Sie sind auf Seite 1von 22

Accepted Manuscript

Method to account for the fiber orientation of the initial charge on the fiber
orientation of finished part in compression molding simulation

Yuyang Song, Umesh Gandhi, Camilo Pérez, Tim Osswald, Srikar Vallury,
Anthony Yang

PII: S1359-835X(17)30204-X
DOI: http://dx.doi.org/10.1016/j.compositesa.2017.05.021
Reference: JCOMA 4673

To appear in: Composites: Part A

Received Date: 21 December 2016


Revised Date: 16 May 2017
Accepted Date: 18 May 2017

Please cite this article as: Song, Y., Gandhi, U., Pérez, C., Osswald, T., Vallury, S., Yang, A., Method to account
for the fiber orientation of the initial charge on the fiber orientation of finished part in compression molding
simulation, Composites: Part A (2017), doi: http://dx.doi.org/10.1016/j.compositesa.2017.05.021

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
PUBLIC 公開

Method to account for the fiber orientation of the initial charge on the fiber
orientation of finished part in compression molding simulation
Yuyang Song1*, Umesh Gandhi1, Camilo Pérez 2, Tim Osswald2, Srikar Vallury3, Anthony Yang3

1. Toyota Research Institute North America


2. Polymer Engineering Center, University of Wisconsin-Madison
3. Moldex3d Northern America Inc

Abstract
In the automotive industry, interest in compression molding of discontinuous fiber reinforced
thermoplastic is rapidly increasing. Compression molding offers fast cycle times, allowing us to
maintain longer fibers. Longer fibers translate into improved mechanical properties, and hence,
lighter parts. The mechanical properties of such components also largely depend on the fiber
orientation. It has been observed that the fiber orientation in the finished part depends on the
initial fiber orientation of the bulk charge being compressed. In this paper, we present methods to
measure initial fiber orientation in the bulk charge and demonstrate how it can be included in the
simulation. Computer tomography is used to scan the bulk charge to estimate the orientation of
the fibers in the charge. The estimated orientation through the charge is mapped as the initial
condition for the charge used in the compression molding simulation. We also demonstrate the
improvement in the predicted fiber orientation resulting from the proposed approach with a
physical example.

Keywords: Initial fiber orientation, Compression molding, Mapping

*Corresponding Author
Toyota Research Institute, NA, 1555 Woodridge, Ann Arbor, MI 48105

Email: Yuyang.song@toyota.com

1
PUBLIC 公開

1. Introduction
Compression molding is a method of plastic molding process in which the molding material,
generally preheated, is placed in an open, heated mold cavity that is suitable for molding
complex, high-strength objects [1-2]. With its short cycle time and high production rate,
compression molding is widely used in the automotive, aerospace, electronics, and sporting
goods industries to produce parts that are large, lightweight, high strength [3-4]. For the fiber
reinforced thermoplastic materials, the initial charge used can broadly be in two forms, bulk or
sheet. Bulk charge is desired to meet the high production volume and lower cost targets that are
typical for the automotive applications [5-6]. Sheet material consists of discontinuous fiber mats
with thermoplastic polymer embedded as copolymers or thin sheets. While such sheets do offer
longer fiber length and improved mechanical properties, the processing is still a challenge. In this
study, we will be focusing on the bulk material only.

Before compression molding, the bulk charge is made using the twin screw or single screw
extruders to compound of fibers and resin. After mixing, the bulk material is placed in the mold
cavity and compressed, as shown in Fig.1. The process offers a higher potential to maintain
longer fibers and better mechanical properties compared to injection molding, as the shear flow
rate through the cavity is quite limited, therefore causing less fiber breakage. The properties of
the finished part are highly dependent on the fiber condition in the part, i.e. fiber length, fiber
orientation and volume fraction distribution through the part. It has also been observed that the
fiber condition varies throughout the part which is a function of process parameters and
geometry of the part.

Fig.1 Typical compression molding process

In order to predict the mechanical properties of the final part, it is important to use the fiber
condition of the final part in considerations. Computation methods based on continuum
mechanics can be used to simulate the compression molding process and estimate the fiber
condition in the final part. Commercial software such as Moldex3D [7] and Moldflow [8] are
available for this purpose. While algorithms to estimate fiber length and fiber concentration are
still evolving [9,10] methods to estimate fiber orientation are well developed. In this study, we
will focus on improving the capability to predict the fiber orientation in the finished part.

2
PUBLIC 公開

Fiber orientation is usually characterized by a second-order orientation tensor. The orientation of


a single fiber can be described as a unitary vector p directed along the fiber axis, the orientation
can be defined by two angles (θ,φ) as illustrated in Fig.2[11].

Fig.2 Single fiber orientation definitions

Usually, there are thousands of fibers in the reinforced part, e.g. in 1 cubic mm, 30% glass fiber
(15-µm diameter) could mean 3000-4000 fibers. Therefore, using the orientation distribution
function ψ(p) is more efficient than taking into account every single fiber. The orientation
function represents the probability of a fiber to be in a certain orientation p. Advani and Tucker
[12] proposed a more general and concise expression, a second order tensor  , for the
representation of fiber orientation distribution using equation 1. Here, the unitary vector is
averaged over the selected volume. This method is widely used in the industry.

(1)
As this method explained, a high probability of fiber alignment in the specified principal
direction will be indicated by a value close to 1, whereas a low probability is indicated by a value
close to 0. A value of 1/3 in all principal direction means that fibers are oriented randomly. This
method of expressing fiber orientation is concise and easy to use with the numerical calculations
used to estimate mechanical properties of fiber reinforced polymers [13-14], it is also widely
adapted in the industry [7,8,15].

The most widely used computational approaches to simulate the compression molding is
Jeffery’s equation [16] with Folgar-Tucker model [17] to estimate the fiber orientations. This
model considers the material’s velocity gradients, strain rates, and fiber-fiber interaction. Using a
fiber orientation algorithm based on the Folgar-Tucker model, Chih-Chung Hsu [18] predicted
thoroughly the fiber flow and fiber orientation after compression molding simulation.

The traditional approach has been to assume that orientation of fibers in the initial compression
charge is random. However, due to the plasticization process in the extruder, a distinctive non-
random fiber orientation emerges within the charge [19]. For an accurate prediction of fiber
orientation in the finished part, it is important to start with an accurate description of initial
orientations. Until now, there has been no method to include the effect of initial fiber orientation
in the bulk charge for the compression molding process. The main purpose of this research is to
develop a method to account for the initial orientation in the compression charge and study the
effect of initial fiber orientation on the final fiber orientation of the compression molded
thermoplastic parts.

3
PUBLIC 公開

2. Development method and steps


2.1. Overall development steps
The overall developed methodology and steps are shown in Fig.3. The methodology includes
two steps. In step 1, a part of the bulk charge or the whole charge is placed in the micro-
computerized tomography (CT) machine. After the µCT scan, the fiber information is stored and
stacked in a series of images. Volume-graphics, an image processing software is used to analyze
the scanned image [15]. In volume-graphics, the scanned image volume is divided into a number
of small segments and then the fiber orientation tensor is estimated for each segment based on
the number of voxels representing fiber in that segment. The detailed steps are discussed in the
following part.

Fig.3 Fiber orientation measurement from test step 1

In Step 2, as Fig.4 shows, the measured orientations are mapped on the finite element model of
the charge. Depending on the shape, some scaling might be required. To help this process, an
algorithm, which maps the estimated fiber orientation measurement to each element of the finite
element model for the charge, has been developed in-house using Matlab [20]. After the mapping
process, the fiber orientation in the initial bulk charge model is defined as initial fiber
orientation. Finally, the compression molding simulation with the defined fiber orientation is
carried out using Moldex3D.

Fig.4 Fiber orientation mapping to the charge in compression molding step 2

4
PUBLIC 公開

2.2. Mapping algorithm


In the following paragraph, the detailed mapping algorithm will be explained. In a first step, the
fiber orientation distributions obtained by the µCT scans are normalized as depicted in Fig.5.

Fig.5 Mapping of µCT data set, Step 1: Normalization

Based on the mesh of the charge defined in Moldex3D, the normalized µCT data set is rescaled
to the dimensions of the mesh. The rescaling continues until the µCT data set comprises the
entire mesh of the charge. This approach leads to points lying outside of the mesh as shown in
Fig.6. In the end, all points outside the mesh are removed from the data set. The dimensions
illustrated here are just used for the explanations of the coding algorithm. For the simplification
of the validation purpose of the code, the part designed for this study is a rectangle bar.

Fig.6 Mapping of µCT data set, Step 2: Removal of overlapping nodes

In order to have a local orientation for every point of the Moldex3D mesh of the charge, an
interpolation algorithm is applied to the orientation tensor at each node of the mesh. The radial
basis function (RBF)[21-24] is a mesh-less interpolation technique and has been shown to be a
powerful method for representing sparse, scattered data. The RBF interpolation represents
interpolated data as a linear combination of the non-linear basis function that is defined at each
control point and has only a radial component

5
PUBLIC 公開

  =
  | −  | =
    (2)
   

where  represents the location of the point where interpolation is performed,  represents the
position of the control point in a specified region of interest,  is the basis function,  is the
weight factor to be determined and N is the number of data points. The distance  is defined
between the point where the function will be interpolated to and the points where the function is
known from the µCT data set. For the basis function,  , the thin plate spline (TPS)[25] was used
because it gives accurate results for scattered data approximations

 | −  | =    =     (3)

The first step in the interpolation is the solving for the weight factors  . A system of equations is
calculated, denoted as

 ⋯   
 ⋮ ⋱ ⋮   …  = … (4)
 ⋯   

where  describes the distances between all possible combinations of points within the µCT data
set. The  vector on the right hand side contains the target values and, in the case of the fiber
orientation distribution,  is one component of the orientation tensor  with

 ⋯    


 ⋮ ⋱ ⋮  … =  … 
 ⋯    
(5)

From the µCT data set, all  and the corresponding  are known. Thus, the system can be
solved for the weight factors. Solving

 r =
    (6)
 

for an arbitrary point of the Moldex3D mesh, the orientation tensor  of that point can be
interpolated. This is done for all nodes of the Moldex3D mesh and components of the orientation
tensor  .After the interpolation, the mapped fiber orientation and its corresponding coordinates
are imported into Moldex3D. The mapping tool was programmed with Matlab[26].

3. Demonstrated effect of initial orientation in the compression charge


In this section, we will demonstrate the benefit of defining the initial orientation for the
compression charge for the simulation. The application of the proposed method in section 2 is
demonstrated for glass fiber reinforced polypropylene, hat section part.

6
PUBLIC 公開

3.1. Fabrication of the part using compression molding process


The typical fabrication process for compression molding used is direct long fiber
thermoplastic(D-LFT). In a typical D-LFT process, depicted in Fig.7, the resin is fed into an
extruder, commonly a twin-screw extruder, where it is melted and mixed together with additives,
pigments, and fillers to obtain a homogeneous material, which is then mixed with continuous
glass fiber strands. This mixture is extruded and cut into a charge with the form of a log, and
then transferred to a mold where it is compressed into the desired shape of the final part [3,27].

Figure 7. Schematic of the typical standard D-LFT process [20,27].

A hat section part is designed and manufactured at a Magna facility in Toronto using the D-LFT
process. The material used for the prototyping is 40% glass fiber reinforced polypropylene
(40GFPP). Fig.8 shows the images for charge located in the tool.

Fig.8 The location of charge in the cavity

Table.1 shows the processing condition used for the part fabrication. The finished sample is presented in
Fig.9.
Process conditions Actual values

Melt Temperature 210 oC

Mold Temperature 70 oC

Compression time 2 secs

Compression pressure 2000KN

Charge weight 170gm

Table.1 Compression molding parameters used for the simulation

7
PUBLIC 公開

Fig.9 Sample made after compression molding, A, B and C are fiber orientation measurement locations

3.2. Fiber orientation measurement for part and charge


3.2.1. Fiber orientation measurement of the part at selected location
The fiber orientation distribution within a fiber reinforced thermoplastic was measured using
µCT. The10x10 mm samples are cut at selected location as shown in Fig.9, and scanned using
industrial CT scanner at Micro-photonics Inc. The resolution of the µCT scanning is 3µ m. It
took about 9 hours to scan one sample. The orientation tensor is estimated using previously
developed method by U. Gandhi [13,28]. In this method, the volume of the scanned sample data
is divided into smaller cells. The fibers are reconstructed in each cell and then the center line of
the fiber is extracted. The end point of the center line is used to estimate the fiber orientation.
Figure.10 shows the scanned section and smaller cells used to reconstruct the fibers. Figure.10b
show 3D image of fiber in a typical cell.

a. µ CT scanned image and selected cells through thickness b. 3D structure of the fibers inside selected cell
Fig.10 Fiber orientation tensor measurement from the µ CT scanning images of a sample cut from part

8
PUBLIC 公開

3.2.2. Fiber orientation measurement of the charge


The initial charge from twin screw used for CT scanning is presented in Fig 11. Due to the nature
of the flow, it is expected that there is periodicity along the length direction for the charge. So
only a 50 mm length of the charge, which comes to be 80x50x20mm is scanned and analyzed.
Also due to the large volume of the charge during µCT scan, a resolution of 50 µ m is used. Since
glass fibers are only 15 µ m in diameter the higher resolution will only capture fiber bundles.
With the D-LFT process, there are lots of fiber bundles. Also in the previous work by Perez [20],
we have shown that the fiber orientation distribution in the charge based on the fiber bundles is
similar to actual fibers. Therefore, the orientation map of fiber bundle is considered acceptable
surrogate for the fiber orientation.

Fig.11 shows the actual scaling process used to map the estimated orientation from the
measurement to the actual finite element model used as initial charge for the compression
molding simulation. Getting CT scan of an actual charge and creating map of orientation is quite
complex and time-consuming process, due to the large size of data (~80GB). To address this
challenge, a smaller charge from similar twin screw was CT scanned and a map of orientation
distribution was created. Next, the orientation from the smaller charge was mapped on the larger
actual charge using radial basis function (RBF) approach. We recognize that this is an
approximation, which is practical and offers improvement over current method assuming the
fibers to be completely random or aligned in only one direction. The details about the
compression molding simulation set-up and boundary conditions are in the next section.

Fig.11 Initial fiber orientation mapping from the real charge to simulation

3.3. Fiber length measurement


To determine the fiber length distribution of the sample, the fibers are extracted from the
composite by burn off test. The glass fibers are analyzed by optical image analysis method [29].
As shown in Fig.12, the samples are cut at location 1 and location 2 with the size of 20x20x2mm.
The frequency distribution shows that more than 60% of the fibers are longer than10 mm, and

9
PUBLIC 公開

the average fiber length by weight is about 8mm. The fibers in the compression molded charge
are highly bundled. Previous work by Perez [20] has reported average fiber length the
compression charge at around 15 mm.

Fig.12 Fiber length distribution from compression molded part at location 1 and location 2

3.4. Compression molding simulation of the part


The Compression Molding (CM) process has several main stages, includes preheat, compression,
cooling, and warpage. Moldex3D is used as the main software for the compression molding
simulation. The true 3D finite volume method (FVM)-based algorithm to simulate the process
has been developed in this company [30]. The compression action is modeled by novel moving
mesh boundary technique, where the computational domain will dynamically deform with time.
Through the whole analysis, the polymer melt is assumed to behave as Generalized Newtonian
Fluid (GNF) with compressibility. Hence the non-isothermal 3D flow motion can be
mathematically described by the following conservation laws for mass, momentum, and energy:
∂ρ
+ ∇ ⋅ ( ρv ) = 0
∂t (7)
∂ρ
( ρv ) + ∇ ⋅ ( ρvv ) − ∇ ⋅ (− pI + τ ) = 0
∂t (8)

 ∂T 
ρC P  + u ⋅ ∇ T  = ∇ (k ∇ T ) + Φ
 ∂t  (9)

C
where vis velocity vector, p is pressure, T is temperature, τ is the stress tensor, and ρ , η, p and k
represents density, viscosity, heat capacity, and thermal conductivity of the plastic material,
respectively.

10
PUBLIC 公開

A volume fraction function f is introduced to track the evolution of the melt front. Where f = 0
is defined as the air phase, f = 1 as the melt phase and the melt front is located within control
volume between 1 and 0. The advancement of f over time is described by the following
transport equation:

f + ∇⋅ fv = 0
∂t (10)

HRIC (High-Resolution Interface Capturing scheme) differencing method [31] is applied to


avoid interface being blurred.

The collocated cell-centered finite volume method proposed by Chang and Yang [32] was
modified to discrete the governing equations (6) -(8). The numerical method is basically a
SIMPLE-like finite volume method with improved numerical stability and convergence for non-
isothermal and non-Newtonian flow. Pressure, velocity, and temperature fields are segregated in
the solver so that the efficiency can be achieved even for three-dimensional calculation. Special
numerical techniques are incorporated in the flow solver so that the numerical convergence and
accuracy can be achieved for the highly non-linear system.

Moldex3D developed the iARD-RPR model [33,34] to predict the fiber orientation change over
time during the filling and compression phase. This is an improvement over the Anisotropic
Rotary Diffusion (ARD) model [16]. The RPR model is indicative of Retarding Principal Rate to
slow down the fast transient orientation rate. iARD-RPR model uses 3 parameters to define fiber
orientations:

 =A
 +A  
A HD IARD (CI , CM ) + ARPR (α ) (11)

Ci: Fiber-Fiber interaction parameter, Cm: Fiber-Matrix interaction parameter

Alpha: Slow down transient fiber response

The iARD-RPR model is used in this compression molding simulation to predict fiber
orientation at the final part.

The geometry of the model used for simulation is presented at Fig.13 and the detail of the model
is listed as well. The total cavity volume is 167cc, and the charge volume is 193cc (about 115%
of the cavity size). The detailed material properties used for the simulation is shown in Table.2.
The process condition used for the simulation followed the same as the manufacturing process
such as compression time, melt temperature, etc. Fiber orientation tensors are extracted after
filling process at locations from point A, B and C—which is exact the same location used from
the actual finished part, showing at Fig.14.

11
PUBLIC 公開

Element type used: tetra and hex mesh

Total element number: 2,501,320

Total node number: 2,598,806

Fig.13 Basic model geometry information for compression molding

Glass fiber Polypropylene

Density 2.55(g/cc) 0.92(g/cc)

Fiber weight percentage 40% N/A

Young’s modulus E1 70,000 Mpa 2,000 Mpa

Young’s modulus E2 70,000 Mpa 2,000 Mpa

Poisson’s Ratio 0.2 0.3

Table.2 Material properties used for the simulation

Fig.14 Point A, B and C for fiber orientation measurement

3.5. Validation of the mapping algorithm


To validate the mapping algorithm, a series of mapping is conducted for the charge we defined
for the compression molding process. As Fig.15 shows, two scenarios are proposed and executed.
First, using unidirectional fiber orientation in the charge; second, using random fiber orientation
in the charge. In these two scenarios, each case has two methods of defining initial fiber
orientation. At one hand, the initial fiber orientation is defined in Moldex3D interface internally,
on the other hand, the initial fiber orientation is defined using the mapping algorithm. At the end,
the final fiber orientation is compared between these two approaches for both of the scenarios.
For the comparison of the final fiber orientation results in point B, the difference between the
mapped initial fiber orientation and non-mapped fiber orientation is minor. From Fig.16, this
result also shows that the final fiber orientation tensor A11 is almost random. It is also showing

12
PUBLIC 公開

in Fig.16, that the fiber orientation in point B has two different results between these two
scenarios. The case when initial fiber orientation is unidirectional shows higher fiber orientation
results than the case when the initial fiber orientation is random. After this comparison, the
validation of the Matlab code is completed, the validation process builds up confidence for the
next part for the real fiber orientation mapping.

Fig.15 Validation process for the fiber orientation tensor results between the mapped process and non-mapped
process

Fig.16 Fiber orientation tensor results between the mapped process and non-mapped process

4. Results and discussion


Three different initial fiber conditions are used for the initial charge and compression molding is
simulated using Moldex3D. The results are compared with the fiber orientation measurements
on the actual part. The three initial conditions used in the simulation (CAE) are as follows,

13
PUBLIC 公開

• CAE initial random: All fibers in the initial charge are random
• CAE initial mapped: Fiber orientation in the initial charge is mapped from the measured data
using the proposed method.
• CAE initial aligned in A11 direction: All fibers are assumed to be aligned along the length of
part
4.1. Fiber orientation tensor distribution comparison

Fig.17 Comparison of the fiber orientation tenor(A11) for Point B by using different CAE input with measurement

The predicted fiber orientation (CAE) in the finished part at location B, for different initial fiber
conditions, is compared in Fig.17. The results, based on the visual comparison, show that when
the initial fiber orientations are mapped, the prediction is closer to the actual measurement, while
other parameters such as fiber to fiber interaction parameter (Ci) remain the same at 0.01.

Fiber interaction coefficient (Ci) is an empirical constant that characterize the effect of fiber-
fiber interaction in the concentrated suspensions [7,8]. Variation in the Ci will result in different
level of fiber to fiber interaction and hence different fiber orientation tensor distribution in the
part. One of the approach used to improve simulation, is to change this empirical value of Ci
between 0 to 1, until a good match between the measured and estimated fiber orientation is
obtained. To understand the effect of Ci on the predicted orientation tensor, we also estimated
fiber orientations for two additional values of Ci i.e. 0.005 and 0.02

The fiber orientation tensor components A11, A22 and A33 for the three selected locations A, B
and C at different Ci are presented in Figure 18, 19 and 20. The improvement of using the initial
fiber orientation from measurement is quite evident from Location B and C, see fig 19 (a) and 20
(a). We observe a good match not only in terms of value but shape of the fiber orientation tensor
through the thickness. We believe that this is result of using initial fiber orientation based on the
measured values. Physics of the process also supports this claim. Typically, when resin flows
the fiber tends to align along the flow in a laminar flow pattern. The time or length of flow to
align all the fibers depend on the viscosity and other flow parameters. In our case of compression

14
PUBLIC 公開

molding, the time for fibers to align is very small and therefore the initial fibers in the charge are
playing important role in how the fibers are distributed in the charge.

For the clarity, the case when initial fibers are aligned along the length direction of the hat
section part is not included in these plots.

(a) (b)

(c)

Fig.18 Comparison of the fiber orientation results (A11, A22, A33) from point A at different Ci

15
PUBLIC 公開

(a) (b)

(c)

Fig.19 Comparison of the fiber orientation results (A11, A22, A33) from point B at different Ci

16
PUBLIC 公開

(a) (b)

(c)

Fig.20 Comparison of the fiber orientation results (A11, A22, A33) from point C at different Ci

4.2. Calculated tensile modulus comparison


The fiber orientation components compared from figure 18 to 20 are plots, which can be
compared visually. Further to compare the effect of fiber orientations numerically, we estimated
the elastic modulus for the composite material using glass and matrix material properties and
fiber micro structure such as volume fraction and fiber orientation tensor. Since all the variables
other than the fiber orientation are assumed to be same, the elastic modulus of composite
material, which is a numeric value, represents average of orientation tensor component in the
selected direction.

The tensile modulus for A11 direction is estimated for 40GF Polypropylene material and
estimated orientation tensor using Mori-Tanaka homogenization [35] based on commercial
software Digimat [36], see Table 3. The modulus calculated are based on the fiber orientation

17
PUBLIC 公開

obtained by using Ci=0.01 for the baseline condition. We have also included the measured value
of elastic modulus at location B and C, as reference. The comparison shows that the predicted
elastic modulus from CAE using initial mapped tensor get closer match to the modulus from
measured modulus using MTS machine.

CAE CAE Modulus


Analytic modulus measured using
E11 Modulus(MPa) Initial Initial calculated from MTS machine
Locations random mapped measured tensors (for reference)

A 5108 6544 6100 N/A

B 5832 5460 5474 5259

C 6364 5807 5905 5202

Table.3 Tensile modulus calculated at three locations for different fiber orientations assigned

Fig.21 Comparison of the calculated E11and measurement for these three locations and the difference between them

5. Conclusion
In summary, the effect of initial fiber orientation of the charge on the final fiber orientation in the
finished part in compression molding is investigated. The results show that for compression
molding, the initial fiber orientation in the charge is an important microstructure detail that must
be considered. A Matlab code is developed to map the fiber orientation measured from µCT scan
18
PUBLIC 公開

to the bulk charge used for compression molding, using radial basis function (RBF). The
compression molding simulation results show that by accounting for initial fiber orientation
mapped from the measurement, the predicted fiber orientation (A11, A22, A33) in the finished
part shows a closer agreement with the measured values.

Therefore, based on this study, we strongly recommend the use of initial fiber orientation for the
charge during the compression molding simulation. We do recognize that measurement of such
charges can be quite time consuming and expensive. One approach to address this may be the
development of a database of generic initial orientations for a given extruder for a range of
processing conditions. Also knowing repeatability of conditions in the charge the size of the
charge that needs to be scanned can be reduced.

Acknowledgements

The authors gratefully acknowledge the support provided by Magana Canada for the materials
/tool to manufacture the part, also for the support from Moldex3D Northern America, Inc. for the
technical support on CAE simulations.

Reference
1. Isayev, A. Injection and compression molding fundamentals. CRC Press, 1987.
2. Beardmore, P., et al. Fiber-reinforced composites- Engineered structural
materials. Science 1980; 208:832-840.
3. Biron, Michel. Thermoplastics and thermoplastic composites. William Andrew, 2012.
4. Ebewele, R. O. Polymer science and technology. Boca Raton, FL: CRC Press,2000.
5. Beardmore, P., & Johnson, C. F. The potential for composites in structural automotive
applications. Composites science and Technology,1986;26(4):251-281.
6. Schuh,T.G. Renewable materials for automotive applications. Daimler-Chrysler AG,
Stuttgart,1999.
7. Manual of Moldex3D software, CoreTech. Inc, Taiwan,2014.
8. Manual of Moldflow software, Autodesk.Inc. New York, USA,2014.
9. Hine, P. J., Lusti, H. R., & Gusev, A. Numerical simulation of the effects of volume fraction,
aspect ratio and fiber length distribution on the elastic and thermoplastic properties of short
fiber composites. Composites science and technology, 2002;62(10):1445-1453.
10. Londoño-Hurtado, A., Osswald, T. A., & Hernandez-Ortíz, J. P. Modeling the behavior of
fiber suspensions in the molding of polymer composites. Journal of Reinforced Plastics and
Composites, 2011;30(9):781-790.

19
PUBLIC 公開

11. T.D.Papathanasiou and D.C.Guell, Flow-Induced alignment in composite materials,


Woodhead publishing limited, 1997.
12. J. H. Phelps and C. L. Tucker III, An anisotropic rotary diffusion model for fiber orientation
in short- and long-fiber thermoplastics, Journal of Non-Newtonian Fluid Mechanics,
2009;156(3):165–176.
13. Gandhi, U., Kunc, V., & Song, Y. Method to measure the orientation of discontinuous fiber
embedded in the polymer matrix from computerized tomography scan data. Journal of
Thermoplastic Composite Materials,2016;29 (12):1696-1709.
14. A.Bernasconi, F. Cosmi, and D. Dreossi, Local anisotropy analysis of injection molded fiber
reinforced polymer composites, Composites Science and Technology,2008;68(12):2574–
2581.
15. Manual of Volume Graphics software, Volume Graphics GmbH.2014.
16. Jeffery, G. B. The motion of ellipsoidal particles immersed in a viscous fluid. Proceedings of
the Royal Society of London A: Mathematical, Physical and Engineering Sciences,
1922;102(715):161-179.
17. S. G. Advani and C. L. Tucker III, The Use of Tensors to Describe and Predict Fiber
Orientation in Short Fiber Composites, Journal of Rheology, 1987;31(8):751–784.
18. Hsu, C. C., Chiu, H. S., Huang, C. T., & Chang, R. Y. Through fiber orientation investigation
to visualize compression molding. Proceedings of The 29th International Conference of the
Polymer Processing Society-Conference Papers. AIP Publishing 2014;1593(1): 619-622.
19. Advani, S. G., & Tucker, C. L. A numerical simulation of short fiber orientation in
compression molding. Polymer composites,1990: 11(3):164-173.
20. Perez, C., Osswald, T. A., & Steffen Ropers. Fiber Length and Orientation Distribution
Measurement of a Charge for D-LFT Compression Molding, ANTEC, 2013, Cincinnati, OH.
21. Park, J., & Sandberg, I. W. Universal approximation using radial-basis-function
networks. Neural computation, 1991;3(2): 246-257.
22. Lopez-Gomez, I., Estrada, O., and Osswald, T.A., Modeling and Simulation of Polymer
Processing Using the Radial Functions Method, Journal of Plastics Technology., 2007;(3), 2.
23. Ramírez, D., López, I., Estrada, O., and Osswald, T.A., Simulation of the Fountain Flow
Effect by Means of the Radial Functions Method (RFM), SPE ANTEC, Orlando, Florida,
2012.
24. Osswald, T.A., and Hernandez, J.P, Polymer Processing - Modeling and Simulation, Chapter
11 - Radial Functions Method, Hanser Verlag, München, 2006.
25. Stayton, C. T. Application of thin-plate spline transformations to finite element
models. Evolution,2009; 63(5):1348-1355.
26. Manual of Matlab software, Mathworks.Inc, New York, 2014.
27. Schemme, M. LFT–development status and perspectives. Reinforced Plastics, 2008;52(1),
32-39.
28. U.Gandhi, Sebastian DeBoodt, V. Kunc, Investigation of fiber orientation in fiber reinforced
polymers, Digimat User Meeting, Munich, Germany,2011.

20
PUBLIC 公開

29. Weber, Charles, and Chris Rannenberg. Method of determining an average length of
reinforcing fiber in a sample of reinforcing fibers. U.S. Patent No. 6,925,857. 9 Aug. 2005.
30. L.Liu, W. H. Yang, D. Hsu and V.Yang, ANTEC 02, San Francisco, CA, 5–9 May 2002.
31. Muzaferija S., Peric M., Sames P., Schelin T., Proc. Twenty-Second Symposium on Naval
Hydrodynamics, Washington, D.C, Aug.14,1998,
32. R. Y. Chang and W. H. Yang, Numerical simulation of mold filling in injection molding
using a three-dimensional finite volume approach. Int. J. Numerical Methods Fluids,2001;
37:125.
33. T.-H. Le, P.J.J. Dumont, L. Orgéas, D. Favier, L. Salvo, E. Boller, X-ray phase contrast
microtomography for the analysis of the fibrous microstructure of SMC composites,
Composites Part A,2008;39(1):91–103.
34. Huan-Chang Tseng, Yuan-Jung Chang, Chia-Hsiang Hsu. Prediction of Fiber Microstructure
for Injection Molding: Orientation, Degradation and Concentration, SPE-ACCE
Conference,2014, Novi, Mi.
35. T. Mori and K. Tanaka. Average stress in the matrix and average elastic energy of materials
with misfitting inclusions. Acta Metall. Mater.,1973;21:571–574.
36. Digimat User Manual, MSC Software corporation, Newport Beach, USA, 2015.

21

Das könnte Ihnen auch gefallen