Acessibilidade / Reportar erro

Concrete modeling using micromechanical multiphase models and multiscale analysis

Modelagem do concreto usando modelos micromecânicos de múltiplas fases e análise multiescala

abstract:

Concrete in its macrostructure is a multiphase cementitious composite material, however, by reducing its scale, it is possible to identify the phases that compose it, among the phases are those embedded in the microscale: the hydrated silicates, in the mesoscale: the cement paste, transition zones and aggregates and in the macro phase: the composite itself. Modeling this type of material with two-phase micromechanical models is common in the literature, but there are already proven limitations that two-phase models can provide high modeling errors and are not recommended for this type of study. Faced with this problem, an alternative would be to use multiple-phase models, combined with a multiscale perspective in an attempt to minimize the error in modeling this material. The present paper models the concrete in two different constructions: without an interfacial transition zone and with the inclusion of the interfacial transition zone, verifying the modeling error when neglecting this important phase. The entire homogenization process is performed using the decoupled multiscale technique, obtaining results that rule out the use of two-phase models and methodologies that do not evaluate the interfacial transition zone in conventional concrete. The results obtained with the use of multiple-phase models reduced the relative error to practically zero (compared to experimental tests), demonstrating that micromechanics can be a concrete modeling tool provided that the multiscale process considers as many as possible phases and robust models that take this nature into account.

Keywords:
multiscale modeling; micromechanics; concrete

resumo:

O concreto em sua macroestrutura é um material compósito cimentício multifásico, contudo ao se reduzir sua escala consegue-se identificar as fases quem compõe o mesmo, dentre as fases estão as embutidas na microescala: os silicatos hidratados, na mesoescala: a pasta de cimento, zonas de transições e agregados e na fase macro: o próprio compósito. Modelagens desse tipo de material com modelos micromecânicos bifásicos são comuns na literatura, porém existe limitações já comprovadas de que modelos de duas fases podem aferir erros altos a modelagem não sendo recomendado para esse tipo de estudo. Diante dessa problemática, uma alternativa seria empregar modelos de múltiplas fases, aliado a uma perspectiva multiescala na tentativa de minimizar o erro na modelagem desse material. O presente trabalho modela o concreto em duas construções distinta: sem zona de transição interfacial e com a inclusão da zona de transição interfacial, verificando o erro de modelagem à negligência essa importante fase. Todo processo de homogeneização é realizado utilizando a técnica multiescala desacoplada obtendo resultados que descartam a utilização de modelos bifásicos e metodologias que não avaliam a zona de transição interfacial em concretos convencionais. Os resultados obtidos com a utilização de modelos de múltiplas fases reduziram o erro relativo a praticamente zero (em comparação com ensaios experimentais), demonstrando que a micromecânica pode ser uma ferramenta de modelagem do concreto desde que o processo multiescala leve em consideração o maior número possíveis de fases e modelos robustos que levem em conta essa natureza.

Palavras-chave:
modelagem multiescala; micromecânica; concreto

1 INTRODUCTION

Multiscale modeling is a technique that has been developed over the last decades with several applications in the most varied fields of science. The idea is to simulate continuous behavior from information obtained at lower scales [11 J. Fish, G. J. Wagner, and S. Keten, “Mesoscopic and multiscale modelling in materials,” Nat. Mater., vol. 20, pp. 777-786, Jun. 2021.].

In his paper Fish et al. [11 J. Fish, G. J. Wagner, and S. Keten, “Mesoscopic and multiscale modelling in materials,” Nat. Mater., vol. 20, pp. 777-786, Jun. 2021.] he presents a review of the most interesting applications of multiscale modeling including the prediction of electrical, magnetic, and chemical properties of complex materials (for example composites).

For each application, it is possible to define the multiscale approach that will be used, explaining its advantages and disadvantages in addition to the degree of complexity inherent to its technique [22 O. Lloberas-Valls, D. J. Rixen, A. Simone, and L. J. Sluys, “Multiscale domain decomposition analysis of quasi-brittle heterogeneous materials,” Int. J. Numer. Methods Eng., vol. 89, pp. 1337-1366, Nov. 2011.]. Multiscale modeling consists of evaluating each level and how they influence each other as the composite is built. Lloberas-Valls et al. [22 O. Lloberas-Valls, D. J. Rixen, A. Simone, and L. J. Sluys, “Multiscale domain decomposition analysis of quasi-brittle heterogeneous materials,” Int. J. Numer. Methods Eng., vol. 89, pp. 1337-1366, Nov. 2011.] defines two types of multiscale modeling, namely: hierarchical multiscale technique and concurrent multiscale technique.

The hierarchical multiscale technique can be divided into decoupled or weak coupled, in the decoupled technique the boundary condition problem is already solved, and the effective properties are carried from one scale to another (common in micromechanical homogenization). When the problem is analyzed by weak coupling, properties can be taken from global to local scale and vice versa, and the boundary problem is not initially solved. In the concurrent multiscale technique, the different scales are solved simultaneously, maintaining the condition of equilibrium and displacements, explaining a strong coupling [22 O. Lloberas-Valls, D. J. Rixen, A. Simone, and L. J. Sluys, “Multiscale domain decomposition analysis of quasi-brittle heterogeneous materials,” Int. J. Numer. Methods Eng., vol. 89, pp. 1337-1366, Nov. 2011.].

Given this premise of expansion of knowledge and since concrete can be considered a complex material (composite) due to its multiphase and cementitious nature, this work proposes to estimate the mechanical properties of concrete using multiscale, micromechanical modeling of mean fields, bringing as a contribution to the application of models involving multiple phases, including the interfacial transition zone, in addition to verification and comparison with classical biphasic models. To validate the multiscale modeling, a comparison is made with experimental tests.

On a macroscopic scale, concrete may be considered as a homogenous material, however, as the scale decreases (meso and microscale), it cannot be considered homogenous - it is heterogenous, explaining its respective phases relate in [33 O. Bernard, F.-J. Ulm, and E. Lemarchand, “A multiscale micromechanics-hydration model for the early-age elastic properties of cement-based materials,” Cement Concr. Res., vol. 33, pp. 1293-1309, Sep. 2003.], [44 G. Constantinides and F.-J. Ulm, “The effect of two types of C-S-H on the elasticity of cement-based materials: results from nanoindentation and micromechanical modeling,” Cement Concr. Res., vol. 34, no. 1, pp. 67-80, Jan. 2004.], [55 J. F. Unger and S. Eckardt, “Multiscale modeling of concrete,” Arch. Comput. Methods Eng., vol. 18, p. 341, Jul. 2011.]. Knowing the phases that compose the concrete, its respective fractions, and how they interact with one another is fundamental to maximizing its use.

Rodrigues [66 E. A. Rodrigues, “Um modelo multiescala concorrente para representar o processo de fissuração do concreto,” Ph.D. dissertation, Poli USP, São Paulo, Brazil, 2015.] explains that the region that comprehends the mesoscale is a target of interest for several researchers: Häfner et al. [77 S. Häfner, S. Eckardt, T. Luther, and C. Könke, “Mesoscale modeling of concrete: geometry and numerics,” Comput. Struct., vol. 84, no. 7, pp. 450-461, Feb. 2006.], Eckardt and Könke [88 S. Eckardt and C. Könke, “Adaptive damage simulation of concrete using heterogeneous multiscale models,” J. Algorithm Comput. Technol., vol. 2, no. 2, pp. 275-298, Jun. 2007.], Eckardt [99 S. Eckardt, “Adaptive heterogeneous multiscale models for the nonlinear simulation of concrete,” Ph.D. dissertation, BUW, Weimar, Germany, 2009.], Nguyen et al. [1010 V. P. Nguyen, M. Stroeven, and L. J. Sluys, “Multiscale failure modeling of concrete: Micromechanical modeling,discontinuous homogenization and parallel computations,” Comput. Methods Appl. Mech. Eng., vol. 201-204, pp. 139-156, Jan. 2012.] since this phase defines the concrete as a composite and multiphase material.

In their study, Pichler and Hellmich [1111 B. Pichler and C. Hellmich, “Upscaling quasi-brittle strength of cement paste and mortar: a multi-scale engineering mechanics model,” Cement Concr. Res., vol. 41, no. 5, pp. 467-476, May. 2011.] characterize well the multiscale modeling of some phases of the concrete (decoupled technique), namely: cement paste and the process of hydration of the paste that contemplates portions of the hydrated silicates, water/capillary pores, and clinker.

Further on, multiscale modeling of concrete was proposed, contemplating an experimental campaign to validate the models previously proposed by Göbel et al. [1212 L. Göbel, M. Königsberger, A. Dimming-Osburg, and B. Picheler, “Viscoelastic behavior of polymer-modified cement pastes: insight from downscaling short-term macroscopic creep tests by means of multiscale modeling,” Appl. Sci., vol. 8, no. 4, p. 487, Mar. 2018.].

Siventhirarajah et al. [1313 K. Siventhirarajah, Y. Yoda, and Y. Elakneswaran, “A two-stage model for the prediction of mechanical properties of cement paste,” Cement Concr. Compos., vol. 115, p. 103853, Jan. 2021.] can go further, through a multiscale analysis reducing the cement paste in some levels, reaching the nanometric scale, evaluating the hydrated calcium silicates HCS that compose the clinker (C3S, C2S, C3A, C4AF), tricalcium silicate, dicalcium silicate, tricalcium aluminate, tetracalciumalumino ferrite, respectively. In 2004 Scrivener [1414 K. L. Scrivener, “Backscattered electron imaging of cementitious microstructures: understanding and quantification,” Cement Concr. Compos., vol. 26, no. 8, pp. 935-945, Nov. 2004.] published an important paper on the characterization and measurement of concrete phases, especially the phases that make up a cement past.

Figure 1 illustrates the multiscale diagram of concrete showing its phases and dimensions from the microscale, through the mesoscale to the macroscale [33 O. Bernard, F.-J. Ulm, and E. Lemarchand, “A multiscale micromechanics-hydration model for the early-age elastic properties of cement-based materials,” Cement Concr. Res., vol. 33, pp. 1293-1309, Sep. 2003.], [44 G. Constantinides and F.-J. Ulm, “The effect of two types of C-S-H on the elasticity of cement-based materials: results from nanoindentation and micromechanical modeling,” Cement Concr. Res., vol. 34, no. 1, pp. 67-80, Jan. 2004.], [99 S. Eckardt, “Adaptive heterogeneous multiscale models for the nonlinear simulation of concrete,” Ph.D. dissertation, BUW, Weimar, Germany, 2009.], [1010 V. P. Nguyen, M. Stroeven, and L. J. Sluys, “Multiscale failure modeling of concrete: Micromechanical modeling,discontinuous homogenization and parallel computations,” Comput. Methods Appl. Mech. Eng., vol. 201-204, pp. 139-156, Jan. 2012.], [1111 B. Pichler and C. Hellmich, “Upscaling quasi-brittle strength of cement paste and mortar: a multi-scale engineering mechanics model,” Cement Concr. Res., vol. 41, no. 5, pp. 467-476, May. 2011.].

Figure 1
Multiscale micromechanical diagram, processes, levels, and stages of concrete. Source: Author, adapted from Bernard et al. [33 O. Bernard, F.-J. Ulm, and E. Lemarchand, “A multiscale micromechanics-hydration model for the early-age elastic properties of cement-based materials,” Cement Concr. Res., vol. 33, pp. 1293-1309, Sep. 2003.], Siventhirarajah et al. [1313 K. Siventhirarajah, Y. Yoda, and Y. Elakneswaran, “A two-stage model for the prediction of mechanical properties of cement paste,” Cement Concr. Compos., vol. 115, p. 103853, Jan. 2021.].

Siventhirarajah et al. [1313 K. Siventhirarajah, Y. Yoda, and Y. Elakneswaran, “A two-stage model for the prediction of mechanical properties of cement paste,” Cement Concr. Compos., vol. 115, p. 103853, Jan. 2021.] subdivides the analysis of concrete into two stages - the first stage being responsible for integrating models of humidity, thermodynamics, and the hydration of the cement paste to evaluate the volumetric fractions of the comprehended phases between levels 1 to 3 (microscale). With the respective fractions and properties of the phases of stage 1, it is possible to estimate the property of the cement paste. The second stage is related to the mechanical evaluation of the composite (meso and macroscale), comprehending the phases of the sand and gravel, the respective interfacial transition zone (ITZ), and mortar.

In this stage (< level 6), the concrete can be understood as a composite, cementitious and multiphase material, which makes it complex, and simplified tests cannot evaluate its behavior.

Still, at level 5, the literature has long tried to understand the participation of each component of this mixture in the final characteristics of the composite, much has already been discovered, since the great influence of inclusions [1515 M. S. Meddah, S. Zitouni, and S. Belâabes, “Effect of content and particle size distribution of coarse aggregate on the compressive strength of concrete,” Constr. Build. Mater., vol. 24, no. 4, pp. 505-512, Apr. 2010.], [1616 K.-R. Wu, B. Chen, W. Yao, and D. Zhang, “Effect of coarse aggregate type on mechanical properties of high-performance concrete,” Cement Concr. Res., vol. 31, no. 10, pp. 1421-1425, Oct. 2011.], [1717 H. Beushausen and T. Dittmer, “The influence of aggregate type on the strength and elastic modulus of high strength concrete,” Constr. Build. Mater., vol. 74, pp. 132-139, Jan. 2015.], [1818 N. V. Silva, “Método de determinação de resistência à tração e módulo de elasticidade de agragados graúdos naturais,” M.S. thesis, USP, São Paulo, Brazil, 2018.], as well as a third, more fragile phase located at the matrix-inclusion interface (interfacial transition zone), which is difficult to measure but which has already been extensively tested by several techniques [1919 G. Prokopskia and J. Halbiniakb, “Interfacial transition zone in cementitious materials,” Cement Concr. Res., vol. 30, no. 4, pp. 579-583, Apr. 2000.], [2020 K. L. Scrivener, A. K. Crumbie, and P. Laugesen, “The Interfacial Transition Zone (ITZ) between cement paste and aggregate in concrete,” Interface Sci., vol. 12, pp. 411-421, Oct. 2004.], [2121 G. Li, Y. Zhao, and S. Pang, “Four-phase sphere modeling of effective bulk modulus of concrete,” Cement Concr. Res., vol. 29, no. 6, pp. 839-845, Feb. 1999.].

Silva [1818 N. V. Silva, “Método de determinação de resistência à tração e módulo de elasticidade de agragados graúdos naturais,” M.S. thesis, USP, São Paulo, Brazil, 2018.] highlight the relationship between these phases and the possible behavior of each one when subject to request. When the inclusion phase is dominant in the stiffness of the composite, the matrix phases and the transition zone deform to a greater extent, which is probably the preferred path for cracking and fracturing of the composite.

However, even if the interfacial transition zone generated by the gravel is included, there are other phases, highlighting: sand pores, interfacial transition zone generated by sand, and even smaller phases such as non-hydrated clinker, hydrated silicates, among others, passing through all the levels shown in Figure 1.

Each of these phases has distinct properties and volumetric fractions and good modeling is directly associated with the characterization of each one of them. This dependence on concrete modeling as a function of the number of phases is remedied with a multiscale modeling using multi-phase micromechanical models, as proposed in this article.

2 INITIAL CONSIDERATIONS ON COMPOSITE HOMOGENIZATION

2.1. Representative Volume Element

A given body admitted homogenous and continuous in its macro-structure when reduced to a sufficiently small scale, does not behave as homogenous material, but as heterogeneous one, besides presenting discontinuities.

The literature brings several definitions about which volume will be representative of its macro-scale as seen in [1111 B. Pichler and C. Hellmich, “Upscaling quasi-brittle strength of cement paste and mortar: a multi-scale engineering mechanics model,” Cement Concr. Res., vol. 41, no. 5, pp. 467-476, May. 2011.], [2222 Z. Hashin, “Analysis of composite materials-a survey,” ASME J. Appl. Mech., vol. 50, pp. 481-505, Sep. 1983.], [2323 W. Drugan and J. Willis, “A micromechanics-based nonlocal constitutive equation and estimates of representative volume element size for elastic compositessize for elastic composites,” J. Mech. Phys. Solids, vol. 44, pp. 497-524, Apr. 1996.], [2424 M. Ostoja-Starzewski, “Microstructural randomness versus representative volume element in thermomechanics,” ASME J. Appl. Mech, vol. 69, pp. 25-35, Jun. 2002.], [2525 M. Stroven, H. Askes, and L. Sluys, “Numerical determination of representative volumes for granular materials,” Comput. Methods Appl. Mech. Eng., vol. 193, no. 30-32, pp. 3221-3238, Jul. 2004.], [2626 A. Drago and M.-J. Pindera, “Micro-macromechanical analysis of heterogeneous materials: macroscopically homogeneous vs periodic microstructures,” Compos. Sci. Technol., vol. 67, no. 6, pp. 1243-1263, May. 2007.].

Stroven et al. [2525 M. Stroven, H. Askes, and L. Sluys, “Numerical determination of representative volumes for granular materials,” Comput. Methods Appl. Mech. Eng., vol. 193, no. 30-32, pp. 3221-3238, Jul. 2004.] discourses about the diverse definitions related to RVE (representative volume element), proposing a form of its measuring, taking into consideration a statistic dispersion associated with a specific property, when its scale is reduced. Still, in the paper [2525 M. Stroven, H. Askes, and L. Sluys, “Numerical determination of representative volumes for granular materials,” Comput. Methods Appl. Mech. Eng., vol. 193, no. 30-32, pp. 3221-3238, Jul. 2004.] the following definitions are highlighted:

Hashin [2222 Z. Hashin, “Analysis of composite materials-a survey,” ASME J. Appl. Mech., vol. 50, pp. 481-505, Sep. 1983.] admits that a volume is representative if it contains all of the phases that characterize the microstructure of the heterogeneous material studied and a sufficient number of these phases for the corresponding mean properties to the RVE to be independent of the applied conditions of contour, once these conditions are macroscopically uniform, or, in other words, the values oscillate around a mean value with a small standard deviation, becoming insignificant at a small distance of the surface.

Drugan and Willis [2323 W. Drugan and J. Willis, “A micromechanics-based nonlocal constitutive equation and estimates of representative volume element size for elastic compositessize for elastic composites,” J. Mech. Phys. Solids, vol. 44, pp. 497-524, Apr. 1996.] indicate that the representative volume element is associated with the smaller volume that can represent the mean properties of the composite.

Ostoja-Starzewski [2424 M. Ostoja-Starzewski, “Microstructural randomness versus representative volume element in thermomechanics,” ASME J. Appl. Mech, vol. 69, pp. 25-35, Jun. 2002.] defines RVE in two situations: when the system is periodic in its micro-structure and there is a cell that represents it, and in the second case when there is a volume (great enough) that can incorporate several phases in its micro-structure, presenting homogenous structures.

The reference quoted above about RVE differs a little, however, RVE is always considered inferior to the macroscale of the composite, which can represent its phases and must have little dispersion in the evaluation of its properties when its volume is disturbed [2525 M. Stroven, H. Askes, and L. Sluys, “Numerical determination of representative volumes for granular materials,” Comput. Methods Appl. Mech. Eng., vol. 193, no. 30-32, pp. 3221-3238, Jul. 2004.].

Drago and Pindera [2626 A. Drago and M.-J. Pindera, “Micro-macromechanical analysis of heterogeneous materials: macroscopically homogeneous vs periodic microstructures,” Compos. Sci. Technol., vol. 67, no. 6, pp. 1243-1263, May. 2007.] bring a definition of RVE as being a heterogenous system in its micro-structure, where when the specific contour conditions are applied for each representative volume, the answer does not differ from the material in its macroscale. Drago and Pindera [2626 A. Drago and M.-J. Pindera, “Micro-macromechanical analysis of heterogeneous materials: macroscopically homogeneous vs periodic microstructures,” Compos. Sci. Technol., vol. 67, no. 6, pp. 1243-1263, May. 2007.] still explained the concept of the unitary cell of repetition, which is widely used when the problem has a condition of periodical contour.

In their review about RVE the authors described several studies that aimed at the characterization of the representative volume element, especially for concretes, once they were trying to know if the samples evaluated in fact could represent the composite. These studies established a parametric analysis of the volume and its phases submitted to several conditions of homogenous contour. These investigations tried to define the number of inclusions contained in the RVE that would be sufficient for the evaluation of the properties of the composite.

Pichler and Hellmich [1111 B. Pichler and C. Hellmich, “Upscaling quasi-brittle strength of cement paste and mortar: a multi-scale engineering mechanics model,” Cement Concr. Res., vol. 41, no. 5, pp. 467-476, May. 2011.] define representative volume element due to the analysis scales. In a simplified manner, the definitions quoted above continue to be valid, however, there may be a representative element inside another representative element to be able to evaluate its properties on an even smaller scale.

Suppose that a representative volume element on scale 𝐴, composed of 𝑛1, 𝑛2, …, 𝑛𝑚

phases, in such a way that when the analysis scale is reduced to 𝐵, it is possible that some of its 𝑛1, 𝑛2, …, 𝑛𝑚, phase have to be represented by other 𝑘1, 𝑘2, …, 𝑘𝑚, phases, expanding the range of phases for the multi-scale analysis, well defined in concrete. e.g.: Pichler and Hellmich [1111 B. Pichler and C. Hellmich, “Upscaling quasi-brittle strength of cement paste and mortar: a multi-scale engineering mechanics model,” Cement Concr. Res., vol. 41, no. 5, pp. 467-476, May. 2011.] and approached in this study.

Micromechanical analyzes are normally conducted based on the concept of a representative volume element (RVE). The composite material can be understood as the sum of the matrix volume with the sum of the volume of all the inclusions, which can be voids, materials, cracks, etc [2727 R. Christensen and K. Lo, “Solutions for effective shear properties in three phase sphere and cylinder models,” J. Mech. Phys. Solids, vol. 27, no. 4, pp. 315-330, Aug. 1979.]. Equation 1, below, represents this method.

V = V m + α = 1 n V i (1)

Where Vm is the volume of the matrix, Vi is the volume of inclusions, and V is the total volume of the representative element. When taking the ratio between the matrix volume and the total volume, as well as the volume of inclusions and the total volume, the following volumetric fractions are obtained, considering Equations 2 and 3, respectively.

f m = V m V (2)
f i = 1 V i = 1 n V i (3)

Volumetric fractions are extremely important in this type of analysis. The simplest proposes a weighted average as a function of fractions, of properties for global analysis, also known as the rule of mixtures. Over time, several methods were developed to solve the micromechanical problem, highlighting the Equivalent Inclusion Method, developed by Eshelby [2828 J. Eshelby “The determination of the elastic field of an ellipsoidal inclusion and related problems,” Proc. Royal Society London Ser. A, vol. 241, pp. 376-396, Mar. 1957.], one of the great contributions to the development of the micromechanics of effective media.

From the method proposed by Eshelby [2828 J. Eshelby “The determination of the elastic field of an ellipsoidal inclusion and related problems,” Proc. Royal Society London Ser. A, vol. 241, pp. 376-396, Mar. 1957.] other methods were developed, highlighting: Self Consistent [2929 Y. Benveniste, “A new approach to the application of Mori-Tanaka’s theory in composite materials,” Mech. Mater., vol. 6, no. 2, pp. 147-157, Jun. 1987.], Mori-Tanaka [3030 Z. Hashin, “The differential scheme and its application to cracked materials,” J. Mech. Phys. Solids, vol. 36, no. 6, pp. 719-734, 1988.], and Differential Scheme [3131 L. Walpole, “On the overall elastic moduli of composite materials,” J. Mech. Phys. Solids, vol. 17, no. 4, pp. 235-251, Sep, 1969.], among others.

The study of the micromechanics of effective media admits the hypothesis that the representative stresses and strains of a composite material (matrix + inclusions) can be represented by the average of stresses and strains in the representative volumes of each phase. Based on this principle, the total average stress, in the matrix and the inclusions, can be expressed by (Equations 4-6):

σ - = 1 V V . σ x d v (4)
σ - m = 1 V m V m . σ x d v (5)
σ - i = 1 V i V i . σ x d v (6)

Knowing that:

V . σ x d v = V m . σ x d v + i = 1 n V i . σ x d v (7)

replacing Equation 7, in Equation 4 there have been (Equation 8):

σ - = f m σ - m + i = 1 n f i σ - i (8)

Arrived at an expression that can determine the total average stress as a function of the stresses in the matrix and the inclusions and their respective volumetric fractions. Assuming that V is a representative volume, it can be stated that the average total stress σ- is equal to the effective stress in the material σ. Similarly, we have the total mean strain, in the matrix and the inclusions are given by the Equations 9-11:

ε - = 1 V V . ε x d v (9)
ε - m = 1 V m V m . ε x d v (10)
ε - i = 1 V i V i . ε x d v (11)

Knowing that:

V . ε x d v = V m . ε x d v + i = 1 n V i . ε x d v (12)

replacing Equation 12, in Equation 9 there have been (Equation 13):

ε - = f m ε - m + i = 1 n f i ε i (13)

Assuming that matrix and inclusions are elastic materials. So, it can be said that (Equations 14-16):

σ = C : - ε (14)
σ m = C m : ε m (15)
σ i = C i : ε i (16)

Where, C,- Cm, and Ci are respectively the global constitutive tensor, the matrix constitutive tensor, and the inclusion constitutive tensor, being the same fourth-order tensors. If the material is linear elastic, the global constitutive tensor is constant.

The concentration tensors of a composite are tensors that relate the average stresses and strains in the composite with the matrix and the inclusions. Substituting Equations 14-16 to Equation 8, we obtain (Equations 17-18):

f m C m : ε m = C - : ε - i = 1 n f i C i : ε i (17)
f m C m : 1 V m V . ε x d v - i = 1 n V i . ε x d v = C - : ε - i = 1 n f i C i : ε i (18)

With a little algebra one can arrive at (Equation 19):

C - - C m : ε = i = 1 n f i ( C i - C m ) : ε i (19)

This equation relates the constitutive tensors of the matrix, the inclusion and the global tensor, the global deformations, and the inclusion.

2.2. The Eshelby Problem

In 1957 Eshelby [2828 J. Eshelby “The determination of the elastic field of an ellipsoidal inclusion and related problems,” Proc. Royal Society London Ser. A, vol. 241, pp. 376-396, Mar. 1957.] proposed to determine the elastic field of an ellipsoidal inclusion in a solid (Figure 2).

Figure 2
Esheby's inclusion problem

It is assumed that the Ω region undergoes a geometric transformation so that in the absence of the material that surrounds the ellipse region it would correspond to a homogeneous deformation. In this hypothesis, Eshelby was able to assess what the elastic fields would be like inside and outside the ellipsoidal region. To exemplify Eshelby's strategy, initially, the ellipsoidal region, where the strain is initially zero, is removed, and a homogeneous strain is applied as shown in Figure 3.

Figure 3
Eshelby's problem strategy.

The next step is to apply external forces to the region so that it returns to its initial volume. When these external forces are applied, the volume of the ellipsoidal region decreases, however, a tension field is associated with this force in the body (Figure 4).

Figure 4
Eshelby's problem strategy.

The next step consists of returning to the ellipsoidal region, now with a tension field in the region of the representative element, removing the external actions that surrounded the ellipsoidal region, as can be seen in Figure 5.

Figure 5
Eshelby's problem strategy.

Eshelby [2828 J. Eshelby “The determination of the elastic field of an ellipsoidal inclusion and related problems,” Proc. Royal Society London Ser. A, vol. 241, pp. 376-396, Mar. 1957.] states that within the ellipsoidal inclusion Ω, the stresses and strains are constant but dependent on the geometry of the inclusion. With this, we define the [2828 J. Eshelby “The determination of the elastic field of an ellipsoidal inclusion and related problems,” Proc. Royal Society London Ser. A, vol. 241, pp. 376-396, Mar. 1957.] relation between the deformations in the ellipsoid and the deformation imposed on the system by a transformation tensor, called the Eshelby tensor (S), in the form:

ε Ω = S Ω ε * (20)

The assembly of the Eshelby tensor can be seen in the original paper [2828 J. Eshelby “The determination of the elastic field of an ellipsoidal inclusion and related problems,” Proc. Royal Society London Ser. A, vol. 241, pp. 376-396, Mar. 1957.]. With the same strategy mentioned above, it is enough to modify the property of the ellipsoid to arrive at the equivalent inclusion method that was the precursor of the mean-fields micromechanics.

2.3. Composite Homogenization Models

There are homogenization models that have the purpose of establishing limits, the best known being the rule of mixtures (Reuss and Voigt Model) [3232 A. Kaw, Mechanics of Composite Materials, 2nd ed. New York, NY, USA: Taylor & Francis, 2006.] and the Hashin-Shtrikman models [3333 Z. Hashin and S. Shtrikman, “On some variational principles in anisotropic and nonhomogeneous elasticity,” J. Mech. Phys. Solids, vol. 10, no. 4, pp. 335-342, Oct./Dec., 1962.]. Models should not be used to estimate the homogenized properties of composites, as they have several simplifying hypotheses such as the constant stress or strain field, or the combination of the two states, another simplification occurs in the non-evaluation of the interaction between the particles, among others.

To circumvent the limitations of limit models, methods based on the Eshelby equivalent inclusion problem stand out, the best known being the Mori-Tanaka model [3434 T. Mori and K. Tanaka “Average stree in matrix and average energy of materials with mis-fitting inclusions,” Act. Metall., vol. 21, no. 5, pp. 571-574, May. 1973.], before the paper of Mori and Tanaka [3434 T. Mori and K. Tanaka “Average stree in matrix and average energy of materials with mis-fitting inclusions,” Act. Metall., vol. 21, no. 5, pp. 571-574, May. 1973.], Hill [3535 R. Hill, “A self-consistent mechanics of composite materials,” J. Mech. Phys. Solids, vol. 13, pp. 213-222, Aug. 1965.] proposed the “Self-Consistent” model, which was later expanded by Christensen and Lo [2727 R. Christensen and K. Lo, “Solutions for effective shear properties in three phase sphere and cylinder models,” J. Mech. Phys. Solids, vol. 27, no. 4, pp. 315-330, Aug. 1979.]. Also, noteworthy are the paper of Benveniste [2929 Y. Benveniste, “A new approach to the application of Mori-Tanaka’s theory in composite materials,” Mech. Mater., vol. 6, no. 2, pp. 147-157, Jun. 1987.], with the model of equivalent inclusion [3030 Z. Hashin, “The differential scheme and its application to cracked materials,” J. Mech. Phys. Solids, vol. 36, no. 6, pp. 719-734, 1988.], and his model called “Differential Scheme” and the model “Double Inclusion” [2121 G. Li, Y. Zhao, and S. Pang, “Four-phase sphere modeling of effective bulk modulus of concrete,” Cement Concr. Res., vol. 29, no. 6, pp. 839-845, Feb. 1999.].

Kaw [3232 A. Kaw, Mechanics of Composite Materials, 2nd ed. New York, NY, USA: Taylor & Francis, 2006.] divides composite homogenization models into the mechanic of materials, semi-empirical models, and elasticity theory models. The present paper adopts as a reference this subdivision proposed by Kaw [3232 A. Kaw, Mechanics of Composite Materials, 2nd ed. New York, NY, USA: Taylor & Francis, 2006.] adding the models that are derived from the mean-fields micromechanical formulation.

The Dilute Suspension model, although mentioned in Table 1, was not used in the micromechanical modeling of concrete due to its limitation when the volumetric fraction is high [3737 H. Zhang, K. Anupam, A. Scarpas, and C. Kasbergen, “Comparison of different micromechanical models for predicting the effective properties of open graded mixes,” Transp. Res. Rec., vol. 2672, no. 28, pp. 404-415, Sep. 2018.].

Table 1
Characteristics of homogenization models.

3 MULTI-SCALE HOMOGENIZATION OF CONCRETE

The procedure of homogenization adopted in this studied case considers the phases of construction of the material. To define the homogenized properties of concrete, the phases of the constitution of the composite (Figure 6) are considered in its meso-macroscale.

Figure 6
Three-level homogenization considering (without [a] /with [b]) interfacial transition zone and three- and multiple-phase models. Source: Author.

This way, each step receives as input a matrix and a random inclusion and provides as output the homogenized material. The elastic properties of the materials used for the composite construction are defined in Table 2. These parameters were quoted by Silva [1818 N. V. Silva, “Método de determinação de resistência à tração e módulo de elasticidade de agragados graúdos naturais,” M.S. thesis, USP, São Paulo, Brazil, 2018.].

Table 2
Property of the composite phases (Experimental).

In the first analysis, the composite is treated in a simplified way (Figure 6a), disregarding the presence of the transition zone, and the effective properties are calculated with two-phase models, in the second analysis, the four-phase and multiphase models including two more phases in the construction of the composite (Figure 6b). For both analyzes, in addition to the modulus of elasticity, the Poisson ratio and the estimated characteristic compressive strength of the concrete are evaluated.

Silva [4141 N. V. Silva, “Propriedades mecânicas das fases e seus efeitos no comportamento mecânico do concreto,” Ph.D. dissertation, Poli USP, São Paulo, Brazil, 2022.] explains that the longitudinal Young modulus (Ec) is a requirement for determining the characteristic strength of concrete (fck), being a parameter for several empirical models of strength prediction found in the literature. Some models can be seen in Table 3.

Table 3
Normative equations to estimate the modulus of elasticity of concrete.

4 RESULTS AND DISCUSSIONS

To assess the calibration of micromechanical models with experimental results, the concept of relative error is used between the result obtained with the micromechanical analysis and the value obtained in the laboratory.

In the analysis of relative errors to obtain the effective elastic properties of the concrete, it was considered that the longitudinal Young modulus of 34.5 GPa, which was obtained experimentally by Li et al. [4040 Y. Li, Y. Li, and R. Wang, “Quantitative evaluation of elastic modulus of concrete with nanoidentation and homogenization method,” Constr. Build. Mater., vol. 212, pp. 295-303, Jul. 2019.], was considered.

One of the hypotheses raised would be the verification of limit models: models of Reuss, Voigt, and limits of Hashin. For this hypothesis, the construction of the composite without a transition zone was used (Figure 6a), with the corrections of the volumetric fractions for each of the three levels adopted in these two initials analyses shown in Table 4.

Table 4
Correction of volumetric fractions in the homogenization steps.

It should be noted that the errors for the analyzed analytical models are above 40% at the last level, which is not adequate. It is also observed that in the last homogenization step, the Reuss and Hashin [-] models are poorly conditioned since the pore modulus of elasticity is null and the relative error calculation would produce negative errors, which would not make physical sense (Table 5).

Table 5
Homogenization results for the limit models.

Continuing the modeling, two-phase models were used, disregarding the interfacial transition zone both in the sand and in the gravel, obtaining the following result in the homogenization steps (Table 6). When considering the two-phase numerical models in the homogenization process, a decrease in the concrete modeling error can be seen.

Table 6
Homogenization results for the biphasic models.

It can be seen that the two-phase homogenization models are not primarily responsible for

the high value of the relative error, with peaks of up to 37.496%, while the limit models had errors of 53.295%. This error is directly associated with the disregard of the interfacial transition zone in the composite construction and analysis process, even though the two-phase micromechanical models have limitations for concrete modeling.

The following analysis consists of modeling the concrete considering the interfacial transition zone, with models capable of evaluating more than two phases. The correction of volumetric fractions for these models can be seen in Table 7.

Table 7
Correction of volumetric fractions in the homogenization steps.

In the last level of homogenization, the Mori-Tanaka model was used to correct the composite with pores. The obtained results reduced the error found by Li et al. [4040 Y. Li, Y. Li, and R. Wang, “Quantitative evaluation of elastic modulus of concrete with nanoidentation and homogenization method,” Constr. Build. Mater., vol. 212, pp. 295-303, Jul. 2019.] from 8% to below 5%, with the multiphase model reaching an insignificant error for the four-phase model, indicating a substantial gain when using a more robust homogenization model. Table 8 illustrates the decrease in error for each homogenization step.

Table 8
Numerical results by homogenization steps.

As expected, more robust homogenization models combined with more experimented phases brought very satisfactory results. The error below 1% indicates that micromechanical models can accurately predict the homogenized properties of cementitious composites.

When evaluating the Poisson ratio homogenized for both the multiphase modeling (with ITZ and without ITZ), there were no significant differences, being in both models within the expected pattern Tables 9-10.

Table 9
Numerical results by homogenization steps for Poisson's ratio (without ITZ).
Table 10
Numerical results by homogenization steps for Poisson's ratio (with ITZ).

An estimate for the characteristic strength of concrete considered the values of Young's modulus found at Level 6 of the multiscale procedure, associated with two-phase models (without ITZ) Table 6 and multiple-phase models (with ITZ) Table 8. In the analysis, it was adopted a = 1 (Quartzite aggregates).

The estimated results for the characteristic strength of concrete, disregarding the interfacial transition zone, can be seen in Table 11.

Table 11
fck estimated using empirical models and homogenization (Level 6).

It is observed that there is an upward growth curve of the relative error, further demonstrating that it is not interesting to use two-phase models to estimate the strength of concrete. When using models that consider the interfacial transition zone, it is noted that the relative error also grows, but they are much smaller about two-phase models, which can be a support tool (Table 12).

Table 12
fck estimated using empirical models and homogenization (Level 6).

It should be noted that the neglect of the transition zone directly influences the modeling of this type of concrete, and therefore, it should be included in the analysis. For conventional concretes, the use of two-phase models is not recommended, as it is not possible to introduce the interfacial transition zone in the analysis even if one wants to.

Increasing the levels of micromechanical multiscale modeling would be a hypothesis for use of two-phase models, but it could lead to problems of the inclusions not being spherical, or the matrix not being infinite which would go against the micromechanical formulation.

5 CONCLUSIONS

With conventional concrete, it is proven that two-phase models need to be well adjusted to being able to evaluate the homogenized properties of concrete [4040 Y. Li, Y. Li, and R. Wang, “Quantitative evaluation of elastic modulus of concrete with nanoidentation and homogenization method,” Constr. Build. Mater., vol. 212, pp. 295-303, Jul. 2019.] using essentially two-phase models with adaptations to include the interfacial transition zones in the analysis since the literature already confirms that their absence interferes drastically in the final result of the modeling. The errors obtained by Li et al. [4040 Y. Li, Y. Li, and R. Wang, “Quantitative evaluation of elastic modulus of concrete with nanoidentation and homogenization method,” Constr. Build. Mater., vol. 212, pp. 295-303, Jul. 2019.] were 8.7% using his strategy.

By modifying the strategy used by Li et al. [4040 Y. Li, Y. Li, and R. Wang, “Quantitative evaluation of elastic modulus of concrete with nanoidentation and homogenization method,” Constr. Build. Mater., vol. 212, pp. 295-303, Jul. 2019.] proposing a multiscale analysis with robust models of multiple phases, satisfactory results are effectively achieved, reducing the error to 4% in the multiphase model and zero when using the four-phase model.

This panorama indicates that in conventional concrete it is essential to evaluate the interfacial transition zone with more complex models, however a robust experimental apparatus to model this cementitious composite is still needed. Any neglected phase can significantly interfere in the micromechanical modeling and the possibility of removing any of its phases must be investigated on a case-by-case basis.

Every micromechanical multiscale procedure was applied to the meso and macro scale of concrete, however, every analysis methodology can be used on the microscale and also on smaller scales such as the nanoscale, always considering the definition of a representative volume element [3737 H. Zhang, K. Anupam, A. Scarpas, and C. Kasbergen, “Comparison of different micromechanical models for predicting the effective properties of open graded mixes,” Transp. Res. Rec., vol. 2672, no. 28, pp. 404-415, Sep. 2018.].

ACKNOWLEDGEMENTS

The authors acknowledge the Laboratory of Scientific Computing and Visualization (LCCV), from the Technology Center (CTEC) at the Federal University of Alagoas (UFAL).

  • Financial support: None.
  • Data Availability: the authors confirm that all data underlying the findings openly available without restriction in “SciELO Data” available at https://doi.org/10.48331/scielodata.UPXLIE.
  • How to cite: R. M. S. Silva and A. S. R. Barboza, “Concrete modeling using micromechanical multiphase models and multiscale analysis,” Rev. IBRACON Estrut. Mater., vol. 16, no. 5, e16501, 2023, https://doi.org/10.1590/S1983-41952023000500001

1 REFERENCES

  • 1
    J. Fish, G. J. Wagner, and S. Keten, “Mesoscopic and multiscale modelling in materials,” Nat. Mater., vol. 20, pp. 777-786, Jun. 2021.
  • 2
    O. Lloberas-Valls, D. J. Rixen, A. Simone, and L. J. Sluys, “Multiscale domain decomposition analysis of quasi-brittle heterogeneous materials,” Int. J. Numer. Methods Eng., vol. 89, pp. 1337-1366, Nov. 2011.
  • 3
    O. Bernard, F.-J. Ulm, and E. Lemarchand, “A multiscale micromechanics-hydration model for the early-age elastic properties of cement-based materials,” Cement Concr. Res., vol. 33, pp. 1293-1309, Sep. 2003.
  • 4
    G. Constantinides and F.-J. Ulm, “The effect of two types of C-S-H on the elasticity of cement-based materials: results from nanoindentation and micromechanical modeling,” Cement Concr. Res., vol. 34, no. 1, pp. 67-80, Jan. 2004.
  • 5
    J. F. Unger and S. Eckardt, “Multiscale modeling of concrete,” Arch. Comput. Methods Eng., vol. 18, p. 341, Jul. 2011.
  • 6
    E. A. Rodrigues, “Um modelo multiescala concorrente para representar o processo de fissuração do concreto,” Ph.D. dissertation, Poli USP, São Paulo, Brazil, 2015.
  • 7
    S. Häfner, S. Eckardt, T. Luther, and C. Könke, “Mesoscale modeling of concrete: geometry and numerics,” Comput. Struct., vol. 84, no. 7, pp. 450-461, Feb. 2006.
  • 8
    S. Eckardt and C. Könke, “Adaptive damage simulation of concrete using heterogeneous multiscale models,” J. Algorithm Comput. Technol., vol. 2, no. 2, pp. 275-298, Jun. 2007.
  • 9
    S. Eckardt, “Adaptive heterogeneous multiscale models for the nonlinear simulation of concrete,” Ph.D. dissertation, BUW, Weimar, Germany, 2009.
  • 10
    V. P. Nguyen, M. Stroeven, and L. J. Sluys, “Multiscale failure modeling of concrete: Micromechanical modeling,discontinuous homogenization and parallel computations,” Comput. Methods Appl. Mech. Eng., vol. 201-204, pp. 139-156, Jan. 2012.
  • 11
    B. Pichler and C. Hellmich, “Upscaling quasi-brittle strength of cement paste and mortar: a multi-scale engineering mechanics model,” Cement Concr. Res., vol. 41, no. 5, pp. 467-476, May. 2011.
  • 12
    L. Göbel, M. Königsberger, A. Dimming-Osburg, and B. Picheler, “Viscoelastic behavior of polymer-modified cement pastes: insight from downscaling short-term macroscopic creep tests by means of multiscale modeling,” Appl. Sci., vol. 8, no. 4, p. 487, Mar. 2018.
  • 13
    K. Siventhirarajah, Y. Yoda, and Y. Elakneswaran, “A two-stage model for the prediction of mechanical properties of cement paste,” Cement Concr. Compos., vol. 115, p. 103853, Jan. 2021.
  • 14
    K. L. Scrivener, “Backscattered electron imaging of cementitious microstructures: understanding and quantification,” Cement Concr. Compos., vol. 26, no. 8, pp. 935-945, Nov. 2004.
  • 15
    M. S. Meddah, S. Zitouni, and S. Belâabes, “Effect of content and particle size distribution of coarse aggregate on the compressive strength of concrete,” Constr. Build. Mater., vol. 24, no. 4, pp. 505-512, Apr. 2010.
  • 16
    K.-R. Wu, B. Chen, W. Yao, and D. Zhang, “Effect of coarse aggregate type on mechanical properties of high-performance concrete,” Cement Concr. Res., vol. 31, no. 10, pp. 1421-1425, Oct. 2011.
  • 17
    H. Beushausen and T. Dittmer, “The influence of aggregate type on the strength and elastic modulus of high strength concrete,” Constr. Build. Mater., vol. 74, pp. 132-139, Jan. 2015.
  • 18
    N. V. Silva, “Método de determinação de resistência à tração e módulo de elasticidade de agragados graúdos naturais,” M.S. thesis, USP, São Paulo, Brazil, 2018.
  • 19
    G. Prokopskia and J. Halbiniakb, “Interfacial transition zone in cementitious materials,” Cement Concr. Res., vol. 30, no. 4, pp. 579-583, Apr. 2000.
  • 20
    K. L. Scrivener, A. K. Crumbie, and P. Laugesen, “The Interfacial Transition Zone (ITZ) between cement paste and aggregate in concrete,” Interface Sci., vol. 12, pp. 411-421, Oct. 2004.
  • 21
    G. Li, Y. Zhao, and S. Pang, “Four-phase sphere modeling of effective bulk modulus of concrete,” Cement Concr. Res., vol. 29, no. 6, pp. 839-845, Feb. 1999.
  • 22
    Z. Hashin, “Analysis of composite materials-a survey,” ASME J. Appl. Mech., vol. 50, pp. 481-505, Sep. 1983.
  • 23
    W. Drugan and J. Willis, “A micromechanics-based nonlocal constitutive equation and estimates of representative volume element size for elastic compositessize for elastic composites,” J. Mech. Phys. Solids, vol. 44, pp. 497-524, Apr. 1996.
  • 24
    M. Ostoja-Starzewski, “Microstructural randomness versus representative volume element in thermomechanics,” ASME J. Appl. Mech, vol. 69, pp. 25-35, Jun. 2002.
  • 25
    M. Stroven, H. Askes, and L. Sluys, “Numerical determination of representative volumes for granular materials,” Comput. Methods Appl. Mech. Eng., vol. 193, no. 30-32, pp. 3221-3238, Jul. 2004.
  • 26
    A. Drago and M.-J. Pindera, “Micro-macromechanical analysis of heterogeneous materials: macroscopically homogeneous vs periodic microstructures,” Compos. Sci. Technol., vol. 67, no. 6, pp. 1243-1263, May. 2007.
  • 27
    R. Christensen and K. Lo, “Solutions for effective shear properties in three phase sphere and cylinder models,” J. Mech. Phys. Solids, vol. 27, no. 4, pp. 315-330, Aug. 1979.
  • 28
    J. Eshelby “The determination of the elastic field of an ellipsoidal inclusion and related problems,” Proc. Royal Society London Ser. A, vol. 241, pp. 376-396, Mar. 1957.
  • 29
    Y. Benveniste, “A new approach to the application of Mori-Tanaka’s theory in composite materials,” Mech. Mater., vol. 6, no. 2, pp. 147-157, Jun. 1987.
  • 30
    Z. Hashin, “The differential scheme and its application to cracked materials,” J. Mech. Phys. Solids, vol. 36, no. 6, pp. 719-734, 1988.
  • 31
    L. Walpole, “On the overall elastic moduli of composite materials,” J. Mech. Phys. Solids, vol. 17, no. 4, pp. 235-251, Sep, 1969.
  • 32
    A. Kaw, Mechanics of Composite Materials, 2nd ed. New York, NY, USA: Taylor & Francis, 2006.
  • 33
    Z. Hashin and S. Shtrikman, “On some variational principles in anisotropic and nonhomogeneous elasticity,” J. Mech. Phys. Solids, vol. 10, no. 4, pp. 335-342, Oct./Dec., 1962.
  • 34
    T. Mori and K. Tanaka “Average stree in matrix and average energy of materials with mis-fitting inclusions,” Act. Metall., vol. 21, no. 5, pp. 571-574, May. 1973.
  • 35
    R. Hill, “A self-consistent mechanics of composite materials,” J. Mech. Phys. Solids, vol. 13, pp. 213-222, Aug. 1965.
  • 36
    A. Reuss, “Berechnung der fließgrenze von mischkristallen auf grund der plastizitätsbedingung für einkristalle,” J. Appl. Math. Mechanics, vol. 9, pp. 49-58, 1929.
  • 37
    H. Zhang, K. Anupam, A. Scarpas, and C. Kasbergen, “Comparison of different micromechanical models for predicting the effective properties of open graded mixes,” Transp. Res. Rec., vol. 2672, no. 28, pp. 404-415, Sep. 2018.
  • 38
    M. Hori and S. Nemat-Nasser, “Double-inclusion model and overall moduli,” Mech. Mater., vol. 14, no. 3, pp. 189-206, Jan. 1993.
  • 39
    C. Shi, H. Fan, and S. Li, “Interphase model for effective moduli of nanoparticle-reinforced composites,” J. Eng. Mechanic, vol. 141, no. 12, pp. 141-153, Dec. 2020.
  • 40
    Y. Li, Y. Li, and R. Wang, “Quantitative evaluation of elastic modulus of concrete with nanoidentation and homogenization method,” Constr. Build. Mater., vol. 212, pp. 295-303, Jul. 2019.
  • 41
    N. V. Silva, “Propriedades mecânicas das fases e seus efeitos no comportamento mecânico do concreto,” Ph.D. dissertation, Poli USP, São Paulo, Brazil, 2022.

Edited by

Editors: Osvaldo Manzoli, Guilherme Aris Parsekian.

Publication Dates

  • Publication in this collection
    16 Jan 2023
  • Date of issue
    2023

History

  • Received
    13 Aug 2022
  • Accepted
    25 Nov 2022
IBRACON - Instituto Brasileiro do Concreto Instituto Brasileiro do Concreto (IBRACON), Av. Queiroz Filho, nº 1700 sala 407/408 Torre D, Villa Lobos Office Park, CEP 05319-000, São Paulo, SP - Brasil, Tel. (55 11) 3735-0202, Fax: (55 11) 3733-2190 - São Paulo - SP - Brazil
E-mail: arlene@ibracon.org.br