Nab Ova Ti 09

Embed Size (px)

Citation preview

  • 8/6/2019 Nab Ova Ti 09

    1/10

    A general model for the permeability of fibrous porous media based on fluid

    flow simulations using the lattice Boltzmann method

    Aydin Nabovati a,*, Edward W. Llewellin b, Antonio C.M. Sousa a,c

    a Department of Mechanical Engineering, University of New Brunswick, Fredericton, NB, Canadab Department of Earth Sciences, Durham University, Durham, UKc Department of Mechanical Engineering, University of Aveiro, Aveiro, Portugal

    a r t i c l e i n f o

    Article history:

    Received 22 September 2008

    Received in revised form 24 February 2009

    Accepted 12 April 2009

    Keywords:

    A. Fibres

    B. Mechanical properties

    C. Computational modelling

    E. Resin transfer moulding (RTM)

    a b s t r a c t

    Fluid flow analyses for porous media are of great importance in a wide range of industrial applications

    including, but not limited to, resin transfer moulding, filter analysis, transport of underground water

    and pollutants, and hydrocarbon recovery. Permeability is perhaps the most important property that

    characterizes porous media; however, its determination for different types of porous media is challenging

    due its complex dependence on the pore-level structure of the media. In the present work, fluid flow in

    three-dimensional random fibrous media is simulated using the lattice Boltzmann method. We deter-

    mine the permeability of the medium using the Darcy law across a wide range of void fractions

    (0.086 /6 0.99) and find that the values for the permeability that we obtain are consistent with avail-

    able experimental data. We use our numerical data to develop a semi-empirical constitutive model for

    the permeability of fibrous media as a function of their porosity and of the fibre diameter. The model,

    which is underpinned by the theoretical analysis of flow through cylinder arrays presented by [Gebart

    BR. Permeability of unidirectional reinforcements for RTM. J Compos Mater 1992; 26(8): 110033], gives

    an excellent fit to these data across the range of/. We perform further simulations to determine the

    impact of the curvature and aspect ratio of the fibres on the permeability. We find that curvature has

    a negligible effect, and that aspect ratio is only important for fibres with aspect ratio smaller than 6:1,

    in which case the permeability increases with increasing aspect ratio. Finally, we calculate the permeabil-

    ity tensor for the fibrous media studied and confirm numerically that, for an isotropic medium, the per-

    meability tensor reduces to a scalar value.

    2009 Elsevier Ltd. All rights reserved.

    1. Introduction

    Permeability prediction, and more generally, the investigation

    of the effect of pore structure on the bulk properties of porous

    media, has posed a major challenge to researchers and engineers

    in a wide range of industrial and academic disciplines. These in-

    clude, but are not limited to, resin transfer moulding [1,2], biomed-

    ical engineering [35], subsurface flow of oil and groundwater[6,7], filter simulation [8,9] and fuel cell simulations [1012].

    Macroscopic approaches for fluid flow simulation in porous

    media, either using the Darcy law [13] or more complicated mod-

    els [14], require as an input the permeability; however, the analy-

    sis of the effect of pore-scale parameters on the macroscopic bulk

    properties is a cumbersome task. The pore structure in porous

    media is often complex, and complicated flow patterns exist within

    the pores and between the grains. Consequently, permeability is

    found to be highly medium-specific, hence there is no general

    model for permeability as a function of the bulk properties of a

    medium. The determination of permeability for a specific material

    typically requires time-consuming experimental work.

    Most experimental methods of permeability prediction apply a

    constant pressure gradient to the porous medium and determine

    the average flow velocity from the measured fluid flow-rate. The

    mediums permeability is subsequently determined using the

    Darcy law [13], as follows:

    hui K

    l:rp 1

    where hui, K, rp and l are the volume-averaged flow velocity, per-meability tensor, pressure gradient vector and the dynamic viscos-

    ity of the fluid, respectively. This relationship is valid in the

    creeping-flow regime (Reynolds number ( 1).Earlier studies of porous media flow were conducted experi-

    mentally and some of the best-known models for the permeability

    of porous media are based on experimental data [1315]. In these

    studies, primarily due to the macroscopic nature of the experimen-

    tal approach, the details of the pore-scale flow-pattern in the

    1359-835X/$ - see front matter 2009 Elsevier Ltd. All rights reserved.doi:10.1016/j.compositesa.2009.04.009

    * Corresponding author. Tel.: +1 506 452 6128; fax: +1 506 453 5025.

    E-mail address: [email protected] (A. Nabovati).

    Composites: Part A 40 (2009) 860869

    Contents lists available at ScienceDirect

    Composites: Part A

    j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / c o m p o s i t e s a

    mailto:[email protected]://www.sciencedirect.com/science/journal/1359835Xhttp://www.elsevier.com/locate/compositesahttp://www.elsevier.com/locate/compositesahttp://www.sciencedirect.com/science/journal/1359835Xmailto:[email protected]
  • 8/6/2019 Nab Ova Ti 09

    2/10

    porous medium cannot be captured. In general, studying the effect

    of pore-scale parameters on bulk properties requires a large exper-

    imental dataset, which is time-consuming and expensive to

    generate.

    As alternative approaches, analytical and numerical methods

    aim to predict the permeability by solving the fluid flow inside

    the pores of the porous medium. The numerical procedure echoes

    the experimental approach: a pressure gradient is applied; the

    fluid flow inside the pores of the medium is solved; the permeabil-

    ity of the mediumis then determined using the Darcy law (Eq. (1)).

    This approach has two main advantages over the experimental ap-

    proach. First, the geometry of the digitally constructed medium

    can be varied rapidly and arbitrarily. Second, simulated experi-

    ments are typically much quicker to run than their laboratory

    counterparts. Together, these benefits allow a more-rapid and

    more-thorough exploration of parameter space than can be

    achieved in the laboratory.

    The very first challenge in the numerical simulation of permeat-

    ing flow is that the pore-level structure of the medium is required

    as input. Imaging techniques, such as computed tomography

    [16,17], have been developed to capture the complex structure of

    real porous media; however, these methods may be costly and

    time consuming and impose limits on the resolution that can be

    achieved. The alternative approach, which has been widely used

    in the literature, is to reconstruct the pore-level structure virtually.

    The level of the reconstructed structures complexity depends on

    the computational resources available and the nature of the prob-

    lem under study. The reconstructed medium can be in the form of

    ordered or random arrangements of two and three-dimensional

    obstacles.

    Analytical studies of the pore-level flow, in general, employ the

    Stokes equation (a simplified form of the NavierStokes equation,

    which is valid for creeping flow) for a specified domain with peri-

    odic boundary conditions. Due to the limitations of the methods

    applied, the computational domain, which is the building block

    of the pore structure, is in the form of a simplified, well-defined

    structure in which the grains of the porous medium are repre-sented in the form of two dimensional obstacles [18,19], ordered

    sphere packing [20], or ordered packing of cylinders [2124].

    Rapid increase in available computing power and the develop-

    ment of advanced numerical algorithms mean that detailed

    numerical simulations of flow in porous media are now feasible.

    Removing the constraints of the analytical approaches, more com-

    plex pore geometries, which resemble the real porous-media struc-

    tures more closely, can be used in fluid flow simulations. Ordered

    or random packing of different geometric configurations, such as

    square blocks, spheres, cylinders, and parallelepipeds [25] have

    been used in the literature to reconstruct the pore structure. The

    choice of the constructing elements depends on the nature and

    application of the porous medium to be modeled. Random arrange-

    ments of spheres with mono-dispersed, bi-dispersed, or distrib-uted diameter are often used in simulations of flow in geological

    materials, including studies of groundwater flow and hydrocarbon

    recovery, and flow in packed beds, and rocks [6,7,2628]. Simula-

    tions of flow in the preform matrix in resin transfer moulding

    (RTM) [24,29,30], and through paper fibres [31,32] and woven

    materials [9,3336] often use random arrangements of fibres.

    1.1. Porositypermeability relationships for fibrous materials

    The key parameter controlling the permeability of fibrous mate-

    rials and, indeed, all porous materials is the porosity / = Vpore/

    Vsample, where Vsample is the total volume of the sample and Vpore is

    the volume not occupied by solid fibres. Several workers have pub-

    lished relationships for the permeability of fibrous materials as afunction of their porosity.

    Gebart [24] presents a combined theoretical, numerical and

    experimental investigation of the permeability of ordered arrays

    of fibres. The analytical treatment of creeping flow perpendicular

    to the long axis of the fibres is predicated upon the assumptions

    that permeability is controlled by the narrow slots formed between

    the fibres at their closest approach and that the widthof these slots

    varies only slowly. These assumptions are most valid in the limit of

    close-packed fibres. Gebart derives the following functional form

    for K(/):

    K

    a2 C

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffi1 /c1 /

    s 1

    !5=22

    where a is the fibre radius, /c is the critical value of porosity below

    which there is no permeating flow (the percolation threshold) and C

    is a geometric factor (Gebart calculates C 16=9ffiffiffi

    2p

    p and/c = 1 p/4 for a square array, C 16=9

    ffiffiffi

    6p

    p and /c 1 p=2ffiffiffi

    3p

    for a hexagonal array). Gebart presents numerical results, obtained

    using a finite difference solution of the NavierStokes equations,

    that show excellent agreement with the relationship up to at least

    / = 0.65.

    Koponen et al. [31] used the LBM on a D3Q19 lattice to studycreeping flow through three-dimensional random-fibre sheets,

    analogous to paper and non-woven fabrics. They report that the

    permeability of such materials is exponentially dependent on the

    porosity and independent of whether the fibres were placed ran-

    domly or not. They present an empirical relationship for the per-

    meability as a function of the porosity, based on a fit to their

    data, for porosities in the range 0.4 < / < 0.95:

    K

    a2

    5:55

    e10:11/ 13

    Clague et al. [37] also studied the permeability of three-dimensional

    ordered and disordered fibrous media. They used the lattice Boltz-

    mann method (LBM) on a D3Q15 and D3Q19 lattice to simulate

    creeping flow through fully three-dimensional random fibre net-

    works, in which free overlapping of the fibres was allowed. Both

    wall-bounded and unbounded media were considered and the ef-

    fect of the wall on the overall permeability of the fibrous media

    was investigated. They use a scaling analysis to develop a phenom-

    enological relationship between permeability and porosity for both

    the bounded and unbounded fibrous media. For the case of an un-

    bounded medium, they find:

    K

    a2 b1

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffi1 /c1 /

    s 1

    !2eb2 1/ 4

    where b1 and b2 are curve fitting parameters. For a disordered (ran-

    dom) arrangement of fibres, Clague et al. calculate a typical value of

    /c = 1p/4 % 0.21 but with a 1r variation that gives minimum andmaximum values of /c = 0 and /c 2 ffiffiffiffipp p % 0:4, respectively.They show an excellent fit to their data, which span the range

    0.33 < / < 0.95 (i.e. they do not cover the region near the percola-

    tion threshold).

    In the present work, fluid flow is simulated in 3D random fi-

    brous media at the pore level. We employ the lattice Boltzmann

    method for the fluid flow simulation and calculate the permeabil-

    ity of the medium using the Darcy law. We cover a wide range of

    porosity, from near the percolation threshold to very dilute sys-

    tems (0:08 / 0:99). Based on curve fitting of our numericalexperimental results, we propose a semi-empirical constitutive

    relationship for the permeability as a function of porosity. We also

    investigate the effect of various other pore-level parameters,

    including the curvature, diameter and aspect ratio of the fibres,

    on the predicted permeability.

    A. Nabovati et al./ Composites: Part A 40 (2009) 860869 861

  • 8/6/2019 Nab Ova Ti 09

    3/10

    2. Methodology

    The complex structure of the pore-level geometry, especially in

    media of low porosity, yields small pores and narrow flow pas-

    sages. As a rule of thumb, the local flow velocity in these narrow

    pores is proportional to the volume-averaged flow velocity divided

    by the porosity. The narrow flow passages and locally high veloci-

    ties limit the applicability of the conventional computational fluiddynamics approaches. Mesoscopic methods such as the smoothed

    particle hydrodynamics (SPH) [38,39], lattice gas automata (LGA)

    [40,41] and lattice Boltzmann method (LBM) [4244] have been

    successfully used for macroscopic fluid flow simulations, which re-

    quire the mesoscopic details of the flow to be considered. The LGA

    and SPH methods, in their current state-of-the-art, tend to be com-

    putationally costly to perform three-dimensional simulations of

    flow in porous media that are of a size adequate to yield physically

    meaningful results. Although the early versions of LBM [45] suf-

    fered from similar difficulties, later developments of the LBM have

    seen dramatic improvements in the computational efficiency, mak-

    ing it a suitable tool for mesoscopic, three-dimensional simulations

    [42,46,47]. Other attractive features of the LBM are that numerical

    operations are spatially local, easing implementation, while solid

    boundaries of arbitrary complexity can be included without perfor-

    mance penalty. Furthermore, the LBM is well suited to flow simu-

    lation at the mesoscopic scale, and is amenable to parallelization.

    These characteristics have made the LBM the most popular method

    for numerical pore-level analysis; indeed this was one of the first

    applications of LBM [48]. The LBM has been shown to be a more

    efficient tool for flow simulation in such complex geometries than

    conventional fluid dynamic approaches [49,50].

    In the present work, three-dimensional fluid flow was simu-

    lated in fibrous porous media using LBflow,1 an implementation

    of the single-relaxation-time (SRT) LBM on a D3Q15 lattice [51,52].

    As in all such implementations, the flow is represented by the prop-

    agation of fluid mass through a lattice. The lattice is a discrete

    representation of physical space; in the present case a three-dimen-

    sional cubic lattice. At any time t, fluid mass can propagate from anode with position r to any of its six orthogonal neighbours or eight

    long-diagonal neighbours, or it can remain at the present node. Since

    time is also discrete, the propagating fluid arrives at its new location

    at time t+ 1, hence, there are i = 15 possible fluid velocities at each

    node, represented by the vector ei. The spatial discretization Dx

    and the time step Dt define the units of the simulation. The total

    density at each node is given by:

    qr; t X

    i

    fir; t 5

    where q is also in simulation units (typically initialized to q = 1throughout the lattice). Similarly, the average fluid velocity at each

    node is given by:

    ur; t P

    ifieir; t

    qr; t6

    In addition to the propagation step, at each node r, at each time step

    t, the incoming fluid masses fi undergo collision, in which they re-

    lax towards the equilibrium distribution:

    feqi qxi 1 ei u

    c2su u

    2c2s

    ei u2

    2c4s

    " #7

    where the weightsx0 = 2/9,x1. . . 6 = 2/9 and x7. . .14 = 1/72, and thelattice pseudo-sound speed cs Dx=

    ffiffiffi

    3p

    Dt. This equilibrium distri-

    bution is a discrete analogue of the Maxwell-Boltzmann distribu-

    tion for a population of fluid particles having the same density q

    and average velocity u as the incoming fluid masses in the simula-

    tion. The propagation and collision steps are encapsulated in the

    lattice Boltzmann equation:

    fir eiDt; t Dt

    |fflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}propagation

    fir; t fir; t f

    eqi r; t

    s

    |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}collision

    8

    where s is a relaxation parameter, related to the fluid viscosity. Inthe implementation of the LBM adopted in the present work, flowis driven by imposing a constant, uniform body force G on the fluid

    at every point, which is physically analogous to a gravitational force

    acting on the fluid. This is achieved by adding an extra term gi to

    each of the mass components fi prior to propagation. The term giis formulated in such a way that the total mass is conserved but

    the momentum is adjusted to account for the force acting on the

    mass at the node for the duration of the time step:

    gir; t xiDtc2s

    G ei 9

    This term is added to the right hand side of Eq. (8). The halfway

    bounceback method [44] was used to implement the solid wall

    boundary condition.

    LBflow uses a scripting language to set up the flow simulation;

    via an interpreted text file, the user can specify the geometry and

    parameters of the simulation. In this study, the geometries of the

    porous media of interest are specified as a three-dimensional bin-

    ary mask of the simulation lattice. Dimensional quantities for the

    simulation are specified in SI units, rather than simulation units.

    In this study, the working fluid has the properties of water at

    20 C: kinematic viscosity m = 1.004 106 m2 s1, densityq = 998.29kg m3. The driving pressure gradient rp is specifiedin units of Pa m1. Once the flow has settled to steady state, theaverage velocity of the fluid nodes u is output in SI units. The vol-

    ume-averaged fluid velocity hui /u is determined allowing thepermeability of the medium to be calculated using the Darcy law,

    Eq. (1). Note that the choice of the working fluid properties is

    somewhat arbitrary, since the permeability is independent of vis-

    cosity, density and driving pressure gradient for creeping flow

    (i.e. Reynolds number Re ( 1). This was confirmed numericallyby repeating a typical simulation multiple times for a range of

    rp spanning several orders of magnitude, all at low Reynolds num-ber; the calculated permeability was indeed independent of pres-

    sure gradient.

    The fluid viscosity is related to the relaxation parameters of theSRT LBM:

    m s1

    2

    c2sDt 10

    The relaxation parameter represents the degree to which the fluid

    populations are relaxed towards the equilibrium value during thecollision step (Eq. (8)). Consequently, the larger the value of the

    relaxation parameter, the more rapid is the flow settlement. How-

    ever, we should note the use of the halfway bounceback method

    for the solid wall boundary condition with the SRT LBM, yields pre-

    dictions of the flow field in porous media, and therefore of their per-

    meability, that are dependent on the choice of the relaxation

    parameter. Consequently, the accuracy of the simulation is depen-

    dent on the relaxation parameter. Pan et al. [53] evaluated different

    LB methods and solid wall boundary condition treatment methods

    for flow simulations in porous media. They showed that results ob-

    tained using the SRT LBM with the halfway bounceback method for

    the solid boundaries are in good agreement with the results of the

    MRT LBM with interpolated bounceback boundaries, and that the

    effect of the relaxation parameter dependency is negligible for

    s = 1. We set the relaxation parameter equal to 1 in all the1 Available from http://www.lbflow.co.uk.

    862 A. Nabovati et al./ Composites: Part A 40 (2009) 860869

    http://www.lbflow.co.uk/http://www.lbflow.co.uk/
  • 8/6/2019 Nab Ova Ti 09

    4/10

    simulations performed. This value represents the best compromise

    between simulation accuracy and the rate of flow settlement.

    3. Results and discussion

    3.1. Validation

    To validate our methodology, we simulate creeping fluid flow in

    a hexagonal array of infinite, parallel cylinders. The radius of the

    cylinders was kept constant and equal to 40 lattice units in a do-

    main of l ffiffiffi

    3p

    l 1, where l was varied between 82 and 230 latticeunits; this yields a medium of porosity between 0.15 and 0.89.

    Fig. 1 depicts the schematic of the computational domain unit cell

    with periodic boundary conditions on both sides; the gray colour

    identifies the simulation domain. Due to the invariance of the do-

    main geometry under translation in thez-direction, a fully 3D sim-

    ulation can be achieved with a domain 1 lattice unit thick in that

    direction by using periodic boundary conditions.

    The results of these simulations are presented in Fig. 2, which

    compares the permeability, calculated according to the methodol-

    ogy presented in Section 2, with the value determined analytically

    using the relationship for this geometry presented in [24] (Eq. (2)).

    This model, described in Section 1.1, has been widely used for flow

    in regular arrays of fibres [54,55]. The agreement over the range of

    porosities for which that relationship is valid, i.e. 0.10 < / < 0.65, is

    excellent. It is noteworthy that the fit between the data and the

    analytical solution is good even beyond the upper limit of validity

    claimed by Gebart (dotted line in Fig. 2).

    3.2. Flow in fibrous media

    To create the fibrous medium structure, randomly oriented,

    straight, cylindrical fibres of constant diameter are randomly

    placed in a cubic domain with free overlapping. By allowing the fi-

    bres to overlap freely, we are able to investigate flow in fibrous

    media with porosities across the full range, right down to the per-

    colation threshold. Media in which the fibres are not able to over-

    lap have a minimum porosity, which is higher than the percolation

    threshold. Koponen et al. [31] studied the fluid flow in three

    dimensional fibre webs, where flexible fibres were placed ran-

    domly in the computational domain without overlapping, the min-

    imum reported porosity was higher than 0.4. Nabovati and Sousa

    [56] investigated the permeability of sphere packs with and with-

    out free overlapping. They found that overlapping has a negligible

    impact on permeability for media with porosities higher than 0.85

    and leads to a decrease of less than 35% in permeability for low

    porosity media. The minimum reported porosity for random pack-

    ing of spheres without overlapping was 0.55.

    The computational domain comprises a cube of 128 lattice sites

    on each side and the periodic boundary condition was applied to

    all faces of the cube. The fibres extend to the boundaries of the do-

    main. The fibres are placed in the computational domain by using

    the following algorithm: (1) a random position vector is chosen

    within the computational domain, or one fibre radius of it, as the

    origin of the fibres core, (2) a random vector representing the spa-

    tial orientation of the fibre core is generated, (3) the fibre core line

    is extended from the origin point along the randomly determined

    orientation in both directions until it intersects the domain bound-

    aries, and (4) lattice sites that are closer to the core line than the

    radius of the fibre are designated as solid. The radius of the fibres

    is constant along their length and equal to 2, 3, 4, 5 or 6 lattice

    units, Dx, depending on the experimental run. Fig. 3 shows a high

    porosity sample of the reconstructed fibrous medium; the porosity

    of the sample is 0.80.

    Fig. 1. Schematic of two dimensional section through the computational unit cell

    used for the permeability prediction in hexagonal arrangement of cylinders; flow

    domain with periodic boundary conditions on both sides is shown in gray colour.

    0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.910

    -6

    10-4

    10-2

    100

    102

    Porosity ()

    Present Work

    Gebart [24]

    K/a2

    Fig. 2. Normalized permeability, calculated using the methodology presented in Section 2, is plotted with the analytically-determined relationship (Eq. (2)) presented byGebart in [24] for a hexagonal arrangement of solid cylinders. Gebarts relationship is shown dotted outside the range of values of / for which Gebart claims validity.

    Fig. 3. Reconstructed medium with straight fibres and porosity/ = 0.80; the radius

    of the fibres is four lattice units and their length is such that they span the

    computational domain.

    A. Nabovati et al./ Composites: Part A 40 (2009) 860869 863

  • 8/6/2019 Nab Ova Ti 09

    5/10

    A pressure gradient is applied in thex-, y- orz-direction and the

    velocity field is calculated in the pores of the reconstructed med-

    ium. Fig. 4 shows the velocity vectors in a slice of a three-dimen-

    sional fibrous medium which has a porosity equal to 0.2; the

    pressure gradient, and the mean flow direction, are in the positive

    x-direction. We should note that, whilst some pores appear to be

    dead-end or isolated in the two-dimensional slice, in the third

    dimension, these pores are effectively connected to each other

    and they contribute to the fluid flow pattern. As a result, the per-

    colation threshold reported to be equal to 0.33 for two-dimen-

    sional media made up of random arrangements of squareobstacles [57], in fact, takes a lower value for three-dimensional

    media. In the present study, permeating pathways for fluid flow

    exist for porosity as low as 0.08.

    3.2.1. Effect of fibre radius

    Permeability has dimensions of length squared, and it is usual

    to normalize permeability by the square of a length scale that is

    characteristic of the system; for fibrous media, this is typically

    the fibre radius. To validate this approach, we determine the per-

    meability of random networks of fibres with radii of 2, 3, 4, 5

    and 6 lattice units. Results are shown in Fig. 5. Fig 5 demonstrates:

    (1) permeability increases as porosity increases for constant fibre

    radius; (2) permeability increases as fibre radius increases for con-

    stant porosity. Fig. 5b demonstrates that normalizing the perme-ability by the square of the fibre radius causes the curves for

    different fibre radius to collapse onto a single curve, hence, that

    this non-dimensionalization is appropriate. Note that the normal-

    ized permeability is not completely independent of fibre radius;

    the two show an inverse relationship with the normalized perme-

    ability for the thinnest fibre typically about 20% higher than for the

    fattest on average. We postulate that this dependence is due to the

    finite length of the fibres in our simulations giving rise to a variable

    aspect ratio as the fibre radius is increased. This is supported by the

    results presented later in Section 3.2.4.

    3.2.2. Constitutive permeabilityporosity relationship for random fibre

    networks

    We determine the permeability of around 50 random fibre net-works in the x-, y- and z-direction (giving $150 data points in to-

    tal). For each network, the fibres all have the same radius: either 2

    or 4 lattice nodes. The permeability of the networks is in the range

    0.08 < / < 0.99. Results are plotted in Fig. 6 (note the logarithmic

    scale on the permeability axis). The normalized permeability is

    approximately exponentially dependent on porosity in the range

    0.2 < / < 0.9. At lower and higher porosities, the dependence is

    stronger. As expected, the permeability tends towards infinity in

    the limit /? 1 and drops towards zero at low (but finite) porosity.

    We note that the variation in the data is greatest at very low poros-

    ities, where random differences in the placement of the fibres be-

    tween different domains lead to large relative changes in thepredicted permeability.

    We find that a modified version of the Gebart [24] relationship

    provides an excellent fit to data across the full range of porosity

    (Fig. 6). We adapt Gebarts original relationship (Eq. (2)) by allow-

    ing the three constants it contains to vary:

    K

    a2 C1

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffi1 /c1 /

    s 1

    !C211

    where /c is the critical value of porosity above which permeating

    flow can occur (the percolation threshold). C1 and C2 are related

    to the geometry of the network (compare with the values deter-

    mined by Gebart for a regular array of fibres, presented in Section

    1.1). We use the freely-available statistical analysis package SimFit2

    to fit Eq. (11) to the data presented in Fig. 6. The fit is carried out in

    logarithmic space to avoid biasing the fit towards large values of per-

    meability at high porosity. Table 1 shows the best fit values obtained

    (R2 = 0.999).

    The values are consistent with those determined analytically by

    Gebart for a hexagonal array of aligned fibres (Section 1.1); differ-

    ences are due to the different geometries of the networks. The va-

    lue of/c falls within the range calculated, on statistical grounds, by

    Clague et al. [37], for a random network of fibres (Section 1.1).

    The published relationships of Koponen et al. [31] and Clague

    et al. [37] Eqs. (3) and (4), respectively in Section 1.1 were sim-

    ilarly adapted and fitted to the data in Fig. 6. The Koponen relation-

    ship provided a very poor best fit to data, and was abandoned. The

    Fig. 4. Velocity vectors in a slice of the three-dimensional fibrous medium; the porosity of the sample is 0.2 and the pressure gradient drives flow in the positive x direction.

    2 Available from: http://www.simfit.man.ac.uk.

    864 A. Nabovati et al./ Composites: Part A 40 (2009) 860869

    http://www.simfit.man.ac.uk/http://www.simfit.man.ac.uk/http://www.simfit.man.ac.uk/
  • 8/6/2019 Nab Ova Ti 09

    6/10

    best fit of the Clague relationship was as good as the Gebart fit. We

    favour the Gebart relationship because it has a sound theoretical

    basis, whereas the Clague relationship is purely phenomenological.

    To assess the variation of the predicted permeability values due

    to the random nature of the reconstructed fibrous porous media,

    we performed multiple repeated simulations for media at three

    different porosities: $0.1, $0.5 and $0.95, respectively. Due tothe discrete nature of the fibre placement procedure, it is not pos-

    sible to create different random media with exactly the same

    porosity; hence, there is a variation of around 0.5% for each pre-

    scribed porosity. For each porosity, we created 21 media with dif-

    ferent random fibre placement and simulated flow in x, y, and z

    directions, resulting in 63 permeability determinations for each

    porosity. Each dataset is separately plotted against porosity inFig. 7; the permeability predicted using Eq. (11), with the parame-

    ter values presented in Table 1, is shown on each plot as a solid

    line. Table 2 shows the mean permeability for each dataset and

    the permeability normalized by the value predicted using Eq.

    (11). The standard deviation is also presented, expressed as a per-

    centage of the mean permeability.

    From Fig. 7 and Table 2, it can be seen that both the extent of

    the variation in permeability and the quality of fit of Eq. (11) de-

    pend on the porosity. The standard deviation is small (around

    10% of the mean) for mid and high-range porosity; at very lowporosity, the standard deviation is larger (around 50% for /

    % 0.1), reflecting the large changes in permeability that arise fromsmall structural differences near the percolation threshold. Eq. (11)

    has a tendency to slightly under-predict permeability at lowporos-

    ity and to over-predict at high porosity. Given the$6 order of mag-nitude difference in permeability between / % 0.1 and / % 0.95,however, we consider Eq. (11) to provide an accurate and flexible

    tool for permeability prediction across the porosity range and note

    that the model value is within one standard deviation of the

    numerical data across the porosity range.

    The literature contains permeabilityporosity data from a num-

    ber of laboratory investigations into fibrous media; Jackson and

    James [23] provide a summary. In Fig. 8, we compare our perme-

    abilityporosity relationship (Eq. (11)) with experimental datafor high porosity fibrous media with randomly oriented straight

    0.2 0.4 0.6 0.8 110

    -10

    10-9

    10

    -8

    10-7

    10-6

    10-5

    Porosity ()

    K

    (m2)

    a = 2 lu

    3

    4

    5

    6

    0.2 0.4 0.6 0.8 110

    -3

    10-2

    10-1

    100

    101

    Porosity ()

    K/a2

    (b)(a)

    Fig. 5. (a) Numerically-determined permeability as a function of theporosity for fibres with a range of radii. Within each simulation suite, all fibres have thesame radius; (b)

    numerically-determined permeability normalized by fibre radius as a function of the porosity.

    0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 110

    -6

    10-4

    10-2

    100

    102

    K/a2

    Numerically Predicted Values

    Proposed Relation, Eq. 11

    Porosity ()

    Fig. 6. Dimensionless permeabilityas a function of porosity for randomnetworks of straight fibres. Fibre radius is 2 or 4 lattice units in allcases. Thesolid line shows thefit of

    our semi-empirical relationship (Eq. (11)) which is adapted from Gebart [24]. Best fit parameters are given in Table 1. The fit (R2 = 0.999) indicates a percolation threshold of

    /c% 7.4%.

    Table 1

    Best fit parameters of Eq. (11) to data presented in Fig. 6; regression coefficient

    R2 = 0.999. See main text for details of fitting procedure.

    Parameter Best fit value

    /c 0.0743

    C1 0.491

    C2 2.31

    A. Nabovati et al./ Composites: Part A 40 (2009) 860869 865

  • 8/6/2019 Nab Ova Ti 09

    7/10

    cylindrical fibres of constant diameter presented in that work. The

    data were obtained using a range of experimental methods, a

    broad spectrum of fibrous materials including filter pads, nylon fi-

    bres, Kapron fibres, collagen, metal fibres and polymer fibres, and

    various working fluids including water, glycerol and air. Despite

    the diversity of these investigations, there is a broad agreement

    (within an order of magnitude) among the resulting datasets. Over-

    all the agreement between Eq. (11) and the experimental data is

    good, especially for porosities higher than approximately 0.75.

    3.2.3. Effect of fibre curvature

    We investigate the effect of fibre curvature on the permeability

    of fibrous media, by replacing straight fibres with randomly curved

    fibres as the constituting elements of the medium. We generate the

    fibres by constructing a cylinder of constant diameter around a

    randomly curved fibre core. The origin of the fibres is chosen ran-

    domly. Three different third-order polynomials represent the three

    coordinates defining the orientation of the fibre core and they are a

    function of a single variable, t:

    xi ai;3t3 ai;2t

    2 ai;1t ai;0 i 1; 2; 3 12

    where the coefficients of these polynomials are chosen randomly in

    the range of [1,1], this ensures that fibres are smoothly curved and

    avoids tight spirals. The variable tis incremented and decrementedin appropriate steps to extend the fibre in both directions until the

    fibre reaches the domain boundaries, producing a smoothly curved

    fibre, which crosses the domain.

    Fig. 9, shows the permeability we determine from simulations

    of flow in media composed of curved fibres. The results are almost

    indistinguishable from the results for straight fibres; and the effect

    of the fibres curvature on the overall permeability of the medium

    can be considered negligible.

    3.2.4. Effect of fibre aspect ratio

    We investigate the effect of fibre aspect ratio a (length to diam-eter ratio of straight cylindrical fibres of finite length) on perme-

    ability for a range of aspect ratios between 1 and 20, for twodifferent values of porosity, namely 0.51 and 0.73. The fibres diam-

    4.0

    5.0

    6.0

    7.0

    8.0

    9.0

    0.946 0.947 0.948 0.949 0.950

    Porosity ( )

    K/a2

    0.030

    0.035

    0.040

    0.045

    0.050

    0.055

    0.060

    0.497 0.498 0.499 0.500

    Porosity ( )

    K/a2

    (a) (b)

    0.0E+00

    2.0E-05

    4.0E-05

    6.0E-05

    8.0E-05

    1.0E-04

    1.2E-04

    1.4E-04

    0.0995 0.0996 0.0997 0.0998 0.0999 0.1

    Porosity ( )

    K/a2

    (c)

    Fig. 7. Dimensionless permeability for three different porosities: (a) $0.95, (b) $0.5, and (c) $0.1. At each porosity, the permeability was determined for 21 different randomnetworks of straight fibres to determine the random variation in permeability. The solid line represents Eq. (11) with the parameters given in Table 1. Note the dramatically

    expanded scales compared with Fig. 6. See Section 3.2.2 for discussion.

    10-3

    10-2

    10-1

    100

    10-4

    10-3

    10-2

    10-1

    100

    101

    102

    103

    Solid Fraction (1-)

    K/a2

    Chen (1995)

    Ingmanson et al. (1959)

    Kirsch & Fuchs(1967)

    Stenzel et al. (1971)

    Kostornov & Shevchuck (1977)

    Jackson and James (1982)

    Proposed relation, Eq. 11

    Fig. 8. Comparison of dimensionless permeability of fibrous media, calculated

    using Eq. (11) with the parameters given in Table 1, with the experimental data

    reported by Jackson and James [23]. Readers are directed to that work for full

    references to the original experimental studies.

    Table 2

    Mean permeability and standard deviation for three different porosities: $0.1, $0.5,

    and $0.95. See Fig. 7.

    Porosity Mean permeability (K/a2) Mean permeability

    (normalized to Eq. (11))

    Standard

    deviation (%)

    $0.1 2.58

    105 2.09 52

    $0.5 4.62 102 1.01 10$0.95 7.50 0.877 11

    866 A. Nabovati et al./ Composites: Part A 40 (2009) 860869

  • 8/6/2019 Nab Ova Ti 09

    8/10

    eter is constant and equal to 6 lattice units and the aspect ratio is

    varied by changing the fibres length. The numerically-determined

    values of the permeability are shown in Fig. 10 as a function of the

    fibre aspect ratio; the permeability values are normalized by the

    permeability calculated using Eq. (11). It can be seen that perme-

    ability increases with aspect ratio for a < 6; for a > 6, permeabilityis largely independent of aspect ratio. This finding explains the

    imperfect collapse of the data for fibres of varying radius, pre-

    sented in Fig. 2, when normalized by the fibre radius. Since the

    length of the fibres is finite limited by intersection with the do-

    main boundaries the aspect ratio of the fibres in that suite of sim-

    ulations varies somewhat with fibre radius.

    For a cylinder of diameter d, the ratio of the surface area to the

    volume of the fibre (specific surface area) is equal to 1d4 1a. An

    increase in aspect ratio therefore yields a decrease of the specific

    surface area of the fibres, leading to a reduction in the frictional

    drag force and higher values for the permeability.

    3.2.5. Permeability tensor

    Any experimental or numerical attempt to determine the per-meability tensor and the principle directions of a porous medium

    should be performed in three dimensions. This is an expensive

    and time consuming task. Neuman [58] proved analytically that

    the permeability tensor is a symmetric second order tensor, which

    has six distinct elements in general. For isotropic materials, the

    three diagonal elements are equal and non-zero and all off-diago-

    nal elements are zero. In this case, the permeability can be repre-

    sented as a single, scalar value.

    Ahn et al. [59], Weitzenbck et al. [60,61] and Parnas et al. [62]

    discuss experimental methods and approaches for permeability

    tensor measurements. Kolodziej et al. [63] analytically investigated

    the permeability tensor of high porosity fibrous porous media. In

    their study, fibres had a unidirectional arrangement with non-uni-

    form spacing. Nedanov and Advani [64] studied fluid flow in two-

    scale fibrous porous media numerically; governing equations of

    the fluid flow around and inside the solid and permeable fibres

    were developed using the homogenization method and were

    solved numerically. They report the diagonal elements of the per-

    meability tensor for single and three-ply fabrics. Song et al. [35]calculated the permeability tensor of a three-dimensional, woven

    fibrous medium using the finite volume method.

    In this study, to evaluate the viability of the LBM for permeabil-

    ity tensor determination, the permeability tensor elements of three

    samples of different porosities of 0.90, 0.70 and 0.50, respectively,

    are determinedbased on the simulated flowfield. For eachporosity,

    a pressure gradient is applied in one of the three principle direc-

    tions. The mean flow is in the direction of the applied pressure gra-

    dient, which builds the diagonal element of the permeability tensor

    in the specified direction. We calculated the off-diagonal elements

    of the permeability tensor using the net flow in each of the other

    two directions, and repeated this procedure for each of the three

    directions to obtain the nine elements of the permeability tensor.

    The predicted permeability tensors for these media are as follows:

    K0:90

    1:0866 0:0073 0:0203

    0:0073 1:0526 0:0138

    0:0206 0:0138 1:0820

    264

    375 107 m2

    K0:70

    1:1643 0:0262 0:0021

    0:0262 1:1726 0:0006

    0:0021 0:0006 1:1632

    264

    375 108 m2

    K0:50

    1:8851 0:0478 0:0395

    0:0478 1:9017 0:0029

    0:0395 0:0029 1:8484

    264

    375 109 m2

    As the media are created with randomly-oriented fibres, there is no

    preferential flow direction and the media are expected to be isotro-

    pic. In each case, we find that the diagonal elements of the perme-

    ability tensor are the same to within $3%; this small difference caneasily be accounted for by the random nature of the network. Fur-

    thermore, the off-diagonal elements are approximately symmetri-

    cal, and are smaller than the diagonal elements by around two

    orders of magnitude. These results support the analysis of Neuman

    [58] and demonstrate that it is valid to report a scalar value for per-

    meability of these random porous media, and a determination of

    the full second order tensor is unnecessary.

    4. Conclusion

    Three-dimensional fluid flow simulations in fibrous media are

    conducted using the SRT LBM; the fibrous media are recon-structed by random placement of cylindrical fibres, with random

    0 5 10 15 200.5

    0.6

    0.7

    0.8

    0.9

    1

    Fiber Aspect Ratio

    K/Kmodel(Eq.

    11)

    Phi = 0.505

    Phi = 0.730

    Fig. 10. Effect of the fibre aspect ratio on the numerically-determined permeability

    for two different porosities, 0.51 and 0.73, respectively; permeability values are

    normalized by the permeability calculated using Eq. (11) with theparameters givenin Table 1.

    0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 110

    -4

    10-2

    100

    102

    Straight Fibres - Equation 11

    Curved Fibres

    K/a2

    Porosity ()

    Fig. 9. Numerically-determined permeabilityof fibrous media as a function of the porosity for media formed of straight (solidline, Eq. (11) with parameters given in Table 1)

    and randomly curved fibres (circles, simulation results).

    A. Nabovati et al./ Composites: Part A 40 (2009) 860869 867

  • 8/6/2019 Nab Ova Ti 09

    9/10

    orientations, within the computational domain. The radius, curva-

    ture and length of the fibres are varied systematically. We find that

    dividing the permeability by the square of the fibre radius yields an

    appropriate non-dimensionalization. We also find that fibre curva-

    ture has a negligible impact on the permeability of the medium.

    For fibres of finite length with aspect ratios smaller than $6,permeability increases with increasing aspect ratio; this effect is

    negligible for values of the aspect ratio greater than 6.

    The permeability values that we obtain are compatible with the

    available experimental data and, based on the determined values,

    we develop a relationship (Eq. (11)) for the permeability as a

    function of the materials porosity and the fibre diameter. The

    relationship that we present results from a semi-empirical

    parameterization of a published analytical relationship for ordered

    arrays of fibres [24]. The fit of the relationship to the data is excel-

    lent across the range of porosities investigated (0.08 < / < 0.99).

    Our analysis shows that the percolation threshold for a three-

    dimensional network of randomly oriented fibres is /c = 0.0743 .

    The media in this study were created randomly, hence there

    was no preferential direction for the flow through the medium;

    therefore, the media can be assumed isotropic. Prediction of the

    permeability tensor for media of three different porosities sup-

    ported this assumption, as the diagonal elements of the numeri-

    cally-determined permeability tensors differ by less than $3%,and the off-diagonal elements were two orders of magnitude smal-

    ler than the average value of the diagonal elements. This finding

    indicates that it is valid to report a scalar value, instead of a second

    order tensor, for the permeability of the fibrous media studied.

    The results obtained in this study and the general relationship

    proposed for the permeability, can be fed to the macroscopic flow

    modelling approaches for the industrial applications, e.g. resin

    transfer moulding process, where the pore level approach is not

    applicable due to the large computational resources requirement.

    Acknowledgment

    The authors acknowledge the support received from NSERC(Natural Sciences and Engineering Research Council of Canada)

    Discovery Grant 12875 (ACMS) and from the Foundation for Sci-

    ence and Technology (FCT, Portugal) through the research grant

    POCTI/EME/59728/2004 (ACMS). EWL is supported by NERC (UK)

    Research Fellowship NE/D009758/2. We thank two anonymous

    reviewers for their helpful comments.

    References

    [1] Phelan Jr FR. Simulation of the injection process in resin transfer molding.

    Polym Compos 1997;18(4):46076.

    [2] Abrate S. Resin flow in fiber preforms. Appl Mech Rev 2002;55(6):57999.

    [3] Khanafer K, Vafai K. The role of porous media in biomedical engineering as

    related to magnetic resonance imaging and drug delivery. Heat Mass Transfer

    2006;42(10):93953.

    [4] Vafai K, Ai L. A coupling model for macromolecule transport in a stenosed

    arterial wall. Int J Heat Mass Transfer 2006;49(910):156891.

    [5] Yang N, Vafai K. Modeling of low-density lipoprotein (LDL) transport in the

    artery-effects of hypertension. Int J Heat Mass Transfer 2006;49(56):85067.

    [6] Pan C, Hilpert M, Miller CT. Pore-scale modeling of saturated permeabilities in

    random sphere packings. Phys Rev E 2001;64(6):066702-1.

    [7] Pan C. Use of pore-scale modeling to understand transport phenomena in

    porous media. PhD dissertation in environmental science and engineering.

    University of North Carolina at Chapel Hill; 2003.

    [8] Lu WM, Tung KL, Hwang KJ. Fluid flow through basic weaves of monofilament

    filter cloth. Textile Res J 1996;66(5):31123.

    [9] Tung KL, Shiau J, Chuang C. CFD analysis on fluid flow through multifilament

    woven filter cloths. Separat Sci Technol 2002;37(4):799821.

    [10] Koido T, Furusawa T, Moriyama K. An approach to modeling two-phase

    transport in the gas diffusion layer of a proton exchange membrane fuel cell. J

    Power Sources 2008;175(1):12736.

    [11] Djilali N. Computational modelling of polymer electrolyte membrane (PEM)

    fuel cells: challenges and opportunities. Energy 2007;32(4):26980.

    [12] Wang L, Afsharpoya B. Modeling fluid flow in fuel cells using the lattice-Boltzmann approach. Math Comput Simul 2006;72(2-6):2428.

    [13] Vafai K. Handbook of porous media. Taylor & Francis; 2005.

    [14] Vafai K, Tien CL. Boundary and inertia effects on flow and heat transfer in

    porous media. Int J Heat Mass Transfer 1981;24(2):195203.

    [15] Bear J, Bachmat Y. Introductionto modeling of transport phenomena in porous

    media. Dordrecht: Kluwer Academic Publishers; 1990.

    [16] Hazlett RD. Simulation of capillary-dominated displacements in

    microtomographic images of reservoir rocks. Transport Porous Med

    1995;20(12):2135.

    [17] Kerckhofs G, Schrooten J, Van Cleynenbreugel T. Validation of X-ray

    microfocus computed tomography as an imaging tool for porous structures.

    Rev Sci Instrum 2008;79(1):013711.[18] Larson RE, Higdon JJL. Microscopic flow near the surface of two-dimensional

    porous media. Part 1: axial Flow. J Fluid Mech 1986;166:44972.

    [19] Larson RE, Higdon JJL. Microscopic flow near the surface of two-dimensional

    porous media. Part 2: transverse flow. J Fluid Mech 1987;178:11936.

    [20] Sangani AS, Acrivos A. Slow flow through a periodic array of spheres. Int J

    Multiphase Flow 1982;8(4):34360.

    [21] Sangani AS, Acrivos A. Slow flow past periodic arrays of cylinders with

    application to heat transfer. Int J Multiphase Flow 1982;8(3):193206.

    [22] Drummond JE, Tahir MI. Laminar viscous flow through regular arrays of

    parallel solid cylinders. Int J Multiphase Flow 1984;10(5):51540.

    [23] Jackson GW, James DF. Permeability of fibrous porous media. Can J Chem Eng

    1986;64(3):36474.

    [24] Gebart BR. Permeability of unidirectional reinforcements for RTM. J Compos

    Mater 1992;26(8):110033.

    [25] Nabovati A, Sousa ACM. Fluid flow simulation at open-porous medium

    interface using the lattice Boltzmann method. Int J Numer Method Fluids

    2008;56(8):144956.

    [26] Van Der Hoef MA, Beetstra R, Kuipers JAM. Lattice-Boltzmann simulations

    of low-Reynolds-number flow past mono- and bidisperse arrays of

    spheres: results for the permeability and drag force. J Fluid Mech

    2005;528:23354.

    [27] Yang A, Miller CT, Turcoliver LD. Simulation of correlated and uncorrelated

    packing of random size spheres. Phys Rev E 1996;53(2):1516.

    [28] Martys NS, Torquato S, Bentz DP. Universal scaling of fluid permeability for

    sphere packings. Phys Rev E 1994;50(1):4038.

    [29] Yang J, Jia Y, Sun S. Mesoscopic simulation of the impregnating process of

    unidirectional fibrous preform in resin transfer molding. Mater Sci Eng A

    2006:51520.

    [30] Shojaei A, Ghaffarian SR, Karimian SMH. Modeling and simulation approaches

    in the resin transfer molding process: a review. Polym Compos 2003;24(4):

    52544.

    [31] Koponen A, Kandhai D, Hellen E. Permeability of three-dimensional random

    fiber webs. Phys Rev Lett 1998;80(4):716.

    [32] Koponen A. Simulations of fluid flow in porous media by lattice gas and lattice

    Boltzmann methods. PhD dissertation, Jyvaskyla, Department of Physics:

    University of Jyvaskyla; 1998.

    [33] Wang Q, Maze B, Tafreshi HV. A note on permeability simulation ofmultifilament woven fabrics. Chem Eng Sci 2006;61(24):80858.

    [34] Wang Q, Maze B, Tafreshi HV. On the pressure drop modeling of

    monofilament-woven fabrics. Chem Eng Sci 2007;62(17):481721.

    [35] Song YS, Chung K, Kang TJ. Prediction of permeability tensor for three

    dimensional circular braided preform by applying a finite volume method to a

    unit cell. Compos Sci Technol 2004;64(1011):6291636.

    [36] Song YS, Youn JR. Asymptotic expansion homogenization of permeability

    tensor for plain woven fabrics. Compos A: Appl Sci Manuf

    2006;37(11):20807.

    [37] Clague DS, Kandhai BD, Zhang R. Hydraulic permeability of (un)bounded

    fibrous media using the lattice Boltzmann method. Phys Rev E 2000;61(1):

    61625.

    [38] Jiang F, Sousa ACM. Smoothed particle hydrodynamics modeling of transverse

    flow in randomly aligned fibrous porous media. Transport Porous Med

    2008;75(1):1733.

    [39] Jiang FM, Oliveira MSA, Sousa ACM. Mesoscale SPH modeling of fluid flow in

    isotropic porous media. Comput Phys Commun 2007;176(7):47180.

    [40] Frisch U, Hasslacher B, Pomeau Y. Lattice-gas automata for the NavierStokes

    equation. Phys Rev Lett 1986;56(14):15058.[41] Krafczyk M, Schulz M, Rank E. Lattice-gas simulations of two-phase flow in

    porous media. Commun Numer Methods Eng 1998;14(8):70917.

    [42] Qian YH, DHumieres D, Lallemand P. Lattice BGK models for NavierStokes

    equation. Europhys Lett 1992;17(6):47984.

    [43] Chen S, Doolen GD. Lattice Boltzmann method for fluid flows. Ann Rev Fluid

    Mech 1998;30:32964.

    [44] Succi S. The lattice Boltzmann equation for fluid dynamics and

    beyond. Oxford: Oxford University Press; 2001.

    [45] McNamara GR, Zanetti G. Use of the Boltzmann equation to simulate lattice-

    gas automata. Phys Rev Lett 1988;61(20):23325.

    [46] Higuera FJ, Succi S, Benzi R. Lattice gas dynamics with enhanced collisions.

    Europhys Lett 1989;9(4):3459.

    [47] Higuera FJ, Jimenez J. Boltzmann approach to lattice gas simulations. Europhys

    Lett 1989;9(7):6638.

    [48] Succi S, Foti E, Higuera F. Three-dimensional flows in complex geometries with

    the lattice Boltzmann method. Europhys Lett 1989;10(5):4338.

    [49] Bernsdorf J, Durst F, Schfer M. Comparison of cellular automata and finite

    volume techniques for simulation of incompressible flows in complexgeometries. Int J Numerical Method Fluids 1999;29(3):25164.

    868 A. Nabovati et al./ Composites: Part A 40 (2009) 860869

  • 8/6/2019 Nab Ova Ti 09

    10/10

    [50] Breuer M, Bernsdorf J, Zeiser T. Accurate computations of thelaminar flowpast

    a square cylinder based on two different methods: lattice-Boltzmann and

    finite-volume. Int J Heat Fluid Flow 2000;21(2):18696.

    [51] Llewellin EW. LB flow: an extensible lattice Boltzmann framework for the

    simulation of geophysical flows. Part I: theory and implementation. Comput

    Geosci, accepted for publication.

    [52] Llewellin EW. LBflow: an extensible lattice-Boltzmann framework for the

    simulation of geophysical flows. Part II: usage and validation. Comput Geosci,

    accepted for publication.

    [53] Pan C, Luo L-S, Miller CT. An evaluation of lattice Boltzmann schemes

    for porous medium flow simulation. Comput Fluids 2006;35(89):898909.

    [54] Wang JF, Hwang WR. Permeability prediction of fibrous porous media in a bi-

    periodic domain. J Compos Mater 2008;42(9):90929.

    [55] Grujicic M, Chittajallu KM, Walsh S. Lattice Boltzmann method based

    computation of the permeability of the orthogonal plain-weave fabric

    preforms. J Mater Sci 2006;41(23):79898000.

    [56] Nabovati A, Sousa ACM. LBM mesoscale modeling of porous media. Int Rev

    Mech Eng 2008;2(4).

    [57] Koponen A, Kataja M, Timonen J. Permeability and effective porosity of porous

    media. Phys Rev E 1997;56(3):331925.

    [58] Neuman SP. Theoretical derivation of Darcys law. Acta Mech 1977;25(34):

    15370.

    [59] Ahn SH, Lee WI, Springer GS. Measurement of the three-dimensional

    permeability of fiber preforms using embedded fiber optic sensors. J Compos

    Mater 1995;29(6):71433.

    [60] Weitzenbock JR, Shenoi RA, Wilson PA. Measurement of three-dimensional

    permeability. Compos Part A: Appl Sci Manuf 1998;29(12):15969.

    [61] Weitzenbock JR, Shenoi RA, Wilson PA. A unified approach to determine

    principal permeability of fibrous porous media. Polym Compos2002;23(6):113250.

    [62] Parnas RS, Liu Q, Giffard HS. New Set-Up for in-plane permeability

    measurement. Compos Part A (Appl Sci Manuf) 2007;38(3):95462.

    [63] Kolodziej JA, Dziecielak R, Konczak Z. Permeability tensor for heterogeneous

    porous medium of fibre type. Transport Porous Med 1998;32(1):119.

    [64] Nedanov PB, Advani SG. Numerical computation of the fiber preform

    permeability tensor by the homogenization method. Polym Compos

    2002;23(5):75870.

    A. Nabovati et al./ Composites: Part A 40 (2009) 860869 869