17
International Journal for Multiscale Computational Engineering, 7(2)xxx–xxx(2009) Multiscale Model for Temperature Distribution in Hydrating Concrete V. ˇ Smilauer & T. Krejˇ ı Faculty of Civil Engineering, Department of Mechanics, Czech Technical University in Prague, Th´ akurova 7, 166 29 Prague 6, Czech Republic ABSTRACT Temperature rise in hydrating concrete presents a formidable problem that may lead to sig- nificant acceleration of hydration kinetics, early-age cracking, and decreased durability. Mul- tiscale formulation is proposed, coupling a cement hydration model on the microscale with the finite element method (FEM) heat conduction problem on the macroscale. Although dis- crete hydration model predicts heat evolution controlled by macroscale temperature, the FEM satisfies heat balance equation during thermal conduction. Two- and three-dimensional vali- dations show a reasonable temperature conformity with an access to the local quantities, such as a degree of hydration. KEYWORDS concrete temperature, hydration heat, heat conduction, cement hydration model *Address all correspondence to [email protected], Phone 420-224-354-369 1543-1649/08/$35.00 c 2009 by Begell House, Inc. 1

Multiscale Model for Temperature Distribution in Hydrating Concrete

Embed Size (px)

Citation preview

Page 1: Multiscale Model for Temperature Distribution in Hydrating Concrete

International Journal for Multiscale Computational Engineering, 7(2)xxx–xxx(2009)

Multiscale Model for Temperature Distributionin Hydrating Concrete

V. Smilauer & T. KrejcıFaculty of Civil Engineering, Department of Mechanics, Czech Technical University in Prague, Thakurova 7, 166

29 Prague 6, Czech Republic

ABSTRACT

Temperature rise in hydrating concrete presents a formidable problem that may lead to sig-nificant acceleration of hydration kinetics, early-age cracking, and decreased durability. Mul-tiscale formulation is proposed, coupling a cement hydration model on the microscale withthe finite element method (FEM) heat conduction problem on the macroscale. Although dis-crete hydration model predicts heat evolution controlled by macroscale temperature, the FEMsatisfies heat balance equation during thermal conduction. Two- and three-dimensional vali-dations show a reasonable temperature conformity with an access to the local quantities, suchas a degree of hydration.

KEYWORDS

concrete temperature, hydration heat, heat conduction, cement hydration model

*Address all correspondence to [email protected], Phone 420-224-354-369

1543-1649/08/$35.00 c© 2009 by Begell House, Inc. 1

Administrator
Note
Author please check up the text and If you find errors - please, specify where its in text. There aren't other questions. Thank you.
Page 2: Multiscale Model for Temperature Distribution in Hydrating Concrete

2 SMILAUER AND KREJCI

1. INTRODUCTION

Concrete performance is affected by temperaturehistory in many ways. With no regard to well-controlled laboratory conditions, the mixing, plac-ing, and curing of concrete occur worldwide undervarious environmental conditions. The understand-ing of underlying phenomena during concrete ma-turing and performance follows different attitudesof material and civil engineers and, in fact, presentsa multiscale and interdisciplinary approach.

From a material point of view, concrete is a liv-ing heterogeneous material whose microstructureformation is considerably affected by temperature.It was recognized that higher curing temperaturespeeds up the rate of reactions and promotes highearly strength [1]. However, such microstructureis characteristic by inhomogeneous placing of hy-drates, coarser porosity, and higher density of cal-cium silicate hydrates (C-S-H) [2], resulting in adecreased long-term strength and durability. Therecent advancement in high-strength concrete, uti-lization of secondary cementitious materials (SCM),and the use of admixtures has made the issue of con-crete material even more complex [3].

The prediction of thermal stresses based on tem-perature history is a well-established concept in theframework of structural mechanics. A low thermalconductivity of concrete, the exothermic hydrationprocess, and surrounding temperature cause tem-perature gradients accompanied by strains, espe-cially in massive concrete elements. During the con-crete cooling, tensile stresses often appear due to mi-nor restraints in the system, which may lead to a mi-crocrack formation and decreased durability [4, 5].

Several researchers formulated predictive mod-els for temperature evolution under adiabatic con-dition, in the form of exponential function [6] orhydration model [7]. Proper validation became astepping stone for a multiscale approach, recogniz-ing the heterogeneous temperature distribution onthe macroscale. The phenomenological hydrationmodel [8] or later the more sophisticated affinity hy-dration model [9] have been previously employedin the coupled micro/macrothermal analysis.

The formulation of fully coupled thermochemo-mechanical models was nevertheless proposed [10–13] and is remarkable for the amount of input pa-rameters. Several of them have to be obtained ex-

perimentally, thus weakening the potential for ageneral-purpose tool.

Instead of direct modeling of the crack formationusing highly sophisticated models, engineers oftenproposed guidelines to produce high-quality con-crete satisfying elementary rules [1, 14]. Such an ap-proach does not require detailed modeling and re-lies more on experience and expertise. For example,the temperature anywhere in the concrete elementshould not exceed 70C during hardening to avoidpossible risk of delayed ettringite formation accom-panied by expansion [15]. The cooling rate shouldnot exceed 3C/h during the first 24 h and specialcaution is necessary in cold weather when concretetemperature falls bellow 4C [14].

Proposed simulation methodology relies ona multiscale concept, introducing two separablescales, as follows:

1. Cement paste level found on the scale between1 and 100 µm. Clinker minerals, gypsum, CH1,CSH, other minor phases, and capillary poros-ity are present,

2. Structural level represents the span from cen-timeter up to several meters. Concrete foundat this level is comprised mainly from cementpaste, SCM as well as fine and coarse ag-gregates. The topology of concrete structurewith assigned boundary conditions comple-ments material information.

A thermochemical multiscale model is formu-lated to predict realistic temperature evolution anddistribution in arbitrary concrete member undervarying boundary conditions. The emphasis isgiven on the material scale, taking into account ce-ment chemical and physical properties. A discreteCEMHYD3D hydration model based on cellular au-tomata is used to simulate microstructure evolutionof hydrating cement paste [16]. A conventional tran-sient heat conduction problem is solved through thefinite element method (FEM) at the macroscale level[17]. The formulation relies on intrinsic materialdata, such as reaction enthalpy, resulting in a ver-satile and extensible engineering tool.

1 Using standard notation of cement chemistry.

International Journal for Multiscale Computational Engineering

Page 3: Multiscale Model for Temperature Distribution in Hydrating Concrete

MULTISCALE MODEL FOR TEMPERATURE DISTRIBUTION 3

2. CEMENT HYDRATION MODEL ON THEMICROSCALE

The microscale of cement paste is a stepping stonefor multiscale analysis, relying on the assumptionthat aggregates do not contribute significantly to thecement hydration process. Therefore, the heat evo-lution of concrete may be linked directly to the ce-ment paste level. Any hydration model designedfor the prediction of heat release has to capture un-derlying chemical reactions as well as the topol-ogy of cement grains. The latter controls the ini-tial surface area of cement grains with the conse-quences to hydration kinetics and later stages ofhydration due to the deposition of hydrates typ-ically around cement grains, preventing capillarywater from further reaction. Hydration modeling isbased on an open-source, discrete hydration modelCEMHYD3D [16], allowing one to include directlythe following:

• particle size distribution of cement

• chemical composition of cement (C3S, C2S,C3A, C4AF, gypsum)

• actual temperature and its history

• water regime (under saturated and sealed con-ditions)

• water-to-cement ratio (w/c) and the amount ofcement in concrete

• supplementary cementitious materials (silica,slag, fly ash)

The microstructure of cement paste, referred toas a representative volume element (RVE) [18], isapproximated in the CEMHYD3D model by voxelswith the size of 1 × 1 × 1 µm. Each of them rep-resents one chemical phase, either in solid or dis-solved state. Figure 1 shows an example of initialand well-hydrated RVE of 50× 50× 50 µm. The ini-tial periodic RVE is reconstructed by throwing dig-ital spheres following cement particle size distribu-tion (PSD). Assessment of real cement PSD is usu-ally not available; hence, it becomes more conve-nient to express PSD curve by a single and accessiblevariable–Blaine fineness. Rosin-Rammler cumula-tive distribution G(d) was fitted to a NIST referencedatabase for Dyckerhoff cements, see [16], with theresults

G(d) = 1− e−bdn

, limd→∞

G(d) = 1 (1)

n = −8.333× 10−4 fineness + 1.1175 (2)

b = 7.54× 10−4 fineness− 0.143 (3)

where d (in microns) is the particle diameter andBlaine fineness is expressed in meters squared perkilogram. Figure 2 displays cumulative distribu-tion for coarse and fine cements with correspondingBlaine fineness of 250 and 500 m2/kg, according toEqs. (1)–(3). The amount of cement in the RVE iscontrolled by w/c and gypsum content.

FIGURE 1. RVE 50 × 50 × 50 µm, w/c = 0.25, initial (left) and at the degree of hydration of 0.63 (right). Dominantphases include water-filled porosity (black), C3S (red), C2S, C3A, C4AF, gypsum, CH (blue), and C-S-H (violet)

Volume 7, Number 2, 2009

Page 4: Multiscale Model for Temperature Distribution in Hydrating Concrete

4 SMILAUER AND KREJCI

0

0.2

0.4

0.6

0.8

1

0 20 40 60 80 100

Cum

ulat

ive

mas

s fr

actio

n [-

]

Cement grain diameter [µm]

500 m2/kg300 m2/kg250 m2/kg

FIGURE 2. Assumed Rosin-Rammler cumulative distri-bution for cements with three Blaine finenesses accordingto Eqs. (1)–(3)

The chemical reactions are modeled on the ba-sis of cellular automata rules, describing the pro-cess of dissolution, transport, nucleation, and re-action. Liberated heat is calculated directly fromevolving microstructure assigning enthalpy to eachchemical reaction [7]. For example, the hydration ofthe most common mineral C3S of Portland cementyields 517 J/g of dissolved C3S

C3S(1) +5.3H(1.34) → C1.7SH4(1.52)

+1.3CH(0.61)(4)

where the numbers in parenthesis correspond to re-action volumes at 20C. In the model, saturated orsealed curing conditions are considered with no fur-ther moisture effect on hydration. Such simplifica-tion is supported by the fact that relative humid-ity (RH) seldom drops below 90% even in exteriorparts within a few days when exposed to ambient60% RH [1]. Therefore, a sealed curing boundarycondition is a realistic assumption when no con-siderable water evaporation takes place. If the RHdrops below approximately 80% RH, them the hy-dration ceases.

The temperature effect on hydration kinetics isdetermined from Arrhenius equations simultane-ously for all implemented reactions. The charac-teristic time τ is modified according to the maturityprinciple [19]

τ(T ) = τ(T0) exp[Ea

R

(1T0− 1

T

)](5)

where T0 = 298.15 K is a reference temperature,T corresponds to arbitrary homogeneous tempera-ture, R = 8.314 Jmol−1K−1 represents the univer-sal gas constant and Ea stands for an apparent ac-tivation energy, for ordinary Portland cement typi-cally around 40 kJmol−1 [20]. Equation (5) proves astrong temperature effect on heat release rate (e.g.,going from 25 to 50C is a speedup factor of 3.5 forhydration progress and heat release rate).

The cycles in cellular automata represent nomeaningful scale of time. Original linear mappingof cycles showed considerable disagreement withexperiments [21]. Parabolic mapping, based on theKnudsen’s parabolic dispersion model, was foundappropriate under different curing conditions andcement types [22]. Knudsen’s model assumes thatthe diffusion of ions takes the control over the hy-dration rate

τ = t0 + β cycle2 (6)

The parameter β is usually found in the interval1 × 10−4 ≤ β ≤ 1.1 × 10−3 h/cycle2 [22, 23], whilet0 represents extra dormant time. Both parameterscontain the influence of admixtures, alkalis, cementimpurities, etc. A standard induction period of Port-land cement is included directly in the hydrationmodeling and sets t0 = 0 h.

The size of RVE plays a fundamental role in thecalculations [24]. For the cement hydration, one hasto consider at least the effect of truncated PSD, ce-ment fineness, and w/c. The maximum accommo-dated particle diameter must be truncated to ap-proximately half of RVE edge to allow finer grainplacement (i.e., to maximum diameter 5 µm at 10 ×10 × 10, 25 µm at 50 × 50 × 50 µm, and 50 µm at100 × 100 × 100 µm). Coarse cement with high w/cis the most susceptible combination, where trunca-tion and PSD renormalization creates in fact a finecement (Fig. 2).

To shed a light on the effect of RVE size, five ran-dom realizations were generated for ordinary coarsePortland cement with Blaine fineness 250 m2/kgand w/c = 0.5. Figure 3 displays the evolution ofdegree of hydration, tightly related to released heat.Hydration at 20C shows that 10× 10× 10 µm pro-ceeds faster due to truncated large particles than ref-erence at 100× 100× 100 µm. The reasonable selec-tion is 50 × 50 × 50 µmfor the scatter and computa-tional speed.

International Journal for Multiscale Computational Engineering

Page 5: Multiscale Model for Temperature Distribution in Hydrating Concrete

MULTISCALE MODEL FOR TEMPERATURE DISTRIBUTION 5

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0 20 40 60 80 100

Deg

ree

of h

ydra

tion

[-]

Time of hydration [h]

10 x 10 x 10 µm50 x 50 x 50 µm

100 x 100 x 100 µm

FIGURE 3. Effect of RVE sizes on degree of hydration,each with five random realizations

Nowadays, SCM such as slag or fly ash areblended with Portland cement due to economi-cal, ecological, and durability reasons. The com-mon level of substitution reaches up to 35% bymass, although much higher replacements may ex-ist [1]. The consequence is decreased potential heatof blended cement with slow transformation of sec-ondary binders to hydration products. Because ofvery low heat contribution, if any, from the reactionof SCM generally occurs during the first hydrationdays [25–28], the SCM are considered as an inertfiller in the simulations.

3. HEAT TRANSPORT ON THE STRUCTURALSCALE

Transient three-dimensional (3D) heat conduction issolved on the structural scale. Energy balance on adifferential element yields [17, 29]

−∇T q(x) + Q(x, t) = ρ(x)cp(x)∂T (x, t)

∂t(7)

where q(x) (in Watts per meters squared) is a heatflux originating in conduction, convection, or radi-ation, Q(x, t) [W/m3] represents given heat sourcefrom the CEMHYD3D model, ρ(x) (in kilogramsper cubic meter) stands for a concrete density, cp(x)[Jkg−1K−1] is a specific heat capacity and T (x, t) (indegrees Kelvin) represents unknown temperaturefield.

The introduction of Fourier’s law into Eq. (7)leads to the formulation of 3D heat conductionequation in terms of temperature field

λ(x)∆T (x) + Q(x, t) = ρ(x)cp(x)∂T (x, t)

∂t(8)

where λ(x) (in Watts per meter per degree Kelvin) isa thermal conductivity of isotropic material.

Dirichlet’s, Neumann’s, or Cauchy’s boundaryconditions can be associated with Eqs. (7) and (8).Treatment of heat convection on a concrete surfaceis conveniently described by Newton’s law

n(x)T q(x) = h(x) (T (x)− T∞) , x ∈ Γc (9)

where n(x) is a unit normal vector pointing out-wards from the surface, h(x) is a heat transfer coeffi-cient and T∞ is the ambient temperature far enoughfrom the boundary layer. The solution of Eq. (8)requires prescribed initial temperature T (x, 0) overthe body

T (x, 0) = T (x, 0), x ∈ Ω (10)

The analytical solution of Eq. (8) is not amenablefor more complicated cases. Instead, FEM is em-ployed, starting with the formulation of a weakproblem [17]. A linear set of algebraic equations willbe obtained in the form

Cr + Kr = p (11)

where r represents temperature nodal derivativeswith respect to time, C is a capacity matrix, andK is a conductivity matrix [17]. The heat load vec-tor p comprises heat flow on the boundary and heatsource in the domain. The solution of the first-ordertransient problem of Eq. (11) can be accomplishedby many techniques (e.g., finite differences or modesuperposition). Here, the time domain is discretizedby a time step

∆tn = tn+1−tn, n=1, 2 . . . (time steps−1) (12)

and the solution proceeds by the v-form of trape-zoidal scheme [30]. Crank-Nicolson integrationscheme is used further. It should be noted that heatsource Q(x, tn+1) is determined from average tem-perature at the previous time step tn thus bringingsmall inaccuracy which remains unbalanced in thecalculation.

Volume 7, Number 2, 2009

Page 6: Multiscale Model for Temperature Distribution in Hydrating Concrete

6 SMILAUER AND KREJCI

Figure 4 displays adopted coupling between twoscales. The heat source Q(x, t) represents the hy-dration heat of cement paste determined from theCEMHYD3D model. Because the temperature has astrong effect on released heat, it becomes necessaryto couple hydration model on the microscale withthe structural model on the macroscale in terms oftemperature and liberated hydration heat.

Instead of having separate microstructures ineach Gauss point, they can be assigned directly tofinite element(s) or a broader selection. The tem-perature for CEMHYD3D is taken as an averagefrom corresponding Gauss points. Such simplifi-cation is advantageous in reduced computationaltime when finite elements are expected to exhibitsimilar temperature field. However, the same heatpower Q(x, t) passed to associated macroscopic do-main generally leads to underprediction of maxi-mum temperature and overprediction of minimumtemperature. Proper identification of elements withexpected similar field is an important point whenreduction in computational time is required.

The multiscale model is implemented in a gen-eral FEM open-source package TRFEL,2 suited fortransport problems.

4. THERMAL PROPERTIES OF CONCRETE

The simulation of heat conduction problems re-quires concrete heat conductivity, capacity and heattransfer coefficient. The hydration process changesthe volume fraction of evaporable water, unhy-drated cement, and hydrates. Therefore, thermalproperties vary with concrete composition and de-gree of hydration. Figure 5 displays concrete con-ductivity for mature ordinary concrete dependingon its unit mass and saturation conditions accord-ing to [1] and former Czech standard CSN 731208.The latter considers 1.5 for dry and 1.7 Wm−1K−1

for saturated concrete.Concrete can be considered at the first approxi-

mation as a two-phase medium comprising cementpaste and aggregates. Heat capacity of cement pasteis significantly affected by the amount of water withheat capacity 4.18 Jg−1K−1. The cement powderhas typically 0.75 Jg−1K−1 [31]. Detailed studydiscovered that fresh cement paste corresponds

2 http://mech.fsv.cvut.cz/ ˜sifel

FIGURE 4. Coupling between the level of cement pasteand structure

0

0.5

1

1.5

2

2.5

3

3.5

1500 1750 2000 2250 2500

Con

cret

e co

nduc

tivity

[Wm

-1K

-1]

Concrete unit mass [kgm-3]

Saturated concreteExposed to weatherProtected from weatherCSN 731208

FIGURE 5. Thermal conductivity of mature concreteaccording to a Czech standard and [1]

to heat capacities 1.55, 1.73, and 1.89 Jg−1K−1 forw/c = 0.3, 0.4, and 0.5, respectively [31]. The law ofmixture can be used for the estimation of heat ca-pacity for concrete. Figure 6 shows heat capacityfor cement paste according to [31] and for concretecontaining 85% of granite aggregate by mass. For-mer Czech standard CSN 731208 declares 0.84 and0.87 Jg−1K−1 for dry and saturated mature concrete,respectively.

Different heat transfer coefficients were re-ported in the literature, e.g., [9] considered h =11.63 Wm−2K−1 for concrete exposed to air while

International Journal for Multiscale Computational Engineering

Page 7: Multiscale Model for Temperature Distribution in Hydrating Concrete

MULTISCALE MODEL FOR TEMPERATURE DISTRIBUTION 7

0

0.5

1

1.5

2

0 0.2 0.4 0.6 0.8 1

Hea

t cap

acity

[Jg-1

K-1

]

Degree of hydration [-]

Cement paste

Concrete

W/c=0.5W/c=0.4W/c=0.3

FIGURE 6. Heat capacity for cement paste and concretepredicted by the law of mixture

h = 9.3 Wm−2K−1 for the surface surrounded by aplywood formwork. [8] used h = 9.7 Wm−2K−1

for the heat transfer between formwork and am-bient air. Table 1 summarizes heat transfer coeffi-cient for concrete according to former Czech stan-dard CSN 731208. When concrete surface is coveredwith a thin layer of other material, such as a form-work, heat transfer coefficients add inversely. Theoverall heat transfer coefficient is approximated bya harmonic mean

hi =λi

ti(13)

1h

=n∑

i=1

1hi

=1h1

+t2λ2

+ . . . (14)

where ti represents the thickness of a layer.

5. VALIDATION

5.1 Isothermal Calorimetry of Cement Paste

Validation of hydration model starts on a paste pre-pared from ordinary Portland cement CEM I 42.5 R,produced at location Mokra, Czech Republic. Ce-ment was mixed with water at w/c = 0.5 and in-serted in an isothermal calorimeter at 25C. TheCEMHYD3D model was calibrated for parametersβ and t0 according to Eq. (6). Initial RVE 50 × 50 ×50 µm was reconstructed and hydrated with inputparameters summarized in Table 2.

TABLE 1. Heat transfer coefficient for concrete exposedto air or water according to the standard CSN 731208

Concrete in contact with h (Wm−2K−1)air in enclosed environment 8air during summer 15air during winter 23water 31

The calibration provided β = 7.8×10−4 h/cycle2

and t0 = 1.2 h. Both the reference model and ex-perimental temperatures are identical; hence, the ac-tivation energy is irrelevant. Results in Fig. 7 ex-hibit a good overall match. However, the discrep-ancy up to 10 hours is a consequence of simplifiedassumption in which the dissolution of clinker min-erals is controlled by a power function of C-S-H [23].An effect of electric double-layer, ion concentration,and autocatalytic nucleation of C-S-H is not incor-porated directly and surely presents one of the mostchallenging tasks in the hydration modeling. An ob-served difference is acceptable for multiscale simu-lation because the majority of heat is released afterinitial period. Second discrepancy at later hydrationstage is attributed to the sensitivity and stability ofthe calorimeter unit. Potential hydration heat Qpot,determined from t = 130 h, corresponds to realistic520.5 J/g of cement.

0

50

100

150

200

250

300

350

400

450

1 10 100 1000

Rel

ease

d he

at [J

/gce

m]

Time of hydration [h]

t=130 hQ=330.02 J/gDoH=0.634Qpot=520.5 J/g

CalorimeterCEMHYD3D

FIGURE 7. Isothermal calorimetry and CEMHYD3Dsimulation at 25C, w/c = 0.5

Volume 7, Number 2, 2009

Page 8: Multiscale Model for Temperature Distribution in Hydrating Concrete

8 SMILAUER AND KREJCI

TABLE 2. Cement properties and simulation parameters

Component C3S C2S C3A C4AF GypsumMass content without gypsum 0.612 0.126 0.070 0.10 -Normalized mass content without gypsum 0.683 0.138 0.083 0.096 -Volume content with gypsum 0.649 0.131 0.078 0.092 0.05Blaine fineness 306 m2/kgAutocorrelation NIST files cm115x2f

5.2 Cube of Self-Compacting Concrete

The temperature evolution of self-compacting con-crete was measured at Czech Technical Universityin Prague. A cube of 1 × 1 × 1 m was casted at onestep in the shelter, protecting the specimen from sunradiation and rain. Nine equally spaced tempera-ture gauges were embedded in the horizontal direc-tion across the cube thickness, the outer sensor wasplaced 30 mm from the surface (Fig. 8). Two tem-perature gauges measured ambient air temperatureand all data were automatically gathered for the du-ration of at least six days.

The model description is given by Fig. 8. Thetop surface was in direct contact with atmosphere,therefore the heat transfer coefficient was chosenh = 19 Wm−2K−1 as the average between winterand summer season, Table 1. The thickness of form-work plywood in the vertical direction was 15 mm,therefore,

h =(

119

+0.0150.13

)−1

= 5.95 Wm−2K−1 (15)

During pilot simulation, the difference between thecenter and surface temperatures revealed the needto lower the heat transfer coefficient to the valueof 2.0 Wm−2K−1. One of the reasons could be thedisturbance of air boundary layer by the formworksteel ribs.

The cube was placed on an old massive concreteseparated by 2 mm thick impregnated paper, there-fore

h =λ

t=

0.10.002

= 50.0 Wm−2K−1 (16)

and the old concrete was considered to have thetemperature of ambient air.

Symmetrical configuration allowed the simula-tion of a cube quarter only. The prism was meshedfor 2×2×4 = 16 quadratic brick elements, the meshhad 141 DoF. One CEMHYD3D model was assignedto each finite element. The time integration step wasfixed as 10 min.

FIGURE 8. Experimental setup of a temperature measurement and corresponding quarter model (photo J. Litos)

International Journal for Multiscale Computational Engineering

Page 9: Multiscale Model for Temperature Distribution in Hydrating Concrete

MULTISCALE MODEL FOR TEMPERATURE DISTRIBUTION 9

The concrete was composed from CEM I 42.5 Rby the amount of 310 kg/m3 with w/c = 0.4. Fly ashas a filler was added further to the mixture but wasdisregarded for the heat contribution. Because ce-ment mineralogical properties were unknown, datafor initial microstructure were taken from Table 2from an ordinary Portland cement of the same gradewith the activation energy Ea = 38.3 kJ/mol. Ther-mal conductivity of concrete was assumed constant1.7 Wm−1K−1, its capacity constant 0.84 Jg−1K−1,and density 2.5 t/m3.

Figure 9 shows the temperature evolution in themiddle of the cubic specimen and ambient air tem-perature. Following parameters were calibrated tomatch experimental data; t0 = 0 h and β = 6.3 ×10−4 h/cycle2. The maximum temperature in thecenter reached 50.17C at 25.6 h and the simulationpredicted 50.3C. Maximum temperature achieved30 mm from the surface was 45.23C, and the simu-lation predicted 44.3C. The cooling stage after ap-proximately two days shows overprediction by thesimulation and emphasizes the role of heat trans-fer coefficients. Some heat could have dissipatedfrom water evaporation, which was not taken intoaccount.

Figure 10 shows the contour fill of the prismviewed from the central direction during the max-imum center temperature at 25.6 h. Total compu-tational time took 70 min on 3.2 GHz PC with op-timized g++ compilation. Profiling showed that98.5% of the whole computational time is taken bythe CEMHYD3D routine. In such a case, the refine-ment of an integration time step does not create anytime savings.

0

10

20

30

40

50

60

0 1 2 3 4 5 6

Tem

pera

ture

[o C]

Hydration time [days]

Center

Surface

Ambient air

Experiment

FIGURE 9. Validated heat evolution in the 1 × 1 × 1 mspecimen

FIGURE 10. Contour fill of temperature field at 25.6 hwhen maximum temperature was achieved. A view fromthe central part of the quarter model

5.3 Prestressed Bridge

A new scaffold bridge was built during 2006–2007in Prague. The bridge is a continuous beam of12 spans made from prestressed concrete with thetotal length of 443 m.

The first simulation deals with the casting of bot-tom slab of the thickness of 0.697 m with a pre-cast shoulder. The mesh was generated by T3D3

software with resulting 215 nodes, 4 triangular, and179 quadrilateral elements, together with 215 DoF(Fig. 11). The horizontal layered mesh structure waspreferred due to CEMHYD3D assignment. Eachlayer of 20 elements has a similar temperature field;therefore, these elements are grouped to eight inde-pendent CEMHYD3D models according to Fig. 11.

The casting began at 10:00 on June 6, 2006. Initialtemperature of casted concrete was measured inthe interval 14.1–15.7C and therefore was set toan initial 15C together with the precast shoulder.Both the casted concrete and the precast shoulder

3 http://mech.fsv.cvut.cz/˜dr/t3d.html

Volume 7, Number 2, 2009

Page 10: Multiscale Model for Temperature Distribution in Hydrating Concrete

10 SMILAUER AND KREJCI

FIGURE 11. Mesh and boundary conditions of bottom slab. Different colors represent eight CEMHYD3D materialmodels, the right shoulder is of mature precast concrete

were given the constant material parameters, ther-mal conductivity 1.7 Wm−1K−1 and heat capac-ity 0.84 Jg−1K−1. Used concrete composition withw/b = 0.425 is summarized in Table 3.

Pilot simulation revealed that the amount of Port-land cement is excessive and has to be reducedto the minimum value according to EN 197-1 (i.e.,280 kg of pure Portland cement and 20 kg of in-ert SCM admixtures, yielding w/c = 170/380 =0.4473). Input parameters were assumed accordingto Table 4. Mapping of hydration cycles to real timewas calibrated to the temperature evolution of cen-tral T4 yielding t0 = 0 h and β = 5.6 ·10−4 h/cycle2.

Boundary conditions were set according toFig. 11. While the bottom part was set to mea-sured air ambient temperature, the temperature ofthe top surface, exposed to sun radiation, was ap-proximated by

T (t) = 23 +[10 + 5H

(sin

2πt +

24

)]

× sin

2πt +4π

24

(17)

where H is a Heaviside step function and time t isin days relatively to the beginning of casting. Themaximum air temperature was considered at 14.00 haccording to the measured data.

Similar location of gauges T1–T3 and T2–T5 givesvery similar evolution of temperatures. Results forinner gauge T4 and ambient air temperature areplotted in Fig. 12; Fig. 13 shows gauges T3, T5. Theinner part of casted massive concrete is close to adi-abatic condition and influenced to a minor extent

TABLE 3. Concrete composition used in the simulationof the bridge, w/b= 0.425

Concrete parameter ValueType C35/45-XF2Cement (CEM I 42.5 R + SCM) 380+20 kg/m3

Water 170 kg/m3

Fine aggregates 1100 kg/m3

Coarse aggregates 450 kg/m3

Fly ash 200 kg/m3

Superplasticizer 3.6 kgBulk density 2.32 kg/m3

0

10

20

30

40

50

60

0 1 2 3 4 5 6

Jun07 Jun08 Jun09 Jun10 Jun11 Jun12

Tem

pera

ture

[o C]

Hydration time [days]

Ambient air (incomplete data)

T4 simulationT4 experiment

FIGURE 12. Temperature of central gauge T4 and am-bient air temperature

by boundary conditions. Therefore, the correspon-dence between simulation and experiment revealsappropriate selection of material parameters. Thedata from ambient air temperature are incomplete,

International Journal for Multiscale Computational Engineering

Page 11: Multiscale Model for Temperature Distribution in Hydrating Concrete

MULTISCALE MODEL FOR TEMPERATURE DISTRIBUTION 11

TABLE 4. Cement properties and material simulation parameters

Component C3S C2S C3A C4AF GypsumMass content without gypsum 0.65 0.15 0.10 0.10 -Volume content with gypsum 0.6175 0.1425 0.095 0.095 0.05Blaine fineness 300 m2/kgAutocorrelation NIST files 16130s1Activation energy Ea 38.3 kJ/mol

0

10

20

30

40

50

60

0 1 2 3 4 5 6

Jun07 Jun08 Jun09 Jun10 Jun11 Jun12

Tem

pera

ture

[o C]

Hydration time [days]

T3 simulationT3 experimentT5 simulation

T5 experiment

FIGURE 13. Temperature evolution in gauges T3 andT5

and the results between two points are interpolatedin the simulation.

The maximum temperature 56.2C was reachedin gauge T4 at 1.53 day, and the temperature field isplotted in Fig. 14. Whole simulation took 57 min on3.2 GHz PC and occupied 29 MB of RAM. Total sim-ulation covered six days of hydration period withthe constant time integration step 10 min.

Generally speaking, the difficulties of experimen-tal data acquisition on rush construction site or sev-eral missing parameters tacitly or explicitly postu-

lated in the simulation lead logically to a scatter ofthe results. The disagreement in Figs. 12 and 13 upto 0.5 days rises speculations about gauge surround-ing environment, gauge calibration, true initial mixtemperature casted from several batches, etc. Noneof them can be proved, and the results are left asthey were gathered. Another striking fact is a highermaximum measured temperature at the gauge T3compared to the central gauge T4. Although suchambiguities might seem to falsify the simulation, re-sults from a well-controlled small-scale experimentsfrom Figs. 7 and 9 are convincing.

The new bridge beams had been casted on thebottom slab since July 12, 2006, 6:39 a.m. (i.e., af-ter 35 days of bottom slab casting). The maximumbeam width was 1.65 m and the maximum heightreached 2.15 m. The generated mesh had 239 nodes,39 triangular, and 177 quadrilateral elements withtotal 239 DoF. A simulated, typical cross section isdepicted in Fig. 15 and allows another considerationof bridge symmetry vertical axis. Although the datafrom gauges T1–T5 were gathered from the bottomslab, only gauge T5 was considered important frommaturing concrete slab. The mesh was adjusted insuch a way that gauges T5–T10 were forced to beat a nodal position. The assignment of five hydra-tion models with expected similar temperature fieldis displayed in Fig. 15.

FIGURE 14. Contour fill at maximum temperature 56.2C in gauge T4 after 1.53 day

Volume 7, Number 2, 2009

Page 12: Multiscale Model for Temperature Distribution in Hydrating Concrete

12 SMILAUER AND KREJCI

FIGURE 15. Topology and boundary conditions of the beam. Different colors represent different CEMHYD3Dmaterial models, the right shoulder and bottom slab are considered nonhydrating

The initial ambient air temperature was 26C,while mature and fresh casted concrete reached25C. The bottom ambient air temperature was ap-proximated by the function

T (t) = 26 + 10 sin

2πt− 2.7π

24

(18)

and the upper and vertical surfaces

T (t) =26+[10+5H

(sin

2πt− 2.7π

24

)]

× sin

2πt− 2.7π

24

(19)

Concrete and simulation parameters were set ac-cording to Tables 3 and 4. Further calibration mod-ified t0 = 3 h and β = 4.0 × 10−4 h/cycle2 tomatch better results from central gauge T7. The pairgauges T6–T7, T9–T10 gave very similar experimen-tal results; therefore, only gauges T5, T7, T10 werevalidated.

Results for the central gauge T7 are shown inFig. 16. Maximum temperature was not mea-sured directly due to power disconnection; how-

0 10

20 30

40 50

60 70

80 90

0 2 4 6 8 10

Jul13 Jul15 Jul17 Jul19 Jul21

Tem

pera

ture

[o C]

Hydration time [days]

Ambient air

T7 simulationT7 experiment

FIGURE 16. Temperature of central gauge T7 and ap-proximated ambient air

ever, the simulated temperature reached a maxi-mum of 86.8C at 1.367 days (Figs. 16 and 18).Good accordance is achieved in T5 and T10 as well(Fig. 17). Both figures show a temperature underes-timation probably due to lower heat transfer coeffi-cients on the surfaces. Whole beam simulation took68 min on 3.2 GHz PC and occupied 19 MB of RAM.

International Journal for Multiscale Computational Engineering

Page 13: Multiscale Model for Temperature Distribution in Hydrating Concrete

MULTISCALE MODEL FOR TEMPERATURE DISTRIBUTION 13

0 10

20 30

40 50

60 70

80 90

0 2 4 6 8 10

Jul13 Jul15 Jul17 Jul19 Jul21

Tem

pera

ture

[o C]

Hydration time [days]

T10 simulationT10 experiment

T5 simulationT5 experiment

FIGURE 17. Temperature evolution in gauges T5 and T10

FIGURE 18. Temperature field at 1.367 days when the maximum in T7 is achieved

Calculations covered 10 days of the hydration pe-riod with a constant time integration step of 10 minusing the Crank-Nicolson scheme.

The multiscale formulation allows accessing localquantities. The evolution of hydration degrees is de-picted in Fig. 19 and demonstrates the effect of tem-perature history. The ultimate hydration degree ap-proaches 0.8 due to low cement fineness, hence, theinability to hydrate large cement grains completely.Young’s modulus can be predicted on the scale ofcement paste and concrete. Here, the adopted an-alytical homogenization methodology is similar to[32], relying on intrinsic elastic properties of clinkerminerals, hydration products and aggregates. In ad-

dition, homogenization includes the effect of inter-facial transition zone, two types of C-S-H [33] andentrained air. Typical values complementing knownconcrete composition in Table 3 were used and theresults are plotted in Fig. 20.

5.4 Role of Input Parameters at the StructuralLevel

The effect of individual input parameters on the re-sulting temperature field is explored in a closer de-tail. For this purpose, half of the bridge cross sec-tion in Fig. 21 was taken, capturing different ther-mal evolution in a massive beam and a slender

Volume 7, Number 2, 2009

Page 14: Multiscale Model for Temperature Distribution in Hydrating Concrete

14 SMILAUER AND KREJCI

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.1 1 10

Deg

ree

of h

ydra

tion

[-]

Hydration time [days]

T7 simulationT10 simulation

FIGURE 19. Evolution of hydration degree in gauges T7 and T10

0

5

10

15

20

25

30

0.1 1 10

You

ng’s

mod

ulus

[GP

a]

Hydration time [days]

Cement paste

Concrete

T7 simulationT10 simulation

FIGURE 20. Predicted evolution of Young’s modulus in gauges T7 and T10 on the scale of cement paste and concrete,no experimental data

FIGURE 21. Topology of bridge cross section with the mesh, boundary conditions, and important node positions

International Journal for Multiscale Computational Engineering

Page 15: Multiscale Model for Temperature Distribution in Hydrating Concrete

MULTISCALE MODEL FOR TEMPERATURE DISTRIBUTION 15

slab. Each of 84 quadrilateral finite elements withlinear interpolation corresponds to 84 independentCEMHYD3D material models. Because no greataccuracy is expected, only 25 × 25 × 25 µm mi-crostructures were used. The macroscopic modelhas 109 DoF.

The starting “normal” concrete composition ismade from 300 kg of CEM I 42.5 R accordingto Table 2 with assigned activation energy Ea =38.3 kJ/mol and w/c = 0.5. Besides, the variationof other parameters covered as follows:

• 400 and 200 kg of cement CEM I 42.5 R

• increased Blaine fineness from 306 to500 m2/kg

• C3S increased from original 68.30% C3S to 73%by mass without gypsum, preserving the ratioof other clinker minerals and maintaining 5% ofgypsum addition by volume

• w/c decreased from 0.5 to 0.3

• initial mix temperature in the range from 0 to30C surrounded by 20C ambient air temper-ature

Figure 22 indicates maximum temperatures inthe beam and slab with corresponding elapsed timewhen β = 7.8 × 10−4 h/cycle2 and t0 = 0.0 h.As expected, massive beam with the dimensions1.2 × 2.1 m is close to adiabatic curing conditions

while the slab is progressively cooled down by am-bient air. Cement in the amount of 200 kg/m3

may represent not only pure Portland cement buta hydrating part of blended cements. Hydration atw/c = 0.3 is slowed down in later stages due to wa-ter demand, thus resulting in decreased reaction ki-netics. A low initial mix temperature is very ben-eficial in preventing excessive overheating and, infact, represents a common method found in variousguidelines. A similar effect is expected from the sur-rounding ambient air temperature which would re-flect casting in different seasons.

The maximum temperature in the beam isachieved between 40 and 100 h of hydration anddemonstrates high influence of mix and curing con-ditions. Temperature in the slab 0.45 m thick isreached sooner for a given setup, approximatelyin the halftime of the beam (i.e., in the interval 20and 47 h). It should be noted that the cycle-timemapping parameter β could be lower, which wouldcause time shortening of maximum temperature.

The computation of one case took 46 min on3.2 GHz PC and occupied 43 MB of RAM, includingall code, data, and shared libraries. The simulationcovered the period up to five days using a 10 minconstant integration step.

6. CONCLUSION

A general multiscale heat conduction model forcement-based materials was developed and vali-dated on small-size experiments, large bridge slab,

0 10 20 30 40 50 60 70 80 90

Norm

al

400 kg

200 kg

500 m2

/kg

more C

3 Sw

/c = 0.3

0 oC

10 oC

20 oC

30 oC

Max

imum

tem

pera

ture

[o C]

Ambient air temperature

59 30 50 29 62 31 41 21 45 25 54 27 100

47 75 36 59 30 40 20

FIGURE 22. Maximum temperatures in the beam (left columns) and slab (right columns) when varying individualparameters of cement, concrete, or air temperature. Number in the bar represents time in hours when maximumtemperature was reached

Volume 7, Number 2, 2009

Page 16: Multiscale Model for Temperature Distribution in Hydrating Concrete

16 SMILAUER AND KREJCI

and beam. The emphasis was given on the ce-ment physical and chemical properties, utilizing in-trinsic reaction enthalpies as a source for liberatedheat. The model provides a reasonably accurate toolsuited for engineering purposes with appropriateconsumption of computation time regarding multi-scale model complexity.

Effect of water evaporation, RH influence on hy-dration, and changes in concrete thermal conductiv-ity and capacity were not incorporated. Two param-eters had to be calibrated in all experiments, repre-senting cycle-time mapping of the hydration model.They were found in the ranges t0 ∈ 〈0, 3〉 h andβ ∈ 〈4.0, 7.8〉 h/cycle2, covering the role of alka-lies or admixtures and possible inaccurate bound-ary conditions. Several uncertainties remain espe-cially due to a lack of information from constructionsite. The small-scale experiments show much betteragreement due to available details.

ACKNOWLEDGMENT

We greatly appreciate the financial support from thegrant MSM 6840770003. Bentz and Garboczi are ac-knowledged for the CEMHYD3D hydration model.Special thanks belong to J. Litos from CTU in Pragueand M. Vokac from the Klokner Institute of CTU inPrague for the provision of experimental data.

REFERENCES

1. Neville, A. M. Properties of Concrete. Wiley,Hoboken, NJ, 1997.

2. Lothenbach, B., Matschei, T., Moschner, G., andGlasser, F. Thermodynamic modelling of the ef-fect of temperature on the hydration and poros-ity of Portland cement. Cem. Concr. Res. 38(1):1–18, 2008.

3. Bentur, A., and Mitchell, D. Material perfor-mance lessons. Cem. Concr. Res. 38(2):259–272,2008.

4. Nilsson, M. Thermal Cracking of Young Con-crete. Partial Coefficients, Restraint Effects andInfluence of Casting Joints. Licentiate thesis,LuleaUniveristy of Technology, 2000.

5. Schrage, I., and Summer, T. Factors influenc-ing early cracking of high strength concrete,Proceedings of International RILEM Symposium.

Thermal Cracking in Concrete at Early Ages, pp.237–244, 1994.

6. Wang, C., and Dilger, D. Prediction of tem-perature distribution in hardening concrete.Proceedings of International RILEM Symposium.Thermal Cracking in Concrete at Early Ages, pp.21–28, 1994.

7. Bentz, D. P., Waller, V., and de Larrard, F. Pre-diction of Adiabatic Temperature Rise in Con-ventional and High-Performance ConcretesUsing a 3-D Microstructural Model. Cem. Concr.Res. 28(2):285–297, 1998.

8. Estrada, C. F., Godoy, L. A., and Prato, T.Thermo-mechanical behavior of a thin concreteshell during its early age. Thin-Walled Struct.44(5):483–495, 2006.

9. Park, K.-B., Jee, N.-Y., Yoon, I.-S., and Lee, H.-S. Prediction of Temperature Distributionin High-Strength Concrete Using HydrationModel. ACI Mater. J. 105(2):180–186, 2008.

10. Hellmich, Ch., Mang, H. A., and Ulm, F.-J. Hy-brid method for quantification of stress statesin shotcrete tunnel shells: combination of 3Din situ displacement measurements and ther-mochemoplastic material law. Comput. Struct.79:2103–2115, 2001.

11. Gawin, D., Pesavento, F., and Schrefler, B.Hygro-thermo-chemo-mechanical modellingof concrete at early ages and beyond. Part I:Hydration and hygro-thermal phenomena. Int.J. Numer. Methods Eng. 67(3):299–331, 2006a.

12. Gawin, D., Pesavento, F., and Schrefler, B.Hygro-thermo-chemo-mechanical modellingof concrete at early ages and beyond. Part II:shrinkage and creep of concrete. Int. J. Numer.Methods Eng., 67(3):332–363, 2006b.

13. Ulm, F.-J., and Coussy, O. Couplings in early-age concrete: From material modeling to struc-tural design. Int. J. Solids Struct. 35(31–32):4295–4311, 1998.

14. NRMCA. Concrete In Practice. Technical report,National Ready Mixed Concrete Association,1978–2007.

15. Barbarulo, R., Peycelon, H., Prene, S., andMarchand, J. Delayed ettringite formationsymptoms on mortars induced by high temper-ature due to cement heat of hydration or late

International Journal for Multiscale Computational Engineering

Page 17: Multiscale Model for Temperature Distribution in Hydrating Concrete

MULTISCALE MODEL FOR TEMPERATURE DISTRIBUTION 17

thermal cycle. Cem. Concr. Res. 35(1):125–131,2005.

16. Bentz, D. P. CEMHYD3D: A Three-DimensionalCement Hydration and Microstructure Develop-ment Modeling Package. Version 3.0. TechnicalReport, NIST Building and Fire Research Lab-oratory, Gaithersburg, MD, 2005.

17. Wang, B.-L., and Mai, Y.-W. Transient one-dimensional heat conduction problems solvedby finite element. Int. J. Mech. Sci. 47(2):303–317,2005.

18. Zeman, J., and Sejnoha, M. From randommicrostructures to representative volumeelements. Modell. Simul. Mater. Sci. Eng.15(4):S325–S335, 2007.

19. Maekawa, K., Chaube, R. P., and Kishi, T. Mod-eling of Concrete Performance. E & FN SPON,London, first edition, 1999.

20. Kada-Benameur, H., Wirquin, E., andDuthoit, B. Determination of apparent ac-tivation energy of concrete by isothermalcalorimetry. Cem. Concr. Res. 30(2):301–305,2000.

21. Bentz, D. CEMHYD3D: A Three-Dimensional Ce-ment Hydration and Microstructure DevelopmentModeling Package. Version 2.0. Technical Report,NIST Building and Fire Research Laboratory,Gaithersburg, MD, 2000.

22. Bentz, D. A Three-Dimensional Cement Hydrationand Microstructure Program. I. Hydration Rate,Heat of Hydration, and Chemical Shrinkage. Tech-nical Report, NIST Building and Fire ResearchLaboratory, Gaithersburg, MD, 1995.

23. Bentz, D. P., Garboczi, E., Haecker, C., andJensen, O. Effect of cement particle size dis-tribution on performance properties of Port-land cement-based materials. Cem. Concr. Res.29:1663–1671, 1999.

24. Kanit, T., Forest, S.,Galliet, I., Mounoury, V.,

and Jeulin, D. Determination of the size ofthe representative volume element for ran-dom composites: Statistical and numericalapproach. Int. J. Solids Struct. 40:3647–3679,2003.

25. Pane, I., and Hansen, W. Investigation ofblended cement hydration by isothermalcalorimetry and thermal analysis. Cem. Concr.Res. 35(6):1155–1164, 2005.

26. Papadakis, V. Effect of fly ash on Portland ce-ment systems: Part I. Low-calcium fly ash. Cem.Concr. Res. 29(11):1727–1736, 1999.

27. Robeyst, N., Gruyaert, E., and Belie, N. D. Ad-vances in Construction Materials 2007: Ultrasonicand Calorimetric Measurements on Fresh Concretewith Blast-Furnace Slag, Springer, Berlin, pp.497–504, 2007.

28. Snelson, D. G., Wild, S., and O’Farrell, M. Heatof hydration of Portland Cement-Metakaolin-Fly ash, (PC-MK-PFA) blends. Cem. Concr. Res.38(6):832–840, 2008.

29. Lienhard, J., and Lienhard, J. A Heat TransferTextbook. Phlogiston Press, Cambridge, 2008.

30. Hughes, T. The Finite Element Method–LinearStatic and Dynamic Finite Element Analysis.Dover Publishers, New York, 2000.

31. Bentz, D. Transient Plane Source Measure-ments of the Thermal Properties of HydratingCement Pastes. Mater. Struct. 40(10):1073–1080,2007.

32. Bernard, O., Ulm, F.-J., and Lemarchand, E.A multiscale micromechanics-hydration modelfor the early-age elastic properties of cement-based materials. Cem. Concr. Res. 33(9):1293–1309, 2003.

33. Smilauer, V., and Bittnar, Z. Microstructure-based micromechanical prediction of elasticproperties in hydrating cement paste. Cem.Concr. Res. 36(9):1708–1718, 2006.

Volume 7, Number 2, 2009