Cohesive Crack Propagation in a Random Elastic Medium

Embed Size (px)

Citation preview

  • 8/13/2019 Cohesive Crack Propagation in a Random Elastic Medium

    1/13

    Probabilistic Engineering Mechanics 23 (2008) 2335

    www.elsevier.com/locate/probengmech

    Cohesive crack propagation in a random elastic medium

    M. Bruggia, S. Casciatib, L. Faravellia,

    aDepartment of Structural Mechanics, University of Pavia, via Ferrata 1, 27100 Pavia, ItalybASTRA Department, School of Architecture, University of Catania, via Maestranze 99, 96100 Siracusa, Italy

    Received 2 April 2007; received in revised form 28 September 2007; accepted 1 October 2007

    Available online 12 October 2007

    Abstract

    The issue of generating non-Gaussian, multivariate and correlated random fields, while preserving the internal auto-correlation structure of each

    single-parameter field, is discussed with reference to the problem of cohesive crack propagation. Three different fields are introduced to model

    the spatial variability of the Young modulus, the tensile strength of the material, and the fracture energy, respectively. Within a finite-element

    context, the crack-propagation phenomenon is analyzed by coupling a Monte Carlo simulation scheme with an iterative solution algorithm based

    on a truly-mixed variational formulation which is derived from the HellingerReissner principle. The selected approach presents the advantage

    of exploiting the finite-element technology without the need to introduce additional modes to model the displacement discontinuity along the

    crack boundaries. Furthermore, the accuracy of the stress estimate pursued by the truly-mixed approach is highly desirable, the direction of crack

    propagation being determined on the basis of the principal-stress criterion. The numerical example of a plain concrete beam with initial crack

    under a three-point bending test is considered. The statistics of the response is analyzed in terms of peak load and loadmid-deflection curves, in

    order to investigate the effects of the uncertainties on both the carrying capacity and the post-peak behaviour. A sensitivity analysis is preliminarily

    performed and its results emphasize the negative effects of not accounting for the auto-correlation structure of each random field. A probabilistic

    method is then applied to enforce the auto-correlation without significantly altering the target marginal distributions. The novelty of the proposed

    approach with respect to other methods found in the literature consists of not requiring the a priori knowledge of the global correlation structure

    of the multivariate random field.c 2007 Elsevier Ltd. All rights reserved.

    Keywords: Multivariate non-Gaussian random fields; Auto-correlation; Cohesive crack propagation; Truly-mixed finite-element method; Monte Carlo simulations

    1. Introduction

    The cohesive crack propagation problem is considered

    as a suitable example of having to generate non-Gaussian

    correlated random fields when considering the uncertainties

    of the physical parameters. The issue arises from observing

    that the simulation of non-Gaussian, multivariate random fields

    with a cross-correlation structure cannot be conceived but inan approximated manner [1]. Indeed, the task of matching

    the target marginal distributions conflicts with the one of

    preserving the spatial auto-correlation of each single-parameter

    random field. A probabilistic method for the generation

    of the non-Gaussian random fields is developed starting

    from the availability of traditional Gaussian random field

    Corresponding author.E-mail address: [email protected](L. Faravelli).

    realizations for each physical parameter, initially considered

    as independent of the others. In contrast to other methods

    found in the literature [24], the proposed procedure avoids

    the computational burden of directly considering the global

    correlation structure of the multivariate random field. In

    particular, the Gaussian realizations obtained by assigning

    each spectral density function are used to statistically estimate

    the covariance matrix of each corresponding random field.

    Thence, the auto-correlation structure of each random field is

    obtained in an already discretized manner, as an alternative

    to the common practice of first assigning an auto-correlation

    function of exponential type and then discretizing it [2]. The

    eigenvectors of each covariance matrix are then applied to

    the cross-correlated, non-Gaussian entries resulting from the

    inverse Nataf transform, in order to restore the auto-correlation

    structure of each field. Although the latter eigenvector mapping

    slightly alters the marginal distributions of the random

    0266-8920/$ - see front matter c

    2007 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.probengmech.2007.10.001

    http://www.elsevier.com/locate/probengmechmailto:[email protected]://dx.doi.org/10.1016/j.probengmech.2007.10.001http://dx.doi.org/10.1016/j.probengmech.2007.10.001mailto:[email protected]://www.elsevier.com/locate/probengmech
  • 8/13/2019 Cohesive Crack Propagation in a Random Elastic Medium

    2/13

    24 M. Bruggi et al. / Probabilistic Engineering Mechanics 23 (2008) 2335

    variables, the results of a sensitivity analysis show that

    accounting for the internal auto-correlation is fundamental in

    order to obtain reliable results in terms of statistics of the

    response.

    The developed probabilistic method is first validated using

    a simple numerical example (whose results are reported in the

    Appendix), and is then applied to the problem of cohesive crackpropagation. Three different fields are generated in order to

    model the spatial variability of the Young modulus, the tensile

    strength of the material, and the fracture energy, respectively,

    associated with the crack development. It is worth noting that

    a scalar representation of the Young modulus at any point of

    the body implicitly introduces an isotropy assumption, which

    is in conflict with the inherent anisotropic nature of a random

    medium. The finite-element discretization allows, however, to

    conceive an anisotropic medium as the assemblage of isotropic

    finite elements. Being a full simulation of all anisotropic elastic

    and failure parameters beyond the scope of this study, a scalar

    definition of the Young modulus is assigned at each point for

    the sake of convenience.

    The numerical study of a crack-propagation phenomenon

    requires a mechanical model able to follow the crack-path

    evolution, which is a priori unknown and not aligned with

    the body discretization of the initial un-cracked domain.

    In [5], an adaptive remeshing strategy was proposed. Further

    studies aimed to limit and possibly avoid the computationally

    expensive remeshing phase. The procedures developed from

    these studies usually rely on either one of two alternative

    strategies: the XFEM (extended finite-element method), or

    the meshless approach. The XFEM method is based on a

    continuous displacement formulation which needs to be locally

    enriched with discontinuous modes in order to be able to cross

    the existing mesh by exploiting the partition-of-unit property

    of the shape functions [6]. The meshless strategy [7] seems

    to be ideally tailored to handle crack-propagation problems,

    but must overcome some numerical difficulties, such as the

    quadrature formulas and the boundary conditions assignment,

    which are easily solved by a finite-element approach.

    The approach adopted in the present paper cannot be

    grouped in any of the two afore-mentioned categories, since

    it is based on an extension of the truly-mixed variational

    formulation developed in [8] from the HellingerReissner

    principle. The associated solution algorithm, which is able

    to follow the a priori unknown crack path in both the pre-and post-peak regimes, was proposed and verified in [9].

    This approach is chosen to be coupled with a Monte Carlo

    simulation scheme because it presents several advantages.

    By using equilibrated stress fields, with square-integrable

    divergence and inherently discontinuous displacements, the

    stress-flux continuity is imposed in an exact manner at each

    load step. Furthermore, the potentially active discontinuity of

    the displacements at each crack-interface element allows the

    direct inclusion of a cohesive law. Within this framework,

    the stress element of Johnson and Mercier [10] is selected, it

    being one of the very few elements able to pass the infsup

    condition required for the convergence of the method.

    The problem of cohesive crack propagation in elastic media

    is investigated by coupling the above-mentioned mechanical

    model based on the truly-mixed formulation, and the newly

    proposed probabilistic method for the generation of 2D and

    multiparameter cross-correlated random fields of non-Gaussian

    nature. In the following, the theoretical backgrounds of the

    random field generation procedure and the truly-mixed finite-element formulation are discussed in separate sections, and

    are then jointly applied to a numerical example. Within this

    example, Monte Carlo simulations with finite elements are

    carried out to determine the statistics of the response of a

    plain concrete beam undergoing a three-point bending test.

    A sensitivity analysis is performed on a limited number

    of samples in order to preliminarily check the influence of

    different probabilistic assumptions on the results. Finally, the

    effects of the uncertainties on both the carrying capacity and the

    post-peak behaviour are quantified by considering a significant

    number of random field realizations.

    The methodology developed in this work can be applied

    for further developments within fracture mechanics of quasi-brittle materials. Indeed, both the energetic size effect (of

    a deterministic nature) and the randomness in the material

    properties affect the maximum load-carrying capacity of a

    structural component. According to Ref. [11], for a certain class

    of structures, the first factor governs the deterministic mean

    of the nominal strength, while the second is responsible for

    the higher order moments. As such, a probabilistic approach

    is needed in order to evaluate the probability density function

    of the response, in view, for example, of an estimate of the

    reliability of the structure [12]. Furthermore, when considering

    the microscopic origin of the crack formation, homogenization

    techniques are usually applied to model a standard continuumwhich behaves like the originally micro-cracked body

    [13,14]. It is, therefore, of interest to check the hypothesis of

    a homogenized random field by evaluating the influence of the

    spatial variability of the material properties on the response.

    2. The probabilistic approach

    2.1. Framing the problem

    In the literature, random fields were first introduced as a 2D

    natural extension of stochastic processes [1517]. Moreover,

    the simulation of their realizations provided the support for the

    development of stochastic finite elements [1821]. Advancedtopics include the generation of spatial-temporal wind velocity

    fields[22,23]and the discretization issues in crack-propagation

    analysis[24]. An extended state-of-the-art report can be found

    in Ref. [25]. Refs. [14] are devoted to the development

    of simulation methods for non-Gaussian processes. Very

    few authors [2,4] discuss the simulation of multivariate and

    cross-correlated non-Gaussian random fields and the existing

    approaches are all, to some extent, approximate. For this

    reason, the issue of providing a non-Gaussian nature to a cross-

    correlated, multiparameter random field while preserving its

    internal auto-correlation structure, is still considered an open

    area of research.

  • 8/13/2019 Cohesive Crack Propagation in a Random Elastic Medium

    3/13

    M. Bruggi et al. / Probabilistic Engineering Mechanics 23 (2008) 2335 25

    The selective list of references mentioned above does

    not aim to entirely capture the broad spectrum of results

    provided by the research activities in the related areas, but

    only to represent those that directly influenced the authors

    in developing the probabilistic approach followed in this

    paper. In particular, one assumes that a commercial software

    for the simulation of Gaussian random fields with assignedspectral density function is readily available. The task is then

    to investigate how it can be conveniently exploited when

    simulating cross-correlated random fields of non-Gaussian

    nature. The answer found in this work consists of being able

    to statistically estimate, from several independently generated

    Gaussian random fields, the corresponding auto-correlation

    matrices whose eigenvectors are then used to restore the

    spatial auto-correlation destroyed when applying a non-linear

    transformation, such as the inverse Nataf transform. This

    approach avoids the computational burden of building the

    global auto-correlation structure of the multidimensional and

    multivariate random field. Furthermore, the auto-correlation

    of each considered random field is assigned in an alreadydiscretized manner, thus avoiding the following of the standard

    procedure of first assigning an auto-correlation function of

    exponential type, and then discretizing it. The steps that must

    be taken in order to first match the given marginal probability

    distributions and cross-correlation structure, and then enforce

    the spatial auto-correlation are discussed in the following sub-

    section.

    2.2. The proposed algorithm for the simulation of cross-

    correlated non-Gaussian random fields

    A multivariate (m= 3) and multidimensional (n= 2) non-Gaussian stochastic field is defined over a rectangular domainH(x), x Rn . (1)The domain is discretized into an rby s grid consistent with the

    finite-element mesh, so that N= r s is the number of points towhich each random field sample of size m is assigned.

    The first step is to generate a discrete sample of uncorrelated

    standard Gaussian vectors. For computational convenience, the

    random field vectors are organized in a matrix ofNrows andm

    columns. Therefore, each row vector, U(x i ), is the uncorrelated

    multiparameter Gaussian sample associated to the i th node of

    the grid,i=

    1, . . . ,N.

    A mapping of each vector U(x i )to the correlated Gaussian

    space is performed as follows

    z= L U (2)L being the Cholesky decomposition of the correlation matrix

    obtained by applying the regression formulas in Ref.[26] to the

    values of the correlation coefficients originally prescribed for

    H(x) in the non-Gaussian space. The resulting sample, z, of

    correlated Gaussian variables is then transformed into a non-

    Gaussian vector, by applying the inverse Nataf transformation

    to each scalar component of the random field

    Hh(x i )=F1

    H h[(zi )] (3)

    for h = 1, . . . ,m and i = 1, . . . ,N. Due to the non-linearity of the transformation, the single Hh(x)vector (i.e., thevalues assumed by each parameter in different nodes) is made

    of uncorrelated random variables. Hence, a transformation

    restoring the internal correlation must be performed, and it is

    given by

    Hh(x)=N

    i=1

    iHh(x i ) (4)

    where iare the eigenvectors of the target auto-correlation

    matrix of the single h th discretized random field. This matrix

    can be either obtained by discretization of the auto-correlation

    function, which is usually assumed to be of exponential type

    with given correlation length, or by assigning the spectral

    density function of each single-parameter random field. In the

    latter manner, one can ignore the global structure of the cross-spectral density matrix and consider only each single spectral

    density function, together with the correlation coefficients of

    the vector of sizem assigned to each node by Eq.(2).It is worth noting that, in the above summarized procedure,

    the order in which the operations in Eqs. (3) and (4)

    are performed is fundamental. Indeed, if the eigenvector

    mapping takes place before the inverse Nataf transformation,

    the non-linearity of the latter transformation destroys the

    correlated results expected from the first operation. As a

    consequence, non-Gaussian cross-correlated fields made of

    auto-uncorrelated entries would be generated. An iterative

    procedure could be pursued by altering the auto-correlationstructure, but the inverse Nataf transformation would still

    produce internally uncorrelated variables. Instead, when

    the inverse Nataf transformation precedes the eigenvector

    mapping, the second transformation reaches its expected goalof correlating the internal variables, whereas the marginal

    distributions pursued by the Nataf transformation are just

    slightly altered. This alteration is minimal if the starting point

    is made of internally independent realizations. In other words,

    the sequence eigenvectorNatafeigenvector would produce

    less accurate results than the simplest path Natafeigenvector

    does. The above statements are supported by the results of the

    numerical analyses carried out in theAppendix.

    3. The mechanical model

    3.1. Truly-mixed finite-element formulation for cracked media

    The crack-propagation algorithm used in this paper was

    originally developed in [9], and it is intimately tied to the

    truly-mixed finite-element formulation which is here only

    briefly recalled. In the following, the governing equations are

    presented in a mixed weak continuous form, which is soon after

    discretized by resorting to the JohnsonMercier stress element.Adopting a classical notation, one defines, within the domain

    R2, the square-integrable vector of the body loads, g , andindicates with and the unknown stress and strain tensors,

    respectively.The compatibility equation (written in terms of stresses

    through the elastic constitutive law, =

    C

    1, with C the

  • 8/13/2019 Cohesive Crack Propagation in a Random Elastic Medium

    4/13

    26 M. Bruggi et al. / Probabilistic Engineering Mechanics 23 (2008) 2335

    fourth-order tensor of the elastic constants) is tested for a virtual

    stress field, . The equilibrium equation is tested by means of

    a virtual displacement field, v. The resulting weak variational

    formulation reads: find( , u)S(div;)22sym[L2()]2 suchthat:

    C

    1

    : dx+ div udx

    [[u]] ( n)dx=0, S(div;),

    div vdx=

    g vdx= 0, v [L2()]2(5)

    where one defines S(div;)22sym = { : i j = ji L2(),div [L2()]2}, andL 2()is the space of the functions thatare square integrable on the domain .

    When a macro-crack propagates through the medium, the

    line integral in the first row of Eq. (5) accounts for the

    (cohesive) energy dissipated across the fracture, it being derived

    from the GaussGreen theorem

    u dx=

    div udx+

    u ( n)d (6)

    where = 1 2, with 1,2 the two opposite sides of thecrack, and( n)is the stress flux acting on them.

    In Eq.(5), the operator[[]]denotes the strong discontinuityof the quantity to which it is applied, i.e., the displacement jump

    between the two opposite edges of the crack, which, under the

    small displacements assumption, can be written as

    u ( n)d=1

    u ( n)d2

    u ( n)d

    =

    [[u]] ( n)d. (7)

    By introducing the rate independent and piece-wise linear

    idealization in Fig. 1 as a cohesive law, the displacement

    jumps, [[u]], can be expressed as functions of the tractionstress fluxes, ( n). Three regions (labeled A, B and C,respectively, in Fig. 1) need to be distinguished in order to

    model the global behaviour of the cohesive interface. The first

    (zone A) is representative of a regime where the medium is

    un-cracked and the resistance, t , is larger than the currentnormal traction. After the crack initiation (zone B), an energy

    release takes place driving the current normal traction to a value

    that is smaller than the resistance, t. However, as long as thedisplacement jump,[[u]]n , in the normal direction with respectto the crack, is smaller than a critical value, [[u]]n , there is still aresidual cohesion between the two sides of the crack. Once this

    threshold value is reached (zone C), no cohesion between the

    two sides of the crack is experienced.

    When deploying the described truly-mixed approach within

    a finite-element discretization, the main difficulty to cope

    with is due to the symmetry of the stress tensors. The two

    interpolation fields of stresses and displacements must satisfy

    the so-called infsup condition in order for the method to be

    globally convergent. Among the few available approaches able

    to pass the infsup condition in a truly-mixed setting, the

    Fig. 1. Pure mode I cohesive law.

    Fig. 2. The JM triangular element.

    Fig. 3. Stress (circles) and displacement (squares) degrees of freedom.

    Fig. 4. Deterministic geometry assumed for the numerical example of a three-

    point bending test on a plain concrete beam with initial crack (dimensions inmm).

    Table 1

    Defining the marginal distributions of the three-variate random field

    Random variable Distribution

    type

    Mean Coefficient of variation

    Young modulus, E Lognormal

    (L)

    36.5

    (GPa)

    0.2

    Tensile strength, f Weibull (W) 3.19

    (MPa)

    0.2

    Specific fracture

    energy,G

    Weibull (W) 100

    (N/m)

    0.2

  • 8/13/2019 Cohesive Crack Propagation in a Random Elastic Medium

    5/13

    M. Bruggi et al. / Probabilistic Engineering Mechanics 23 (2008) 2335 27

    Fig. 5. Statistics of the 1000 realizations in the first node of the grid, before

    forcing the internal correlation.

    JM element (by Johnson and Mercier) is selected. It consists

    of a composite triangular element, which is made of three

    sub-triangles (Fig. 2). Stress shape functions with complete

    first-order polynomial bases are defined in each sub-triangle,

    while shape functions of the same polynomial type model the

    globally discontinuous displacements over the whole triangle.

    This discretization results into a number of degrees of freedom

    per element equal to six for the displacements, and equal to

    fifteen for the stresses (Fig. 3).

    The main advantage of this approach consists of not

    requiring the introduction of any extra mode or shape function

    in order to handle the energy dissipation across the fracture.

    In fact, the displacement field is discontinuous per se and it is

    Fig. 6. Statistics of the1000 realizationsin thefirst node of thegrid,accounting

    for the internal correlation.

    sufficient to evaluate the line integral of Eq.(5)across the crack

    to take into account the entire phenomenon.

    3.2. The solution algorithm

    The fracture path, , being a priori unknown, the problem

    is inherently non-linear. In the present paper, the solution

    process is handled by an iterative algorithm, which works on a

    linearization of the original problem at each load step. First, the

    current mixed matrix is computed and the following linearized

    matrix equationA() Bu

    Bu 0

    u

    =

    0

    g

    , (8)

    is solved, with the blockA()evaluated by updating the line

    integral in Eq.(5).

  • 8/13/2019 Cohesive Crack Propagation in a Random Elastic Medium

    6/13

    28 M. Bruggi et al. / Probabilistic Engineering Mechanics 23 (2008) 2335

    Fig. 7. Example of one realization of the three-parameter random field: (a) Elastic modulusE(GPa), (b) Tensile strength f(MPa), (c) Fracture energy G (N/m).

    A check on the elements in the regimes [B] or [C] ofFig. 1

    is then performed. For each edge element on the crack, thedisplacement jump is evaluated as the difference between the

    displacements of the nodes facing each other on the opposite

    sides of the crack itself. If the jump is less than the critical value,

    an update of type [B] is added to the mixed governing matrix.

    Conversely, if the threshold jump value is exceeded, the relevant

    node is added to the tail of the crack where no residual cohesion

    is detected, and the two sides of the crack are independent of

    each other.

    This procedure must be repeated at fixed load until

    convergence is achieved, meaning that the cohesive constitutive

    law in Fig. 1 is exactly imposed over the entire crack. Once

    achieved, a (positive) load increment is further applied to thestructure, until a proper stress average on a small contour

    centered at the crack tip reaches the limit stress value, t. Whenthis occurs, the principal-tensile-stress criterion is adopted, and

    the crack is allowed to propagate in the direction normal to the

    maximum tensile stress.

    The presented steps are repeated until no equilibrium

    configurations are found. If this is the case, the maximum

    load sustainable by the specimen is reached and negative

    loading increments (decrements) are applied to the structure so

    that it exploits its post-peak softening regime. The procedure

    is stopped when failure occurs, i.e. when the residual load-

    carrying capacity of the structure is null.

    4. Numerical example

    The classical problem of a three-point bending test on a plain

    concrete beam with initial crack (Fig. 4) is considered. This ex-

    ample was deterministically analyzed in [27]and probabilisti-

    cally approached in [2]. In the latter reference, a non-Gaussian,

    multiparameter random field was generated by building an ex-

    tended covariance matrix and it was coupled with a meshless

    strategy to run the Monte Carlo simulations. In the present

    work, a random field generation method that avoids consider-

    ing the global correlation structure is instead preferred, as em-

    phasized in Section2. Furthermore, the Monte Carlo simula-

    tions are performed within a finite-element scheme exploiting

    the truly-mixed variational formulation described in Section3.Using the JM element inFig. 2,the values of each random

    field sample are assigned to the vertices of the squares formed

    by four elements. The random fields are accordingly discretized

    into a grid of 10 by 40 nodes. At each node, the realizations

    of three random variables corresponding to the material Young

    modulus, its tensile strength and the specific fracture energy,

    respectively, are specified.

    Table 1 collects the statistics of the random variables

    considered in the problem. In particular, the same coefficient of

    variation, equal to 0.2, and the same correlation coefficient of

    0.8 are assumed for all the random variables having, however,

    different distribution types and mean values. The latter ones

  • 8/13/2019 Cohesive Crack Propagation in a Random Elastic Medium

    7/13

    M. Bruggi et al. / Probabilistic Engineering Mechanics 23 (2008) 2335 29

    Fig. 8. Local remeshing for crack propagation: (a) crack path at the specimen collapse and (b) relevant stress map for stress normal to the edge (MPa).

    Fig. 9. Load-relative mid-deflection statistical and deterministic curves computed over the 100 MC samples obtained from the probabilistic assumptions of cases:

    (a) (a) and (b) (b) ofTable 2.

    Table 2

    Sensitivity analysis of the statistical assumptions, on the basis of only 100 realizations of the multiparameter random field

    Case Distribution types Coefficients of variation Auto-correlation Correlation coefficients Peak load statistics

    E f G E f G E f E G f G Mean (kN) Variance(kN2)

    (a) L W W 0.20 0.20 0.20 Yes 0.80 0.80 0.80 62.802 36.5276

    (b) L W W 0.20 0.20 0.20 No 0.80 0.80 0.80 62.6317 14.8563

    (c) L W W 0.15 0.18 0.20 No 0.80 0.80 0.80 62.5745 12.8776

    (d) L W W 0.20 0.20 0.20 No 0.70 0.50 0.90 62.6624 17.0197

    (e) L W W 0.15 0.18 0.20 No 0.70 0.50 0.90 62.6088 14.8543

    (f) L L L 0.20 0.20 0.20 No 0.80 0.80 0.80 62.6337 15.1436

    (g) L L L 0.20 0.20 0.20 No 0.00 0.00 0.00 62.2682 12.1846

    (h) L L L 0.20 0.20 0.20 Yes 0.00 0.00 0.00 61.4451 24.7527

    reflect the physical quantities used for the deterministic study

    in [27], while the second-order moments are selected as in [2]

    to allow for a comparison of the probabilistic results. In [2] this

    choice was motivated by a higher computational simplicity. It

    is worth noting that the probabilistic approach proposed in the

    present work does not have these computational limitations,

    because it avoids building the global covariance matrix of

    the multivariate random field. Furthermore, Ref. [2] assumes

    all the variables to be lognormally distributed, while here

    a Weibull distribution assumption is introduced for both the

    tensile strength and the fracture energy, as suggested by most

    of the general literature. The sensitivity of the results to

    different probabilistic assumptions is verified in the following

    Section4.2.

    In order to generate each random field, the corresponding

    spectral density matrix is given. The covariance matrix and the

  • 8/13/2019 Cohesive Crack Propagation in a Random Elastic Medium

    8/13

    30 M. Bruggi et al. / Probabilistic Engineering Mechanics 23 (2008) 2335

    spectral density matrix are then handled by simply assigning

    one function to each variable, rather than considering a matrix

    of functions. Following the reasoning in [19,28], the spectral

    density function is assumed to be of the form

    G(kx , ky)

    =2

    dx dy

    4 exp

    kx dx

    2

    2

    + ky dy

    2

    2

    (9)where kx and ky are the wave numbers defined in the

    interval (,+), and is the standard deviation. The twoparametersdx anddy must be selected in such a way that they

    are consistent with the finite-element mesh. For this purpose,

    one first estimates the minimal lag of the mesh in the two

    directions [28], then the Nyquist cut-off values, kxu and kyu ,

    and thence

    dx=

    2

    kxu /3; dy=

    2

    kyu/3. (10)

    For the specific example under investigation, Eq.(10)leads to avalue of 20.257 mm in both directions, so that the associated

    exponential auto-correlation function shows a length which

    covers nearly three elements. The target auto-correlation matrix

    of each non-Gaussian random field could then be obtained

    from the discretization of the afore-mentioned auto-correlation

    function. As an alternative, the following strategy is instead

    adopted: 1000 realizations of the standard Gaussian process

    are simulated following the spectral density scheme in Eq. (9),

    and their auto-correlation matrix is afterwards estimated on a

    statistical basis. This operation provides the 400 by 400 matrix

    whose eigenvectors are used in Eq. (4), to give an internal

    correlation to the realizations achieved after the inverse Nataf

    transformation of Eq.(3).

    By applying the stochastic modelling procedure proposed

    in Section 2.2, a total of 1000 realizations are generated for

    each random field.Figs. 5and6provide a comparison between

    the assigned marginal distributions of each parameter and

    those estimated from the corresponding 1000 realizations, as

    achieved before and after forcing the spatial auto-correlation

    by means of the eigenvectors mapping in Eq. (4), respectively.

    In particular,Fig. 5provides a synthesis of the statistics of the

    three variables after the inverse Nataf transformation of Eq.

    (3),but before forcing the internal correlation of each random

    field by Eq.(4). Instead,Fig. 6provides the final statistics of

    the input parameters, thus accounting for their spatial auto-correlation. By comparing the two figures, one can observe how

    the last step of the stochastic modelling procedure slightly alters

    the marginal probability distributions of the parameters, but the

    entity of this effect is negligible. The sensitivity analysis carried

    out in the following Section 4.2 emphasizes the importance

    of accounting for the auto-correlation, whose absence leads

    to unreliable results in terms of statistics of the response.

    Within an approximated framework, the slight alteration of the

    marginal distributions is, therefore, acceptable with respect to

    the advantage of accounting for the spatial auto-correlation.

    Finally, the realizations generated by accounting for the spatial

    auto-correlation are assigned to the corresponding nodes of

    Fig. 10. Mean values of peak loads calculated on increasing number of MC

    realizations compared to the deterministic value (dotted line).

    Fig. 11. Standard deviation of the peak loads calculated for an increasing

    number of MC realizations.

    the finite-element mesh. Fig. 7 shows, as an example, one

    realization of the resulting multiparameter random field.

    4.1. Remark on the solution algorithm

    In principle, the algorithm outlined in Section2.2can follow

    any crack-path geometry, by choosing the crack-propagation

    direction on the basis of the maximum tensile-stress criterion.

    For this purpose, the accuracy of the stress estimate provided by

    the mixed approach is higher than the other methods. However,

    the computation of the line integral in Eq. (5) calls for a

    slight local remeshing to align the evolving crack path with

    the boundaries of the adjacent finite elements involved in thefracture. This procedure is automatic, but it slows down the

    structural analysis. In view of the high computational effort

    always demanded when adopting a Monte Carlo simulation

    scheme, the assumption that the system is symmetric with

    respect to both the geometry and the load is made.

    Sufficient checks are performed to ensure the validity of

    this assumption. For this purpose, the algorithm in its general

    formulation, able to represent any crack trajectory, is initially

    applied to random field samples of reduced size (just 100

    replicates). As an example of the achieved results, Fig. 8(a)

    shows, at the end of a simulation, a very small deviation of

    the crack path from the straight vertical line. Fig. 8(b) plots

  • 8/13/2019 Cohesive Crack Propagation in a Random Elastic Medium

    9/13

    M. Bruggi et al. / Probabilistic Engineering Mechanics 23 (2008) 2335 31

    Fig. 12. Load-relative mid-deflection diagrams obtained over the 1000 MC samples referring to case (a) inTable 2: (a) statistical and deterministic curves; (b)

    envelope.

    Fig. 13. Histogram and fitting normal probability distribution for peak loads

    calculated on 1000 MC samples.

    the relevant stress map at the same load step. Similar results

    are obtained when using different random field realizations as

    input. It can be concluded that the removal of the symmetry

    assumption does not significantly improve the accuracy of the

    method, but only augments its computational time. Therefore,

    in order to rely on a faster algorithm, all the following

    computations are carried out by a priori approximating the

    crack path with a straight line. It is worth noting that, under this

    assumption, the mixed approach still presents advantages with

    respect to the XFEM method, since the inherent displacement

    discontinuity prevents us from having to introduce extra

    discontinuous modes in order to allow the crack propagation.

    4.2. Preliminary studies of the uncertainty propagation

    A sensitivity analysis is performed with respect to the

    statistical assumptions made for the random variables involved

    in the problem under consideration. The effects of changing the

    data as reported inTable 2are investigated by calculating, for

    each considered case, the statistics of the response. In order

    to cover all the cases envisioned in Table 2 with a moderate

    computational effort, the analyses are performed on only 100

    realizations of each parameter random field.

    The first row ofTable 2 denotes the original assumptions

    inTable 1 as case (a). Case (b) differs only in not having an

    auto-correlation structure. Cases (c) through (e), together with

    not having an auto-correlation structure, also consider different

    values of the coefficients of variation (case (c)), the correlation

    coefficients (case (d)), or both (case (e)). Finally, cases (f)through (h) assume a lognormal distribution for all the variables

    which are first considered as mutually correlated but without

    any auto-correlation (case (e)), as independent and not auto-

    correlated (case (f)), and lastly the introduction of the auto-

    correlation in the independent case is evaluated (case (h)).

    The results in terms of peak load statistics are reported in

    the last two columns of Table 2. It can be noted that, while

    the mean value is not significantly affected by the changes of

    the initial assumptions, the higher order moments seem to be

    more sensitive to the actual statistics of the random variables.

    In particular, when the internal auto-correlation within the

    realizations of each random field is not considered, a drop inthe variance of the response is observed.

    Fig. 9(a) and (b) illustrate the deterministic and statistical

    (mean and mean standard deviation) load-relative mid-deflection curves obtained from the cases (a) and (b) ofTable 2,

    respectively. The difference between the values of the peak

    load variance in the two cases is evident along the entire

    curves. Moreover, the curve of the mean values resulting

    from neglecting the auto-correlation structure (case (b)) does

    not match the deterministic curve as good as it does when

    accounting for the auto-correlation (case (a)), especially in the

    softening branch.

    In conclusion, the results of the sensitivity analyses justify

    a certain freedom in selecting the statistical properties of the

    input parameters (e.g., same coefficients of variation and same

    correlation coefficients), but also emphasize the importance of

    simulating each random field with an internal structure that is

    in agreement with an assigned auto-correlation matrix.

    4.3. Statistical analysis of the response

    Figs. 10and11provide evidence that the results obtained

    in the last two columns of Table 2 by considering only 100

    replicates are not representative estimates of the mean and

    variance of the response, respectively. Indeed, to reach a

    convergence in the response statistical properties, at least 500

  • 8/13/2019 Cohesive Crack Propagation in a Random Elastic Medium

    10/13

    32 M. Bruggi et al. / Probabilistic Engineering Mechanics 23 (2008) 2335

    Fig. A.1. Statistic elaboration over a sample of size 1000: (a) after the Nataf transformation, and (b) after the Natafeigenvector sequence.

    realizations of each parameter random field must be considered.

    For this reason, the analyses are now repeated for samples of

    1000 realizations. These realizations are generated according

    to the probabilistic assumptions of case (a) in Table 2, thus

    accounting for the spatial auto-correlation structure of each

    parameter random field.The Monte Carlo finite-element analyses are carried out by

    running the solution algorithm described in Section 3.2 for

    the 1000 realizations obtained from the simulation strategy of

    Section2.2,with the marginal distributions as given in Table 1

    and the spectral density function of Eq.(9).At the end of each analysis, the loaddisplacement curve

    is calculated. Its statistics (mean and standard deviation)

    are then computed over its 1000 realizations. In Fig. 12(a),

    the resulting statistical loaddisplacement curves (mean and

    mean standard deviation) are plotted together with thedeterministic diagram. A good agreement between the curve

    of the mean values and the deterministic one is obtained. In

    Fig. 12(b), the envelope of the curves obtained from each Monte

    Carlo simulation shows that the uncertainties have remarkable

    effects in the zone next to the load-carrying capacity and in the

    softening branch.

    The histogram in Fig. 13 gives the probability density

    function (pdf) of the peak load. Among the known distribution

    models, a Normal pdf with a mean of 62.50 kN and a standard

    deviation of 6.81 kN is also drawn in Fig. 13as the curve that

    best fits the histogram.

  • 8/13/2019 Cohesive Crack Propagation in a Random Elastic Medium

    11/13

    M. Bruggi et al. / Probabilistic Engineering Mechanics 23 (2008) 2335 33

    Fig. A.2. Influence on the statistics of the realizations of the starting auto-correlation: (a) wished (i.e., assigned as inTable A.1), (b) fully correlated, and (c)

    uncorrelated.

    Table A.1

    Assigning the three-parameters, 2D field of theAppendixexample: (a) marginal distribution and central moments, (b) local cross-correlation coefficients, and (c)auto-correlation matrix

    Random variable Marginal distribution Mean Standard deviation

    (a)

    X1 Lognormal 36.5 7.3

    X2 Weibull 3.19 0.638

    X3 Weibull 100 20

    (b)

    Cross-correlation coefficients, i j X1 X2 X3X1 1 0.8 0.8

    X2 0.8 1 0.8

    X3 0.8 0.8 1

    (c)

    Ra=

    1 0.8 0.5 0.2 0 0.5 0.4 0.25 0.1 0 0 0 0 0 00.8 1 0.8 0.5 0.2 0.4 0.5 0.4 0.25 0.1 0 0 0 0 0

    0.5 0.8 1 0.8 0.5 0.25 0.4 0.5 0.4 0.25 0 0 0 0 0

    0.2 0.5 0.8 1 0.8 0.1 0.25 0.4 0.5 0.4 0 0 0 0 0

    0 0.2 0.5 0.8 1 0 0.1 0.25 0.4 0.5 0 0 0 0 0

    0.5 0.4 0.25 0.1 0 1 0.8 0.5 0.2 0 0.5 0.4 0.25 0.1 0

    0.4 0.5 0.4 0.25 0.1 0.8 1 0.8 0.5 0.2 0.4 0.5 0.4 0.25 0.1

    0.25 0.4 0.5 0.4 0.25 0.5 0.8 1 0.8 0.5 0.25 0.4 0.5 0.4 0.25

    0.1 0.25 0.4 0.5 0.4 0.2 0.5 0.8 1 0.8 0.1 0.25 0.4 0.5 0.4

    0 0.1 0.25 0.4 0.5 0 0.2 0.5 0.8 1 0 0.1 0.25 0.4 0.5

    0 0 0 0 0 0.5 0.4 0.25 0.1 0 1 0.8 0.5 0.2 0

    0 0 0 0 0 0.4 0.5 0.4 0.25 0.1 0.8 1 0.8 0.5 0.2

    0 0 0 0 0 0.25 0.4 0.5 0.4 0.25 0.5 0.8 1 0.8 0

    0 0 0 0 0 0.1 0.25 0.4 0.5 0.4 0.2 0.5 0.8 1 0

    0 0 0 0 0 0 0.1 0.25 0.4 0.5 0 0.2 0.5 0.8 1

  • 8/13/2019 Cohesive Crack Propagation in a Random Elastic Medium

    12/13

    34 M. Bruggi et al. / Probabilistic Engineering Mechanics 23 (2008) 2335

    5. Conclusions

    The problem of stochastic cohesive crack propagation is

    numerically investigated by coupling a truly-mixed finite-

    element approach with a Monte Carlo simulation scheme.

    The formulation manages the fracture problem thanks to the

    peculiar nature of the adopted discretizing fields. Discontinuous

    displacements and continuous stress fluxes directly allow the

    simulation of crack propagation along element boundaries.

    Furthermore, a multiparameter stochastic field is introduced to

    model the material properties. For this purpose, a methodology

    that pursues to assign an auto-correlation structure to the

    generated non-Gaussian, cross-correlated fields is developed,

    without having to consider the global structure of the

    multiparameter covariance matrix. A classical three-point

    bending specimen made of concrete is used to perform

    numerical tests on the proposed methodology. Results from

    Monte Carlo analyses are shown to determine the statistics

    of the response, with peculiar attention to load-crack mouth

    opening diagrams and load-carrying capacity.

    Acknowledgements

    The authors acknowledge the grants received from the

    Athenaeum research funds of the University of Catania and the

    University of Pavia.

    Appendix

    In order to support the development of the probabilistic

    approach proposed in Section 2 for the generation of cross-

    correlated, non-Gaussian random fields, a simplified example

    is used here to test the consequences of adopting different

    options. The objective is to simulate a three-parameter, 2D

    random field, with the marginal distributions, the local cross-

    correlation structure, and the spatial auto-correlation assigned

    inTable A.1.In particular, the auto-correlation is specified by a

    15 15 matrix, corresponding to a 5 3 nodal discretization ofthe field.

    The simplest way to generate the realizations of the

    multiparameter random field consists of simulating a sequence

    of three times 5 3, independent standardized Gaussiannumbers, which can be regarded as a realization of three

    independent Gaussian fields. They must then be transformed

    into non-Gaussian cross-correlated quantities, and subsequentlyinto auto-correlated fields. The two operations are performed

    by the Nataf transform and by the classical mapping using the

    eigenvectors of the auto-correlation matrix, respectively. The

    first transformation is summarized in Eq.(6);the eigenvectors

    mapping in Eq.(7).The only selectable option is in the order of

    the operations.

    (1) Eigenvector mapping first and then Nataf: the second

    transformation destroys the expected result of the first

    operation; as a consequence, one has three non-Gaussian

    cross-correlated fields made of uncorrelated entries.

    Since iterations are often introduced when dealing with

    non-linear transformations, the auto-correlation structure

    is altered to check for the consequences of a stronger

    correlation. Once again the Nataf transformation results

    into internally uncorrelated variables, thus showing that the

    iteration path cannot be pursued within this framework.

    (2) Nataf first and then the eigenvectors mapping. Fig. A.1

    shows the statistic elaborations over a sample of size

    1000, when only the Nataf transformation is applied (a)and after the Natafeigenvector sequence (b). The second

    transformation reaches its expected goal, whereas the

    marginal distributions pursued by the Nataf transformations

    are just slightly altered. This alteration is minimal if

    the starting point is made of internally independent

    realizations. Fig. A.2 shows the influence of the starting

    auto-correlation on the statistics of the realizations. It is

    evident that the sequence eigenvectorsNatafeigenvectors

    would produce less accurate results than the simplest path

    Natafeigenvectors do.

    References

    [1] Grigoriu M. Applied non-Gaussian processes. Englewood Cliffs (NJ,

    USA): Prentice Hall; 1995.

    [2] Most T, Bucher C. Stochastic simulation of cracking in concrete structures

    using multiparameter random fields. Reliability and Safety 2006;1(12):

    16886.

    [3] Yamazaki F, Shinozuka M. Digital simulation of non-Gaussian stochastic

    fields. Journal of Engineering Mechanics 1988;114(7):118397.

    [4] Popescu R, Deodatis G, Prevost JH. Simulation of homogeneous non-

    Gaussian stochastic vector fields. Probabilistic Engineering Mechanics

    1998;13(1):113.

    [5] Carpinteri A. Post-peak and post-bifurcation analysis of cohesive crack

    propagation. Engineering Fracture Mechanics 1989;32:26578.

    [6] Babuska I, Melenk JM. The partition of unity method. International

    Journal for Numerical Methods in Engineering 1997;40:72758.[7] Farrell DE, Park HS, Liu WK. Implementation aspects of the

    bridging scale method and application to intersonic crack propagation.

    International Journal for Numerical Methods in Engineering 2007;71(5):

    583605.

    [8] Brezzi F, Fortin M. Mixed and hybrid finite element methods. New York:

    Springer-Verlag; 1991.

    [9] Bruggi M, Venini P. A truly-mixed approach for cohesive-crack

    propagation in functionally graded materials. Mechanics of Advanced

    Materials and Structures 2007. doi:10.1080/15376490701672849.

    [10] Johnson C, Mercier B. Some equilibrium finite elements methods for

    two dimensional elasticity problems. Numerische Mathematik 1978;30:

    10316.

    [11] Bazant ZP. Probability distribution of energetic-statistical size effect

    in quasibrittle fracture. Probabilistic Engineering Mechanics 2004;19:30719.

    [12] Haldar A, Mahadaven S. Reliability assessment using stochastic finite

    element analysis. New York (USA): John Wiley; 2000.

    [13] Mariano PM. Multifield theories in mechanics of solids. Advanced

    Applied Mechanics 2001;38:193.

    [14] Mariano PM, Gioffre M, Stazi FL, Augusti G. Elastic microcracked

    bodies with random properties. Probabilistic Engineering Mechanics

    2004;19:12743.

    [15] Yang JN. On the normality and accuracy of simulated random processes.

    Journal of Sound and Vibration 1973;26(3):41728.

    [16] Vanmarcke E. Random fields. Cambridge (MA, USA): The MIT Press;

    1983.

    [17] Casciati F, Faravelli L. Fragility analysis of complex structural systems.

    Taunton (UK): Research Studies Press; 1991.

    http://dx.doi.org/doi:10.1080/15376490701672849http://dx.doi.org/doi:10.1080/15376490701672849http://dx.doi.org/doi:10.1080/15376490701672849
  • 8/13/2019 Cohesive Crack Propagation in a Random Elastic Medium

    13/13

    M. Bruggi et al. / Probabilistic Engineering Mechanics 23 (2008) 2335 35

    [18] Chamis CC. Probabilistic structural analysis methods for spacepropulsion

    system components. Probabilistic Engineering Mechanics 1987;2(2):

    10010.

    [19] Faravelli L, Bigi D. Stochastic finite elements for crash problems.

    Structural Safety 1989;7(34):11330.

    [20] Casciati F, Faravelli L. Randomness effects in crashworthiness analysis.

    Journal of Nonlinear Mechanics 1991;82734.

    [21] Ghanem RG, Spanos PD. Stochastic finite elements: A spectral approach.New York (USA): Springer-Verlag; 1991.

    [22] Di Paola M. Digital simulation of wind field velocity. Journal of Wind

    Engineering & Industrial Aerodynamics 1996;7476:91109.

    [23] Carassale L, Solari G. Monte Carlo simulation of wind velocity fields on

    complex structures. In: Augusti G, Schueller GI, Ciampoli M, editors.

    Proceedings ICOSSAR 2005. Rotterdam (The Netherlands): Millpress;

    2005.

    [24] Ditlevsen O. Extremes of random fields over arbitrary domains with

    application to concrete rupture stresses. Probabilistic Engineering

    Mechanics 2004;19:37384.

    [25] Ditlevsen O. Dimension reduction and discretization in stochastic

    problems by regression method. In: Casciati F, Roberts JB, editors.

    Mathematical models for structural reliability analysis. Boca Raton (FL,

    USA): CRC Press; 1996.

    [26] Liu PL, Der Kiureghian A. Multivariate distribution models with pre-

    scribed marginals and covariances. Probabilistic Engineering Mechanics

    1989;1:10512.[27] Carpinteri A, DiTommaso A, Fanelli M. Influence of material parameters

    and geometry on cohesive crack propagation. In: Wittmann F, editor.

    Fracture toughness and fracture energy of concrete proceedings of the

    international conference on fracture mechanics of concrete. Amsterdam:

    Elsevier; 1986.

    [28] Shinozuka M. Stochastic fields and their digital simulation.

    In: Schueller GI, Shinozuka M, editors. Stochastic methods in structural

    dynamics. Dordrecht (The Netherlands): Martinus Nijhoff Publishers;

    1987. p. 93133.