15
Numerical Simulation of Fiber Orientation in Injection Molding of Short-Fiber-Reinforced Thermoplastics S. T. CHUNG and T. H. WON* Department of Mechanical Engineering Pohang University of Science and Technology Pohang, Korea The present study develops a numerical simulation program to predict the transient behavior of fiber orientations together with a mold filling simulation for short-fiber-reinforced thermoplastics in arbitrary three-dimensional injection mold cavities. The Dinh-Armstrong model including an additional stress due to the existence of fibers is incorporated into the Hele-Shaw equation to result in a new pressure equation governing the filling process. The mold filling simulation is performed by solving the new pressure equation and energy equation via a finite element/finite difference method a s well as evolution equations for the second- order orientation tensor via the fourth-order Runge-Kutta method. The fiber orientation tensor is determined at every layer of each element across the thick- ness of molded parts with appropriate tensor transformations for arbitrary three- dimensional cavity space. INTRODUCTION n injection molding of short-fiber-reinforced ther- I moplastics, the prediction of fiber orientation is important for obtaining the desired mechanical prop erties of injection molded parts. It is well known that anisotropic mechanical properties are significantly af- fected by the final fiber orientation state. The flow during a mold filling process plays an important role in forming a flow-induced fiber orientation field. How- ever, the flow field is also affected by the orientation of fibers. Therefore a mold filling simulation of fiber- filled polymer must include an anisotropic constitu- tive equation in order to describe the system completely. Jeffery’s model (1). which was developed for the motion of a single ellipsoidal particle immersed in a viscous fluid, has been a basis of most studies in this field. Since this model does not account for the fiber- fiber interaction and does not consider the effect of fiber orientation on the velocity field, this is valid only for sufficiently dilute suspension. Givler et aL (2) calculated fiber orientations from the numerical inte- gration of Jeffery’s equation along streamlines via the finite element method. Folgar and Tucker (3) devel- oped a model for the orientation behavior of concen- trated suspensions by adding a diffusion term to the Jeffery’s equations in order to account for fiber-fiber interactions. Dinh and Armstrong (4) developed a rheological equation of state for semiconcentrated ‘To whom correspondence should be addressed. 604 POL suspensions. They replaced the multiparticle problem with a single-particle one with the subsequent use of Batchelor’s cell model (5) to estimate the drag on a test fiber to account approximately for fiber-fiber in- teractions, Bibbo et aL (6) showed a good agreement between computed and measured rheological proper- ties using the Dinh and Armstrong model. In recent years several different methods have been used to describe the fiber orientation states. Advani and Tucker (7) developed the use of tensors to describe and predict fiber orientation. These tensors offer a concise representation of orientation state but a cle sure approximation is required to obtain a closed set of evolution equations for orientation tensors. Altan, et aL (8). studied tensorial approximations for the description of orientation for fiber suspensions in homogeneous flows. A closure approximation for three-dimensional structure tensors was proposed by Advani and Tucker (9) modifying the scalar measure of fiber alignment. Recently, Advani and Tucker (10) developed a numerical method to predict the orienta- tion of fibers in a thin, flat part. They combined a finite element/control volume method of the mold filling flow and a finite element calculation for the transient orientation problem. But their orientation calculation was limited to flat parts. Altan et a1 (1 1) presented a numerical method to predict the trans- ient three-dimensional fiber orientations in Hele- Shaw flows in irregularly shaped planar cavities. Each fiber location was traced during mold filling, and along these paths, the independent components of fourth-order orientation tensors were solved at zero .YMER ENGINEERINGAND SCIENCE, MID-APRIL 1995, Vol. 35, NO. 7

Numerical simulation of fiber orientation in injection molding of short-fiber-reinforced thermoplastics

Embed Size (px)

Citation preview

Numerical Simulation of Fiber Orientation in Injection Molding

of Short-Fiber-Reinforced Thermoplastics S . T. CHUNG and T. H. W O N *

Department of Mechanical Engineering Pohang University of Science and Technology

Pohang, Korea

The present study develops a numerical simulation program to predict the transient behavior of fiber orientations together with a mold filling simulation for short-fiber-reinforced thermoplastics in arbitrary three-dimensional injection mold cavities. The Dinh-Armstrong model including an additional stress due to the existence of fibers is incorporated into the Hele-Shaw equation to result in a new pressure equation governing the filling process. The mold filling simulation is performed by solving the new pressure equation and energy equation via a finite element/finite difference method as well as evolution equations for the second- order orientation tensor via the fourth-order Runge-Kutta method. The fiber orientation tensor is determined at every layer of each element across the thick- ness of molded parts with appropriate tensor transformations for arbitrary three- dimensional cavity space.

INTRODUCTION

n injection molding of short-fiber-reinforced ther- I moplastics, the prediction of fiber orientation is important for obtaining the desired mechanical prop erties of injection molded parts. I t is well known that anisotropic mechanical properties are significantly af- fected by the final fiber orientation state. The flow during a mold filling process plays an important role in forming a flow-induced fiber orientation field. How- ever, the flow field is also affected by the orientation of fibers. Therefore a mold filling simulation of fiber- filled polymer must include an anisotropic constitu- tive equation in order to describe the system completely.

Jeffery’s model (1). which was developed for the motion of a single ellipsoidal particle immersed in a viscous fluid, has been a basis of most studies in this field. Since this model does not account for the fiber- fiber interaction and does not consider the effect of fiber orientation on the velocity field, this is valid only for sufficiently dilute suspension. Givler et aL (2) calculated fiber orientations from the numerical inte- gration of Jeffery’s equation along streamlines via the finite element method. Folgar and Tucker (3) devel- oped a model for the orientation behavior of concen- trated suspensions by adding a diffusion term to the Jeffery’s equations in order to account for fiber-fiber interactions. Dinh and Armstrong (4) developed a rheological equation of state for semiconcentrated

‘To whom correspondence should be addressed.

604 POL

suspensions. They replaced the multiparticle problem with a single-particle one with the subsequent use of Batchelor’s cell model (5) to estimate the drag on a test fiber to account approximately for fiber-fiber in- teractions, Bibbo et aL (6) showed a good agreement between computed and measured rheological proper- ties using the Dinh and Armstrong model. In recent years several different methods have been used to describe the fiber orientation states. Advani and Tucker (7) developed the use of tensors to describe and predict fiber orientation. These tensors offer a concise representation of orientation state but a c le sure approximation is required to obtain a closed set of evolution equations for orientation tensors. Altan, et aL (8). studied tensorial approximations for the description of orientation for fiber suspensions in homogeneous flows. A closure approximation for three-dimensional structure tensors was proposed by Advani and Tucker (9) modifying the scalar measure of fiber alignment. Recently, Advani and Tucker (10) developed a numerical method to predict the orienta- tion of fibers in a thin, flat part. They combined a finite element/control volume method of the mold filling flow and a finite element calculation for the transient orientation problem. But their orientation calculation was limited to flat parts. Altan et a1 (1 1) presented a numerical method to predict the trans- ient three-dimensional fiber orientations in Hele- Shaw flows in irregularly shaped planar cavities. Each fiber location was traced during mold filling, and along these paths, the independent components of fourth-order orientation tensors were solved at zero

.YMER ENGINEERINGAND SCIENCE, MID-APRIL 1995, Vol. 35, NO. 7

Numerical Simulation of Fiber Orientation

POLYMER ENGINEERING AND SCIENCE, MID-APRIL 1995, VOl. 35, NO. 7 605

volume fraction limit. And Frahan et aL (12) devel- oped a numerical method for calculating fiber orien- tation in the midsurface of a molded part of small thickness in injection molding. Two-dimensional fiber orientation was predicted in three-dimensional ge- ometries in the absence of coupling between the flow calculation and the fiber orientation.

This paper presents a numerical simulation that couples a nonisothermal mold filling flow and a tran- sient three-dimensional fiber 0rientation.h arbitrary three-dimensional thin cavities in injection molding. The first section reviews the theory describing the fiber orientation states and the equation of change for the orientation tensors. Presented next is a model- ing of the coupled mold filling process with Dinh and Armstrong's rheological equation of state for semi- concentrated fiber suspensions. The numerical analy- sis section describes the development of a computer simulation program to predict the mold filling flow field modified by the presence of fibers and the flow induced fiber orientation states. The program com- bines finite element/finite difference methods for mold filling simulation with a numerical method for an integration of the evolution equations for the fiber orientation tensor. Finally, several numerical results are presented.

GOVERNING EQUATIONS

Fiber Orientation

There are several different approaches to describe the fiber orientation. The simplest form is the scalar angle between the fiber and one of the reference axes. In any injection molded part there are fibers oriented in many different directions. A single orientation an- gle cannot describe the true orientation state of fibers. The most complete, albeit complex, way to describe the orientation state is the use of a probability distri- bution function, 9, which is associated with a unit vector p along the axis of the fiber indicating the orientation. The distribution function, +(p, t), is de- fined such that it gives the probability of a fiber having an alignment in the direction p . The probabil- ity distribution function must be periodic and nor- malized, and it must satisfy the continuity condition (7). Using the distribution function to calculate the fiber orientation state everywhere in the part would make the calculation too long and expensive for a practical use. A more compact representation of fiber orientation state is thus needed.

Advani and Tucker (7) showed the use of even order tensors as a more concise description for the orienta- tion state in a suspension or composite system. The second- and fourth-order orientation tensors are

The second- and fourth-order orientation tensors are completely symmetric and the trace of aij is unity, which is the normalization condition. An advantage of

using orientation tensors is computational efficiency. We will use the second-order three-dimensional ori- entation tensor to describe the three-dimensional ori- entation state. In this case there are only five inde pendent tensor components.

To predict the flow-induced orientation state by using compact orientation tensors, we need the equa- tion of change (i.e., evolution equation) for orientation tensors. One takes the material derivative of Eq I with the help of Jeffrey's equation and continuity equation for I), then obtain the equations of change for orientation tensors (7).

In practical injection molded parts of short-fiber- reinforced thermoplastics, there are too many fibers to consider the fluid as a dilute suspension, so Jef- frey's equation becomes invalid. The orientation state of concentrated suspensions is similar to that of di- lute suspensions. However, as the volume fraction of fibers increases, the orientation state is influenced by fiber-fiber interactions. It is very difficult to model phyically the fiber-fiber interactions. So Folgar and Tucker (3) suggested a phenomenological model. They added a diffusive type of term to Jeffrey's equation by introducing a phenomenological coefficient for model- ing the randomizing effect of mechanical interactions between fibers.

Folgar and Tucker's equation (7) of change for the second-order orientation tensor in a concentrated suspension is

-- Daij 1

Dt 2 - - - ( w i k a k j - aikwkj)

+2C,j(Sij-aaij)

where 8, is a unit tensor and (Y equals 3 for three- dimensional orientation. wij and yij are the velocity and the rate of deformation tensors, respectively, de- fined in terms of velocity gradients as

aui aui aui aui w , , = d - - y =2+-

d X i a x j ' iJ axi ax,

y is the generalized shear rate defined by I 1

And A is a parameter related to the particle shape. I t is given as

where re is the aspect ratio of the fiber. The dimen- sionless interaction coefficient C, describes the strength of fiber-fiber interactions. In fact, C, could be variable depending on the fiber orientation state, fiber aspect ratio, fiber volume fraction, etc. Without a detailed model of interactions there is no way to predict C,: it is assumed to be constant in the pres- ent study.

S. T. Chung and T. H. Kwon

To solve Eq 2 in terms of aij we must introduce a suitable closure approximation for the fourth-order tensor qjkt. There are several types of closure approx- imation is better for a random distribution of orienta- tion, while the quadratic closure approximation is exact for a perfectly aligned orientation state. A hy- brid closure approximation, which mixes the linear and quadratic forms according to a scalar measure of orientation f, performed best for a wider range of orientation states (9). We will use the hybrid closure approximation given for the three-dimensional orien- tation in Ref. 9.

a y k [ = (1 - f ) ai jkl +fa,jkl (3)

where aijkl and hijkr denote the linear closure a p proximation and the quadratic closure approximation respectively. The scalar measure of orientation for three-dimensional orientation field is defined by

f= 1-27 det[ a,,]

f varies from zero (in case of a random orientation) to unity (in case of a perfectly aligned orientation), so that f can be regarded as a measure of orientation.

Injection Molding Filling Flow In injection molding filling flow of short-fiber-rein-

forced thermoplastics, there exists coupling between the flow field and the fiber orientation. Several models are available for calculating flows coupled with fiber orientation. Dinh and Armstrong (4) developed a rhe- ological equation of state for semiconcentrated fiber suspensions. Their model treats slender cylindrical fibers and ignores the fiber thickness effect. The model is valid for a semiconcentrated region where the aver- age spacing between neighboring fibers is greater than the fiber diameter D but less than its length L. Within this concentration range, fiber-fiber interac- tions are taken into account by utilizing a self-con- sistent cell model and considering a representative fiber immersed in an effective medium. The constitu- tive equation of this model can be described as follows:

( nL) - '" (aligned)

( nL2)- (random) h = 1

where T ~ - ~ = total stress tensor

q = viscosity of the thermoplastic medium d U (

rl Xj ui,; = - : velocity gradient tensor

n = number of fibers per unit volume h = average distance from a given fiber

In the present simulation, it is assumed that the average distance from a given fiber to its nearest

to its nearest neighbors

neighbors, h is linear in terms of the scalar measure of orientation f. When the number of fibers per unit volume n is greater than 1/(DL2), the average dis- tance between fibers is assumed to be the same as haliyned for the aligned orientation state. That is,

1 1 for - DLZ o2 L

(5)

The thickness of injection molded parts is usually much smaller than other characteristic dimensions of the part. So, the injection molding filling flow can be approximated by the use of the lubrication approx- imation, based upon the fact that the velocity gradi- ents in the thickness are much larger than those in the in-plane directions. One then obtains the Hele- Shaw approximation of the viscous flow (13). A foun- tain flow is observed at the flow front where the flow field is fully three-dimensional. However, it is known that the size of this region is of the order of a few gap widths, and thus the fountain flow effect is neglected in the present study. The flow of fiber suspensions in narrow gaps was extensively investigated by Tucker (15). When the gap is very thin but interaction (or diffusion) effects between fibers introduce some slight out-of-plane orientation, the gapwise shear stress term still dominates the momentum balance. Also, pressure p is still independent of gapwise coordi- nates z. Thus, the essence of the lubrication approxi- mation is retained. The appropriate simplification of the constitutive equation in such a case is

T nL3

(6)

where 01 takes the value 1 and 2. The coordinate system in Eq 6 is chosen such that 3 is along the thickness direction, 1 and 2 lying on the plane, as indicated in Fig. 1 .

Under the above simplification, the injection mold- ing filling flow for nondilute fiber suspensions is gov-

- w x 1)

Fig. 1 . Schematic diagram of a cavity with coordinate sys tern.

606 POLYMER ENGINEERING AND SCIENCE, MID-APRIL 1995, Vol. 35, No. 7

Numerical Simulation of Fiber Orientation

erned by the momentum equation

where

?T nL3

6 ln(2 h / D ) N G

and by the mass conservation equation

d a -( bii) + -( bZ) = 0 (8) d X J Y

where b is the half gap thickness of the part, which may be a function of x and y. " " denotes an average over z, the gapwise coordinate. And 77, a function of shear rate and temperature T, is the viscosity of the thermoplastic medium. In the present study, a modi- fied cross model has been employed as follows:

-

where n, B, C, and Tb are constants for a given material and q0 is zero-shear-rate viscosity. Appre priate boundary conditions in the z-direction are given by

u = O v = O at z = b du/dz=O dv/dz=O at z=O (10)

Integration of Eq 7 with the help of E q 10 results in

(11)

(12)

where

which is different from a fundamental characteristic of Hele-Shaw type flows of thermoplastics without fibers. The gapwise-average velocities are obtained by integration of E q 12:

s a p S" d p s x y d p

- s d p s y d p S " Y d p

- u=---+--+-- b d x b d x b d y

u = - _ - +--+-- b d y b dy b d x

(14)

where

(15)

S X Y = l t ' N X Y - 22 &

0 77

Hence, substituting E q 14 into Eq 8 gives the pres- sure equation as follows:

Boundary conditions on the pressure equation are

p = 0 on the moving flow front u' n, + 3. ny = 0 on the impermeable boundaries -

(17)

In order to deal with the nonisothermal flow, the energy equation is also to be solved during the filling process. In a thin cavity mold, the energy equation can be simplified as

Appropriate boundary conditions in the z-direction for the above governing equation are given by

T=T, at z = b (19)

dT/dz=O at z=O

NUMERICAL ANALYSIS Mold Filling Flow

The injection molding filling flow is, in its nature, nonisothermal and non-Newtonian with a moving boundary (the flow front). It is difficult to treat the

(13)

moving flow front in cavities of complex geometry. Here, a finite element method is used to solve Eq 16 for the pressure distribution over the flow domain while the gapwise temperature distribution is solved using a finite difference technique (13, 14).

The two last terms in Eq 12 indicate the coupling effect of the pressure gradient to the velocity in the x and y directions, respectively. It should be noted that the velocity vector is not in the direction of -Vp,

POLYMER ENGINEERING AND SCIENCE, MID-APRIL 7995, Vol. 35, NO. 7 607

S. T. Chung and T. H. Kwon

608 POLYMER ENGINEERING AND SCIENCE, MID-APRIL 1995, VOl. 35, NO. 7

In the present treatment, we employ thin shell triangular elements for the cavity and linear shape functions for p . Accordingly, on each element, e, the pressure field is of the form:

3

(20)

where j corresponds to the three vertex nodes with qj being linear interpolation functions and pj being the nodal pressures. A Galerkin method is employed on Eq 16 such that the resulting governing equation for any interior node is given by

The above nonlinear system of equation can be solved by iteration.

The melt front tracking algorithm with fxed mesh associates a control volume with each node. The nodal fill factor, which is defined as the filled fraction of nodal control volume and varies from zero (empty node) to unity (completely filled node), is assigned to each node. Partially filled nodes, which have the fill factor greater than zero and less than unity, are considered as the melt flow front nodes. Pressure values are calculated at the completely filled nodes by solving the fiiite element equations. And the finite difference form is also used to solve Eq 18 for gap wise temperature distribution in the present study. The flow front is advanced at each filling time step by updating the nodal fill factor according to mass bal- ance between connected nodes. The procedure is re- peated until the entire mold is filled. The finite ele- ment/finite difference scheme can simulate the mold filling flow with nonisothermal conditions in very complex geometries ( 13, 14).

Fiber Orientation

I t is well known that gapwise variation in shearing and stretching deformation of fluid produces distinct orientation structure along the gapwise direction. The evolution equation for the second-order orientation tensor, Eq 2, in three dimensions produces a system of five coupled, nonlinear, differential equations. There are many numerical methods to solve these kinds of evolution equations (2, 10. 1 1 , 12). In the present study these are solved at the centroids of the same finite element mesh as is used for the mold filling analysis, based on its local coordinate system, as shown in Fig. 1 . This method is useful because the velocity gradient field is calculated based on the local coordinate system from the filling analysis, and the flowability constants in Eq. 15 are determined at the centroid of each element. Therefore, Eq 2 can be

rewritten as follows: d aij d aij d aij - + u- + u-

d t ax d y

Fiber orientation calculations are performed concur- rently with the mold filling simulation. The orienta- tion tensor components are calculated at each layer subdividing local part thickness, for each element in the mesh, by a fourth-order Runge-Kutta method for the time integration.

Equation 22 requires the velocity gradient at the centroid of element at each layer:

u. x u, y u, 2

u. 1-J . = [ 7 2 \2] (23)

However, the filling simulation provides only nodal pressures and shear velocity gradients, u, and u, z ,

from Eq 1 1 , at the centroid of element at each layer. In order to evaluate the spatial derivatives of velocity in x and y directions, one needs the nodal velocities. Because the velocity is proportional to the pressure gradient, one can find the velocities u and u using Eq 12 at the centroid of element at each layer. Then, nodal velocities are evaluated by a global averaging procedure, i.e., weighted by the associated subvolume of neighboring elements.

Equation 22 also requires the spatial gradients of orientation tensor. An ad hoc procedure was devel- oped to evaluate the spatial gradients of the orienta- tion tensor in view of the upwinding scheme. In Fig. 2, the directional derivatives of the orientation tensor components, based on the local coordinate system of element (e), between centroids of element ( e ) and element (e,) are given by

Fg. 2. 'Reatrnent of convection term forfiber orientation.

Numerical Simulation of Fiber Orientation

where fk is a unit vector of a vector (Axk;+ A y5). directed from the centroid of element ( e ) to that of element ( ek), k = 1.2. This is rewritten as follows:

., (25)

As far as the upwinding scheme is concerned, two elements for k = 1 and 2 are chosen from the upwind- ing direction associated with the element ( e), as indi- cated in Fig. 2. From the above two linear algebraic equations ( E q 25), one can find the spatial deriva- tives of orientation tensor, d a , /dx and d a,/d y. Here one needs the tensor transformation of orientation state of element (e,) to the local coordinate system of element (el. As a special case, when the plane of element ( e ) is different from that of element ( ek), it is assumed that the orientation state of fibers is pre- served across the folds. The assumption is that the orientation state in the bent plate is the same that of flat plate. For this assumption one needs an addi- tional tensor transformation which rotates the fibers by the angle 8 between the folds.

To start the numerical simulation it is necessary to specify the initial orientation of fibers at the gates. I t may be noted that the fiber orientation at the gate is usually not known. Thus the initial orientation state at the gate may be assigned as either a random distribution or an aligned distribution in the flow direction. As the melt front advances during the mold filling process, more elements will be included in the computational domain. In each time step, elements that include at least one fully filled vertex node are considered as new elements to be added for that step. Then, the initial orientation state at the new element is required for the subsequent simulation of fiber orientation. To be rigorous, one has to take into ac- count the fountain flow effect to find the initial orien- tation state at those new elements in the melt front. According to the fountain flow effect, the outer layers close to the mold wall are fed by fibers originating from the central layer, and the fiber from the central region will be reoriented while it moves from the central layer to the outer layer. However, this will require tremendously complicated three-dimensional analysis for three-dimensional thin cavity mold, which is beyond our capability at this moment. Therefore, several different schemes to deal with the initial orientation state were introduced in the pres- ent study: i) by convecting the orientation state at each layer from the neighboring element 0.e. by inte- grating Eq 22 from the state of the neighboring ele- ment to obtain the orientation state at the new ele- ment position), ii) by applying the convected fiber orientation at the central layer to all layers, and iii) by introducing random orientation to all layers except the central layer, where the convected fiber orienta- tion is applied. As will be discussed later, fiber orien- tation states obtained from numerical simulation based on the three different schemes were found to be indistinguishable from each other. This numerical

result indicates that the overall fiber orientation is not significantly affected by the fountain flow effect. Therefore any scheme seems to be good enough for the initial state at the melt front region.

The overall numerical scheme for the mold filling flow together with the fiber orientation is schemati- cally summarized in Fig. 3. The computer program developed in the present study can calculate the cou- pled flow field and the flow-induced fiber orientation in the injection molding of arbitrary thin three- dimensional geometries.

RESULTS OF NUMERICAL ANALYSIS

Numerical results for two different cavity geome- tries will be discussed in this section. The thermal properties and polymer melt viscosity of polypro- pylene used in the following numerical simulations are as follows:

density, specfic heat, thermal conductivity, coefficient of viscosity

coefficient of viscosity

coefficient of viscosity

coefficient of viscosity

p = 7.70 x lo2 kg/m3 C, = 3.46 X lo3 J/(kg.K)

k = 1.51 x lo-' W/(m.K)

function, B = 9.60 x Pa.s

function, n= 3.50X lo-'

function, c= 4 . 8 0 ~ 10-4

function, T~ = 5.90 x 103 K

Read the input data F Initialization P

-QL@ Is the mold full ?

No I

Relaxation pressure field and determine the time step

Find velocity gradlent field

R g . 3. Overall numerical scheme.

POLYMER ENGINEERING AND SCIENCE, MID-APRIL 7995, VOl. 35, NO. 7 609

S. T. Chung and T. H. Kwon

Numerical Results With a Tension Test Specimen

Numerical simulations have been camed out for a cavity geometry of a tension test specimen with a pin end gate. The dimensions of a tension test specimen of ASTM standard are as follows:

length over-all = 183 mm width over-all = 19 mm length of narrow section = 57 mm width of narrow section = 6 mm thickness = 2 mm

and the input data are as follows:

inlet flow rate, inlet melt temperature, mold wall temperature, the number of fibers

length of fiber, L= 1.0mm diameter of fiber, phenomenological coefficient,

Q = 5.0 x lo3 mm”/s T, = 200 “C T, = 30 “C

per unit volume, n = i/mm3

D = 0.04 mm C, = 0.00 1

F’igure 4 shows the finite element mesh of tension test specimen, which contains 520 elements, 327 nodes, and 14 grid points through the thickness di- rection. Figures 5a and b show the successive flow fronts and the velocity fields, respectively. when the cavity gets filled. One can see the diverging flow around the gate followed by parallel flow, converging-parallel-diverging and parallel flow. Figure 6 shows the orientation fields at the end of filling at four different layers when a random orientation is imposed at the gate as an initial condition and the initial orientation state at the new element in the melt front is obtained by convecting the orientation state at each layer from the neighboring element. Three eigenvectors with the magnitude of each eigen- value of the second-order orientation tensor are plot- ted to represent the orientation state. Figure 6a is the orientation distribution at the grid point 1, core layer (i.e., the center layer). Here the flow around the gate is extensional and fibers are seen to align in the direction perpendicular to the flow direction. Figures 6b and c are the orientation distribution at the grid points 3 and 5, at which the normalized half gap thickness equals to 0.2 and 0.4, respectively. Here the flow involves both shear and extensional effects. Fig ure 6d presents the orientation distribution at the grid point 13, skin layer (i.e. surface layer), at which the normalized half gap thickness equals to 0.9. It is clear that shear effects are dominant here with fibers tending to align in the flow direction. Figure 7 shows the longitudinal fiber orientation state at cross sec- tion A-A’ indicated in FLg. 4. Most of the fibers are aligned to the flow direction with a slight inclination toward the center layer. Figure 8 shows the trans- verse fiber orientation state at several cross sections indicated in Fig. 4. Here, one can see the complex fiber orientation state. But at a converging flow region

c... . ..

B ....

FYg. 4. Finite element mesh for a tension test specimen with 520 elements a n d 327 nodes.

cross section C-C’, fibers are inclined to the inward. While at a diverging flow region, cross section E-E’, fibers are inclined outward.

Eflect of the Concentration of Fibers

Figure 9 shows the effect of fiber volume fraction on the clamping force. This effect results from the calcu- lation of the coupled flow field in the present study. One can find that the clamping force increases asymptotically as the number of fibers increases as expected. In this particular case, when the fiber vol- ume fraction is greater than 3.1 %, n becomes larger than 1/( DL2) so that the average distance between fibers for aligned orientation state, haligned. is used for h, as explained in Eq 5. It may be mentioned that most of fibers are found to be aligned to the flow direction except near the gate according to numerical analysis results. Therefore, this ad hoc treatment for determining h seems to be acceptable.

Figure 1 0 shows the variation of the velocity com- ponents u and u across the gapwise direction at two

61 0 POLYMER ENGINEERING AND SCIENCE, MID-APRIL 1995, VOl. 35, NO. 7

Numerical Simulation of Fiber Orientation

different elements for two cases of fiber volume frac- tions. At the element 6 at which the random initial orientation state is imposed at every layer, the differ- ence of the magnitude of x and y directional velocity components decreases as the volume fraction in- creases. This tendency might be explained by the fact that the coupling effect, as indicated in Eq 12, b e comes stronger as the number of fibers per unit volume n, and thus N increases. At the element 20 near the gate, the velocity profile tends to be flattened as the volume fraction increases. It may be men- tioned that in the narrow section of the tension test specimen, the velocity profile remains almost un-

(a) (b) Fig. 5. Filling simulation resultsJor a tension test Specimen: (d jlling pattern fstep increment = 0.0464 s, $[ling time 0.8807 s) and (b) oefocityjeld.

changed with the volume fraction. The out-of-plane component of the fiber orientation tensor a, in the narrow region is almost zero at every layer (see Fig. 7). resulting in insignificant effect of fibers on the velocity field according to Eqs 12 and 13.

Figure I I shows the variation of the temperature across the gapwise direction at the element 20. One can find that the volume fraction of fibers has little effect on the variation of temperature.

Eflect of the Fiber-Fiber Interaction Coefficient C,

The dimensionless coefficient C, describes the strength of the randomizing effect of mechanical in-

Flg. 6. Fiber orientation state when the cavity is .filled.for a tension test specimen: fd at z = 0, (hi at z = 0.2b, (c) at z = 0.4b. and (d, at z = 0.9b.

Wall - - - -- - ----- -- --- -- --- --- -- - - -

_ _ _ _ - - - - - - - - - - - - - - - - - - - -

Center Layer Flg. 7. Longitudinal (on y - z planelftber orientation state with a view direction along x-direction-for a tension test specimen at cross section A-A indicated in Flg. 4.

POLYMER ENGINEERING AND SCIENCE, MID-APRIL 7995, VOI. 35, NO. 7 61 1

F

I / I t \ I I I I I t I

+ + + +

I + + + + E'

I / / / \ \ / / / \ \

B l %

Fig. 8. Transverse (on x-z p1ane)Jber orientation state with a view direction along y-direction for a tension test specimen at cross section B-B', C-C', D-D', E-E' and F-F' indicated in l%g. 4.

45:

5 - - 4 0 1 u 0:

$ 30 z

0 35

25

40 -

5 - u 0: 0 35-

z

20 , 0 5 10 15 20 25 30 35 40 45

FIBER VOLUME FRACTION ( % )

l%g. 9. Clamping force us. fiber volume fraction with a con stant inletgow rate, 5 . 0 ~ lo3 mm3/s, for a tension test specimen.

teraction between fibers. Figure 12 shows the fiber orientation state for C, = 0.05, which should be com- pared with Fig. 6 for C, = 0.001. It is clear that the fiber orientation state approaches the randomized orientation state as the magnitude of C, increases. The flow field is also affected by the magnitude of C,. Figure 13 shows the effect of C, on the clamping force with the fiber volume fraction, 3.1%, fixed. It is found

that the clamping force increases as the magnitude of C, increases because the extra stress term increases as the fiber orientation becomes more randomized.

In the case when the phenomenological coefficient C, is zero, a stable orientation state is observed near the center layer and is found to be almost the same as those in Fig. 6 with C, = 0.001. But near the surface layer one can observe an unstable oscillatory orientation state. For instance, Fig. 14a shows the transient behavior of orientation tensor component in the case when C, is zero at the element 292 indicated in Fig. 4 especially near the surface layer. In such a case, it is well known that in the simple shear flow, fiber oscillates with a constant period. This behavior depends on the fiber aspect ratio. Espe cially for fibers with an infinite aspect ratio, no oscil- latory behavior is observed (16). In Fig. 14b it can be seen that the small interaction coefficient C, of 0.001 has the effect of suppressing the oscillatory behavior.

Dependence on Initial Conditions a t the Gate

To investigate the effect of the initial orientation state at the gate, an aligned orientation state was intentionally introduced at the gate with the same input data as the case of Figs. 5 through 8. Figure 15 shows the transient fiber orientation states at the grid point 1, center layer, with the aligned initial condition at the gate as injection mold filling prc-

61 2 POLYMER ENGINEERfNG AND SCIENCE, MID-APRIL 1995, Vol. 35, NO. 7

Numerical Simulation of Fiber Orientation

1

m lJl W

h 0.8 0 I t-

9 0.6 0 LL a

0.4 n N

I 0.2

W

< P a:

0

1

v) v) W

$ 0.8

F 0

9 0.6 0 U _J

a I 0.4 n N

z 0.2 a

W

<

0

ELEMENT # 6

U (n = 1, 0.1%) +- U (n = 25, 3.1%) -+---

V (n = 1, 0.1%) -o-- V (n = 25, 3.1%) -I ..

~~ ~

0 10 20 30 40 50 60 70

VELOCITY( cm I sec )

(a)

ELEMENT # 20

. . 0 2 4 6 8 10 12 14 16 18

VELOCITY( cm / sec )

(b) Fg. 10. Variation of velocity components along the gapwise direction for a tension test specimen: (a) at element 6 and (b) at element 20 indicated in Fg. 4.

1

v) lJl W

0.8 0 F 4 0.6 0 U J

a 0.4

e a B

N

z 0.2 (r

0

ELEMENT # 20

I

t 20 40 60 80 100 120 140 160 180 200 -_I

TEMPERATURE( deg )

R g . 1 I . Variation of temperature along the gapwise direction for a tension test specimen at element 20 indicated in Fig. 4.

(a)

Fg. 12. Fiber orientation state with C, = 0.05 when the caw ity is fXed for a tension test specimen: (a) at z = 0, (b) at z = 0.2b. (c) at z = 0.4b and (d) at z = 0.9b.

36

241/ ' ' ' ' ' ' ' ' ' 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 0.11

INTERACTION COEFFICIENT ( CI )

Fg. 13. Clamping force us. Jber-tozfiber interaction coeflt cient C, with the Jber volume fraction, 3.1%. f l e d and a constant inletjlow rate, 5.0 X 1 O3 mm3/ s, for a tension test specimen.

POLYMER ENGINEERING AND SCIENCE, MID-APRIL 1995, Vol. 35, No. 7 61 3

S. T. Chung and T. H. Kwon

1

0.8

0.6

:= 0.4

0.2

ELEMENT#292 ( z - 085 b ) , CI - 0 001

- . - - ___ - - - - I ,

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4

TIME ( sec )

(bl Flg. 14. Effect of the dimenswnlessJberzfiber interaction c e efficient C, for a tension test specimen a t element 292 (z = 0.85b) indicated in Fig. 4: (a) C, = 0.0 and (b) C, = 0.001.

ceeds. At the initial stage, fibers around the gate are aligned to be perpendicular to the flow direction be- cause of the strong diverging flow, as shown in FUJ. 15a. But, as the melt front advances, fibers near the gate tend to be aligned with the initial orientation, as indicated in Figs. 1%-d This change of orientation of fibers near the gate as melt flow proceeds might be explained as follows: The diverging flow around the gate becomes weaker owing to the geometrical effect of the tension test specimen, and the convective transport from the gate becomes relatively dominant. It may be noted that fiber orientation away from the gate is not affected by the initial orientation. Also, it should be mentioned that the difference between the case with an aligned initial condition and that with the random initial condition at the gate is observed only at layers close to the center layer. At layers close to the wall, the magnitude of velocity is smaller than the center layer, so the shear flows become dominant

Fig. 15. Pans ien tmer orientation state a t core layer (z = 0) with aligned initial condition a t thegate: (d a t t = 0.081 s, (b) at t = 0.255 s. (c) at t = 0.557 s, and (d a t t = 0.881 s.

compared with the convective transport from the gate. One may recall that the fiber tends to be aligned with the velocity direction when the shear flow is domi- nant like in the skin layer. In this regard, the differ- ence between the two cases becomes smaller as the layer becomes closer to the wall.

Dependence on Initial Conditions at the Flow Front

To investigate the effect of the initial orientation state at the flow front upon the entire orientation state, two other schemes were adopted, to be com- pared with the scheme used to obtain Figs. 5 through 8 with the same input data as in Figs. 5 through 8. The first scheme is to obtain the initial orientation state by applying the convected fiber orientation at the central layer to all layers, resulting in J%J. 16, which shows the transient fiber orientation states at the skin layer, as injection mold filling proceeds. The second scheme applies the random orientation state to all layers except the central layer, where the con- vected fiber orientation is applied Figure 17, obtained from this scheme, is to be compared with Rg. 6d and

614 POLYMER ENGINEERING AND SCIENCE, MID-APRIL 1995, VOl. 35, NO. 7

Numerical Simulation of Fiber Orientation

\

(a)

\ \

FQ. 16. Transient fiber orientation state at skin layer (z = 0.9b) when the initial orientation state at the melt front is obtained by applying the convected fiber orientation at the central layer to all layers: (a) at t = 0.081 s. (b) at t = 0.255 s, (c) at t = 0.557 s, and (d) at t = 0.881 s.

Fig. 16. It can be easily seen that the entire orienta- tion state is almost the same except at the flow front region. I t should also be mentioned that the same trends were observed at the other layers.

Because of the fountain flow effect in the melt front, fibers from the central region are deposited to certain layers and subsequently stay at the same layers as the melt front advances. As far as the fiber deposited to a certain layer is concerned, the orienta- tion of the fiber just arriving at the layer should be regarded as an initial orientation with respect to the evolution equation of the orientation tensor. The fiber orientation state is subsequently governed by the velocity gradient field. The fact that the final fiber orientation over most of the domain is not signifi- cantly affected by the initial condition implies that the fibers get oriented by the velocity gradient field in a relatively short time period. In this respect, any scheme mentioned in the present study with regard to the fountain flow effect seems to be satisfactory.

7 I

Fig. 17. Transient fiber orientation state at skin layer (z =

0.9b) when random initial orientation is applied at theflow front element except the center layer, where the convected fiber orientation is applied: (a) at t = 0.081 s, (b) at t = 0.255 s, (c) at t = 0.557 s, and (d) at t = 0.881 s.

Three-Dimensional Example

Podium With a Square Hole

In order to demonstrate the capability of the pres- ent work for three-dimensional simulations, a three- dimensional thin cavity geometry of a podium with a square hole was chosen. The dimension of a podium are maximum width = 100 mm, height = 30 mm, width of a square hole = 20 mm, and thickness = 2 mm. The input data are as follows; inlet flow rate =

5.0 X lo4 mm3/s, inlet melt temperature = 200°C. mold wall temperature = 30"C, the number of fibers per unit volume = l/mm3, length of fiber = 1.0 mm, diameter of fiber = 0.04 mm, and phenomenological coefficient C, = 0.0 1. Figures 18 and 19 are numeri- cal simulation results for the filling pattern and fiber orientations, respectively. In Figs. 18 and 19, two graphical results from two different view directions are displayed to show the overall behavior results from two different view directions are displayed to show the overall behavior in this three-dimensional

POLYMER ENGINEERING A N D SCIENCE, MID-APRIL 1995, Vol. 35, N o . 7 61 5

S. T. Chung and T. H. Kwon

- rear view

front view Fg. 18. Filling pattern for a podium with a square hole for constant inletflow rate, 5.0 X 1 O4 mm3 / s (step increment =

0.031 s, Jlling time = 0.59 s).

cavity geometry. In Fig. 18, it is easily found that the weldline is formed behind the square hole. Figure 19 shows the fiber orientation state of the three different layers including the core and the skin layer. One should observe the complex orientation state along the weldline behind the obstacle. And the orientation state does not change across the edges since we assumed the preserved orientation state across the folds. This result is favorably compared with Frahan’s report (12) in the center layer. It may be noted that Frahan’s work has calculated the fiber orientation only at the center layer, as opposed to the present study.

Finally, it should be mentioned that there is a little difference between the present numerical predictions and the experimental studies (17-20). Probably, these differences might be attributed to the inexact consid-

rear view

front view

Fig. 19. Fiber orientation state for a podium with a square hole: (cd at 1 layer (z = O), (b) at 2 layer (z = 0.27b) and fc) at 7 layer (z = 0.96b).

eration of the fiber-fiber mechanical interactions. In the Folgar and Tucker’s phenomenological model (3) used in the present study, fiber-fiber interaction is proportional to the shear rate and the departure from the random orientation state. Thus the proportional constant C, is the important parameter, which should be determined by fitting numerical results to experi- mental data for each processing condition. In fact, C, could be variable depending on the suspension pa- rameters such as fiber orientation state, fiber aspect ratio, and fiber volume fraction. Recently, an a p proach has been suggested for relating C, with the average spacing between fibers, which depends on such suspension parameters (2 1). The evaluation of the average interfiber spacing requires numerical in- tegrations of the orientation distribution function $,

616 POLYMER ENGINEERING AND SCIENCE, MID-APRIL 1995, VO~. 35, NO. 7

Numerical Simulation of Fiber Orientation

rear view

(b) Fig. 19. Continued.

front view

which is, however, too cumbersome to be introduced in the injection molding simulation. The dependency of C, on the average interfiber spacing needs to be further investigated, and an easier method to deter- mine the average interfiber spacing is needed for practical application to the injection molding of short-fiber-reinforced thermoplastics. With such de- velopments in the future, the present modeling could be improved accordingly.

CONCLUSIONS

A numerical simulation method has been success- fully developed for predicting the three-dimensional fiber orientations of short-fiber-reinforced thermo- plastics during the injection molding filling process in three-dimensional thin cavity molds. The mold filling simulation via a FEM/FDM is based on a new pres sure field equation incorporating the additional stress

rear view

Fig. 19. Continued.

front view

due to the presence of fibers. A compact tensor repre- sentation is adopted that describes the fiber orienta- tion state at each grid point of every element centroid across the thickness of the part. A fourth-order Runge-Kutta method is used to solve the evolution equations of the second-order orientation tensor, us- ing an upwinding scheme for the convection terms. The calculation of the orientation tensor is based on the local coordinate system of each element. So, ten- sor transformation of the orientation tensor is re- quired between the neighboring elements in order to deal with convection terms. Since it is assumed that the fiber orientation state is preserved across the folds, additional tensor transformation is required to evaluate the convection terms when the melt flows across the fold in an arbitrary three-dimensional cav- ity space. In this way it is possible to predict the transient fiber orientation state in an arbitrary

POLYMER ENGINEERING AND SCIENCE, MID-APRIL 1995, Vol. 35, No. 7 61 7

S. T. Chung and T. H . Kwon

three-dimensional cavity geometry during the entire injection mold filling process.

REFERENCES

1. G. B. Jeffrey, Roc. R. Soc., A102. 161 (1922). 2. R. C. Givler, M. J. Crochet, and R. B. Pipes, J . Compos.

3. F. Folgar and C. L. Tucker, J. ReinJ Plast. Compos., 3.

4. S . M. Dinh and R. C. Armstrong, J. RheoL, 28. 207

5. G. K. Batchelor, J. Fluid Mech, 46. 813 (1971). 6. M. A. Bibbo, S. M. Dinh, and R. C. Armstrong, J. Rheol,

7. S. G. Advani and C. L. Tucker, J. RheoL, 31. 751 (1987). 8. M. C. Altan, S. G. Advani, S. I. Giiqeri, and R. B. Pipes, J .

9. S. G. Advani and C. L. Tucker, J. Rheol, 34, 367 (1990). 10. S. G. Advani and C. L. Tucker, Polyrn Cornpos., 11. 164

11. M. C. Altan, S. Subbiah, S. I . GiiGeri, and R. B. Pipes,

Mater., 17. 330 (1983).

98 ( 1984).

(1984).

29, 905 (1985).

RheoL, 33. 1129 (1989).

(1990).

Polym Eng. Sci., 30. 848 (1990).

12. H. Henry De Frahan, V. Verley, F. Dupret, and M. J.

13. C. A. Hieber and S. F. Shen, J. Non-Newt. Ruid Mech, 7 ,

14. V. W. Wang, C. A. Hieber, and K. K. Wang, SPE AN'l'EC

15. C. L. Tucker, J . Non-Newt. Fluid Mech, 39, 239 (1991). 16. S. Akbar and M. C. Altan, Polym Eng. Sci., 32. 810

Crochet, Potyrn Eng. Sci., 32, 254 (1992).

1 (1980).

Tech Papers, 31, 826 (1985).

( 1922). 17. M. W. Darlington and P. L. Mcginley, J. Mater. Sci., 10,

906 (1975). 18. M. W. Darlington, P. L. Mcginley and G. R. Smith, J ,

19. P. F. Bright, R. J. Crowson and M. J. Folkes, J. Mater.

20. M. J. Folkes and D. A. M. Russell, Polymer, 21, 1252

21. S. Ranganathan and S. G. Advani, J. RheoL, 35. 1499

Received May 1 7, 1993 Reuision received November 1993

Mater. Sci., 11, 877 (1976).

Sci., 13, 2497 (1978).

( 1980).

(1991).

61 8 POLYMER ENGlNEERlNGAND SCIENCE, MID-APRIL 1995, YO/. 35, NO. 7