through a prescribed flux per unit area qT15

    formulated as [46]:16

    T cr b eq h T T (3)

    where hcr is a mixed convection-radiation boundary transfer coefficient, Tb is the boundary17

    surface temperature and Teis the environmental temperature.18

    The simulation of thermal inputs associated to solar radiation in concrete structures can be19

    made with significant accuracy through the adoption of models that are readily used in20

    meteorological sciences [31, 47]. Such models can take into account the effects of the spatial21

  • 8/12/2019 Revised Engstruct d 13 00590r1

    14/63

    - 6 -

    relationship between the earth and sun at a given time of the day/year and thus predict the solar1

    radiation that reaches a certain point on earth at sea level (i.e. low atmosphere). It is further2

    possible to compute the angle between the sunbeam and any arbitrarily oriented/inclined surface,3

    and evaluate the intake of energy throughout the day of such surface.4

    The calculation of the solar energy that reaches earth at sea level, qm, is based on the solar5

    constant, q0, which represents the total radiation energy received from the sun at a distance6

    corresponding to 1 Astronomical Unit. Even though q0 varies slightly throughout the year by7

    ~7%, it is usually acceptable to consider q0=1367 Wm-2. The estimation of qm can be done8

    through the following empirical equation [47, 48]:9

    )sin(4.99.0

    0

    h

    T

    m

    l

    eqq

    (4)

    where Tl is the Linke turbidity factor that summarizes the turbidity of the atmosphere10

    (attenuation of the direct beam solar radiation) and his the solar elevation that corresponds to the11

    angle between the direction of the sunbeam and the idealized horizon. Tl is known to usually12

    vary between 3 and 7, whereas hcan be calculated by taking into account latitude, date and time.13

    Further geometrical considerations allow the calculation of the angle between an incident14

    sunbeam and the vector orthogonal to an arbitrarily oriented/inclined surface, termed as i (see15

    detailed description of models to calculate hand iin [44, 47]).16

    Based on the knowledge of qmand iat a given instant, and considering the absorvity of the17

    material of the target surface (S), it is possible to calculate the radiation energy qs that is18

    actually absorbed:19

    )cos(iqq mSS (5)

    Another particularity of the present application in regard to previous works [30]is the use of20

    prestressing ducts acting as cooling pipes. A formulation is thus necessary to describe the added21

    internal heat fluxes that are caused by the presence of an embedded cooling pipe, which can be22

  • 8/12/2019 Revised Engstruct d 13 00590r1

    15/63

    - 7 -

    expressed in the following energy balance equation, to be applied throughout the length zalong1

    the pipe [49]:2

    cp c c wdT

    m c h P T T dz

    (6)

    where m is the mass flow rate of the coolant (air in this case), cp is the specific heat of the3

    coolant, hcis the boundary transfer coefficient between the coolant and the surrounding concrete,4

    Pis the perimeter of the cooling pipe, Tcis the temperature in the cooling pipe and Twis the bulk5

    temperature of concrete around the cooling pipe. The mass flow rate of the coolant m can be6

    estimated through the product of the cooling fluid density () by the fluid mean velocity (m) and7

    by the cross sectional area of the cooling pipe (Ac). The implementation of equation (6) into a8

    finite element software [43] brings further nonlinearities due to the interaction between the9

    cooling fluid and the surrounding concrete, which results in progressive heating of the cooling10

    fluid along the pipe.11

    For updating age-dependent properties along time in the mechanical model, the equivalent12

    age of concrete teqis adopted. Its formulation is based on an Arrhenius type equation established13

    for a reference temperature Tref (usually 293.15K) [50]. For a given instant t, the equivalent age14

    can be calculated as:15

    dett

    TTR

    E

    eq

    ref

    a

    0

    11

    (7)

    2.2

    Mechanical model16The mechanical model is relatively similar to those adopted for time-dependent mechanical17

    analysis of hardened concrete, except for some particularities associated to the facts that: (i) it is18

    being preceded by a thermal analysis (de-coupled), with imposition of strains that are calculated19

    with basis on the thermal dilation coefficient of concrete (T) and the previously calculated20

    temperature field; (ii) there is a strong evolution of mechanical properties throughout the21

    analysis, dully taken into account through the equivalent age concept; (iii) the strong viscoelastic22

  • 8/12/2019 Revised Engstruct d 13 00590r1

    16/63

    - 8 -

    behaviour of concrete at early ages makes it necessary to use creep formulations that can provide1

    adequate estimates within such time span. In specific regard to the last point (iii), basic creep of2

    concrete was accounted for through the use of the Double Power Law (DPL), which has a3

    reasonably good performance on both early age and long term time spans [51]:4

    nm ttttEtE

    ttJ ),

    (),

    (),

    (),

    (

    1),

    ,(

    0

    1

    0

    (8)

    where ),,( ttJ is the compliance function at time t for a load applied at instant,

    t , ),(0 tE is the5

    asymptotic elastic modulus, and 1, m and n are material parameters. Since drying creep is6

    negligible for an application that only envisages early age behaviour, it was disregarded [52].7

    Bearing in mind that the aim of the thermo-mechanical simulations is to assess the risk of8

    cracking, the post-cracking behaviour is not considered relevant and it is thus not simulated.9

    Thus, linear elastic behaviour (with creep) is considered for concrete both in compression and in10

    tension.11

    3 THE PARADELA DAM SPILLWAY: DESCRIPTION AND MONITORING123.1 Overview13

    The Paradela dam, located in the North of Portugal, is a rockfill gravity dam built in the14

    1950s,with 540m longitudinal development and maximum height of 112m from foundation.15

    Due to recent hydraulic problems in one of the dams spillways, it was necessary to build a new16

    complementary ski-jump spillway on the right margin of the river[53]. The case study reported17

    in this paper concerns the cooling measures and assessment of cracking risk in the construction18

    of the central wall of the spillway entrance.19

  • 8/12/2019 Revised Engstruct d 13 00590r1

    17/63

    - 9 -

    3.2 Description of the spillway entrance13.2.1 Geometry and construction phasing2

    The spillway functions in free surface conditions, and has two main entrances at the top level3

    of the dam, each with 5.5m width, being separated by a hydro-dynamic shaped wall with4

    maximum width of 2.8m and 17.4m height. A three dimensional representation of the entrance5

    region of the dam spillway is shown in Figure 1a, whereas its corresponding plan view at6

    approximately mid-height of the wall is depicted inFigure 1b. The reinforcement of the middle7

    wall can be generally characterized by 16//200mm placed vertically and 12//200mm placed8

    horizontally near each surface with a concrete cover of 60mm.9The construction of the wall was generally performed with 1.2m tall construction phases,10

    with empirically defined waiting periods being defined by a target temperature in the core11

    regions during the cooling period (approximately 27C, which corresponded to 17C above12

    average daily temperature during construction). In order to minimize such waiting periods, an13

    air-cooling system based on ventilated prestressing ducts placed horizontally was implemented,14

    allowing lower peak temperatures and faster return to temperature equilibrium with the15

    surrounding environment. The main scope of the present paper is the study of a specific16

    construction phase that corresponds to the zone of embedment of the fixation parts of the sluice17

    gates. Such fixation parts were approximately 2.5m tall, and it was thus desirable to perform a18

    2.5m tall construction phase, labelled as 9thphase inFigure 2a. Due to its larger thickness, this19

    construction phase is the critical one in terms of peak temperatures and cracking risk, being20

    therefore the object of analysis.21

    3.2.2 Materials22The wall of the spillway entrance was generally cast using concrete of class C30/37 [54]with23

    the composition labelled as S1-D32 inTable 1. In the 9thconstruction phase, due to increased24

    complexity of reinforcement near the downstream extremity of the wall (related to the salient25

  • 8/12/2019 Revised Engstruct d 13 00590r1

    18/63

    - 10 -

    concrete blocks), two slightly different compositions with higher fluidity were used in the1

    vicinity of such region, as shown inTable 1 (S3-D32 and S3-D16). In spite of this, the areas of2

    most interest to this study (thickest regions of the wall and monitored sections) correspond to the3

    upstream region. Therefore, and also taking into account the fact that the compositions have4

    similarities, all the characterizations and modelling in the scope of this work pertain to mix S1-5

    D32. Steel reinforcement was S400C [54], with characteristic yield stress of 400MPa.6

    3.2.3 Cooling system7In view of the work reported by Hedlund and Groth [8, 9], which proposed the possibility of8

    using ventilated prestressing ducts for concrete cooling, and bearing in mind the easy availability9

    of the corresponding necessary equipment in the construction site of the wall, it was decided to10

    test the feasibility of this kind of cooling technique. However, in view of practical limitations11

    posed by contractor/owner, this pilot application of air-cooling system was slightly different12

    from that of Hedlund and Groth [8, 9]. Instead of placing the tubes vertically along the wall, they13

    were placed horizontally, even though this implied more limited cooling capacity as the length of14

    tube along freshly cast concrete is longer. In the particular case of the 9thconstruction phase, a15

    total of 6 prestressing ducts of 90mm diameter have been used, with their air intake being made16

    horizontally at the downstream extremity of the wall, and the outtake made near the upstream17

    extremity, on the top surface of the casting phase, in order to avoid a direct upstream-18

    downstream potential leakage channel after construction. The overall path of the ducts is shown19

    in the schemes of Figure 33, with the ducts labelled from T1 to T6. For ventilation, a 0.60m20

    diameter fan was used, with 1200m3/h ventilation capacity, that collected air from the21

    environment and blowed it into the ducts at an internal air speed of approximately 8.6m/s22

    (measured with anemometer at the outtake of the ducts). The fan had to be placed in the23

    downstream extremity of the wall due to practical constraints of the contractor. The cooling24

    system was only started at the age of 14h after the end of casting operations to avoid introducing25

    potentially undesirable vibrations to the freshly cast concrete before its structural setting time.26

  • 8/12/2019 Revised Engstruct d 13 00590r1

    19/63

    - 11 -

    The ventilation system was disconnected 8.6 days after casting in view of the similarity of1

    temperature between the walls core and the surrounding environment. After that, the2

    prestressing ducts were filled with mortar using standard procedures [55].3

    It is remarked that this cooling system had been previously tested in the 8th

    phase of casting,4

    with three prestressing ducts placed at mid-height. Details on this test can be found elsewhere5

    [56].6

    3.3 Monitoring and material characterization73.3.1 General remarks8

    In order to better understand the effectiveness of the cooling system, its influence on the9

    cracking risk, and assess the capabilities of the adopted numerical simulation strategy, an10

    extensive set of actions has been carried out, comprising in-situ internal monitoring of11

    temperatures/strains of concrete, in-situ validation tests, as well as laboratory material12

    characterization.13

    3.3.2 Temperature monitoring14Temperatures inside the 9th construction phase have been monitored with K-type15

    thermocouples, aiming particularly at assessing temperature profiles in a region near the16

    maximum width of the wall (i.e. at approximately 5.3m from the upstream extremity: section A-17

    Aas identified inFigure 3a). The placement of temperature sensors in section A-Ais depicted18

    in the scheme ofFigure 4a, where thermocouples are identified by the prefix TC. Temperatures19

    in the locations labelled as VW inFigure 4ahave also been monitored with resistive temperature20

    sensors, as these are the locations of vibrating wire strain gauges, which contain internal21

    temperature sensors for strain compensation. The internal air temperature of ducts T1-T3 has22

    been monitored both in section A-A and in neighbouring areas, as shown in Figure 4b.23

    Environmental temperature (dry-bulb) has been assessed with a thermocouple.24

  • 8/12/2019 Revised Engstruct d 13 00590r1

    20/63

    - 12 -

    Monitoring was carried out since the instant of casting during a period of 10 days, and the1

    measurement frequency was set to 1 reading per each 30 minutes. The internally monitored2

    temperatures in concrete are shown in Figure 5 for a vertical and a horizontal alignment of3

    sensors that pass through sensor VW3. It is remarked that the results of this figure and all4

    upcoming figures of the paper (either concerning experimental results or numerical simulation5

    results) have their corresponding time axis zeroed in regard to the instant at which the casting6

    operations of the studied phase were finished. From Figure 5 it can be seen that the initial7

    temperature of concrete was ~15C, and the peak temperature was approximately 42C in the8

    core regions (VW3 and VW5). Furthermore, it can be observed that the ascending branch of9

    temperature development is clearly affected at the age of 14 hours, when the cooling system is10

    activated. In specific regard to the vertical profile of temperatures shown in Figure 5a, the11

    expectable behaviour was captured: the core region has the highest peak temperatures (VW3,12

    VW5), whereas a decrease trend is seen towards the top surface. In fact, sensors TC6 and TC713

    exhibit maximum temperatures of ~36C, while VW6 (near the top surface) has the lowest peak14

    temperature (circa 27C). Near the bottom surface of this construction phase, sensor VW115

    highlights the importance of the heat storage effect caused by the previously cast concrete: in16

    fact, even though the temperature peak is lower than that of the core regions, it occurs later and17

    the heat loss rate observed afterwards is lower than in other regions. It should also be remarked18

    that all sensors are almost in equilibrium with environmental temperature by the age of 8 days.19

    In regard to the temperature development in the sensors located along a horizontal alignment,20

    shown in Figure 5b, it can be noticed that the sensors located in the vicinity of vertical21

    boundaries exhibit lower temperature variations (VW4 and TC5), with temperature peaks under22

    35C. It is interesting to observe that TC5 is placed nearer the surface (15cm apart) than the case23

    of VW4 (20cm apart), and consequently the temperature peak of VW4 is slightly higher. By24

    looking at the temperature evolution after the age of 8 days (heat of hydration has been25

    dissipated), it can be seen that temperatures in VW4 remain higher than those of TC5 due to a26

  • 8/12/2019 Revised Engstruct d 13 00590r1

    21/63

    - 13 -

    combination of two main reasons: VW4 is located in the vicinity of a surface oriented to1

    southeast, which is bound to receive more energy through solar radiation than the surface near2

    TC5, which is oriented to northwest; VW4 is 5cm deeper than TC5, thus being slightly nearer3

    the inner core (higher thermal inertia).4

    The temperature evolution along the air inside duct T3 measured by sensors TC3, TC9 and5

    TC10 (Figure 4b) is shown in Figure 6, where the environmental temperature and the6

    temperature in VW5 (hottest region in concrete) are also represented for comparative purposes.7

    The information provided by such figure allows the clear identification of the instant at which air8

    ventilation began (14 hours), as the rate of temperature rise is clearly disrupted inside the duct.9

    Furthermore, the rising temperature tendency along the ducts length is identifiable, as the10

    temperature is consistently higher in TC3 in regard to TC10, and in TC9 in regard to TC3.11

    Taking as example the temperatures recorded at the instant of peak temperature (1.88 days), TC912

    measured a temperature of 32.2C, whereas TC10 indicated a temperature of 29.4C. This13

    represents a shift in temperature of approximately 3C in 3.5m length of duct. At three instants of14

    this study, temperatures were measured also at the entrance of T3 (x=0) and at x=1m through the15

    use of a handheld temperature probe (PT1000). By joining such data with the results of TC3,16

    TC9 and TC10, it was possible to plot a temperature profile along the duct for t=3.84d, t=4.56d17

    and t=6.58d see Figure 6b. It can be observed that the temperature at the inlet of the tube18

    matches the environmental temperature, and that the heating of the air along the tube is strongly19

    dependent on the combination of environmental temperature and internal temperature of20

    concrete. In fact, in the most unfavourable situation shown by Figure 6b, air was heated from21

    ~11C at the air inlet to ~26C at a point located 25m away from the air inlet (t=4.56d). This22

    increase of air temperature is bound to reduce its cooling capacity. However, in spite of such23

    diminishment of cooling capacity, the temperature of the air in the hotter regions of the duct24

    remained at least 10C below that of concrete during the periods at which temperature in25

  • 8/12/2019 Revised Engstruct d 13 00590r1

    22/63

    - 14 -

    concrete was near its peak (see t=1d to t=3d in Figure 6a), showing that the heat removal1

    potential was not negligible at all.2

    The observed diminishment of cooling capacity along the length of the duct highlights the3

    fact that the adopted configuration for the tubes does not maximize cooling capacity, which4

    would be conversely maximized if the length of tube inside concrete had been minimized. Such5

    goal could have been achieved by providing a vertical arrangement for the tubes and introducing6

    more individual smaller tubes.7

    3.3.3 Strain monitoring8Strain measurement was carried out with vibrating wire strain gauges of metallic casing with9

    14cm reference length (TES/5.5/T Gage Technique). Past laboratory tests and in-situ10

    applications [44,57, 58]have shown that this kind of sensor is robust and adequate for strain11

    measurement in concrete at early ages. The strain gauges were placed at the locations identified12

    in Figure 4a (VW1 to VW6), dully positioned in order to measure strains in the longitudinal13

    direction of the wall. VW7 has distinct intents and shall be specifically addressed later.14

    Measurements were taken with the same datalogger and at the same sampling rate as adopted for15

    the temperature sensors.16

    One important issue to tackle is the zeroing of the measured strains. In fact, before concrete17

    sets and has enough stiffness to drive the sensor into the same deformation state, the18

    measurements taken by the sensor do not have any relevant physical meaning. It is thus19

    necessary to assess the instant of solidarization (i.e. the full bond) between concrete and the20

    sensor. In a previous work [57], the zeroing operation has been made by assessing the instant at21

    which two sensors with different casing (plastic and metallic), placed under the same conditions22

    inside concrete, started yielding the same results. This means that both sensors are solidarized (as23

    the plastic sensor is bound to solidarize earlier due to its smaller stiffness). Since the plastic24

    cased sensor was not available, an alternative methodology for zeroing the data was25

    implemented. By interpretation of the findings reported in [57] it can be considered that the26

  • 8/12/2019 Revised Engstruct d 13 00590r1

    23/63

    - 15 -

    solidarization instant coincides with a progressive change in the derivate of strain variation1

    detected by the sensor, that can be obtained by geometrical intersection of tangents of measured2

    strains, as shown inFigure 7.It was decided to use such zeroingcriterion in the scope of this3

    research work. As a result of the application of such rule, the solidarization instant of each sensor4

    VW1 to VW5 was, respectively: 0.17, 0.19, 0.19, 0.20 and 0.17 days (due to malfunctioning of5

    the VW6, the strain results from this sensor are not available). The solidarization instants seem6

    coherent, since they have a trend to increase with the distance of the sensor from the bottom7

    surface of the casting block, due to the natural delay in its involvement by concrete during the8

    casting process.9

    The measured strains in sensors VW1 to VW5 are shown inFigure 8.Even though the strain10

    output is dependent on several factors that interact with each other (thermal deformation,11

    restraint, creep), it is possible to find a set of common points and reasoning. Overall, all12

    deformations are strongly commanded by the temperature variation, following the same kinetics.13

    Sensors VW2, VW3 and VW5, which are located in regions near the core of the walls cross14

    section and had similar temperature development histories, also have similar strain15

    developments. This is bound to be caused by similar thermal deformations and restraints for16

    these locations of measurement. The smallest deformations are recorded in VW1, which is17

    located in the bottom of the casting phase (i.e. near the existing concrete) and thus having less18

    temperature rise (thus less expansion), while being more restrained by the existing concrete19

    below at lower temperatures. Finally VW4, which is near the surface and thus has lower20

    temperature rises (maximum temperature of ~33C), also has a smaller deformation variation21

    when compared to VW2, VW3 and VW5.22

    In order to assess free deformations of the concrete used in the construction (associated to23

    unrestrained autogenous shrinkage and thermal deformations), strain was measured in a concrete24

    cylinder placed in specially devised conditions. The concrete cylinder (150mm diameter and25

    300mm tall), cast simultaneously with the studied construction phase and by using the same26

  • 8/12/2019 Revised Engstruct d 13 00590r1

    24/63

    - 16 -

    concrete, was cast into a special mould internally coated with a soft membrane (with lids also1

    coated with such material). A strain gauge was placed inside the mould to measure longitudinal2

    strainssee photo of the open mould inFigure 9a. After casting concrete into such mould, it was3

    placed horizontally inside the studied construction phase (during its casting procedures) at the4

    location that is identified as VW7 inFigure 4a. This kind of procedure, here termed as use of a5

    no-stress specimen, has been reported in Choi et al. [59], and it allows to measure free6

    deformations of concrete, which can in turn be used to assess the thermal dilation coefficient,7

    TDC (provided that the temperature inside the concrete cylinder is relatively uniform, and8

    autogeneous shrinkage deformations are known). Unfortunately, due to undetermined causes, the9

    output of the sensor could not be read during the first 0.8 days, and thus the reported data only10

    starts at such age, as shown inFigure 9b.11

    3.3.4 Heat generation and activation energy12In an extensive experimental program for the characterization of the cements marketed in13

    Portugal, Azenha [44] has reported a library of heat generation obtained through isothermal14

    calorimetry under several temperatures for plain cement pastes with w/c=0.5. The cement used in15

    the construction concerned in this paper was also characterized (same supplier and16

    manufacturing plant), and the resulting information for calorimetry tests under 20C, 30C, 40C17

    and 50C is shown inFigure 10.A reasonable estimate of the heat generated by concrete can be18

    obtained by multiplying the heat generation reported inFigure 10by the volumetric content of19

    cement, which is of 224kg/m3

    . By using the speed method algorithm [44,60], the necessary data20

    for the numerical simulation of heat generation according to equation 2 was obtained:21

    Ea= 37.31 kJ/mol, A = 4.989109W/m3, Qpot= 8.295107J/m

    3, and function f() characterized22

    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;23

    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],24

    [1.00; 0.00]. Even though this data pertains to CEM I 42.5R of the same company that supplied25

    the cement to this construction site, there may be deviations caused by inevitable variations in26

  • 8/12/2019 Revised Engstruct d 13 00590r1

    25/63

    - 17 -

    the characteristics of the cement. Also, the extrapolation procedure mentioned above did not take1

    into account the presence of fly ash in the mix (96kg/m3), which may have non-negligible effects2

    on the heat generation potential and hydration kinetics. Therefore, in order to assess the potential3

    importance of such deviations, a semi-adiabatic test was conducted in-situ (simultaneously with4

    the casting operations) in a 30cm edge concrete cube, duly isolated by 2.1cm thick plywood and5

    12cm of polystyrene boards. The results of such semi-adiabatic calorimetry test shall be6

    addressed in section 4, upon the simulation of its temperature development through the finite7

    element method.8

    3.3.5 Complementary laboratory characterization9Compressive strength evolution was assessed with concrete cubes (150mm edge) at the ages10

    of 1, 3, 7 and 29 days, whereas tensile strength was measured with splitting tests on cylinders11

    (150mm diameter and 300mm tall) and at the same ages. It is remarked that testing at the12

    reference age of 28 days was not possible due to laboratorial constraints. The evolution of both13

    tensile and compressive strength for concrete cured at 20C (saturated conditions) is shown in14

    Figure 11a (average results of three specimens at each age). In order to assess the activation15

    energy suitable for compressive strength maturity estimations, a set of concrete cubes was cured16

    at 40C with the compressive strength measured at the same ages. The corresponding results are17

    also shown in Figure 11a. By applying the equivalent age concept [61], together with the18

    superposition method [60], it was possible to asses that the activation energy based on19

    mechanical testing has the value of 37 kJ/mol, which is rather consistent with the activation20

    energy obtained through isothermal calorimetry for the same cement (yet without fly ash),21

    37.31 kJ/mol [44]. Such coincidence in activation energy for thermal and mechanical phenomena22

    had already been reported by Ulm and Coussy [62].23

    Basic creep was assessed in creep rigs on prismatic specimens (sealed) with dimensions24

    15cm15cm60cm, loaded at 30%~40% of the concrete compressive strength and internally25

  • 8/12/2019 Revised Engstruct d 13 00590r1

    26/63

    - 18 -

    monitored with vibrating wire strain gages. Such creep tests were conducted at the ages of 1.1,1

    3.3 and 7.3 days, and the corresponding specific creep curves are shown inFigure 11b.2

    The experimental program also included a single specimen for measurement of autogenous3

    shrinkage. Such specimen was a 150mm diameter and 300mm long cylinder, internally4

    instrumented with a vibrating wire strain gage, which was kept in its formwork during the5

    experiment and sealed with a plastic film on the top surface. Unfortunately, two factors6

    contributed to render the results of this specimen unusable for this research: on one hand, the7

    monitoring only could be started at the age of 2 days in the laboratory due to unavailability of8

    datalogging system; on the other hand, the measurements of autogenous taken since t=2 days9

    were disturbed by an inefficient sealing, which promoted undesired drying of the specimen. This10

    was not considered a critical problem in view of the low values of autogenous shrinkage that are11

    usually expectable in concretes of low cement content and high w/c ratio [63-65].12

    3.3.6 Continuous monitoring of concrete stiffness13The evolution of elasticity modulus along time was measured through compressive cyclic14

    testing in concrete cylinders (150mm diameter and 300m tall) at the ages of 1, 3, 7, 15 and 2915

    days. Concomitantly, E-modulus of concrete was continuously assessed through a methodology16

    termed as EMM-ARM (Elasticity Modulus Measurement through Ambient Response Method).17

    This methodology has been developed by Azenha et al.[42]and consists in a variant to classic18

    resonant frequencies that allows the quantification of E-modulus continuously since the instant19

    of casting of the specimen inside the testing mould. The basic principle of EMM-ARM is the20

    following: the specimen is cast inside the testing mould, which is in turn placed in simply21

    supported conditions and continuously subject to modal identification (using accelerometers)22

    without any explicit excitation of the beam, as ambient vibration suffices. As concrete hardens,23

    the first resonant frequency of the composite beam evolves, and the stiffness of concrete can be24

    inferred by applying the equations of motion. Details about the testing setup and procedure25

    applied for the concrete of this spillway application can be found in [56, 66, 67], as it26

  • 8/12/2019 Revised Engstruct d 13 00590r1

    27/63

    - 19 -

    corresponds to an improved version of the originally devised test (a steel mould is used). The1

    collected results with EMM-ARM and cyclic compression tests on cylinders are shown inFigure2

    12,where the feasibility of EMM-ARM is confirmed in view of the resemblance of results. Also,3

    the richness of information that can be obtained through EMM-ARM represents an added value4

    for the numerical simulation.5

    4 NUMERICAL MODELLING64.1 Geometry, mesh, materials, initial/boundary conditions and time integration74.1.1 Geometry of the model and finite element mesh8

    A cross-sectional scheme of the model for simulation is shown in Figure 13a, where the9

    construction stages considered in the analysis can be observed. The first stage of the model10

    encompasses all concrete until the 7thphase of concreting (inclusive), considered as hardened11

    concrete, together with the 8thphase of concrete evaluated as freshly cast concrete. The second12

    and third stages correspond to the 9 thand 10thphases of concreting respectively. This strategy13

    diminishes the computational cost of the model without relevant effect on the accuracy of results.14

    For similar reasons, the underlying subgrade is not modelled, as it is far from the construction15

    phases of interest. Due to the geometrical symmetry of the wall, a longitudinal plane of16

    symmetry is considered, identifiable by a ZX plane inFigure 13.17

    The simulation was made with a 3D finite element model comprising rectangular brick FE of18

    8 nodes (222 integration scheme) for concrete in the thermal model, and coincident 20 nodes19

    brick FE (333 integration scheme) in the mechanical analysis. Convective boundaries were20

    modelled with 4 node planar elements (22 integration scheme), and the cooling ducts were21

    considered with linear elements of 2 nodes (2 point integration scheme) [43]. The schematic22

    representation of geometry, casting phases and cooling duct location for section A -A is shown23

    in Figure 13b. The reader is reminded that phase 8 also had cooling ducts, according to the24

  • 8/12/2019 Revised Engstruct d 13 00590r1

    28/63

    - 20 -

    description of section 3.2.3.The longitudinal layout of all the ducts was considered straight on1

    an horizontal plane, in correspondence to simplifications in the vicinity of the extremities of the2

    wall.3

    It should be remarked that due to the phased analysis, the mesh evolved along time with4

    some elements/boundaries being activated/de-activated (e.g. the convective top boundary of a5

    given phase is de-activated upon the beginning of the next casting phase). The mesh adopted for6

    this simulation is shown in Figure 13c, with a total of 18738 elements and 66037 nodes. As7

    consequence of the symmetry simplification, it was considered that the perimeter of the tube8

    elements located in the symmetry plane of the model was equal to half the actual perimeter of the9

    tubes. As solar radiation does not represent a symmetrical energy input to the structure, the10

    symmetry simplification adopted is not truly valid. However, as solar radiation has most of its11

    effect near the surface, it was decided to keep the symmetry simplification by considering the12

    southeast half of the wall, which is most subject to solar radiation effects.13

    4.1.2 Materials (thermal and mechanical properties)14The thermal conductivity and specific heat of concrete were estimated with basis on the15

    pondered average of the corresponding thermal properties of the constituent materials of the mix16

    [44, 68]. The adopted values for k and c for concrete were, respectively 2.40 W/mK and17

    2.4106 J/m3K. Even though it is known that these thermal properties suffer variations during18

    early ages [69-75], the adopted modelling approach considers them constant in view of the19

    conclusions of the parametric analyses reported by Azenha [44], where a relatively small impact20

    of considering evolving kand cwas found on computed temperatures in hardening concrete.21

    In regard to the data for heat of hydration, the parameters mentioned in section3.3.4 were used.22

    The adequacy of these parameters was evaluated through the semi-adiabatic calorimeter23

    described in the same section, whose behaviour was simulated through a FE simulation model24

    that explicitly considered the extruded polystyrene (XPS) and wood walls of the calorimeter. The25

    material modelling parameters for concrete coincide with those herein described, whereas26

  • 8/12/2019 Revised Engstruct d 13 00590r1

    29/63

    - 21 -

    additional information and mesh representation are shown in Figure 14a. The results of the1

    simulation of the calorimeter were quite coherent with those collected experimentally, as seen in2

    Figure 14b, leading to the confirmation that the strategy described in section3.3.4 to determine3

    the heat generation and activation energy was adequate for defining the heat of hydration4

    modelling parameters adopted in the simulations.5

    The thermal dilation coefficient (TDC) of concrete was assessed with basis on the no-stress6

    specimen described in section 3.3.3. However, in order to obtain the thermal deformation and7

    calculate the TDC, it was necessary to subtract the autogenous shrinkage deformation from the8

    total deformation. As data on autogenous shrinkage was not available, an estimate of the9

    autogenous shrinkage evolution based on Eurocode 2 [54]was used. Another issue to take into10

    account is the fact that the TDC of concrete is not constant during the first hours of age. In fact,11

    several authors have dealt with this subject, and it generally agreed that the initial TDC tends to12

    be larger than that of hardened concrete, and tends to decrease sharply within the first 12 to 2413

    hours of age, reaching then the plateau level corresponding to hardened concrete [69,73,76].14

    The no-stress specimen cast in the scope of this research cannot be used to estimate TDC at15

    concrete early ages, due to the absence of data in the first 8 hours reported in Figure 9b.16

    Nonetheless, as the peak temperature occurred later than 24 hours age, and full data is available17

    for temperatures and strains occurred in such period, calculations could be made under the18

    assumption that TDC was already at its plateau value. The TDC was estimated between instants19

    t=2.0d (peak temperature) and t=8.0d (local minimum) as shown in Figure 9b, and the20

    autogenous shrinkage strain variation in such period was estimated to be of 8.45, (considering21

    fcm=42.3MPa in Eurocode 2 [54]). The estimated constant TDC to be used in the numerical22

    simulation was 11.07C. Nonetheless, since it is known that TDC varies during the first 2423

    hours of age, an alternative formulation for the evaluation of the TDC was considered, based on24

    experimental evidence reported by Laplante and Boulay [77]. Therefore, this alternative25

  • 8/12/2019 Revised Engstruct d 13 00590r1

    30/63

    - 22 -

    formulation considers that during the first 16h the TDC varies according to1

    2( ) 0.16 4.88 48.93TDC t t t (tin hours), and remains constant after such age.2

    The E-modulus evolution of concrete for the simulation model was directly extracted from3

    EMM-ARM data reported in Figure 12, whereas creep modelling was made through the4

    adjustment of DPL parameters to the creep data experiments. The best-fit creep parameters and5

    their adjustment to the experimental data are shown in Figure 11b. Prestressing ducts were6

    modelled with consideration of their inner perimeter of 283mm (90mm inner diameter). Steel7

    reinforcement was disregarded in temperature calculations due to its low interference in8

    temperature development [44, 52]. Regarding mechanical field simulations, steel was not9

    considered because post-cracking behaviour was not sought. In the non-cracked stage, the10

    similitude in TDC of steel in regard to that of hardened concrete strongly minimizes the restraint11

    to concrete thermal deformation, thus rendering the effect of reinforcement negligible for the12

    computation of thermal stresses at early ages [52].13

    4.1.3 Initial/boundary conditions and construction phasing14In what concerns the boundary conditions in the thermal problem, it was assumed that an15

    average wind speed of 2.5m/s occurred (confirmed during 3 days with an anemometer) and the16

    resulting convection/radiation coefficient for concrete surfaces in contact with the environment17

    was estimated to be 15.25 2W m Kin accordance to the predictive formula of Branco et al.[78].18

    For the particular case of formworks surrounding concrete, an equivalent boundary convection19

    coefficient was adopted according to an electrical analogy [46]. Bearing in mind that the20

    formworks were made of wood (kwood=0.175W/m2K) and their thickness was 18mm, the21

    resulting equivalent boundary coefficient was 6W/m2K. Formwork was applied to the vertical22

    surfaces of each casting stage during the first 7 days of age, and removed afterwards. All23

    convective boundaries were subjected throughout the analysis to the environmental temperature24

  • 8/12/2019 Revised Engstruct d 13 00590r1

    31/63

    - 23 -

    that was monitored in-situseeFigure 5.The symmetry plane of the model was considered as1

    an adiabatic boundary.2

    In what regards to solar radiation intake, the several surface directions of the model were3

    taken into account, and the absorbed radiation was calculated according to the model described4

    in section 2.1. Absorvity of concrete was considered as 0.6 [79], the latitude was 41.77 N, and5

    the casting date was 28/03/2011. The Linke turbidity factor was considered as 2.5, and its6

    feasibility was confirmed by comparing computed solar radiation on horizontal surfaces with7

    solar radiation data from a nearby weather station (Rio Torto Station) in conditions of clear8

    skies. The effect of cloudiness was taken into account by normalizing the predictions of the9

    adopted solar radiation model according to information obtained from the piranometer of the10

    neighbouring weather station. The diminishment of solar absorption caused by shadows cast by11

    neighbouring objects was considered negligible throughout the entire day, as the wall was one of12

    the tallest elements in the landscape.13

    Bearing in mind the construction phases taken into account in this calculation model14

    (identified in Figure 13a), the initial temperatures were considered as follows. For the existing15

    concrete at the beginning of analysis, it was assumed that concrete was already in thermal16

    equilibrium with the environment, and so, the average daily temperature of the preceding week17

    (14.5C) was considered for the existing concrete. For the subsequent stages of construction, the18

    initial temperature of concrete was obtained from in-situ monitoring (average value), which was19

    also of approximately 14.5C.20

    In regard to the cooling ducts, due to a limitation of the adopted software, the inlet21

    temperature had to be considered constant, equal to the average environmental temperature22

    during the time in which the tubes were active. The following temperatures were considered for23

    each phase: 8th: 7.13C and 9th: 13.91C. The internal convection coefficient in the ducts was24

    obtained with basis on their internal air speed of 8.6m/s, which according to the studies of25

    Hedlund and Groth [8] should correspond to a convection coefficient of 30.0W/m2K. The26

  • 8/12/2019 Revised Engstruct d 13 00590r1

    32/63

    - 24 -

    cooling duct elements were activated at each construction phase when the surrounding concrete1

    had age of 14 hours.2

    Taking into account the directions of the axes of the coordinate system presented in Figure3

    13a, the mechanical boundary conditions consisted in placing Z direction supports on the bottom4

    surface of the wall and Y supports in all the elements of the symmetry plane.5

    The time step strategy adopted for the model consisted in considering the initial time6

    coincident with the casting instant of 8th phase. Casting of the 9th and 10th phases were7

    considered at the relative instants t=24 days and t=41 days in accordance to the actual8

    construction. All analyses were conducted with a constant time step of 1h duration, even though9

    some localized adjustments were necessary in view of construction phasing and ventilation10

    activation. Nonetheless, all adjustments were carefully made to assure that the duration of all11

    time steps remained under 1h.12

    4.2 Results and discussion13The presentation and discussion of results is centred in the 9th construction phase, with14

    particular emphasis for comparisons between monitored and simulated temperatures/strains. The15

    temperature simulations in concrete were quite coherent with the monitored ones, as it can be16

    confirmed for a set of representative locations (VW1, VW2, VW4 and VW5), whose results are17

    shown inFigure 15.In fact the largest deviations regarding the monitored temperatures during18

    the entire calculation always remain under 4C, thus providing confirmation of the feasibility of19

    the modelling strategy, particularly in regard to the cooling capacity of the ventilated ducts.20

    Based on the confidence gained on the thermal simulation model, a further numerical simulation21

    was made, in which the effect of the cooling ducts was disregarded. The corresponding results22

    for the hottest region of both models are shown in Figure 15b. It can be confirmed that the23

    inclusion of the duct had a twofold effect: not only was the peak temperature diminished by 5C24

  • 8/12/2019 Revised Engstruct d 13 00590r1

    33/63

    - 25 -

    with benefits for cracking safety, but also the return of the internal temperature to thermal1

    equilibrium with the environment was accelerated, with advantages for construction phasing.2

    The fact that the calculated temperatures matched well the monitored ones is a solid starting3

    point for the analysis of results of the mechanical simulation, as any detected deviations are4

    bound to be solely attributed to issues in the mechanical simulation itself. The calculated and5

    measured strains for the same set of sensors that has just been discussed for temperature6

    development are shown in Figure 16. The experimentally measured strains in this figure are7

    represented by their value according to the zeroing procedure mentioned in Section 3.3.3 (Figure8

    7), but also with a lower and upper bound related to possible uncertainties in the instant for9

    zeroing of the sensors output of 2 hours. It can be seen that all the computed strains with10

    consideration of constant TDC underestimate the peak strain at 1.96 day, but the post-peak11

    kinetics seems to have been well captured. As the constant TDC assumption may lead to12

    underestimations of early strain development [57], a further calculation was made using a13

    plausible TDC evolution during the first 24 hours, as discussed in 4.1.2. The corresponding14

    simulation results are shown inFigure 16,where a better overall fit is seen between experimental15

    and calculation data (particularly for core regions). Even though the variable TDC was not based16

    on experimental evidence obtained in the scope of this research, it is feasible to assume that a17

    significant part of the strain deviations regarding experimental results can be explained by the18

    variable TDC at early ages. It has to be kept into consideration that another possible source of19

    deviation of results may be related to the instant at which measured strains were zeroed, which20

    can be debatable. Nonetheless, the adequate prediction of strains that was attained is a good21

    indication of the feasibility of the computed stresses which are to be analysed.22

    The discussion of cracking risk is now addressed by comparing the computed principal23

    tensile stress at the most unfavourable part of the model (located in the core region: x=8.5m,24

    y=0.36m, z=11.7m), as shown in Figure 17. This figure contains the equivalent age-adjusted25

    measured evolution of the tensile strength, and the calculation results for the cases of constant26

  • 8/12/2019 Revised Engstruct d 13 00590r1

    34/63

    - 26 -

    and variable TDC, as well as the case of inexistent cooling ducts and constant TDC. The first1

    comment that can be made fromFigure 17,is that the consideration of constant or variable TDC2

    had very marginal effect on the results. The reason for the very small difference is bound to be3

    related with the very low stress level that is induced during the first 24 hours, as the E-modulus4

    of concrete is still very small and creep/relaxation is very high. Regardless of the comparison5

    between these two models, it can be observed fromFigure 17 that the ratio between the tensile6

    stress and the tensile strength of concrete at the most unfavourable instant is of approximately7

    0.9, which corresponds to a significant cracking risk. Nonetheless, even though the cracking risk8

    was higher than the desirable one, the structure did not present thermal cracks neither later9

    through-cracks (evaluations made until 2 years after casting). It should be remarked that this10

    point of stress analysis was the most unfavourable one within the structure, and significantly11

    lower cracking risks were calculated for distinct regions, resulting in a global scenario of much12

    more cracking safety (compared to a single-point analysis).13

    Interestingly, a simulation of the same construction situation without consideration of14

    cooling ducts would yield to higher cracking risk at the same location (as seen inFigure 17),15

    with the ratio between the tensile stress and the tensile strength of concrete reaching 1.2.16

    Therefore, if the calculations made here are considered trustworthy, the use of the cooling ducts17

    may have been the differentiating factor that avoided a cracking scenario in this concrete lift.18

    5 CONCLUSIONS19A case study regarding the assessment of the cracking risk of a thick wall in the entrance of a20

    dam spillway, internally cooled with air-filled prestressing ducts, was presented in this paper.21

    The use of ventilated prestressing ducts is considered more straightforward than water ducts due22

    to the often easy availability of ducts and fans in construction works associated to massive23

    concrete structures.24

  • 8/12/2019 Revised Engstruct d 13 00590r1

    35/63

    - 27 -

    The case study has involved a comprehensive experimental part, including laboratory1

    characterization of materials and in-situmonitoring for temperatures and strains. A 3D thermo-2

    mechanical simulation of the construction phasing was shown, with input data duly based in the3

    laboratory characterization program.4

    In regard to previous works reported in the literature, the work presented in this paper has its5

    main original contributions in the following fields: (i) the EMM-ARM methodology for6

    continuous monitoring of concrete E-modulus since casting was applied for the first time as a7

    characterization tool for stress simulation in concrete at early ages, thus enhancing the quality of8

    input data; (ii) this is the first reported application of horizontally placed air-cooling pipes, with9

    its efficiency being assessed and numerically simulated; (iii) the work reported here is relatively10

    unprecedented in view of its holistic approach, with the authors being involved in all tasks of11

    laboratory characterization (allowing sustainable estimates of material properties and modelling12

    strategy for numerical simulations), field monitoring and numerical simulation with thermo-13

    mechanical analysis.14

    It is nonetheless acknowledged that a significant research effort is still necessary in regard to15

    the creep monitoring and modelling at early ages, both in view of the effects of early hydration16

    and in view of temperature effects on creep. Even though in-depth analyses of these particular17

    issues of creep have not been included in this paper, the authors consider them to be of critical18

    importance for adequate stress simulation in hardening concrete, thus demanding further19

    research works.20

    The numerical simulation results were compared to those collected by the in-situ monitoring21

    and good coherences were observed both in terms of temperature and strain, providing good22

    prospects in regard to the simulation capabilities of the models and the soundness of the23

    experimentally obtained data. The monitoring/simulation results allowed concluding that the24

    effectiveness of the air cooling system with horizontally placed pipes is limited in view of the25

    significant heating that air suffers along the first meters of tube, thus diminishing its capacity of26

  • 8/12/2019 Revised Engstruct d 13 00590r1

    36/63

    - 28 -

    cooling further regions of concrete. Also, the duct efficiency ends up being quite dependent of1

    the environmental temperature, which cannot be easily anticipated during planning. Nonetheless,2

    in spite of the acknowledged limitations of air cooling, it may prove quite feasible in relatively3

    cool climates and small lengths of embedment (e.g.: 10m or less).4

    Furthermore, the risk of cracking on the studied construction phase has resulted acceptable in5

    most of its regions, even though a non-negligible risk of internal cracking was observed in some6

    regions. The fact that no surface or through cracking was observable in the construction7

    corroborates the cracking risk evaluation. It was also concluded that the same construction8

    phasing without the use of the cooling ducts had a significantly higher cracking probability, thus9

    confirming the usefulness of the cooling system.10

    Finally, it is also worth remarking that a parametric analysis regarding the possibility of11

    considering variable thermal dilation coefficient of concrete along hydration has been carried12

    out. The outcome of such parametric analysis seems to point out a relatively low impact of13

    admitting the thermal dilation coefficient as constant along hydration on the corresponding14

    computed stresses, thus validating such simplification in this case study.15

    6 ACKNOWLEDGEMENTS16Funding provided by the Portuguese Foundation for Science and Technology to the Research17

    Unit ISISE, to the second author through the PhD grant SFRH/BD/64415/2009, and to the18

    research projects PTDC/ECM/099250/2008 and QREN number 5387, LEGOUSE, is gratefully19

    acknowledged. The kind assistance of the contractor (Teixeira Duarte S.A.) and the owner (EDP20

    Eletricidade de Portugal) are also deeply appreciated. The contribution of ngelo Costa to the21

    experimental work here reported is also gratefully acknowledged.22

  • 8/12/2019 Revised Engstruct d 13 00590r1

    37/63

    - 29 -

    7 REFERENCES1[1] Bernander S. Practical Measures to Avoiding Early Age Thermal Cracking in concrete2

    structures. In: Springenschmid R, editor. Prevention of Thermal Cracking in Concrete at Early3

    Ages : State-of-the-art Report Prepared By RILEM Technical Committee 119 (RILEM Report4

    15): E & FN Spon; 1998. p. 255-314.5

    [2] Springenschmid R. Thermal Cracking in Concrete At Early Ages RILEM Proceedings6

    (RILEM Report 25): Taylor & Francis Routledge; 1995.7

    [3] Du C. Dam construction - concrete temperature control using fly ash. Concrete International.8

    1996;18:34-6.9

    [4] Utsi S, Jonasson J-E. Estimation of the risk for early thermal cracking for SCC containing fly10

    ash. Materials and Structures. 2012;45:153-69.11

    [5] Kurita M, Goto S, Minegishi K, Negami Y, Kuwahara T. Precooling concrete using frozen12sand. Concrete International. 1990;12:60-5.13

    [6] Takeuchi H, Tsuji Y, Nanni A. Concrete precooling method by means of dry ice. Concrete14

    International. 1993;15:52-6.15

    [7] St. John J. Construction of the Hoover Dam Bypass. Concrete International. 2011;33:30-5.16

    [8] Hedlund H, Groth P. Air cooling of concrete by means of embedded cooling pipes-Part I:17

    Laboratory tests of heat transfer coefficients. Materials and Structures. 1998;31:329-34.18

    [9] Groth P, Hedlund H. Air cooling of concrete by means of embedded cooling pipesPart II:19Application in design. Mater Struct. 1998;31:387-92.20

    [10] Roush K, O'Leary. Cooling concrete with embedded pipes. Concrete International.21

    2005;27:30-2.22

    [11] Maggenti R. From passive to active thermal control. Concrete International. 2007;29:24-30.23

    [12] Ishikawa S, Matsukawa K, Nakanishi S, Kawai H. Air pipe cooling system. Concrete24

    International. 2007;29:45-9.25

    [13] Nobuhiro M, Kazuo U. Nonlinear thermal stress analysis of a massive concrete structure.26Computers & Structures. 1987;26:287-96.27

    [14] Emborg M. Thermal Stresses in Concrete Structures at Early Age. Doctoral Thesis thesis.:28

    Lule University of Technology; 1989.29

    [15] Ishikawa M. Thermal stress analysis of a concrete dam. Computers & Structures.30

    1991;40:347-52.31

    [16] de Borst R, van den Boogaard AH. Finite-Element Modeling of Deformation and Cracking32

    in Early-Age Concrete. Journal of Engineering Mechanics. 1994;120:2519-34.33

  • 8/12/2019 Revised Engstruct d 13 00590r1

    38/63

    - 30 -

    [17] Torrenti JM, De Larrard F, Guerrier F, Acker F, Grenier F. Numerical simation of1

    temperatures and stresses in concrete at early ages: the French experience. In: R.Springenschmid,2

    editor. Thermal Cracking in Concrete at Early Ages (RILEM Proceedings 25). Paris. 1994. p.3

    281-8.4

    [18] Fairbairn EMR, Ferreira IA, Cordeiro GC, Silvoso MM, Filho RDT, Ribeiro FLB.5Numerical simulation of dam construction using low-CO2-emission concrete. Materials and6

    Structures/Materiaux et Constructions. 2010;43:1061-74.7

    [19] De Schutter G, Vuylsteke M. Minimisation of early age thermal cracking in a J-shaped non-8

    reinforced massive concrete quay wall. Engineering Structures. 2004;26:801-8.9

    [20] Dolmatov AP, Neidlin SZ. Temperature control of the massive concrete of the Krasnoyarsk10

    hydroelectric station dam. Hydrotechnical Construction. 1968;2:956-60.11

    [21] Blinov IF, Shaikin YP, Gal'perin IR. Field studies of the temperature regime and stress state12of the concrete in arch dams subjected to various methods of cooling. Hydrotechnical13

    Construction. 1980;14:19-23.14

    [22] Torrenti J-M, Buffo-Lacarrire L, Barr F. CEOS.FR experiments for crack control of15

    concrete at early age. RILEM-JCI International Workshop on Crack Control of Mass Concrete16

    and Related Issues concerning Early-Age of Concrete Structures (ConCrack 3). Paris. 2012. p. 3-17

    10.18

    [23] Cussigh F. Experience in limiting early age concrete temperature for DEF prevention.19

    RILEM-JCI International Workshop on Crack Control of Mass Concrete and Related Issues20

    concerning Early-Age of Concrete Structures (ConCrack 3). Paris. 2012. p. 79-88.21

    [24] Benboudjema F, Torrenti JM. Early-age behaviour of concrete nuclear containments.22

    Nuclear Engineering and Design. 2008;238:2495-506.23

    [25] Craeye B, De Schutter G, Van Humbeeck H, Van Cotthem A. Early age behaviour of24

    concrete supercontainers for radioactive waste disposal. Nuclear Engineering and Design.25

    2009;239:23-35.26

    [26] Briffaut M, Benboudjema F, Torrenti JM, Nahas G. Numerical analysis of the thermal27

    active restrained shrinkage ring test to study the early age behavior of massive concrete28

    structures. Engineering Structures. 2011;33:1390-401.29

    27 Waller V, dAloa L, Cussigh F, Lecrux S. Using the maturity method in concrete cracking30

    control at early ages. Cement and Concrete Composites. 2004;26:589-99.31

    [28] Xiang Y, Zhang Z, He S, Dong G. Thermalmechanical analysis of a newly cast concrete32

    wall of a subway structure. Tunnelling and Underground Space Technology. 2005;20:442-51.33

  • 8/12/2019 Revised Engstruct d 13 00590r1

    39/63

    - 31 -

    [29] Faria R, M. A, J.A. F. Modelling of concrete at early ages: application to an externally1

    restrained slab. Cement & Concrete Composites. 2006;28(6):572-85.2

    [30] Azenha M, Faria R. Temperatures and stresses due to cement hydration on the R/C3

    foundation of a wind tower-A case study. Engineering Structures. 2008;30:2392-400.4

    [31] Boutillon L, Linger L, Kolani B, Meyer E. Effects of sun irradiation on the temperature and5early age stress distribution in external concrete structure. In: Toutlemonde F, Torrenti J-M,6

    editors. RILEM-JCI International Workshop on: Crack control of mass concrete and related7

    issues concerning early-age of concrete structures ConCrack 3. Paris. 2012. p. 181-92.8

    [32] Zreiki J, Bouchelaghem F, Chaouche M. Early-age behaviour of concrete in massive9

    structures, experimentation and modelling. Nuclear Engineering and Design. 2010;240:2643-54.10

    [33] Zhu Y-M, Liu Y-Z, Xiao Z-Q, He J-R, Lin Z-H, Ma Y-F. Analysis of pipe-cooling system11

    in mass concrete. Nanjing, China: College of Water Conservancy & Hydropower Eng., Hohai12Univ. 2004.13

    [34] James RJ, Dollar DA. Thermal Engineering for the construction of large concrete arch14

    dams. The 6th ASME-JSME Thermal Engineering Joint Conference. Hapena Beach, Hawaii.15

    2003.16

    [35] Xie H, Chen Y. Influence of the different pipe cooling scheme on temperature distribution17

    in RCC arch dams. Communications in Numerical Methods in Engineering. 2005;21:769-78.18

    [36] Myers TG, Fowkes ND, Ballim Y. Modeling the cooling of concrete by piped water. ASCE19

    Journal of Engineering Mechanics. 2009:1375-83.20

    [37] Yamamoto T, Ohtomo T. Pratices for crack control of concrete in Japan. In: Toutlemonde F,21

    Torrenti JM, editors. RILEM-JCI International Workshop on Crack Control of Mass Concrete22

    and Related Issues Concerning Early-Age of Concrete Structures (ConCrack 3). ParisRILEM23

    Publications. 2012. p. 193-200.24

    [38] Tanabe T-a, Yamakawa H, Watanabe A. Determination of convection coefficient at cooling25

    pipe surface and analysis of cooling effect. Proceedings of JSCE. 1984;34:171-9.26

    [39] Yang J-K, Lee Y, Kim J-K. Heat Transfer Coefficient in Flow Convection of Pipe-Cooling27

    System in Massive Concrete. Journal of Advanced Concrete Technology. 2011;9(1):103-14.28

    [40] Yang J, Hu Y, Zuo Z, Jin F, Li Q. Thermal analysis of mass concrete embedded with29

    double-layer staggered heterogeneous cooling water pipes. Applied Thermal Engineering.30

    2012;35:145-56.31

    [41] Kim J