Lin Yang Early-age cracking of concrete and calorimetry

download Lin Yang Early-age cracking of concrete and calorimetry

of 13

Transcript of Lin Yang Early-age cracking of concrete and calorimetry

  • 8/17/2019 Lin Yang Early-age cracking of concrete and calorimetry

    1/13

    Thermal analysis and adiabatic calorimetry for early-age concretemembers

    Part 2. Evaluation of thermally induced stresses

    Yun Lin1 • Hung-Liang Chen1

    Received: 20 July 2015 / Accepted: 24 October 2015 / Published online: 18 November 2015  Akadémiai Kiadó, Budapest, Hungary 2015

    Abstract   In this study, a finite element model was

    developed to perform the stress analysis on early-ageconcrete members to predict the thermally induced stresses

    and the associated cracking risk. FORTRAN subroutines

    were created for ABAQUS finite element program to

    enable solution-dependent material properties in the ther-

    mal stress calculation. Young’s modulus development,

    strength development, tensile creep, and compressive creep

    behaviors at early age were experimentally obtained, and

    these material properties were incorporated in the subrou-

    tines. Two 1.2-m concrete cubes were constructed with

    embedded temperature sensors and vibrating wire strain

    gages to verify the simulation results. Results showed that

    the calculated temperature and strain values correlated wellwith the measured field data. Additionally, visual cracks

    were confirmed at the predicted locations on the concrete

    cube. It is concluded that the method developed in this

    study is capable of determining the thermally induced

    stresses of early-age concrete members.

    Keywords   Mass concrete Heat of hydration Early-ageconcrete    Thermal stress   ABAQUS

    Introduction

    The heat generation from cement hydration leads to a

    temperature rise, especially at the core of a large concrete

    member. At concrete surfaces, the temperatures are rela-

    tively lower due to surface heat loss from external ambient

    cooling. The created temperature differential may lead to

    high tensile stresses at the concrete surfaces and produce

    surface cracking. The most common thermal control

    practice is to limit the temperature differential between the

    center and the surface of the concrete structures. However,

    temperature differential is not conclusive enough to

    determine the cracking risks due to thermal stresses. Nagy

    and Thelandersson [1] pointed out that the development of concrete Young’s modulus is very important in thermal

    stress modeling. Gutsch and Rosatasy [2] suggested the

    importance of tensile strength development and the tensile

    creep behavior in terms of cracking potentials.

    Lawrence et al. [3] reported that temperature differential

    alone was not sufficient to determine thermal stresses.

    Instead, a thermal stress analysis considering the changes

    of concrete material properties, such as thermal expansion

    coefficient, Young’s modulus and viscoelasticity should be

    used.

    During early age, the nonuniform temperature profile

    distribution causes disproportionate thermal expansions

    within the concrete body. The surface of concrete in lower

    temperatures can be under high tensile stresses due to

    relative thermal expansions from internal concrete. The

    heating effect due to hydration and the cooling effect due

    to surface heat loss occur simultaneously. Therefore, the

    surface of concrete is under tension once concrete is set

    until the hydration heat is fully dissipated to the environ-

    ment. The reversal of stress may occur beneath the surface

    of concrete when the concrete passes from the heating

    &   Yun [email protected]

    Hung-Liang [email protected]

    1 Department of Civil and Environmental Engineering, WestVirginia University, P.O. Box 6103, Morgantown,WV 26506-6103, USA

     1 3

    J Therm Anal Calorim (2016) 124:227–239

    DOI 10.1007/s10973-015-5131-x

    http://crossmark.crossref.org/dialog/?doi=10.1007/s10973-015-5131-x&domain=pdfhttp://crossmark.crossref.org/dialog/?doi=10.1007/s10973-015-5131-x&domain=pdfhttp://orcid.org/0000-0002-2450-222X

  • 8/17/2019 Lin Yang Early-age cracking of concrete and calorimetry

    2/13

    phase to the cooling phase. Whether the high surface ten-

    sile stresses can cause cracking is depending on the stress-

    to-strength ratio at the critical locations. During the

    hydration process of the early-age concrete, both the

    thermally induced stresses and the concrete strength are

    being developed but at different rates. Cracks are most

    likely to occur at the critical locations where tensile stress

    exceeds the tensile strength. Figure 1  originally presentedby Tia et al. [4] depicts an example of thermal stress and

    concrete tensile strength development. The cracking zone

    in the figure refers to the time when tensile stress exceeds

    tensile strength. In practice, this cracking time zone is most

    likely to occur within 1–2 days after concrete placement,

    depending on the member geometry, size, boundary

    restraint and the ambient temperature variations.

    The development of thermally induced stresses is a

    complicated phenomenon which includes the variability of 

    temperature distribution, concrete thermal and mechanical

    properties, and the viscoelastic behavior of early-age con-

    crete. In recent years, finite element models have been usedto predict the thermally induced stresses of early-age

    concrete members. Waller et al. [5] presented a model

    using CESAR-LCPC which included two modules, TEXO

    and MEXO, to perform the thermal analysis and stress

    analysis on concrete structures. Wu et al. [6] described the

    procedures calculating thermally induced stresses for a

    wall element using ANSYS. Tia et al. [7] evaluated bridge

    footing elements with wooden formwork using TNO Diana

    software. Their research findings are very helpful to this

    topic; however, the modeling procedure of the viscoelastic

    behaviors due to tensile or compressive stresses was not

    detailed enough for replication purposes.Researchers have emphasized the importance of con-

    crete’s viscoelasticity, which is crucial in calculating

    thermal stresses. Bažant’s B3 model [8], which was

    designed for long-term creep behaviors, has been widely

    adopted recently to describe creep behaviors of early-age

    concrete. Østergaard et al. [9] improved the B3 model on

    the early-age creep behavior by adjusting the ‘‘aging’’ term,

    while Wei and Hansen [10] made an adjustment on the

    later ages by modifying the ‘‘flow’’ term of B3 model.

    However, the temperature, although often ignored in the

    creep calculation, has a significant effect on early-ageconcrete creep behavior. Using the equivalent age to con-

    sider the temperature effect in creep was suggested by

    Bažant and Baweja [11]. Atrushi [12] also showed the

    usage of equivalent age on the modified double power law

    (DPL) with some experimental proof. Luzio and Cusatis

    [13] validated the solidification–microprestress–mi-

    croplane (SMM) model considering moisture variation and

    moisture diffusion associated with environmental exposure

    and internal water consumption. This paper described the

    thermal stress calculation of early-age concrete using a

    modified B3 model considering the aging and temperature

    effect in a variable loading and temperature environment.A thermal stress calculation algorithm with experimental

    verification is presented in this paper.

    Concrete cube construction

    To study the development of thermally induced stresses,

    two 1.2-m concrete cubes were constructed (Fig.  2). For

    both cubes, temperature measurements were taken at the

    center of the cube and 5 cm from the side surface and the

    top surface. The details about the temperature predictions

    and measurements were presented in Lin and Chen [14].Additionally, Geokon vibrating wire strain gages (Model

    4200, gage length 15.25 cm) were installed to measure the

    strain changes within a concrete member during the early

    Liquid Solid

    Cracking zone

       T  e  n  s   i   l  e  s   t  r  e  s  s  a

      n   d  s   t  r  e  n  g   t   h

    Time

    Tensile strength

    Tensile stress

    Fig. 1  Thermal stress and tensile strength development with crack initiation   Fig. 2   Pictures of internal sensor installations before cube 2 casting

    228 Y. Lin, H.-L. Chen

     1 3

  • 8/17/2019 Lin Yang Early-age cracking of concrete and calorimetry

    3/13

    ages after concrete placement. Strain measurements were

    taken at location A for cube 1 and locations A, B, and C for

    cube 2. The locations of these sensors are illustrated in

    Fig. 2; locations A and B are 10 and 5 cm from the center

    of the side surface, while location C is 2.5 cm from the

    center of the top surface. The mix design and cement

    chemical compositions used for the two cubes are shown in

    Table 1   and Table   2. The cement chemical composition

    shown in Table 2   was analyzed by the material testing

    laboratory of West Virginia Department of Transportation

    (WVDOT). Figure 3   shows the pictures of the vibratingwire strain ages and the temperature loggers used and the

    1.2-m concrete cubes.

    Mechanical properties of early-age concrete

    In order to calculate the thermally induced stresses, accu-

    rate estimations of the concrete tensile strength and the

    modulus of elasticity development are crucial. The degree

    of hydration (a) calculated using Eq. (3.1), used to estimate

    the concrete strength and modulus at any given time, is a

    function of the equivalent age,   t e. The equivalent age can

    be calculated using the Arrhenius equation, Eq. (3.3) [15],

    which is depending on the concrete temperature history and

    the activation energy,   E a. The activation energy of this

    particular mix design from Table 1   was determined to be

    41,800 J mol-1 by Yikici and Chen [16] following ASTM

    C 1074-10 procedures. The ultimate degree of hydration,au, can be calculated using Eq. (3.2)   [17]. The hydration

    parameters,  s  and  b, were two constants depending on the

    mix design. s  = 14.0 and  b  = 0.94 were determined from

    the adiabatic temperature rise tests by Lin and Chen [14].

    a   t eð Þ ¼ au exp     st e

    b !  ð3:1Þ

    au ¼   1:031 w=c0:194 þ w=c   ð3:2Þ

    t e ¼

     r t 

    0

    exp  E a

     R

    1

    273 þ T r   1

    273 þ T c   t ð Þ dt    ð3:3Þwhere s,  b  hydration parameters,  R  universal gas constant,

    T c(t ) concrete temperature at time   t ,   T r   reference temper-

    ature, 23   C,  E a   activation energy/J mole-1.

    Concrete strength testing

    The compressive strength development curves obtained

    from the compressive cylinder test results using 0.15 by

    0.3 m cylinders cured at a constant temperature of 23   C

    Table 1   Concrete mix design/kg m-3

    Material Cement Water CA FA AE/Lm-3 WR/Lm-3

    Quantity 335 139 969 844 0.067 1.0

    CA   coarse aggregates,   FA   fine aggregates,   AE   air entraining agent,WR  High-range water reducer

    Table 2   Cement chemical composition/%

    Components CaO SiO2   Al2O3   Fe2O3   MgO SO3   Na2O K 2O

    Percentages 62.3 20.22 4.8 3.1 2.51 3.0 0.034 0.76

    Fig. 3   Pictures of vibratingwire strain gage andtemperature logger and cube 1and cube 2

    Thermal analysis and adiabatic calorimetry for early-age concrete members 229

     1 3

  • 8/17/2019 Lin Yang Early-age cracking of concrete and calorimetry

    4/13

    (Fig. 4a). Schutter [18] reported that concrete compressive

    strength and degree of hydration (a) had a linear relation-

    ship. Results from the current mix design (Table  1) also

    showed a strong linear correlation between the degree of 

    hydration and the compressive strength (Fig. 4b). This

    linear relationship, Eq. (3.1.1), will be used to describe the

    concrete strength at any given degree of hydration.

     f 0c ¼ 45:53a 1:71   a 0:04;  f 0c 0   ð3:1:1Þ

    Splitting tensile strength development was determined

    according to ASTM C496 using 0.15 by 0.3 m cylinders.

    Wight and MacGregor [19] presented Eq. (3.1.2) obtained

    from the mean split cylinder strength ( f ct) from a massive

    database. The curve described by Eq. (3.1.2) has a high

    correlation with the current splitting tensile experiment

    results; the comparison of the test results and the predicted

    values using Eq. (3.1.2) are shown in Fig.  5. For modeling

    purposes, the splitting tensile strength development can

    also be expressed in terms of degree of hydration shown in

    Eq. (3.1.3) by inserting Eqs. (3.1.1)–(3.1.2).

     f ct ¼ 0:53 ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

     f 0c   inMPað Þq 

      ð3:1:2Þ

     f ct ¼ 0:53 ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    45:53a 1:71p 

      a 0:04;  f ct 0ð Þ ð3:1:3Þ

    Young’s modulus development

    In this study, accurately assigning elastic modulus for the

    concrete under tension is essential for thermal stress calcu-

    lations. The tensile modulus development curve was experi-

    mentally obtained. The specimens used in tensile modulus

    test were 0.9-m-long dog-bone specimens, each with a

    0.1 m  9 0.1 m mid-cross section (Fig. 6a). One vibrating

    wire strain gage was embedded in the middle of the concrete

    specimen. A steel hook was placed at each end in order to

    applytension. The specimen was loaded in direct tension non-

    destructively (Fig. 6b) using a force between 600 and 2400

    Newton (approximately 10 % stress-to-strength ratio)

    depending on theconcrete maturity. The strain due to external

    tensile stress was measured by the vibrating wire strain gage.

    The specimen was loaded four times for each data point to

    ensure the accuracy. Each loading lasted approximately 10 s

    to minimize any creep effect. The tensile modulus values

    shown in Fig. 6c was obtained based on the measured stress-

    to-strain ratios. The relationship between compressive

    strength and Young’s modulus for this particular mix design

    can be determined using curve fitting method. A best-fit

    exponential function as shown in Eq. (3.2.1) is used to

    describe the development of the Young’s modulus. In

    Eq. (3.2.1), the compressive strength   f 0c   can be expressedwith degree of hydration (Eq. 3.1.1). The elastic modulus can

    also be expressed in terms of degree of hydration as in

    Eq. (3.2.2). The Young’s modulus was assumed identical in

    both tensile and compressive directions.

    E c ¼ 5407 f 0c0:492 ð3:2:1ÞE c ¼ 5407 45:53a 1:71ð Þ0:492 a 0:04; E c 0ð Þ

    ð3:2:2Þ

    Thermal expansion coefficient

    After performing the tensile modulus testing, the dog-

    bone sample with embedded vibrating wire strain gage

    0 2 4 6 8 00

    5

    10

    15

    20

    25

    30

    35

    0

    5

    10

    15

    20

    25

    30

    35

    0.2

    y  = 45.53x  – 1.71

    R  2 = 0.9969

    0.4

    Degree of hydrationEquivalent age/day

       C  o  m  p  r  e  s

      s   i  v  e  s   t  r  e  n  g   t   h   /   M   P  a

       C  o  m  p  r  e  s

      s   i  v  e  s   t  r  e  n  g   t   h   /   M   P  a

    0.6 0.8

    (a) (b)Fig. 4   Relationship betweencompressive strength anddegree of hydration

    00

    0.5

    1

    1.5

    2

    2.5

    3

    2 4

    Test result

    Wight & Macgregor

    Eqvalent age/day

       T  e  n  s   i   l  e  s   t  r  e  n  g   t   h   /   M   P  a

    6 8

    Fig. 5   Splitting tensile strength test results in comparison with Wightand Macgregor [19]

    230 Y. Lin, H.-L. Chen

     1 3

  • 8/17/2019 Lin Yang Early-age cracking of concrete and calorimetry

    5/13

    was reused to test the coefficient of thermal expansion

    (CTE). The dog-bone specimen was submerged into a

    temperature controlled water tank and placed on fric-

    tionless base provided by ball bearings. The specimen

    was able to freely expand and contract due to temperature

    changes. The strain data of the dog-bone specimen was

    recorded, while the water temperature was controlled to

    slowly rise and drop. The CTE test was repeated three

    times, and the test results correlated closely with an

    average thermal expansion coefficient of 8.53 micros-trains per   C at 28 days of age. The thermal expansion

    coefficient is assumed to be a constant for simplicity. The

    variation of CTE of concrete is difficult to measure

    because of the temperature influence of the concrete

    maturity especially at early age; it was shown by

    McCullough and Rasmussen [20] that concrete CTE

    variation after 24 h of age could be assumed negligible

    where the CTE before 24 h decreased noticeably. It is

    noted that CTE is also depending on the moisture level

    inside concrete, and it was assumed that the moisture

    level in the current concrete cube is close to 98–100 %

    [14].

    Basic creep of early-age concrete under constant

    load

    The viscoelastic behavior of early-age concrete plays an

    important role in calculating thermal stresses. The tensile

    creep behavior of early-age concrete is complicated.

    Tensile creep tests performed by Gutch and Rostasy [2]

    showed pronounced viscoelasticity when load was applied

    at early age. Umehara and Uehara [21] and Atrushi [12]

    demonstrated the influence of temperature on early-age

    tensile creep. Østergaard et al. [9] and Atrushi [12] showed

    the strong loading age dependency in the early ages.

    Bažant and Baweja [11] presented a mathematical

    expression of structural creep law (B3 model) shown in

    Eq. (3.4.1). With experimentally determined empirical

    constants (q1–q4), B3 model was often found accurate interms of correlating with the experimental data.

     J t ;  t 0ð Þ ¼ e   t ð Þr

    ¼ q1 þ q2Q t ;  t 0ð Þ þ q3 ln 1 þ   t  t 0ð Þ0:1h i

    þ q4ln t t 0

    ð3:4:1Þwhere

    Q t ;   t 0ð Þ ¼ Qf   t 0ð Þ   1 þ   Qf  Z t ;  t 0ð Þ r 1=r

    Qf   t 0ð Þ ¼   0:086   t 0ð Þ29þ1:21   t 0ð Þ49h i1 Z t ;  t 0ð Þ ¼   t 0ð Þ1=2ln 1   t  t 0ð Þ0:1

    h ir  ¼ 1:7   t 0ð Þ0:12þ8:0t  current age in days (t  = 0 is when water is added to the

    mixture), t 0  loading age in days,  q1, q2, q3, and q4 empiricalconstants

    (a)

    (b) (c)

    LoadedConcreteSpecimen

    Vibratingwire gage

    Weights

    RigidFrame

    010000

    14000

    18000

    22000

    26000

    30000

    2 4

    Test results

    Eq. (3.2.2)

    Equivalent age/day

       T  e  n  s   i   l  e  m  o   d  u   l  u  s   /   M   P  a

    6 8

    Fig. 6   Tensile modulus testingsetup and results

    Thermal analysis and adiabatic calorimetry for early-age concrete members 231

     1 3

  • 8/17/2019 Lin Yang Early-age cracking of concrete and calorimetry

    6/13

    Østergaard et al. [9] found that for early-age tensile

    creep, B3 model may underestimate the specific creep. The

    early-age concrete exhibits much greater viscoelasticity.

    They modified the  q2 constant using Eq. (3.4.2) to amplify

    the age dependency for early-age concrete, where   q5   is

    always less than the physical loading age (t 0). In theirresearch, with a very early loading age at 16 h, the best-fit

    value of  q5  was found to be 14 h.

    q02 ¼ q2t 0

    t 0 q5 ð3:4:2Þ

    Temperature is also an important factor, which has two

    different effects on the creep behavior of early-age con-

    crete. From the maturity concept, higher curing tempera-

    ture will accelerate the hydration process and increase the

    concrete maturity at the time of loading and therefore

    decrease the specific creep. However, the creep deforma-

    tion at early age increases significantly as the temperature

    increases. Atrushi [12] stated that the increasing effect is

    much greater than the decreasing effect. In Atrushi’sexperimental results, a significant increase in tensile creep

    was found due to the effect of temperature increase. In

    order to consider the effect of the temperature, the equiv-

    alent age concept was used by Bažant and Baweja [11] and

    Atrushi [12]; the equivalent age was used to replace the

    regular age in the places of the loading age and the loading

    duration, where they found better agreements between the

    theoretical and experimental results. Hence, the modified

    B3 model can be expressed as shown in Eq. (3.4.3).

     J t e; t 0e ¼

    e   t eð Þr

    ¼ q1 þ q2 t 0e

    t 0e q5Q t e; t 

    0e

    þ q3 ln 1 þ   t e t 0e

    0:1h i þ q4ln t et 0e

    ð3:4:3Þ

    To measure the basic tensile creep of early-age concrete,

    surface sealing is important because tensile loading can

    significantly accelerate the drying effect and lead to more

    load-induced drying shrinkage. In this study, during each

    creep testing, two identical 0.9-m dog-bone specimens with

    a 0.1 m  9 0.1 m cross section at the mid-span region were

    used. The concrete molds and sensor installation were the

    same as shown in Fig.  6. Both specimens were sealed withepoxy paint plus four layers of plastic wraps immediately

    after unmolding (1 h prior to the loading). Epoxy paint

    creates an adhesion between the plastic wrap and the

    concrete surfaces to further prevent surface drying. Both

    specimens were kept in the same room with a controlled

    temperature of 23   C and 50 % humidity level. One of the

    specimens was loaded with a tensile stress of 0.13 MPa

    (approximately 10 % stress-to-strength ratio), while the

    other was kept free to deform. Although the loading

    magnitude is small with respect to its tensile strength, the

    specific creep was assumed to be un-affected.

    Hauggaard et al. [22] reported that the specific creep

    response of early-age concrete was found to be unchanged

    when a stress-to-strength ratio is below 60 %. Similar

    conclusion was drawn by Atrushi [12] in his tests up to

    80 % stress-to-strength ratio.

    The strain measurements for both specimens (loadedand free) were taken using Geokon vibrating wire strain

    gages. The difference in the monitored strain between the

    two specimens divided by the loading magnitude is cal-

    culated to show creep compliance. The tensile creep test

    for the specific mix design (Table 1) was performed three

    times at three different loading ages (0.75, 1, and 10 days),

    and the results are shown in Fig.  7. As shown in Fig.  7, all

    of the three test results can be described by the modified B3

    model (Eq. 3.4.3). The best-fit empirical constants are

    shown in Table  3.

    Although tensile and compressive creep models were

    often assumed to be the same for simplicity, differenttensile and compressive creep behaviors were discovered

    by numerous researchers [12,   23,   24]. The compressive

    creep model of the current mix design was obtained from

    an existing model by Atrushi [12] shown in Eq. (3.4.4).The

    double power law (DPL) developed by Bažant and Osman

    [25] has been widely used to model compressive creep

    behavior for hardened concrete. Atrushi [12] modified the

    DPL for early-age concrete by incorporating the tempera-

    ture effect observed in early age. The equation includes

    equivalent age at loading (t e0

    ), the current equivalent age

    (t e), the Young’s modulus at loading (E (t e0

    )), and three

    additional creep parameters (/,   d   and   p). This modified

    0

    0

    10

    20

    30

    40

    50

    60

    70

    80

    90

    100

    1 2 3 4 5 6

    Eq. 3.4.3 (0.75 day)

    Eq. 3.4.3 (1 day)

    Eq. 3.4.3 (10 day)

    Experiment (0.75 day)

    Experiment (1 day)Experiment (10 day)

    Time/day

       C  r  e  e  p  c  o  m  p   l   i  a  n  c  e   /   µ

        M   P  a  –   1

    7 8 9 10 11 12 13

    Fig. 7   Comparison of creep compliance between experimentalresults and Eq. (3.4.3)

    Table 3   Best-fit empirical constants for Eq. (3.4.3)

    Constant q1   q2   q3   q4   q5

    Value 0.3 24.0 65.0 0.5 0.2

    232 Y. Lin, H.-L. Chen

     1 3

  • 8/17/2019 Lin Yang Early-age cracking of concrete and calorimetry

    7/13

    double power law (M-DPL) is shown in Eq. (3.4.4).Since

    the mix design used in this study (Table  1) was similar to

    the ‘‘Base-0 mix’’ from Atrushi [12], same creep parameter

    values were used for the current calculation. The values of 

    /,  d , and p  were 0.75, 0.2, and 0.21, respectively, obtained

    from Atrushi [12]. The M-DPL with these pre-determined

    creep parameters was used to account for compressive

    creep of concrete in the FEM calculations.

     J t e; t 0e

    ¼   1E t 0e   1 þ /t 0de   t e t 0e p   ð3:4:4Þ

    Modeling of viscoelastic behavior

    Before cracking, the concrete material is usually assumed

    linear elastic as in Eq. (3.5.1) for one-dimensional stress;

    the elastic modulus is the ratio of the stress and the

    instantaneous strain (eins). However, the early-age concrete

    exhibits high viscoelastic behavior which causes a change

    in effective modulus, and the analytical response is illus-trated in Fig. 8. To simplify the creep calculation, an

    effective modulus (E eff ) is used in the FEM modeling. The

    effective modulus (Eq. 3.5.2) represents the ratio of the

    stress and the total deformation as plotted in Fig. 8a.

    Figure 8b demonstrates a typical creep compliance   J (t, t 0)of a concrete specimen under constant loading.   J (t, t 0) isdefined as the ratio of the total strain (e(t )  =  etotal  =

    ecr  ?  eins) and the stress ( J (t, t 0)  =  e(t ) / r). The creep

    coefficient (C cr) is defined as the ratio of the creep strain

    (ecr) and the instantaneous strain (eins) due to the loading

    (Eq. 3.5.3). Equation (3.5.4) can be derived according to

    Fig. 8b. The growth of Young’s modulus is considered inthis model for early-age concrete as shown in Fig. 8b.

    E ¼   reins

    ð3:5:1Þ

    E eff  ¼   retotal

    ¼   reins   1 þ C crð Þ ¼

      E 

    1 þ C cr ð3:5:2Þ

    C cr   t ;  t 0ð Þ ¼   ecr   t ð Þ

    eins   t ð Þ ¼etotal   t ð Þ eins   t ð Þ

    eins   t ð Þ   ð3:5:3Þ

    C cr   t ;  t 0ð Þ ¼  J t ;   t 

    0ð Þr eins   t ð Þeins   t ð Þ   ¼

     J t ;   t 0ð Þ   1E t ð Þ

    1E t 

    ð Þ¼ E t ð Þ J t ;   t 0ð Þ 1   ð3:5:4ÞThe creep behavior becomes more complicated when

    the concrete member is under variable loading such as

    thermally induced stresses which would change due to the

    variations of the temperature gradient and the mechanical

    properties. The variable loading problem can be solved

    using the superposition principle. At time n, the total load

    can be decomposed to n small increments (Drt). Each

    loading increment has its individual loading time (t 0  =   i)and loading duration (n  -   t 0  =  n  -   i) (Fig.  8c). Withoutconsidering creep, the total stress at  t  =  n,  r total(n) can be

    expressed as the summation of the load increments(Eq. 3.5.5). When creep is considered, each loading

    increment (Drt_cr) can be derived as in Eq. (3.5.6) and the

    total load can be expressed as shown in Eqs. (3.5.7) or

    (3.5.8). The actual overall stress release percentage due to

    all of the load increments can be expressed as the ratio of 

    rtotal_cr(n) and  rtotal(n). Thus, the overall creep coefficient

    and effective modulus can be derived as in Eqs. (3.5.9) and

    (3.5.10), respectively. In the FEM analysis, only basic

    creep was considered. Basic creep refers to the strain

    observed on sealed specimens due to sustained loading

    [26]. In reality, all the surfaces of the two 1.2 m3 were

    covered for entire 5 days after casting. The influences fromdrying effect are assumed negligible.

    rtotal   nð Þ ¼ Dr1 þ Dr2 þ Dr3 þ . . . þ Drn2 þ Drn1þ Drn

    ð3:5:5Þ

    rtotal   nð Þ ¼Xni¼1

    Dri

    1 1E (t ′) E (t )

    E eff

    σ

    σ  

    σ  

    t ′ t Time

       C  r  e  e  p  c  o  m

      p   l   i  a  n  c  e   J

       (   t ,   t

                      ′   )

    Time

       L  o  a   d

    0 1 2 3 n –2 n  –1 n 

    Load duration of

    σ  n–2

    σ  n

    n–1

    n–2

    3

    2

    1

    ε ins   ε 

    ε 

    total

    ε ins(t  )

    ε cr(t )

    (a) (b)

    (c)

    σ  ∆

    σ  ∆

    σ  ∆

    σ  ∆

    σ  ∆

    Fig. 8   Illustration of effective modulus, the creep compliance andvariable load decomposition

    Thermal analysis and adiabatic calorimetry for early-age concrete members 233

     1 3

  • 8/17/2019 Lin Yang Early-age cracking of concrete and calorimetry

    8/13

    Drt cr ¼   Drt1 þ C cr   n; t 0ð Þ ¼

      Drt

    E nð Þ J n; t 0ð Þ   ð3:5:6Þ

    rtotalcr   nð Þ ¼  Dr1

    E nð Þ J n; 1ð Þ þ  Dr2

    E nð Þ J n; 2ð Þ þ  Dr3

    E nð Þ J n; 3ð Þ þ . . .

    þ   Drn2E nð Þ J n; n 2ð Þ þ

      Drn1E nð Þ J n; n 1ð Þ

    þ   DrnE nð Þ J n; nð Þ   ð3:5:7Þ

    rtotal cr   nð Þ ¼Xni¼1

    Dri

    E nð Þ J t ¼ n; t 0 ¼ ið Þ   ð3:5:8Þ

    C cr overall   nð Þ ¼   rtotal   nð Þrtotal cr   nð Þ 1 ¼

    Pni¼1 DriPn

    i¼1Dri

    E nð Þ J n;ið Þ 1

    ð3:5:9Þ

    E eff   nð Þ ¼   E nð Þ1 þ C cr overall   nð Þ   ð3:5:10Þ

    It is also noted that relationship between the creepcompliances and loading could become nonlinear if the

    loading stress and strength ratio becomes very high [26];

    the creep deformation would be further increased due to

    nonlinear creep behavior at high stress-to-strength ratio

    [12]. In this study, since only linear creep behavior is

    considered, when stress-to-strength ratio is beyond 80 %,

    the current creep model would overestimate the thermal

    stresses, and this will be discussed in Sect. 5.

    Discussion of maturity method on this application

    The maturity method has been used to estimate in situ

    concrete strength since late 1940s. Many researchers have

    discovered that high-temperature curing may have a nega-

    tive effect on the long-term concrete strength gain. Carino

    and Lew [27] described the ‘‘crossover’’ effect due to high-

    temperature curing. They suggested that maturity method is

    more reliable on predicting the relative strength rather than

    absolute strength. Tepke et al. [28] concluded that high-

    temperature curing affect the strength–maturity relation-

    ship. Kim and Rens [29] experimented on three sets of 

    concrete cylinders in three different curing temperatures of 

    40, 50, and 60   C. Their results showed higher-temperature-

    cured samples exhibited lower ultimate strength at the

    equivalent age of 28 days. In this study, three sets of con-crete cylinders with the same mix design (Table 1) were

    cured at 23, 40, and 50   C. The strength development curves

    plotted versus equivalent age (Eq. 3.3) are shown in Fig. 9a.

    Concrete cylinders cured at 23 and 40   C showed very

    similar strength–maturity relationship, while the cylinders

    cured at 50   C showed lower strength. It suggests that

    maturity method works for concrete with curing temperature

    between 23 and 40   C but may not work for 50   C or higher

    temperature. For verification purposes, another batch of 

    concrete with the same mix design was cast and cured at 23,

    30, and 40   C. As shown in Fig. 9b, maturity method

    worked quite well up to 7 days of equivalent age. In massconcrete applications, the concrete temperatures are nor-

    mally higher due to the relatively larger member sizes. At

    the center of a mass concrete member, the temperature can

    be kept higher than 50   C for an extended period. However,

    since only the surface tensile stresses are critical, the tem-

    perature near the surface is of particular concern and the

    temperature is usually much lower due to external heat loss.

    For both 1.2-m concrete cubes constructed, the surface

    maximum temperatures were about 45–46   C and quickly

    decreased after the maximum temperatures were reached.

    To verify whether the concrete surface strength of these

    two cubes can be predicted using the maturity method,another compressive strength test was performed using a

    set of concrete cylinders (0.1 m  9 0.2 m) cured in a tem-

    perature history similar to the surface temperatures expe-

    rienced by the surfaces of the cubes (Fig.  10a). Similar to

    Fig. 4a, the strength development curve from cylinders of 

    35

    30

    25

    20

    15

    10

    5

    00 2 4

    23 °C

    40 °C

    50 °C

    23 °C

    30 °C

    40 °C

    Equivalent age/day

       C  o  m  p  r  e  s  s   i  v  e  s   t  r  e  n  g   t   h   /   M   P  a

    35

    30

    25

    20

    15

    10

    5

    0

       C  o  m  p  r  e  s  s   i  v  e  s   t  r  e  n  g   t   h

       /   M   P  a

    6 8 0 2 4

    Equivalent age/day

    6 8

    (a) (b)Fig. 9  Strength developmentcurves for concrete cylinderscured in three differenttemperatures

    234 Y. Lin, H.-L. Chen

     1 3

  • 8/17/2019 Lin Yang Early-age cracking of concrete and calorimetry

    9/13

    identical concrete cured at 23   C was also obtained. Fig-

    ure 10b shows that the compressive strength of the cylin-

    ders with this variable temperature curing can be predicted

    by the strength–maturity relationship. These results indi-

    cate that maturity method may not accurately predict thestrength for long-duration high-temperature curing at a

    constant 50   C or above, but it is applicable for the strength

    prediction of the concrete experiencing short-duration

    high-temperature curing, such as those experienced on the

    surface of the 1.2 m3.

    Simulation process and results

    The computation of thermally induced stresses for early-

    age concrete contains two parts: thermal analysis and stress

    analysis. Thermal analysis was first conducted using 3-di-

    mensional finite element method (FEM), and the calculated

    temperature compared quite well with the experimental

    measurements [14]. Figure 11   shows the temperature

    comparisons between the FEM predictions and the mea-

    surements at the side surface (5 cm inside the surface) and

    at the center of the 1.2-m concrete cube. The details of the

    temperature calculations were described in Lin and Chen

    [14]. The temperature profile results are used for the stress

    analysis herein.

    For thermal stress analysis, the complexity of variation

    in material properties and viscoelastic behavior required usto develop a subroutine ‘‘USDFLD’’ to account for the

    change of material properties and viscoelastic behavior at

    every time increment. For each individual element, the

    program first calculates the equivalent age (t e) and degree

    of hydration (a) based on the calculated temperature his-

    tory. The compressive elements and tensile elements are

    treated independently. The difference in creep behavior

    between elements in tension and compression is considered

    (Sect. 3.4). The modified B3 model is used to describe the

    tensile creep, and the modified double power law (M-DPL)

    is used to describe the compressive creep behavior. To

    simplify the analysis, it was programmed to check the

    maximum principal stress in each element at every time

    step to identify tensile and compressive elements.

    The creep stress loading magnitude changes because of 

    temperature variation. Load decomposition and superpo-

    sition rules were used to calculate the overall creep coef-

    ficient. The effective modulus was then used to incorporate

    0 1015

    20

    25

    30

    35

    40

    45

    50

    20

    Cube #123 °C

    23 °C (Fig. 4(a))

    Match curingCube #2

    Curing temperature

    30 40

    Time/h

    50 60 70 80 00

    5

    10

    15

    20

    25

    30

    35

    30 60 90 120 150 180

    Equivalent age/h

       C  o  m  p  r  e  s  s

       i  v  e  s   t  r  e  n  g   t   h   /   M   P  a

       T  e  m

      p  e  r  a   t  u  r  e   /   °   C

    (a) (b)Fig. 10 a  Surface temperaturehistories of the two cubes andthe variable curing temperature,b  measured compressivestrength of the specimens curedin different temperaturehistories

    010

    20

    30

    40

    50

    60

    70Cube 1

    Experiment (Center)

    FEM (Center)

    FEM (side)

    Ambient

    Experiment (side)

    Experiment (Center)

    FEM (Center)

    FEM (side)

    Ambient

    Experiment (side)

    Cube 2

    10

    0

    20

    30

    40

    50

    60

    70

    20 40 60 80

    Time/h

    100 120 0 20 40 60 80

    Time/h

    100 120

       T  e  m  p  e  r  a   t  u  r  e   /   °   C

       T  e  m  p  e  r  a   t  u  r  e   /   °   C

    (a) (b)Fig. 11   Center and the sidetemperature predictions of cube1 and cube 2

    Thermal analysis and adiabatic calorimetry for early-age concrete members 235

     1 3

  • 8/17/2019 Lin Yang Early-age cracking of concrete and calorimetry

    10/13

    both elastic deformation and viscoelastic deformation due

    to the creep behavior (Sect.  3.5). At each time step, the

    thermal stress was computed for each element based on the

    calculated thermal gradient, current elastic modulus, and

    overall creep coefficient. Finally, the equivalent age,

    degree of hydration, and stresses in principle directions for

    each element were stored for the next time step. Figure  12

    shows the programming algorithm of the stress analysis.The analysis used a fixed time increment of 1 h. The entire

    algorithm was executed for each individual element at

    every time step. For simplicity, the Poisson’s ratio and

    coefficient of thermal expansion (CTE) were assumed

    constants. Poisson’s ratio was assumed to be 0.2. CTE was

    experimentally obtained to be 8.53 microstrains per   C

    (Sect. 3.3). The frictional interaction between the bottom

    of the concrete cube and the wood base was neglected for

    simplicity.

    Stress analysis was performed for both 1.2-m concrete

    cubes using the ABAQUS program and the above FOR-

    TRAN subroutine. The model had 15,625 nodes and13,824 elements using 3-D 8-node linear element (C3D8R)

    with 5-cm element size. Results showed that due to the

    temperature evolution and the thermal expansion, the inner

    elements were expanding which caused the surface element

    to be in tension. The calculated surface tensile stress con-

    tour patterns for the two cubes are similar. As shown in

    Fig. 13   (cube 1), the tensile strength of the concrete was

    exceeded by maximum thermal stress at the center loca-

    tions of the edges (shown as gray color) at 16 h after

    concrete placement. The predicted tensile stresses and the

    estimated tensile strength at the critical locations for both

    cubes are compared in Fig. 14. The estimated tensile

    strength history was calculated using Eq. (3.1.3) for the

    concrete at that location; concrete strength was temperaturehistory dependent, and hence, location dependent. The

    FEM result showed that cube 1 was likely to crack at these

    locations because the maximum thermal stress exceeded

    the tensile strength. From the experimental observation,

    four large cracks with an approximate cracking length of 

    0.6 m were found at the center of the top edges (0.3 m on

    the top surface and 0.3 m downward to the side surfaces).

    Figure 13  shows the calculated stress distribution and the

    actual crack locations for cube 1. No other crack was found

    at side or bottom edges of cube 1. The reason can be the

    nonuniform strength distribution due to vibration com-

    paction; the top surface is typically found to be the weakestpart of the concrete cube [16]. For cube 2, the predicted

    maximum stress was shown meeting the estimated tensile

    strength at the critical locations (Fig. 14); however, no

    thermal crack could be visually identified on any surface of 

    cube 2. The reason for cube 1 to have larger stress mag-

    nitudes in comparison with cube 2 was because of a larger

    Start (time = n) 

    Return to start

       S  u   b  r  o  u   t   i  n  e  :   U   S   D   F   L   D

    Assemble global stiffness matrix and calculate element stressbased on the temperature profile and material properties.

    Store history values of calculated element equivalent age (t e ) and stress (σ).Time step Advance: n = n+1.

    Input Material properties:Modulus: E eff (n) Thermal expansion coefficient: 8.53 µ °C–1

    (Sect. 3.3)Poisson’s ratio; 0.2 (constant)

    Yes

    Calculate tensile modulus,E t (n)  [Eq.(3.2.2)]. Calculate compressive modulus,E c (n)  [Eq.(3.2.2)].

    Calculate each individual compressive creepcompliance, J(t,t ′ )   (t = 1,2,...,n ) using M-DPLmodel [Eq. (3.4.4)].Calculate the overall creep coefficient, C cr_overall[Eq. (3.5.9)]Calculate compressive effective modulus, E eff(n) 

    using Calculated E c (n) and Eq. (3.5.10).

    Calculate each individual tensile creepcompliance, J(t,t ′ )  (t = 1,2,...,n) usingmodified B3 model [Eq. (3.4.3)].Calculate the overall creep coefficient,Ccr_overall [Eq. (3.5.9)]Calculate tensile effective modulus, E eff(n) using Calculate E t (n) and Eq. (3.5.10).

    NoS(n–1) > 0 

    Load element stress history (S ) in principle direction.

    Load element temperature history and calculate its equivalent

    age (t e ) [Eq. (3.3)].and degree of hydration (α) [Eq. (3.2)].

    Fig. 12   Algorithm of the stressanalysis

    236 Y. Lin, H.-L. Chen

     1 3

  • 8/17/2019 Lin Yang Early-age cracking of concrete and calorimetry

    11/13

    s/MPa2.972.392.352.101.951.801.651.511.371.221.080.930.790.640.500.350.210.06

    Fig. 13  Predicted stress field of cube 1 at 16 h

    00

    0.5

    1

    1.5

    2

    2.5

    3

    3.5

    0

    0.5

    1

    1.5

    2

    2.5

    3

    3.5

    20

    Cube 1 Cube 2

    Tensile strength

    Thermal stress

    Tensile strength

    Thermal stress

    40 60

    Time/h

       T  e  n  s   i   l  e  s   t  r  e  s  s   /  s   t  r  e  n  g   t   h   /   M   P  a

       T  e  n  s   i   l  e  s   t  r  e  s  s   /  s   t  r  e  n  g   t   h   /   M   P  a

    80 100 120 0 20 40 60

    Time/h

    80 100 120

    (a) (b)Fig. 14   Comparison of calculated thermal stress and

    estimated tensile strength

    0 10 20 30

    Time/h

       M   i  c  r  o  -   S   t  r  a   i  n

    40 50

    0 10 20 30

    Time/h

    40 50 0 10 20 30

    Time/h

    40 50

    00

    50

    100

    150

    200

    250

       M   i  c  r  o  -   S   t  r  a   i  n

    0

    50

    100

    Calculated

    Experiment

    Calculated

    Experiment

    Calculated

    Experiment

    Calculated

    Experiment

    150

    200

    250

       M   i  c  r  o  -   S   t  r  a   i  n

    0

    50

    100

    150

    200

    250

       M   i  c  r  o  -   S   t  r  a   i  n

    0

    50

    100

    150

    200

    250

    300

    10 20 30

    Time/h

    40 50

    (a) (b)

    (c) (d)

    Fig. 15   Comparison of calculated and measured strainchanges at the locations near theconcrete surfaces  a  cube 1—location A,  b  cube 2—locationA,  c  cube 2—location B andd  cube 2—location C

    Thermal analysis and adiabatic calorimetry for early-age concrete members 237

     1 3

  • 8/17/2019 Lin Yang Early-age cracking of concrete and calorimetry

    12/13

    ambient temperature drop at the night right after the cube 1

    was constructed (see Fig.  11); an 18   C (Fig. 11) drop in

    ambient temperature at the first night after cube 1 con-

    struction caused a significant increase in thermal stresses.

    It is noted that the current FEM model assumes a creep

    model that is linear to the applied stresses. The nonlinear

    creep behavior due to high stress-to-strength ratio is not

    considered in the current model. It was observed by Atrushi[12] that when applied tensile stress/strength ratio was

    about 80 %, the creep coefficient became nonlinear with

    respect to the applied stress; more creep strain was

    observed at higher stress-to-strength ratio. Therefore, it is

    assumed that the current creep model is only able to esti-

    mate the allowable thermal stress up to 80 % of the tensile

    strength. Because of the linear assumption, the estimation

    of the stress at the level higher than 80 % is considered to

    be conservative (the estimation is higher than the actual

    stress value) using the current model. Although the gray

    region of Fig.  13  shows thermal stress exceeded the tensile

    strength, it can only be used as a qualitative indication of high cracking probability. The current modified B3 model

    is assumed to be only valid to obtain the thermal stress up

    to 80 % of the tensile strength. The nonlinear creep

    behavior due to high stress-to-strength ratio needs further

    investigation.

    The finite element calculated strains were verified with

    the measurements. As mentioned in Sect.  2, concrete strain

    histories were measured at several locations using vibrating

    wire gages. For cube 1, the strain history was measured at

    location A. For cube 2, strain values were measured at

    locations A, B, and C. The locations are marked in Fig.  2.

    The calculated strain histories at these locations were foundto be reasonably close to the measurements as shown in

    Fig. 15. The best match was shown at location C of cube 2

    (2.5 cm in depth at the center of the top surface) which was

    far from any steel reinforcing bar (shown in Fig. 15d). The

    calculated strain histories (Fig.  15a) showed some devia-

    tions from the experimental measurements, which were

    possibly due to the restrains of concrete movement pro-

    vided by the parallel steel reinforcing bars close to the

    sensor during cube 1 testing. On the other hand, during the

    cube 2 testing (shown in Fig.  15b, identical location), there

    was no parallel reinforcing bar attached to the sensor. It is

    noted that the current FEM calculation neglected the effects

    of autogenous shrinkage and external restraint; therefore, it

    was not included in the predicted strains in Fig.  15. The

    autogenous shrinkage of this particular concrete mixture

    was measured to be approximately 10 microstrains in sealed

    condition after the first 7 days. Hence, the influence of the

    autogenous shrinkage was neglected in the calculation. In

    general, mass concrete structures are constructed using

    concrete with low cement contents that would typically

    produce low autogenous shrinkage.

    Conclusions

    This paper describes a method to perform thermal stress

    analysis using ABAQUS program with the aid of user

    subroutines. The developed subroutine program uses the

    degree of hydration to estimate the variable elastic mod-

    ulus and strength developments. Concrete tensile creep

    and compressive creep behavior were included using astep-by-step incremental calculation algorithm. The

    influences from loading age and temperature effects were

    considered in each time increment of the creep models. It

    was assumed that the current creep model is able to cal-

    culate the thermal stress up to 80 % of the concrete

    strength. The finite element simulations were verified by

    the experimental data from two 1.2-m concrete cubes

    testing. Strain deformations at the locations near the

    concrete cube surfaces were measured and correlated

    reasonably well with the calculated results.

    The concrete cubes have high tensile stresses at the

    surfaces, especially at the center of the edges. The tensilestrength development of the concrete at surface locations

    can be estimated using the maturity method, and the

    cracking risk could be assessed using the stress-to-strength

    ratio obtained at the critical locations. Four visible cracks

    were found perpendicular to the top four edges on cube #1

    as predicted, due to a relatively high ambient temperature

    drop at the first night after construction. The method

    developed can be used to estimate the thermally induced

    stress of concrete members so that precautions can be

    implemented prior to concrete casting in order to prevent

    unexpected cracking.

    Acknowledgements   The authors acknowledge the support providedby the West Virginia Transportation Division of Highways andFHWA for the research project WVDOH RP#257. Special thanks areextended to our project monitors, Mike Mance, Donald Williams, andRyan Arnold of WVDOH. The authors also appreciate the assistancefrom Alper Yikici, Zhanxiao Ma and Jared Hershberger and theWVDOH concrete technicians from Material, Control, Soil andTesting Division for the construction of the 1.2-m concrete cubes.

    References

    1. Nagy A, Thelandersson S. Material characterization of youngconcrete to predict thermal stresses. Thermal cracking in concreteat early ages. In: Springenschmid R, editor. E&FN Spon, 2–6,ISBN: 0419187103, 1994.

    2. Gutch A, Rostasy FS, Young concrete under high tensile stres-ses—creep, relaxation and cracking. In: Springenschmidt R,editor. Thermal cracking in concrete at early ages. Proceedings of the international RILEM symposium, E & FN Spon, London,1995. p. 111–118.

    3. Lawrence MA, Tia M, Ferraro CC, Bergin M. Effect of early agestrength on cracking in mass concrete containing different sup-plementary cementitious materials: experiment and finite-elementinvestigation. J Mater Civ Eng. 2012;24(4):362–72.

    238 Y. Lin, H.-L. Chen

     1 3

  • 8/17/2019 Lin Yang Early-age cracking of concrete and calorimetry

    13/13

    4. Tia M, Lawrence A, Ferraro C, Smith S, Ochiai E. Developmentof design parameters for mass concrete using finite elementanalysis. The Florida Department of Transportation, reportnumber: 00054863, 2010.

    5. Waller V, d’Aloı̈a L, Cussign F, Lecrux S. Using the maturitymethod in concrete cracking control at early ages. Cement ConcrCompos. 2004;26:589–99.

    6. Wu S, Huang D, Lin FB, Zhao H, Wang P. Estimation of cracking risk of concrete at early age based on thermal stressanalysis. J Therm Anal Calorim. 2011;105:171–86.

    7. Tia M, Lawrence A, Ferraro C, Do TA, Chen Y. Pilot project formaximum heat of mass concrete. The Florida Department of Transportation, report number: 00093793, 2013.

    8. Bažant ZP. Mathematical modeling of creep and shrinkage of concrete. Chichester: Wiley; 1988.

    9. Østergaard L, Lange D, Altoubat SA, Stang H. Tensile basiccreep of early-age concrete under constant load. Cem Concr Res.2001;31(12):1895–9.

    10. Wei Y, Hansen W. Tensile creep behavior of concrete subject toconstant restraint at very early ages. J Mater Civ Eng.2013;25(9):1277–84.

    11. Bažant ZP, Baweja S. Creep and shrinkage prediction model foranalysis and design of concrete structures: model B3. ACI SP–194. Detroit: American Concrete Institute; 2000. p. 1–83.

    12. Atrushi DS. Tensile and compressive creep of early age. Ph.D.dissertation, Trondheim, Norway: Department of Civil Engineer-ing, The Norwegian University of Science and Technology, 2003.

    13. Luzio GD, Cusatis G. Soli Solidification–microprestress–mi-croplane (SMM) theory for concrete at early age: theory, vali-dation and application. Int J Solids Struct. 2012;50(6):957–75.

    14. Lin Y, Chen HL. Thermal Analysis and Adiabatic Calorimetryfor Early-age Concrete Members. J Therm Anal Calorim.2015;122(2):937–45.

    15. Freiesleben Hansen P, Pedersen EJ. Curing of concrete structures.Draft DEB—guide to durable concrete structures, Appendix 1, 1985.

    16. Yikici A, Chen HL. Effect of temperature-time history on con-crete strength in mass concrete structure.TRB 92nd annual con-ference proceeding, Transportation Research Board of theNational Academies, Washington, DC, 2013.

    17. Mills RH. Factors influencing cessation of hydration in water curedcement pastes. Special Report No. 90, Proceedings of the Sym-posium on the Structure of Portland Cement Paste and Concrete,Highway Research Board, Washington DC, 1966;406–424.

    18. Schutter GD. Fundamental study of early age concrete behavioras a basis for durable concrete structures. Mater Struct.2002;35:15–21.

    19. Wight JK, MacGregor JG. Reinforced concrete: mechanics anddesign. 5th ed. Pearson Prentice Hall: Upper Saddle River; 2009.p. 58–9.

    20. McCullough BF, Rasmussen RO. Fast-track paving: concretetemperature control and traffic opening criteria for bonded con-crete overlays. US: FHWA 1998, Final Report.

    21. Umehara H, Uehara T. Effect of creep in concrete at early ages onthermal stresses, thermal cracking in concrete at early ages. In:Springenschmidt R, editors. Thermal cracking in concrete at earlyages. Proceedings of the international RILEM symposium, E&FNSpon, London, 1995. p. 79–86.

    22. Hauggaard AB, Darkled L, Hansen PF, Hansen JH, ChristensenSL, Nielsen A. HETEK, Control of early age cracking in con-crete, phase 3: creep in concrete. Danish Road Directorate, Lynby, 1997.

    23. Kanstad T, Bj/stergaard Ø, Sellevold EJ. Tensile and compres-sive creep deformations of hardening concrete containing mineraladditives. Mater Struct. 2012;46(7):1167–82.

    24. Hilaire A, Benboudjema F, Darquennes A, Berthaud Y, Nahas G.Modeling basic creep in concrete at early-age under compressiveand tensile loading. The international conference on structuralmechanics in reactor technology, New Dekhi India, 2011;269,p. 222–30.

    25. Bažant ZP, Osman E. Double power law for basic creep of concrete. Mater Struct. 1976;9:49.

    26. Bažant ZP, Xi Y. Continuous retardation spectrum for solidifi-cation theory of concrete creep. J Eng Mech. 1995;121(2):281–8.

    27. Carino NJ, Lew HS. The maturity method: from theory toapplication. Washington: Structure Congress & Exposition; 2001.

    28. Tepke DG, Tikalsky PJ, Scheetz BE. Concrete maturity fieldstudies for highway application. J Transp Res Board.2004;1893:26–36.

    29. Kim T, Rens KL. Concrete maturity method using variabletemperature curing for normal-strength concrete mixes. J MaterCiv Eng. 2008;20:20.

    Thermal analysis and adiabatic calorimetry for early-age concrete members 239

     1 3