cae tutorial

Upload: mohgscribd

Post on 04-Jun-2018

218 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/13/2019 CAE TUTORIAL

    1/8

    Journal of Biomechanics 37 (2004) 127134

    Design and numerical implementation of a 3-D non-linear viscoelastic

    constitutive model for brain tissue during impact

    D.W.A. Brandsa,*, G.W.M. Petersb, P.H.M. Bovendeerdb

    aDepartment of Mechanical Engineering, Division of Computational and Experimental Mechanics, Eindhoven University of Technology, P.O. Box 513,

    5600 MB Eindhoven, The NetherlandsbDepartment of Biomedical Engineering, Eindhoven University of Technology, P.O. Box 513, 5600 MB Eindhoven, The Netherlands

    Abstract

    Finite Element (FE) head models are often used to understand mechanical response of the head and its contents during impactloading in the head. Current FE models do not account for non-linear viscoelastic material behavior of brain tissue. We developed a

    new non-linear viscoelastic material model for brain tissue and implemented it in an explicit FE code. To obtain sufficient numerical

    accuracy for modeling the nearly incompressible brain tissue, deviatoric and volumetric stress contributions are separated.

    Deviatoric stress is modeled in a non-linear viscoelastic differential form. Volumetric behavior is assumed linearly elastic. Linear

    viscoelastic material parameters were derived from published data on oscillatory experiments, and from ultrasonic experiments.

    Additionally, non-linear parameters were derived from stress relaxation (SR) experiments at shear strains up to 20%. The model was

    tested by simulating the transient phase in the SR experiments not used in parameter determination (strains up to 20%, strain rates

    up to 8 s1). Both time- and strain-dependent behavior were predicted accurately R2 >0:96 for strain and strain rates applied.However, the stress was overestimated systematically by approximately 31% independent of strain(rate) applied. This is probably

    caused by limitations of the experimental data at hand.

    r 2003 Elsevier Ltd. All rights reserved.

    Keywords: Finite element modeling; Impact biomechanics; Non-linear viscoelastic constitutive model; Brain modelling

    1. Introduction

    Traumatic brain injury (TBI) caused by a mechanical

    insult on the head, for example during traffic accidents,

    sport accidents or falls, causes high mortality and

    disability (Brooks et al., 1997; Viano et al., 1997;

    Waxweiler et al., 1995). TBI occurs when the local

    mechanical load, exerted on the brain tissue, exceeds

    certain tolerance levels. Understanding how an external

    mechanical load on a head is transferred to a local

    mechanical load in the brain is needed to improve injury

    protecting devices and diagnostic methods. To obtain

    this understanding, finite element (FE) modeling is often

    used (e.g. in Bandak and Eppinger, 1994; Claessens

    et al., 1997; Turquier et al., 1996; Zhang et al., 2001).

    Current FE head models contain a detailed geometrical

    description of the intracranial contents but lack an

    accurate description of brain material behavior.

    At strain and strain rate levels associated with TBI

    (approximately 20% (Bain and Meaney, 2000;Galbraith

    et al., 1993;Schreiber et al., 1997) and typically 20 s1),

    brain tissue behaves as a non-linear, viscoelastic material

    (Bilston et al., 2001;Estes and McElhaney, 1970;Peters

    et al., 1997). In stress relaxation (SR) experiments shear

    softening occurs, i.e. the stiffness decreases as strain

    increases (Arbogast et al., 1995; Bilston et al., 2001;

    Brands et al., 2000; Prange et al., 2000, 2002) while

    Darvish and Crandall (2001)found that shear hardening

    occurred when increasing strains in oscillatory experi-

    ments at frequencies exceeding 44 Hz;indicating full non-linear material behavior. The bulk modulus of brain

    tissue is about 106 times higher than the shear modulus

    (Etoh et al., 1994;Goldman and Hueter, 1956) indicating

    nearly incompressible material behavior. This provides

    the following requirements for a material model for use in

    FE modeling brain tissue during impacts: accurate

    replication of the non-linear viscoelastic behavior in

    shear-like deformations for strain(rate)s up to 20%

    20 s1 and special precautions for accurate modeling

    of the nearly incompressible behavior.

    ARTICLE IN PRESS

    *Corresponding author. Tel.: +31-40-247-3135; fax: +31-40-244-

    7355.

    E-mail address: [email protected] (D.W.A. Brands).

    0021-9290/$- see front matterr 2003 Elsevier Ltd. All rights reserved.

    doi:10.1016/S0021-9290(03)00243-4

  • 8/13/2019 CAE TUTORIAL

    2/8

    In this paper the development and FE implementation

    of a new non-linear viscoelastic constitutive model for

    modeling brain tissue is presented. The model is written

    in a differential formulation as opposed to the QLV

    formulations currently implemented in explicit FE

    packages. Furthermore, the formulation is such that

    the nearly incompressible material behavior is modeledwith sufficient numerical accuracy. Material parameters

    are determined from two sets of shear measurements

    with porcine brain tissue on a rotational viscometer and

    from ultrasonic experiment data. The model and its

    implementation are tested using a three-dimensional FE

    simulation of transient shear experiments.

    2. Methods

    2.1. Kinematics

    The stress in an arbitrary solid material is determined

    by changes of volume and shape, described by the

    deformation gradient tensor F: Deformation gradienttensor Fis split multiplicatively into an elastic part, Fe;and an inelastic part, Fp:

    d~xx F d~xx0; F Fe Fp; 1

    where d~xx0 and d~xx represent a material line element inundeformed stateC0and deformed state Ct;respectively(see Fig. 1). The inelastic contribution refers to the

    deformation (with respect to the undeformed state), of

    the relaxed stress-free configurationCp;which is defined

    as a fictitious state that would be recovered instanta-

    neously when all loads were removed from the material

    element.

    For later use in the constitutive model, we introduce

    the, recoverable, elastic Finger tensor Be; and itsinvariantsI1;2;3:

    Be Fe Fce;

    I1traceBe;

    I2 12traceBe

    2

    traceB2

    e ;I3detBe: 2

    To separate changes in volume from changes in shape,

    we introduce the isochoric elastic Finger tensor %Be and

    its associated invariants as

    %Be I1=33 Be; %I1;2 I1;2 %Be; %I3 1: 3

    The description of the rate of deformation is based on

    the velocity gradient tensor, L which, is decomposed

    additively into an elastic part, Le and an inelastic part,

    Lp;

    L F F1; L Le L

    p;

    Le Fe F1e ; Lp Fe Fp F

    1p F

    1e : 4

    Both parts, Le and Lp; are decomposed additively as

    L D W; D 12L LT; W 1

    2L LT 5

    with D; the symmetric rate of deformation tensor (LT

    denotes the transpose of tensor L) and W; the skew-symmetric spin tensor. To obtain a unique relaxed

    stress-free state, Cp; it is assumed that the inelasticdeformation occurs spin-free,

    We W and Wp 0: 6

    2.2. The constitutive model

    For accurate modeling of the nearly incompressible

    brain tissue, the Cauchy stress, r; is written as the sumof a volumetric part, rv; which depends on volumetricchanges only, and a deviatoric part, rd;which depend onchange of shape only:

    r rv rd: 7

    The volumetric part of the stress, rv; is assumed linearelastic, implying:

    r

    v

    Kffiffiffiffi

    I3p

    1I 8with unitary tensor Iand bulk modulus, K:

    The deviatoric part of the stress is modeled non-linear

    viscoelastic. It is decomposed in a number of viscoelastic

    modes, rdi:

    rd

    XNi1

    rdi: 9

    The number of modes used, N; is determined by thefrequency range for which the model has to be valid. To

    derive rdi we apply Eq. (1) for each mode separately

    obtaining F Fe;i Fp;i: We then define the strain rate

    ARTICLE IN PRESS

    C0C t

    CP

    F

    Fp Fe

    Fig. 1. Graphical representation of the multiplicative decomposition

    of the deformation gradient tensor F: The inelastic part Fp of Ftransforms the undeformed stateC0 to a relaxed stress-free configura-

    tion, Cp; which is a fictitious state that would be recoveredinstantaneously when all loads were removed from the material

    element. The elastic part FeofFtransforms the stress free state, Cp;tothe deformed state Ct: To obtain a unique fictitious stress free state,Cp; it is assumed that the inelastic deformation occurs spin free i.e. allrotations must be accounted for in the constitutive model governing

    the elastic part.

    D.W.A. Brands et al. / Journal of Biomechanics 37 (2004) 127134128

  • 8/13/2019 CAE TUTORIAL

    3/8

    dependent behavior, governed by viscosity, Z i; as

    Dp;i r

    di

    2Zi: 10

    The elastic behavior is modeled using a strain energy

    density function (SEDF), Wi:

    Wi C10;i %I1;i 3 C01;i %I2;i 3 C20;i %I1;i 32

    C02;i %I2;i 32; 11

    where %I1;i and %I2;irepresent invariants of the tensor %Be;i(Eq. (3)).

    The deviatoric part of the Cauchy stress tensor,

    follows fromW:

    rdi

    2ffiffiffiffiffiffiI3;i

    p fC10;i 2C20;i %I1;i 3g %Bde;i

    2

    ffiffiffiffiffiffiI3;ip fC01;i 2C02;i %I2;i 3g %B

    1e;i

    d: 12

    2.3. Numerical implementation

    The time evolution of stress and strain in each mode,

    is described by an evolution law, based on kinematics

    only. For each mode, the inelastic right CauchyGreen

    tensor, Cp; is defined as

    Cp FTp Fp F

    T B1e F: 13

    Taking the time derivative of Cp and using the

    requirement of spin free inelastic deformation, Eq. (6),

    provides a evolution equation insensitive for large rigid-body rotation and translation

    Cp 2Cp F1 Dp F: 14

    This evolution equation is used for updating Cpnumerically in the time integration procedure. At the

    new time increment, the updated elastic Finger tensor

    then follows from Be F C1p F

    T: Application ofBein constitutive equations (8) and (12) yields updated rdiand rv: Finally, the inelastic rate of deformation Dp isdetermined from Eq. (10) and serves as basis for

    proceeding to the next time step using Eq. (14) again.

    This procedure is implemented in an explicit FE Code

    typically used for crash impact simulations (madymo,

    TNO-Automotive, 1999).

    2.4. Determination of material parameters

    The multi-mode model contains one material para-

    meter for the volumetric behavior K; and fiveparameters for each viscoelastic mode C10;i; C01;i; C20;i;C02;i and Zi).

    The bulk modulus, K;is determined from the velocityof dilatational waves,cp;measured in brain tissue (Etohet al., 1994; Goldman and Hueter, 1956) and mass

    density r; using:

    Kc2pr: 15

    Linear parameters C10;i; C01;i and Zi were determinedfrom data of oscillatory shear experiments or dynamic

    frequency sweeps (DFS) with porcine brain tissue on a

    rotational plate-plate viscometer (ARES, RheometricScientific, 1993) published previously (Brands et al.,

    2000). A strain amplitude of 0.01 was applied in a

    frequency range of 1:6216 Hz at various temperaturesand a master curve, valid for frequencies ranging from

    1.6 to 684 Hz at 37C; was constructed using the timetemperature superposition principle (TTS) first applied

    to brain tissue in (Peters et al., 1997). The results of these

    experiments were presented in terms of the storage

    modulus , G0 and the loss modulus, G00;as a function offrequency o: For small shear strains, the non-linearviscoelastic constitutive equations (10) and (12) reduce

    to a linear multimode Maxwell model for which, G0 and,

    G00; are found as:

    G0 Xni0

    Gil2io

    2

    1l2io2; 16

    G00 Xni0

    Gilio

    1l2io2

    with Gi2C10;i C01;i the shear modulus and li

    Zi=Githe relaxation time of mode i:Note that, only the sumC10;i C01;ican be obtained

    from the shear experiments. We uniquely determined

    C10;iand C01;iby assuming them to be equal as inMiller

    and Chinzei (1997).

    Next, the parameters describing non-linear behavior,

    C20;i and C02;i; were derived from large strain stressrelaxation experiments(SR), using the linear parameters

    determined before. These were performed, with the same

    sample, within minutes after finishing the DFS at shear

    strains of 0.05, 0.1 and 0.2. Experiments were finished

    within 4 h after sacrifice.

    In the SR experiments it was found that, once the

    strain had obtained a constant value, (within 0:1 s), thestress could be written as a function of the shear strain

    at the plate edge, gR; a strain-dependent normalized

    stiffness (damping function), hagR; and the linearrelaxation modulus, Gt:

    tat; gR GthagRgRt; 17

    where

    hagR

    tagRgR

    limgR-0tagRgR

    : 18

    Before using these data in the present work, they were

    corrected for the radial inhomogeneity in the strain field,

    according to the procedure described in the appendix, to

    obtain the true damping function htruegR:

    ARTICLE IN PRESS

    D.W.A. Brands et al. / Journal of Biomechanics 37 (2004) 127134 129

  • 8/13/2019 CAE TUTORIAL

    4/8

    Evaluation of Eq. (12) for simple shear strain provides

    the following expression for normalized stiffness for

    each mode:

    hig 1 2g2fnls;i; fnls;i

    C02;i C20;i

    C01;i C10;i: 19

    The non-linear shear parameter, fnls;i describes thenonlinearity in mode i: Since htruegR is independentof time, we assume equal nonlinearity for each mode, i.e.

    hig htruegR providing fnls;ifnls:As for the linear parameters, also for the nonlinear

    parameters we uniquely determined C20;i and C02;i by

    assuming them to be equal.

    2.5. Test of the constitutive model

    The model is tested by a three-dimensional FE

    simulation of the SR experiments on the rotational

    rheometer including the transient strain onset that has

    not been used for material parameter determination. In

    this manner we test the constitutive model and its

    numerical implementation as well as the validity of the

    strain correction method applied to the experimental

    data.

    The geometry of the brain sample is modeled by a

    quarter cylinder with diameter 25 mm and thickness

    2 mm corresponding to the typical sample size used in

    the experiments. Spatial discretisation is obtained using

    1288 brick elements with linear interpolation functions

    (Fig. 2). A preliminary convergence study revealed a

    maximum deviation of 2%. To prevent mesh locking,

    likely to occur due to the nearly incompressible materialbehavior, reduced spatial integration is used. Symmetry

    boundary conditions are applied on the cross-sectional

    planes. The lower plane is rotated according to the

    experimental data. The upper-plane is rigidly supported.

    To compare numerical and experimental results, the

    reaction torque at the upper plate,T; is determined andthe apparent stress, tat; gR; is calculated, as in theexperiments, using

    tat; gR 2T

    pR3 20

    with, R the radius of the sample. Brain tissue material

    parameters shown inTable 1are applied. The time step

    used is limited by the conditional stability of the explicit

    central difference time integration scheme used in the

    code and is set to 1:66107 s:

    3. Results

    3.1. Determination of material parameters

    With cp 155972 m=s (average7range in literaturevalues) and r 1040 kg=m3 Eq. (15) provides a bulkmodulus of 2:3 GPa:

    The small strain DFS experimental data could be

    fitted well over the complete frequency range using four

    viscoelastic modes with first-order material parameters

    shown in Table 1 (see Fig. 3). The error in G0 equals

    2:079:6% (average7standard deviation) and reachesmaximum absolute values of approximately 20% at

    upper and lower end of the frequency spectrum (684 and

    1:6 Hz;respectively) . The loss modulus is fitted with anaverage error of 3:574:6%:The error values remain lessthan 10% and are approximately distributed at random.

    Application of more modes yielded no improvement to

    the results while less modes deteriorated the accuracy of

    the solution. The damping function obtained from the

    SR data shows that shear softening is more prominent in

    the normalized stiffness data which has been corrected

    for strain inhomogeneity than in the raw experimental

    data (Fig. 4). The shear softening equals 37% at 20%

    strain. Fitting Eq. (19) to this data with fnls 4:49provides a relative error less than 2%.

    ARTICLE IN PRESS

    Fig. 2. Graphical representation of the three-dimensional mesh (Top

    view and side view of cross-section shown). A quarter cylinder is

    modeled using 1288 brick elements with 1672 nodes. Symmetry

    boundary conditions are applied on the cross-sectional planes.

    Table 1

    Material parameters obtained from fitting the 4-mode non-linear

    viscoelastic constitutive model on brain tissue DFS and SR shear data

    (sample G2 inBrands et al. (2000))

    Mode MooneyRivlin parameters Viscosity Bulk modulus

    i C10;i C01;i(Pa) C20;i C02;i(Pa) Zi(Pas) K(GPa)

    0 85.96 386:0 N 2.31 67.27 302:0 18.9 2 80.66 362:2 2.46 3 106.8 479:5 0.606 4 824.9 3704 0.0403

    D.W.A. Brands et al. / Journal of Biomechanics 37 (2004) 127134130

  • 8/13/2019 CAE TUTORIAL

    5/8

    3.2. Test of the constitutive model

    The three-dimensional FE simulation of the rota-

    tional viscometer experiments, with material parameters

    in Table 1, provides predicted stress values which are

    systematically higher than experimental results (Fig. 5).

    A linear regression analysis (Statgraphics 5.1), shows

    good correlation between numerical and experimental

    results R2 >96%: The ratio of predicted over experi-mental stress equals 1:3370:28; 1:3470:25; 1:2670:18for 5%, 10% and 20% strain, respectively (average7

    standard deviation) but does not statistically depend on

    strain at the 95% confidence level (Students-T-test).

    The model thus provides good prediction of time-

    dependent behavior but an average over estimation of

    the stress by 31% which does not depend on strain.

    Scaling the simulation results by a single factor of 0.7

    indeed provides much better simulation results (strain

    averaged maximum relative error equals 7% at 0:2 s(Fig. 6)).

    4. Discussion

    A non-linear viscoelastic material law for brain tissue

    is developed and implemented in a FE code, to improve

    capabilities of FE head models to predict TBI in traffic

    accidents, sport accidents or falls.

    To prevent that errors in the computation of the

    hydrostatic stress interfere with deviatoric stress, we

    decoupled hydrostatic and deviatoric behavior by using

    isochoric strain measures for the deviatoric behavior.

    This is important since, due to the nearly incompressible

    material behavior, deviatoric stresses are typically six

    orders of magnitude smaller than hydrostatic stresses. A

    drawback of this approach is that it introduces the

    additional assumption that shear behavior is indepen-

    dent of volumetric compression. However, to our

    knowledge, evidence of such dependency is not reported

    in literature.

    The hydrostatic stress is modeled linearly elastic.

    Linearity seems valid as volumetric strains are small at

    pressures expected during traffic impacts (typically

    105 Pa (Nahum et al., 1977)). Viscous effects are

    neglected, since ultrasonic experiments indicate signifi-

    cant damping due to hydrostatic deformation only at

    ARTICLE IN PRESS

    0 5 10 15 200.4

    0.6

    0.8

    1

    [%]

    h()

    Raw dataCorrected dataFit result

    Fig. 4. Normalized stiffness,hg;of sample G2 inBrands et al. (2000)obtained from the constant strain part of SR experiments on a plate

    plate viscometer. The raw experimental data is compared with data

    corrected for the non-homogeneous radial strain field by applying

    second-order MooneyRivlin (MR2) model in Eq. (A.1). To obtain

    realistic fits of the normalized stiffness, (i.e. limg-0 hg 1), the

    normalized stiffness value at 5% strain is assumed to be valid at 1%

    strain also. Shear softening is more prominent in corrected data whilefitted result shows that the MR2 model provides excellent fit.

    0 0.1 0.2 0.3 0.4

    0

    10

    20

    30

    Time [s]

    Strain[%]

    0 0.1 0.2 0.3 0.4

    0

    20

    40

    60

    80

    100

    Time [s]

    Stress[Pa]

    0=0.05

    0=0.1

    0=0.2

    NumExp

    Fig. 5. Experimental apparent shear stresses (sample G2 in Brands

    et al. (2000)), at 5%, 10% and 20% maximum edge strain, together

    with the FE simulation results. Left: sample edge shear strain histories,

    Right: apparent shear stresses, Num result of three-dimensional FE

    model, Exp experiment: The numerical predictions are system-atically higher than the experimental results regardless of strain and

    time.

    0 0.1 0.2 0.3 0.40

    20

    40

    60

    80

    Time [s]

    Stress[Pa]

    0=0.05

    0=0.1

    0=0.2

    NumExp

    0 0.05 0.10

    20

    40

    60

    80

    0=0.05

    0=0.1

    0=0.2

    Time [s]

    Stress[Pa]

    Fig. 6. Investigation of error source of simulation results versus

    experimental data. Left: simulation results multiplied by a constant

    factorX0:7 showing much better simulation results. Right: closeup of the stress history during transient part of the deformation (scaled

    result).

    100

    101

    102

    10310

    2

    103

    104

    Frequency [Hz]

    G [Pa]

    ExpFit

    100

    101

    102

    10310

    2

    103

    104

    Frequency [Hz]

    G [Pa]

    Fig. 3. Master curves of storage and loss modulus G0 and G00 from

    sample G2 obtained from DFS with TTS applied, at 1% shear strain in

    Brands et al. (2000) together with four mode fit showing good

    resemblance of small strain behavior.

    D.W.A. Brands et al. / Journal of Biomechanics 37 (2004) 127134 131

  • 8/13/2019 CAE TUTORIAL

    6/8

    frequencies above 50 kHz (Etoh et al., 1994), well above

    1000 Hz typical for impacts of interest in this study.

    The deviatoric behavior is described using a differ-

    ential formulation. Differential formulations have been

    proposed before (Bilston et al., 2001; Pamidi and

    Advani, 1978). However, these predict infinite stress at

    instantaneous loading (Pamidi and Advani, 1978) orwhere not implemented in a FE package (Bilston et al.,

    2001; Pamidi and Advani, 1978). Based on the experi-

    mental finding of strain and time separability in stress

    relaxation (SR) experiments, we could have opted

    for a quasi-linear viscoelastic (QLV) model (Mendis

    et al., 1995; Miller and Chinzei, 1997; Prange et al.,

    2002).

    However, at frequencies above 44 Hz fully non-linear

    material behavior might be present (Darvish and

    Crandall, 2001) which cannot be described with QLV

    theory but can be included in the present formulation.

    The elastic behavior was modelled by a hyperelastic

    SEDF written in polynomial form with integer powers

    of first and second invariants of the isochoric Finger

    tensor. We took a second-order model, the simplest

    model which predicts non-linear shear behavior. For

    correct prediction of shear softening observed (Fig. 4)

    negative second-order parameters were required. If the

    model is used outside its range of validity (i.e. strains

    more than 20%) this might result in negative stiffness

    when shear strains exceed 27%.

    Negative stiffness in simulations in a head model, in

    which it is unknown beforehand which strain levels will

    occur, can be avoided by extending the model with a

    third order MooneyRivlin term with suitable para-meter settings as done inBrands (2002)andBrands et al.

    (2002). However, for realistic parameter settings, experi-

    ments at higher strains must be performed. Negative

    stiffness can also be avoided by using an Ogden SEDF

    with fractional powers of stretches Ogden (1972),

    applied to brain tissue in Miller and Chinzei (2002)

    andPrange et al. (2002, 2000).However, we found that

    this SEDF cannot predict the amount of shear softening

    observed in our experiments.

    The inelastic behavior is modeled by a simple linear

    Newtonian law. To describe the viscoelastic behavior in

    a broad frequency range, we used a multi-mode

    approach with discrete time constants as opposed to a

    relaxation spectrum, i.e. some arbitrarily continuous

    function of relaxation time (Macosko, 1994). A draw-

    back of this method that each additional mode

    introduces five new material parameters in the model.

    To reduce the number of independent parameters we

    assumed the ratio between first- and second-order

    MooneyRivlin parameters governing the elastic

    strain-dependent behavior in each mode to be constant,

    based on the observation of (approximately) time-strain

    separability in SR data (Bilston et al., 2001; Brands et al.,

    2000;Prange et al., 2002).

    We presented a new approach for determining

    material parameters of brain tissue on a rotational

    viscometer. Often all material parameters are deter-

    mined from of SR experiments only while assuming

    perfect instantaneous strain application (Bilston et al.,

    2001;Prange et al., 2002). A drawback of this approach

    is that time constants valid during the rise time of thestrain cannot be determined. Instead, we used two data

    sets from a single sample (Brands et al., 2000). Linear

    viscoelastic material parameters were fitted to DFS

    results valid for small strains (0.01) but high frequencies

    (up to 684 Hz) while parameters C20 C02;describingthe strain-dependent decrease of stiffness, were deter-

    mined from the constant strain part of SR experiments.

    This method implies that time constants, determined for

    small strains and high frequencies, are also valid at large

    strains. Evaluating the predicted material behavior

    during transient strain application in the SR serves as

    a test of this assumption.

    During simulation of the SR experiments, stress

    values were overestimated by 31%, independent of the

    strain level applied, also during transient onset of the

    strain. A potential explanation is that we assumed no

    shear softening for strains between 1% (at which the

    DFS is performed) and 5%, the lowest strain value in

    the SR experiments, resulting in a normalized stiffness

    of 1.0 (Fig. 4). As a result, the fitted model predicts only

    2% shear softening at 5% strain. This softening is low

    compared to the 3050% softening reported in litera-

    ture, when increasing shear strains from 1% to 5%

    (Brands et al., 1999;Bilston et al., 2001).

    Correct strain-dependent behavior shows that thecorrection method for the radially inhomogeneous

    strain field is indeed valid. Correct time-dependent

    behavior indicates that time constants obtained from

    the DFS results are valid for frequencies over 20 Hz

    also. The DFS results were composed from isothermal

    DFS results up to 16 Hz; using the time temperaturesuperpositioning (TTS) principle. This provides evi-

    dence on the validity of TTS for determining time

    constants of brain tissue for given model, as well as the

    validity of these time constants determined at small

    strains, for large strains and high frequencies. However,

    strain rates applied range up to 8 s 1 which is less than

    the rate values expected during injurious impacts (15

    21 s1 Brands (2002)). The behavior at higher strain

    rates remains to be investigated.

    It is impossible to fully characterize the three-

    dimensional non-linear viscoelastic behavior of a

    material using simple shear experiments from a rota-

    tional viscometer. Effects of anisotropic material beha-

    vior (Prange et al., 2002) cannot be determined. Also,

    MooneyRivlin parameters C10;i; C01;i; C20;i and C02;icould not be defined uniquely. The ratio ratios C10=C01in a first-order MooneyRivlin model did have a

    significant effect on stress response in free compression

    ARTICLE IN PRESS

    D.W.A. Brands et al. / Journal of Biomechanics 37 (2004) 127134132

  • 8/13/2019 CAE TUTORIAL

    7/8

    experiments byMendis et al. (1995). Adjusting the ratios

    Cj0;i=C0j;i while keeping Cj0;i C0j;iconstant provides apowerful tool to tune free compression results without

    influencing shear behavior. The performance of the

    model for other deformation modes remains to be

    investigated.

    In this paper, we developed a non-linear viscoelasticmaterial law for brain tissue and implemented it in a FE

    code, to improve capabilities of FE head models to

    predict TBI in traffic accidents, falls or sports. We

    provided an approach to determine material parameters

    for the model and found that the model is capable of

    predicting realistic shear material behavior of brain

    tissue observed in SR experiments including the

    transient application of the strain. However, full

    characterization of the material properties is not

    possible using experimental shear data at hand and

    requires extra experiments.

    Acknowledgements

    The authors would like to acknowledge TNO Prins

    Maurits Laboratory and Ford Motor Company for

    their financial support of this research and P. Nauta of

    TNO Automotive for his assistance during the numer-

    ical implementation of the model.

    Appendix. A

    Brands et al. (2000) determined the stress assuminglinear viscoelastic theory, thus neglecting the effect of

    the radial inhomogeneous strain field between the plates

    of the rotational viscometer. For any isotropic material,

    the true experimental normalized stiffness htruegR can

    be obtained when the strain rate is zero, using Soskey

    and Winter (1984),

    htruegR hagR 1q ln hagR

    4q ln gR

    : A:1

    When experimental data at sufficient strain levels is

    present, q ln hagR=4q ln gR can be estimated without

    making assumptions on material behavior. As we havedata at three strain values only, we fitted Eq. (19) to

    experimentally found hag and applied Eq. (21) to

    obtain htrueg which is then fitted to Eq. (19) again

    to obtain the correct material parameters.

    References

    Arbogast, K.B., Meaney, D.F., Thibault, L.E., 1995. Biomechanical

    characterization of the constitutive relationship for the brainstem.

    In: Proceedings of the 39th Stapp Car Crash Conference, SEA

    Paper 952716, pp. 153159.

    Bain, A.C., Meaney, D.F., 2000. Tissue-level thresholds for axonal

    damage in an experimental model of central nervous system white

    matter injury. Journal of Biomechanical Engineering: Transactions

    of the ASME 122, 615622.

    Bandak, F.A., Eppinger, R.H., 1994. A three-dimensional finite

    elements analysis of the human brain under combined rota-

    tional and translational accelerations. In: Proceedings of the

    38th Stapp Car Crash Conference, SAE Paper 942215,pp. 148163.

    Bilston, L.E., Liu, Z., Phan-Thien, N., 2001. Large strain behaviour of

    brain tissue in shear: some experimental data and differential

    constitutive model. Biorheology 38, 335345.

    Brands, D.W.A., 2002. Predicting brain mechanics during closed head

    impact: numerical and constitutive aspects. Ph.D. Dissertation,

    Eindhoven University of Technology, Eindhoven, The Nether-

    lands. Available on internet: http://alexandria.tue.nl/extra2/

    200210809.pdf.

    Brands, D.W.A., Bovendeerd, P.H.M., Peters, G.W.M., Paas, M., van

    Bree, J., Wismans, J.S.H.M., 1999. Comparison of the dynamic

    behaviour of brain tissue and two model materials. In: Proceedings

    of the 43rd Stapp Car Crash Conference, Paper no. 99sc21,

    pp. 5764.

    Brands, D.W.A., Bovendeerd, P.H.M., Peters, G.W.M., Wismans,J.S.H.M., 2000. The large strain dynamic behaviour of in vitro

    porcine brain tissue and a silicone gel model material. Stapp Car

    Crash Journal 44, 249260.

    Brands, D.W.A., Bovendeerd, P.H.M., Wismans, J.S.H.M., 2002. On

    the potential importance of non-linear viscoelastic material

    modelling for numerical prediction of rain tissue response: test

    and application. Stapp Car Crash Journal 46, 103121.

    Brooks, C.A., Gabella, B., Hoffman, R., Sosin, D., Whiteneck, G.,

    1997. Traumatic brain injury: designing and implementing a

    population-based follow-up system. Arch Phys Med Rehabil 78

    (8), S26S30.

    Claessens, M., Sauren, F., Wismans, J., 1997. Modelling of the human

    head under impact conditions: a parametric study. In: Proceedings

    of the 41st Stapp Car Crash Conference, no. SAE 973338,

    pp. 315328.

    Darvish, K.K., Crandall, J.R., 2001. Nonlinear viscoelastic effects in

    oscillatory shear deformation of brain tissue. Medical Engineering

    and Physics 23 (9), 633645.

    Estes, M.S., McElhaney, J.H., 1970. Response of Brain Tissue to

    Compressive Loading. No. ASME Paper, 70-BHF-13.

    Etoh, A., Mitaku, S., Yamamoto, J., Okano, K., 1994. Ultrasonic

    absorption anomaly of brain tissue. Japanese Journal of Applied

    Physics 33, 28742879.

    Galbraith, J.A., Thibault, L.E., Matteson, D.R., 1993. Mechanical and

    electrical responses of the squit giant axon to simple elongation.

    Journal of Biomechanical Engineering 115, 1322.

    Goldman, D.E., Hueter, T.F., 1956. Journal of the Acoustical Society

    of America 28, 35.

    Macosko, C.W., 1994. Rheology: Principles, Measurements, andApplications. Advances in Interfacial Engineering Series. VCH

    Publishers Inc., New York, NY, USA.

    Mendis, K.K., Stalnaker, R.L., Advani, S.H., 1995. A constitutive

    relationship for large deformation finite element modelling of brain

    tissue. Journal of Biomechanical Engineering: Transactions of the

    ASME 117, 279285.

    Miller, K., Chinzei, K., 1997. Constitutive modelling of brain tissue:

    experiment and theory. Journal of Biomechanics 30 (11/12),

    11151121.

    Miller, K., Chinzei, K., 2002. Mechanical properties of brain tissue in

    tension. Journal of Biomechanics 35, 483490.

    Nahum, A.M., Smith, R.W., Ward, C.C., 1977. Intracranial pressure

    dynamics during head impact. In: Proceedings 21st Stapp Car

    Crash Conference, SAE Paper 770922, pp. 339366.

    ARTICLE IN PRESS

    D.W.A. Brands et al. / Journal of Biomechanics 37 (2004) 127134 133

    http://alexandria.tue.nl/extra2/200210809.pdfhttp://alexandria.tue.nl/extra2/200210809.pdfhttp://alexandria.tue.nl/extra2/200210809.pdfhttp://alexandria.tue.nl/extra2/200210809.pdf
  • 8/13/2019 CAE TUTORIAL

    8/8

    Ogden, R.W., 1972. Large deformation isotropic elasticityon the

    correlation of theory and experiment for incompressible rubberlike

    solids. In: Proceedings of the Royal Society London, no. Series A

    326, London, UK, pp. 565584.

    Pamidi, M.R., Advani, S.H., 1978. Nonlinear constitutive equations

    for human brain tissue. Transactions of the ASME: Journal of

    Biomechanical Engineering 100, 4448.

    Peters, G., Meulman, H., Sauren, A., 1997. Application of the timetemperature superposition theory on brain tissue. Biorheology 34,

    127138.

    Prange, M.T., Meaney, D.F., Margulies, S.S., 2000. Defining brain

    mechanical properties: effects of region, direction and species.

    Stapp Car Crash Journal 44, 205214.

    Prange, M.T., Meaney, D.F., Margulies, S.S., 2002. Regional,

    directional, and age-dependent properties of the brain undergoing

    large deformation. Journal of Biomechanical Engineering: Trans-

    actions of the ASME 124 (2), 198207.

    Rheometric Scientific, 1993. Advanced Rheometric Expansion System,

    with 2K FRTN1 Force Transducer and Liquid Nitrogen Air

    Temperature Control, Instrument Manual. Piscataway, NJ, USA.

    Schreiber, D.I., Bain, A.C., Meaney, D.F., 1997. In vivo thresholds for

    mechanical injury to the blood-brain barrier. In: Proceedings of the

    41st Stapp Car Crash Conference, no. SAE 973335, SAE,

    pp. 277291.

    Soskey, P.R., Winter, H.H., 1984. Large step shear strain experiments

    with parallel-disk rotational rheometers. Journal of Rheology 28

    (5), 625645.

    TNO-Automotive, 1999. MADYMO Users Manual 3D. Version 5.4

    Edition. TNO-Automotive, Delft, The Netherlands.

    Turquier, F., Kang, H.S., Trosseille, X., Willinger, R., Lavaste, F.,Tarriere, C., D#omont, A., 1996. Validation study of a 3D finite

    element head model against experimental data. In: Proceedings

    of the 40th Stapp Car Crash Conference, SAE Paper 962431,

    pp. 283294.

    Viano, D., von Holst, H., Gordon, E., 1997. Serious brain injury from

    traffic related causes: priorities for primary prevention. Accident

    Analysis & Prevention 29 (6), 811816.

    Waxweiler, R.J., Thurman, D., Sniezek, J., Sosin, D., Niell, J.O., 1995.

    Monitoring the impact of traumatic brain injury: a review and

    update. Journal of Neurotrauma 12 (4), 509516.

    Zhang, L., Yang, K.H., King, A.I., 2001. Comparison of brain

    responses between frontal and lateral impacts by finite element

    modeling. Journal of Neurotrauma 18 (1), 2130.

    ARTICLE IN PRESS

    D.W.A. Brands et al. / Journal of Biomechanics 37 (2004) 127134134