bartus 2003_thesis

152
i LONG-FIBER-REINFORCED THERMOPLASTIC: PROCESS MODELING AND RESISTANCE TO BLUNT OBJECT IMPACT by SHANE D. BARTUS A THESIS Submitted to the graduate faculty of the University of Alabama at Birmingham, in partial fulfillment of the requirements for the degree of Master of Science BIRMINGHAM, ALABAMA 2003

Upload: shane-bartus-phd

Post on 27-Jul-2015

35 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Bartus 2003_Thesis

i

LONG-FIBER-REINFORCED THERMOPLASTIC: PROCESS MODELING AND RESISTANCE TO BLUNT OBJECT IMPACT

by

SHANE D. BARTUS

A THESIS

Submitted to the graduate faculty of the University of Alabama at Birmingham, in partial fulfillment of the requirements for the degree of

Master of Science

BIRMINGHAM, ALABAMA

2003

Page 2: Bartus 2003_Thesis

ii

ABSTRACT OF THESIS GRADUATE SCHOOL, UNIVERSITY OF ALABAMA AT BIRMINGHAM

Degree: M.S.Mt.E Program Materials Engineering

Name of Candidate Shane D. Bartus

Committee Chair Uday K. Vaidya

Title Long-Fiber-Reinforced Thermoplastic: Process Modeling and Resistance to Blunt

Object Impact

The use of thermoplastic composites has steadily gained favor over such

traditional materials as steel in structural and semi-structural applications due to their

prominent physical and mechanical behavior, such as specific strength, damping,

corrosion resistance, and impact properties. Moreover, closed-molded discontinuous

long-fiber-reinforced thermoplastic composites (LFTs) share the attractive features of

greater strength, stiffness, and impact properties (in contrast to short-fiber-reinforced

thermoplastics), in addition to high volume processability, ability to fill complex

geometries, intrinsic recyclability, and the capacity for part integration. In this work,

three broad aspects in regard to the production and performance of LFTs were studied: I.

The effect of processing conditions and material properties on the extrusion/compression-

molding process used in the manufacture of LFTs through a simulation matrix performed

in Cadpress-TP, II. Damage tolerance of LFTs subjected to transverse blunt object impact

(BOI), treated from an experimental standpoint, in order to characterize energy

dissipation and damage modes and III. Fiber orientation of LFT, predicted from process

simulation, and its relationship to the failure mode under BOI. Through the simulation

matrix, it was determined for the flat plaque geometry investigated, that the mold

temperature, charge location, and melt viscosity had the greatest effect in the study on the

Page 3: Bartus 2003_Thesis

iii

component processing and final material properties, thereby making those parameters of

greatest importance to accurately model the phenomenon. LFT subjected to BOI

exhibited high impact energy dissipation, which increased linearly with increasing areal

density. The average impact energy dissipation at the critical velocity was 167 J and 121

J for a 4.61 g cm-2 specimen impacted by flat and conically shaped projectiles,

respectively. The fiber orientation also played a large role in energy dissipation; failure

appeared to occur along planes of preferential fiber orientation. The fracture paths

correlated well with the predicted fiber orientation in a Cadpress simulation. Impact

mass did not exhibit any appreciable effects on energy dissipation. This overall work

advances the state-of-the-art in LFTs with an automotive focus.

Page 4: Bartus 2003_Thesis

iv

ACKNOWLEDGEMENTS

It is difficult to overstate my eternal gratitude to my advisor, Dr. Uday K. Vaidya,

who encouraged my research while I worked for him as an undergraduate at my alma

mater and then granted me the opportunity to continue working under him as a graduate

student. He has gone far beyond what is required of an advisor and been a friend, as

well. Dr. Vaidya’s ardent interest in the advancement of composite materials motivates

our entire group. His patience and advice have been unfaltering since I first began work

with him almost four years ago.

This work reflects the contributions of many individuals. I thank my esteemed

committee members, Drs. Gregg M. Janowski, Krishan Chawla, and Klaus Gleich, for

their valuable time and effort. Their input and guidance provided invaluable

contributions to the quality research. In addition, I thank the individuals of our research

group, Abhay Raj Singh Guatam, Selvum Pillay, Juan Camilo Serrano, Chad Ulven,

Haibin Ning, Francis Samalot, Rajan Sriram, and Tujuana Shaw. I also thank Saulius

Drukteinis, David Downs, Joseph Puckett, Andrea Rossillon, Sean Boyle, and Paulo

Coelho for their friendship and support over the last two years.

Finally, I thank my family and friends for their encouragement and support during

this time. Their understanding and acceptance allowed me the freedom to pursue this

research, which would not have been possible without them.

Page 5: Bartus 2003_Thesis

v

TABLE OF CONTENTS

Page

ABSTRACT....................................................................................................................... iii

ACKNOWLEDGEMENTS............................................................................................... iv

LIST OF TABLES........................................................................................................... viii

LIST OF FIGURES ........................................................................................................... ix

INTRODUCTION ...............................................................................................................1

Thermoplastic Composites.......................................................................................1

OBJECTIVE ........................................................................................................................4

LITERATURE REVIEW ....................................................................................................5

Discontinuous Long-Fiber Reinforcement ..............................................................5 Material Property Models ....................................................................................... 7

Material Properties...................................................................................... 8 Modeling of tensile strength ........................................................................9 Modeling of tensile modulus .....................................................................11 Cottrell impact model ................................................................................14

CADPRESS-TP BACKGROUND ....................................................................................20

EXPERIMENTAL PROCEDURE: PROCESS MODELING ..........................................28

Simulation Matrix ..................................................................................................28 Process Variables ...................................................................................................29

Mold temperature, Simulations 1-3 ...........................................................31 Charge location, Simulations 4-8...............................................................31 Boundary conditions ..................................................................................33 Material Parameters ...................................................................................33 Fiber interaction coefficient, Simulations 9 and 10 ...................................35 Melt viscosity, Simulations 11-17 .............................................................35 PvT parameters, Simulations 18 and 19 ....................................................36 Simulation verification, Simulation 20 ......................................................37

Page 6: Bartus 2003_Thesis

vi

TABLE OF CONTENTS (Continued)

Page

RESULTS AND DISCUSSION: PROCESS MODELING ..............................................38

Process Parameter Results, Simulations 1-8..........................................................38 Mold temperature effects, Simulation 1-3 .................................................38 Effect of charge location, Simulations 4-8 ................................................39 Fiber interaction coefficient, Simulations 9 and 10 ...................................40

Material Processing Effects, Simulations 9 – 20. ..................................................40 Effect of melt viscosity, Simulations 11-17...............................................41 PvT parameters effects, Simulations 18 and 19.........................................43 Control run, Simulation 20 ........................................................................44

SUMMARY AND CONCLUSIONS: PROCESS MODELING ......................................62

LITERATURE REVIEW: BLUNT OBJECT IMPACT ...................................................65

Categorization of impact........................................................................................65 Impactor mass and geometry .....................................................................66

Impact Energy........................................................................................................67

EXPERIMENTAL PROCEDURE: BLUNT OBJECT IMPACT.....................................69

Impact Test Apparatus ...........................................................................................69 Firing valve and pressure vessel ................................................................70 Firing mechanism.......................................................................................72 Gas gun carriage and barrel .......................................................................72 Pressure data acquisition............................................................................73 Velocity data acquisition............................................................................74 Calibration curves ......................................................................................75 Capture chamber ........................................................................................76 Sample holder and boundary conditions....................................................79 Sample preparation ....................................................................................79 Blunt object impact test matrix ..................................................................81

MATERIAL PROCESSING .............................................................................................82

Celstran®PP-GF40-03 Processing and Material Properties ..................................82 Numeric Results and Analysis of the Impact Data ................................................87 Damage Characterization.......................................................................................87

Effect of projectile geometry .....................................................................89 Areal density effects ..................................................................................91 Fiber orientation effects .............................................................................91

Micrograph Analysis..............................................................................................92 Correlation Between Predicted Fiber Orientation and Impact Failure Mode ........94

Page 7: Bartus 2003_Thesis

vii

TABLE OF CONTENTS (Continued)

Page

CONCLUSION: BLUNT OBJECT IMPACT.................................................................133

LIST OF REFERENCES.................................................................................................135

Page 8: Bartus 2003_Thesis

viii

LIST OF TABLES

Table Page

1 Properties for 40 wt. % E-glass/polypropylene used in the material models .....8

2 Processing parameters ......................................................................................28

3 Fiber material parameters .................................................................................32

4 Composite material parameters ........................................................................33

5 Matrix material parameters...............................................................................33

6 Fiber interaction coefficient parameters investigated.......................................40

7 Fiber interaction coefficient study results ........................................................40

8 Melt viscosity parameters investigated ............................................................41

9 Melt viscosity parameter study results .............................................................41

10 PvT parameters investigated.............................................................................43

11 PvT parameter study results .............................................................................43

12 Blunt object impact projectile types .................................................................69

13 Blunt object impact test matrix.........................................................................80

14 Material properties of Celstran® PP-GF40-03 adopted from the Ticona website .............................................................................................................83

15 Numerical results and analysis for the impact data on the top specimen.........86

16 Numerical results and analysis for the impact data on the center specimen ....86

17 Numerical results and analysis for the impact data on the bottom specimen...86

18 Numeric results and analysis of the velocity data for the top specimen ..........96

19 Numeric results and analysis of the velocity data for the center specimen......97

20 Numeric results and analysis of the velocity data for the bottom specimen ....98

Page 9: Bartus 2003_Thesis

ix

LIST OF FIGURES

Figure Page

1 Illustration of tensile and shear stress in a single fiber above, below and at the critical fiber length, adapted from Chawla [9] ............................................................6

2 Normalized property models; Cox shear-lag modulus, Cottrell impact strength, and Kelly-Tyson strength versus fiber length for 40 wt. % E-glass fiber (14 μm diameter) in a polypropylene matrix ......................................................7

3 Kelly-Tyson tensile strength model showing tensile strength versus log fiber length for discrete aligned E-glass fibers (40 wt. %) and randomly oriented E-glass fibers (40 wt. %) in a polypropylene matrix.....................................................10

4 Cox and Cox-Krenchel models for tensile modulus versus log fiber length of uniaxially aligned fibers (Cox) and randomly oriented fibers (Cox-Krenchel) for 40 wt. % E-Glass fibers in a polypropylene ......................................................13

5 Cottrell impact model for notched impact energy versus log fiber length of uniaxially aligned E-Glass fibers (40 wt. %) in a polypropylene matrix .................16

6 Cottrell impact models (with and without consideration of fiber strain energy) for notched impact energy versus log fiber length of uniaxially aligned E-Glass fibers (40 wt. %) in a polypropylene matrix...................................................18

7 Charge locations. Charge location (a) was used in Simulations 1-3, (b) in Simulation 4, (c) in 5, (d) in 7, (e) in 6, (f) in 8........................................................32

8 Charge location for the simulation verification study and for Simulations 9 – 20 ..............................................................................................................................37

9 Typical flow front at 50% filling for Simulations 1 – 3 ...........................................45

10 Flow fronts at 50% filling for Simulation 4 .............................................................45

11 Flow fronts at 50% filling for Simulation 5 .............................................................45

12 Flow fronts at 50% filling for Simulation 6 .............................................................45

13 Flow fronts for 80% filling showing the formation of knit lines in Simulation 7 ................................................................................................................................46

Page 10: Bartus 2003_Thesis

x

LIST OF FIGURES (Continued)

Figure Page

14 Flow fronts for 95% filling showing the formation of knit lines for two charges placed in Simulation 8 ..................................................................................46

15 Typical flow front for Simulations 9 - 20 showing 50% filling ................................47

16 Typical flow front for Simulations 9 – 20 showing 90% filling ...............................47

17 Force versus time for the control simulation, fiber interaction coeff. = 0.140..........48

18 Force versus time plot for the fiber interaction coefficient high (0.175) simulation ..................................................................................................................48

19 Force versus time plot for the fiber interaction coefficient low (0.105) simulation...................................................................................................................48

20 Force versus time for the control simulation, null viscosity = 4149 Pa s ..................49

21 Force versus time for the low null viscosity (3112 Pa s) simulation .........................49

22 Force versus time for the high null viscosity (5186 Pa s) simulation........................49

23 Force versus time for the control simulation, infinite shear viscosity = 1.0 s ...........50

24 Force versus time for the low infinite shear viscosity (0.90 s) simulation ................50

25 Force versus time for the high infinite shear viscosity (1.10 s) simulation...............50

26 Force versus time for the control simulation, power law index = 0.599 ...................51

27 Force versus time for the low power law index (0.539) simulation..........................51

28 Force versus time for the high power law index (0.569) simulation.........................51

29 Force versus time plots for low and high PvT coefficient simulations respectively ................................................................................................................52

30 Graphical representation of the fiber orientation in the (a) control run (0.14), (b) low fiber interaction coefficient (0.105) simulation and (c) the high fiber interaction coefficient (0.175) simulation .................................................53

31 Illustration of the selected element locations for the fiber orientation distribution function comparison: (a) element 39, (b) element 29, and (c) element 937................................................................................................................54

Page 11: Bartus 2003_Thesis

xi

LIST OF FIGURES (Continued)

Figure Page

32 Fiber distribution function of element 39 in the control simulation, fiber interaction coeff. = 0.140...................................................................................... 55

33 Fiber distribution function of element 39 in the low fiber interaction coefficient (0.105) simulation .............................................................................. 55

34 Fiber distribution function of element 39 in the high fiber interaction coefficient (0.175) simulation .............................................................................. 55

35 Fiber distribution function of element 29 in the control simulation, fiber orientation coeff. = 0.140 ..................................................................................... 56

36 Fiber distribution function of element 29 in the low fiber interaction coefficient (0.105) simulation .............................................................................. 56

37 Fiber distribution function of element 29 in the high fiber interaction coefficient (0.175) simulation .............................................................................. 56

38 Fiber distribution function of element 937 in the control simulation, fiber orientation 0.140................................................................................................... 57

39 Fiber distribution function of element 937 in low fiber interaction coefficient (0.105) simulation................................................................................................. 57

40 Fiber distribution function of element 937 in high fiber interaction coefficient (0.175) simulation................................................................................................. 57

41 Maximum nodal pressure (Pa) for the control simulation, where the maximum nodal pressure is 12.29 MPa, null viscosity = 4149 Pa s .................... 58

42 Maximum nodal pressure (Pa) for the low null viscosity (3112 Pa s) simulation, where the maximum nodal pressure is 12.37 MPa............................ 58

43 Maximum nodal pressure (Pa) for the high null viscosity (5186 Pa s) simulation, where the maximum nodal pressure is 12.26 MPa............................ 58

44 Graphical representation of the fiber orientation in the (a) control run (null viscosity = 4149 Pa s), (b) low zero shear viscosity (3112 Pa s) simulation, and (c) the high zero shear viscosity (5186 Pa s) simulation ............................... 59

45 Graphical representation of the fiber orientation in the (a) control run (infinite shear viscosity = 1.0 s), (b) low infinite shear viscosity (0.90 s) simulation, and (c) the high infinite shear viscosity (1.10 s) simulation.............. 60

Page 12: Bartus 2003_Thesis

xii

LIST OF FIGURES (Continued)

Figure Page

46 Graphical representation of the fiber orientation in the (a) control run (power law index = 0.599), (b) low power law index (0.539) simulation, and (c) the high power law index (0.659) simulation..............................................................61

47 Disassembled poppet valve: (a) two halves of the valve body, (b) valve center carrier, (c) valve, (d) barrel union, and (e) front valve face .......................71

48 Gas gun assembly on the carriage showing the pressure vessel, pressure transducer, firing valve, and firing mechanism....................................................73

49 Typical calibration curves for pressure versus velocity for the incident (1) and residual (2) velocity chronographs shown for a 100 g sabot.........................75

50 Pro/E drawing of the capture chamber with the access door removed..................77

51 Image of the capture chamber showing the (a) light bank, (b) barrel, (c) access door, (d) toggle clamps, (e) polycarbonate data acquisition windows and the (f) incident, and (g) residual velocity chronographs ................77

52 Pro/E drawing of the kinetic deflector showing the path of the projectile ............78

53 (a) Schematic of the top and side views of the tab plaque (not shown to scale) and the representative sections that were cut from it and (b) Pro/E isoperimetric drawing of the tab plaque...............................................................79

54 Critical velocity versus projectile mass for the top specimen, showing the velocity as an exponentially decreasing function of projectile mass ...................96

55 Critical velocity versus projectile mass for the center specimen, showing the velocity as an exponentially decreasing function of projectile mass .............97

56 Critical velocity versus projectile mass for the bottom specimen, showing the velocity as an exponentially decreasing function of projectile mass ...............98

57 Effect of projectile mass, energy versus projectile mass for the top specimen ................................................................................................................99

58 Effect of projectile mass, energy versus projectile mass, for the center specimen ................................................................................................................100

59 Effect of projectile mass, energy versus projectile mass, for the bottom specimen ................................................................................................................101

Page 13: Bartus 2003_Thesis

xiii

LIST OF FIGURES (Continued)

Figure Page

60 Energy (J) versus projectile mass (g) for the bottom, center, and top specimens with a linear regression analysis, fitted though the mean of each data set, indicative of the independent relationship between projectile mass (impact velocity) and the energy dissipation upon impact .................................... 102

61 Energy (J) versus areal density (g cm-2), plotted as a linear regression fit though the mean of the data for the 25 g, 50 g, 100 g, and 160 g flat projectiles illustrating that no significant relationship exits between the projectile mass and energy dissipation in the mass range examined..................... 103

62 Energy (J) versus areal density (g cm-2), plotted as a linear regression fit through the mean of the data for the 25 g, and 50 g conical projectiles illustrating that no significant relationship exits between the projectile mass and energy dissipation in the mass range examined.............................................. 104

63 High-speed image taken at 14,000 frames s-1 showing a 100 g flat tipped projectile exemplifying projectile tilting just after impacting the sample............. 105

64 High-speed images of a BOI illustrating the onset of damage, K.E. 142.3J ......... 106

65 High-speed images taken at 14,000 frames s-1 showing a 100 g flat projectile after the initial impact and while rebounding, K.E. 142.3 J .................................. 106

66 Impactor impressions left on the target for a projectile tilted (a) and normal (b) .......................................................................................................................... 107

67 Energy versus areal density for flat and conically tipped impactors showing a linearly increasing trend via linear regression fit through the mean of all the impact data............................................................................................................. 108

68 Impacted samples showing the location of sections taken for SEM analysis ....... 109

69 SEM of sample 34B, taken normal to the fracture surface, showing the path of fracture following the main fiber orientation angle .......................................... 110

70 SEM image of sample 34B showing fiber pull-out and fiber breakage ................ 111

71 SEM image of sample 39C showing fiber pull-out. .............................................. 112

72 SEM of sample 1C showing fiber pull-out (fiber sliding) and fiber matrix pull away ............................................................................................................... 113

73 Micrograph of sample 1C showing fiber pull-out lengths in excess of approximately 3 mm.............................................................................................. 114

Page 14: Bartus 2003_Thesis

xiv

LIST OF FIGURES (Continued)

Figure Page

74 SEM image of sample 1C showing fiber pull-out ................................................. 115

75 SEM of sample 34T showing a high degree of orientation and ductile pulling of fibrils ................................................................................................................. 116

76 Micrograph of sample 34T illustrating a brittle-matrix fracture with ductile pulling of fibrils .....................................................................................................117

77 SEM image of sample 38T normal to the fracture surface showing fiber pull-out and breakage ...................................................................................................118

78 SEM of sample 39C showing variations in fiber orientation through the thickness of the section, taken normal to the fracture plane..................................119

79 Illustration of a possible failure mechanism of the samples tested showing a transverse view of a test plaque with the fracture path following a preferential plane of fiber orientation .......................................................................................120

80 (a) Bottom specimen (020920-1-12B), impacted with a 50 g flat projectile, exhibiting a typical fracture pattern, (b) graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (B) with the fracture pattern superimposed over the results .................121

81 (a) Bottom specimen (020920-1-63B), impacted with a 25 g flat projectile, exhibiting a typical fracture pattern, (b) graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (B) with the fracture pattern superimposed over the results .................122

82 (a) Center specimen (020920-1-16C), impacted with a 25 g flat projectile, exhibiting a typical fracture pattern, (b) graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (C) with the fracture pattern superimposed over the results .................123

83 (a) Center specimen (020919-1-35C), impacted with a 160 g flat projectile, exhibiting a typical fracture pattern, (b) graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (C) with the fracture pattern superimposed over the results .................124

84 (a) Center specimen (020919-1-32C), impacted with a 160 g flat projectile, exhibiting a typical fracture pattern, (b) graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (C) with the fracture pattern superimposed over the results ................. 125

Page 15: Bartus 2003_Thesis

xv

LIST OF FIGURES (Continued)

Figure Page

85 Top specimen (020919-1-35T), impacted with a 50 g flat projectile, exhibiting a typical fracture pattern, (b) graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (T) with the fracture pattern superimposed over the results..................126

86 (a) Top specimen (020919-1-107T), impacted with a 160 g flat projectile, exhibiting a typical fracture pattern, (b) graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (T) with the fracture pattern superimposed over the results.................. 127

87 (a) Bottom specimen (020920-1-27B), impacted with a 50 g conical projectile, exhibiting a typical fracture pattern, (b) graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (B) with the fracture pattern superimposed over the results ................. 128

88 (a) Bottom specimen (020920-1-22B), impacted with a 50 g conical projectile, exhibiting a typical fracture pattern, (b) graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (B) with the fracture pattern superimposed over the results ................. 129

89 (a) Center specimen (020919-1-61C), impacted with a 25 g flat projectile, exhibiting a typical fracture pattern, (b) graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (C) with the fracture pattern superimposed over the results ................. 130

90 (a) Bottom specimen (020920-1-87T), impacted with a 50 g conical projectile, exhibiting a typical fracture pattern, (b) graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (T) with the fracture pattern superimposed over the results.................. 131

91 (a) Superimposed view of the fracture patterns over the entire tab plaque for the flat projectiles from Figures 75(a), 79(a) and 80(a). (b) Superimposed view of the fracture patterns over the entire tab plaque, for the conical projectiles from Figures 83(a), 84(a) and 85(a) ................................................... 132

Page 16: Bartus 2003_Thesis

1

INTRODUCTION

Thermoplastic Composites

The current U.S. market for such materials is in excess of 4.54 X 108 kg per

annum, half of which is consumed by the automotive industry [1]. Long-fiber-reinforced

thermoplastic (LFT) composites have one of the highest growth rates in the polymer

material areas, sustaining a projected 30% growth from 2000 to 2004 [2]. Thermoplastic

composites typically comprise a cost-effective commodity matrix, such as polypropylene

(PP), polyethylene (PE), or nylon, reinforced with glass, carbon, or aramid fibers. E-glass

is the most common reinforcement since the automotive market niche is driven more by

cost/performance ratio than weight/performance ratio as demanded by the aerospace

industry.

Thermoplastic composites used in these applications can be short-fiber-reinforced

Thermoplastic (SFRT), glass mat thermoplastic (GMT), or LFT. Injection-molded SFRT

composites (starting fiber lengths less than 4 mm) are currently the most prevalent of the

aforementioned composites. However, the full advantage of the reinforcing fiber is not

realized, due to the low fiber aspect ratio. Injection-molded LFTs also suffer from

excessive fiber length degradation in the plastification and injection stages. In addition,

difficulties arise in processing components with high fiber content and starting fiber

lengths in excess of 13 mm, due to the high melt viscosity.

GMTs consist of a chopped or continuous fiber mat reinforcement in a

thermoplastic matrix. Preparation is done by melt impregnation of non-woven glass mat

Page 17: Bartus 2003_Thesis

2

(dry route) or by mixing chopped fiber with polymer powder in a fluid medium (wet

route), both of which are heated in a GMT oven prior to compression-molding [3].

The large fiber aspect ratio of the reinforcement takes full advantage of the fiber

for strengthening, in contrast to its short fiber counterpart. Injection, injection-

compression, and extrusion-compression-molding techniques are employed for LFT

processing. LFT fiber lengths are typically greater than 13 mm and depend on the

desired properties, fiber concentration, and processing technique. LFTs offer several

advantages in contrast to GMTs, such as the possibility to work without semi-finished

mats (e.g. inline extrusion) making it less labor intensive, and lower compression forces

due to a decrease in melt viscosity, which results in capital cost savings in tooling and

machinery. Moreover, LFTs offer higher surface quality; less part rejection, due to an

increase in the ability to fill complex features; and integrated recycleablity. Another

advantage is greater freedom in choosing fiber and matrix materials.

In compression-molding, fibers develop in plane orientations during flow, which

can plague the consolidated component. Preferential orientation during compression-

molding can reduce strength and stiffness in a critical area and will induce warping

through anisotropic contraction upon cooling [4]. A considerable amount of work has

been done with injection and compression-molding short-fiber-reinforced thermoplastics

and thermosets, examining the state of fiber orientation and flow fronts. Very little work

has been on done modeling long-fiber reinforcement and the analytical effect of the

processing conditions and material properties. Work in this area will be beneficial in

component design, material selection, and process variables. In addition, reasonable

goals can be set in obtaining material properties required for process modeling.

Page 18: Bartus 2003_Thesis

3

As the use of LFTs grows in automotive and other industries, the need to

determine the impact properties of these materials increases, in order to ensure the safety

and stability of designed structures [5]. LFT was recently employed for the underside

“belly pan” of Daimler Chrysler’s PT Cruiser [3]. Sheet molding compound (SMC) was

first employed for this application. However, it proved too brittle, since that component

needed to be flexible and withstand impact from stones and other objects [3]. Few

authors have attempted to characterize the impact performance of discontinuous,

randomly oriented LFT thermoplastics.

The fiber architectures inherent in LFTs make an accurate characterization of the

failure mechanisms complex. Most efforts in understanding the impact performance and

failure mechanisms of LFTs have primarily focused on Charpy and Izod impact testing,

and to an even lesser degree, low velocity drop-tower impact testing [6]. Very little work

has looked into the effect of intermediate velocity blunt object impact (BOI) on LFTs.

Intermediate velocities are greater than low velocity drop tower impacts or pendulum

type impacts (10m s-1), yet slower than high velocity ballistic type impacts. The velocity

range for this purpose simulates the effect of blunt objects, such as rocks and debris,

traveling at highway speeds for automotive applications, as well as impact induced by

debris from hurricanes and tornadoes for storm shelter and military housing applications.

This work can also be extended to transverse-loaded energy dissipation under high

loading rate for automobile crash mitigation purposes.

Page 19: Bartus 2003_Thesis

4

OBJECTIVE

The objective of this work can be divided into three categories:

I. Obtain a quantitative understanding of the processing conditions/material

property relationship in the manufacture of compression-molded LFTs.

II. Provide an understanding of the impact behavior of LFTs under various

intermediate velocity BOIs to assist engineers and scientists in obtaining reliable data

relevant to practical applications.

III. Deduce a qualitative relationship between the fiber orientations predicted in a

process simulation and the respective failure modes seen under BOI.

Page 20: Bartus 2003_Thesis

5

LITERATURE REVIEW

Discontinuous Long-Fiber Reinforcement

Traditional processing of LFT begins by hot melt impregnating a tow of

reinforcing fibers with a thermoplastic matrix and subsequently chopping the continuous

tow into pellets of a set length. Hot-melt impregnation is done by wirecoating, cross-

head extrusion, or thermoplastic pultrusion techniques [3]. The LFT pellets are then fed

into a single-screw plasticator where they are fed down the barrel by the screw, heated

above the melting point of the matrix, and extruded as a charge (shot). The shot is then

placed on a tool and compression molded.

Extrusion/compression molding of LFT has been rapidly gaining favor over

traditional injection molding (especially with in-line compounding) and GMT

compression molding, due to superior mechanical properties at a comparable cost [7].

The fiber aspect ratio, defined as the length to diameter ratio, differentiates short fiber

from long-fiber reinforcement. The aspect ratio of a long-fiber is typically an order of

magnitude greater than that of a short fiber [8]. While short-fiber-reinforced

thermoplastics realize substantial gains in mechanical properties over that of the neat

material, the full potential of the reinforcement is not obtained, because the fiber is below

a critical length.

The critical fiber length is given in equation (1):

τσ rLc

max= (1)

Page 21: Bartus 2003_Thesis

6

where Lc is the critical fiber length, r is the fiber radius, σmax is the tensile stress acting on

the fiber, and τ is the interfacial shear strength, equation (2) [9].

This equation is based on several simplifying assumptions, the first of which is

that the strain to failure for the fiber is less than that of the matrix. This is a reasonable

assumption in the case of thermoplastic matrices. A shortcoming of this equation is that

it assumes the interfacial shear stress in constant over the fiber length. It has been shown

that fibers produce higher stresses at the fiber tips, resulting in a lower elongation to

failure [10]. This is illustrated in Figure 1, adapted from Chawla [9].

lr

2στ = (2)

Cdl

dl

⎟⎠⎞

⎜⎝⎛<

Cdl

dl

⎟⎠⎞

⎜⎝⎛=

Cdl

dl

⎟⎠⎞

⎜⎝⎛>

σf

τ

Figure 1 Illustration of tensile and shear stress in a single fiber above, below, and at the critical fiber length, adapted from Chawla [9].

Page 22: Bartus 2003_Thesis

7

Assuming that, below the critical fiber length, the force required for debonding

increases linearly with fiber length, the interfacial shear strength can be determined from

the slope of the load required for pull-out versus fiber length. Above the critical fiber

length, sufficient interfacial shear stress exists for fiber breakage to occur. This is the

basis for the subsequent material models.

Material Property Models

Several models have been developed in order to predict the modulus, tensile

strength, and impact strength for discontinuous fiber-reinforced composites. For the

normalized Kelly and Tyson theoretical model shown in Figure 2, the composite strength

approaches 90% that of a continuous fiber-reinforced composite, as the fiber length

approaches 14 mm [11]. For the Cox modulus model, Figure 2, 90% of the composite

Figure 2 Normalized property models; Cox shear-lag modulus, Cottrell impact strength, and Kelly-Tyson strength versus fiber length for 40 wt. % E-glass fiber (14 μm diameter) in a polypropylene matrix.

Page 23: Bartus 2003_Thesis

8

stiffness is realized at a fiber length of 0.8 mm [12]. The Cottrell impact model shown in

Figure 2 illustrates that the theoretical impact resistance of discontinuous fiber-reinforced

composites is optimum at the critical fiber length, 3.35 mm [13]. These property models

are interrelated and will be discussed in detail below. The material properties used in the

models are given in the next section.

Material Properties. The material properties in Table 1 were chosen to reflect

those of commercial polypropylene and E-glass and were obtained from the literature [3,

6, 14, 15, 27, 33].

Table 1 Properties for 40 wt. % E-glass/polypropylene used in the material models

Nomenclature Value Unit Property

D 14 μm Fiber diameter

Ef 75 GPa Fiber modulus

Lc 3.35 mm Critical fiber length from equation (1)

σfj 1.82 GPa Fiber tensile strength

Gm 6.56 MPa Matrix shear modulus from equation (8)

Em 1.60 GPa Matrix Modulus

τ 3.80 MPa Interfacial shear strength from equation (2)

σm 38.80 MPa Matrix strength

Um 1000 J m-2 Matrix fracture energy

Ud 500 J m-2 Interface fracture energy

τfs 0.910 MPa Static frictional interfacial shear strength

τfd 0.455 MPa Dynamic frictional interfacial shear strength

αm 96 μm m-1 oC-1 Coefficient of thermal expansion (matrix)

αf 5 μm m-1 oC-1 Coefficient of thermal expansion (fiber)

Ts 120 oC Solidification temperature

Page 24: Bartus 2003_Thesis

9

Modeling of tensile strength. The model for the prediction of a polymer composite

strength with discrete aligned fibers, originally developed by Kelly and Tyson [13]

(1965), is well known. The original work was based on copper/tungsten and

copper/molybdenum metal matrix composites but has since found applications in

polymer matrix composites [11]. The theory of strengthening for fiber-reinforced

composites is based on the idea that interfacial shear stresses at the fiber-matrix interface

are limited by the flow stress of the matrix or by the shear strength of the interface [11].

Table 1 (Continued)

Nomenclature Value Unit Property

Tt 20 oC Test temperature

σr 9.10 MPa Radial stresses due to thermal shrinkage from equation (17)

Xi 4 Geometric parameter for square packing of fibers

ηo 1 Orientation factor (discrete aligned fibers)

ηo 0.375 Orientation factor (random fiber orientation) from equation (13)

ηl 0.277 Fiber length efficiency factor from equation (6)

β 22301 Shear-lag parameter from equation (7)

Wf 0.40 Fiber weight fraction

Vi 0.187 Volume faction of subcritical length fibers

Vj 0.187 Volume faction of supercritical length fibers

Vf 0.187 Fiber volume fraction

νf 0.20 Poisson's ratio (fiber)

νm 0.25 Poisson's ratio (matrix)

ρs 0.10 Coefficient of static friction at fiber-matrix interface

ρd 0.05 Coefficient of dynamic friction at fiber-matrix interface

L 0.01-100 Fiber length in millimeters

Page 25: Bartus 2003_Thesis

10

In order for embedded fibers to fracture upon loading, the fiber length must be greater

than the critical length defined in equation (1). If the fibers are below critical length,

pull-out will result. The Kelly-Tyson model is given in equation (3), and the variables

are given in Table 1.

The summation terms in equation (3) arise from the contribution of subcritical and

supercritical fiber lengths. If only fibers of uniform length are considered, the summation

terms cancel, which is the case in the material models shown in Figures 2 and 3.

( ) umfj

cjfj

iiuc V

LLV

DVL σστσ −+

⎥⎥⎦

⎢⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

⎛−Σ+⎥⎦

⎤⎢⎣⎡Σ= 1

21 (3)

Figure 3 Kelly-Tyson tensile strength model showing tensile strength versus log fiber length for discrete aligned E-glass fibers (40 wt. %) and randomly oriented E-glass fibers (40 wt. %) in a polypropylene matrix. The dashed line indicates the critical fiber length, 3.35 mm from equation (1). The fiber diameter was taken as 14 μm.

Page 26: Bartus 2003_Thesis

11

Equation (3) cannot be integrated to take randomly oriented fibers into account,

which gives rise to the need for a fiber efficiency factor. Chou published an equation to

calculate the fiber efficiency factor, ηo, for random, planar laminates containing fibers of

uniform length, equation (4) [15]. The variables in equation (4) are given in Table 1.

Thomason and Vlug investigated the applicability of equation (4) in reference to the

Kelly-Tyson material model. The theoretical prediction of 0.20-0.25 from equation (4)

showed a good correlation to experimental results in which a linear regression gave a

fiber orientation factor of 0.20 for oriented E-glass fibers (10–40 wt. %) in a

polypropylene matrix [14].

A random orientation factor of ηo = 0.375 was used in the Kelly-Tyson tensile

strength model shown in Figure 3 [16]. A dramatic decrease in tensile strength is seen in

the case of randomly oriented fibers, signifying a strong off-axis effect; this is well

known for unidirectional laminates. This indicates that laminate strength is most likely

governed by fibers oriented parallel to the loading direction, thus making knowledge of

the fiber orientation state a primary concern.

Modeling of tensile modulus. The Cox shear-lag model was developed in 1952 to

predict composite stiffness for aligned discontinuous elastic fibers in an elastic matrix

[12]. Krenchel improved the model by incorporating an orientation parameter, ηo, to

account for variations in planar fiber orientations. The Cox-Krenchel model is given in

equation (5), where ηl is the fiber-length efficiency factor described in equation (6) and

( )⎥⎥⎦

⎢⎢⎣

⎟⎟

⎜⎜

−−

−+−×−+= −

2

2122

3 11

11log

21cos12

38

β

βββββ

πηo (4)

Page 27: Bartus 2003_Thesis

12

β is the shear-lag parameter from equation (7). The shear modulus of the matrix, Gm, can

be calculated from equation (8). The r/R factor in equation (7) is related to fiber volume

fraction by equation (9).

The term Xi in equation (9) is dependent on the geometric packing arrangement of the

fibers. A value of Xi = 4 was used in the calculation, which is appropriate for square

packing of fibers [17]. The Krenchel orientation factor can be calculated from equation

(10):

where

For example, four fiber layers with 0o, 90o, 45o, and -45o planar orientations can be

described by equation (12):

nnno a θη 4cosΣ= (10)

1=Σ nna (11)

( ) ( )fiVRr Χ= πlnln (9)

( ) mfffloc EVEVE −+= 1ηη (5)

( )⎥⎦

⎤⎢⎣

⎡−=

22tan1

LL

l ββη (6)

( )21

/ln22

⎥⎥⎦

⎢⎢⎣

⎡=

RrEG

D f

mβ (7)

( )υ+=

12m

mEG (8)

Page 28: Bartus 2003_Thesis

13

Solving equation (10) with the parameters given in equation (12) results in a Krenchel

orientation parameter of 0.375, equation (13):

An orientation factor 0.375 is considered random and can be verified upon integration of

(10) from -900 < θ < 900, with the exact solution given in equation (14):

which is approximately equal to 0.358 [17]. The remaining variables are described in

Table 1. Figure 4 shows the Cox and Cox-Krenchel tensile modulus models as a

442

0

41

4

3

2

1

4321

πθπθπθ

θ

−====

====and

aaaa (12)

( ) 8/3414101410 =+++=η (13)

πθθ

ππ

π 89cos10

2

2

4 == ∫−d (14)

Figure 4 Cox and Cox-Krenchel models for tensile modulus versus log fiber length of uniaxially aligned fibers (Cox) and randomly oriented fibers (Cox-Krenchel) for 40 wt. % E-Glass fibers in a polypropylene matrix.

Page 29: Bartus 2003_Thesis

14

function of fiber length.

As with tensile strength, fiber orientation plays a pivotal role on the tensile

modulus, which is illustrated in Figures 3 and 4. The effect of fiber orientation is much

more predominant than the effect of fiber aspect ratio, since 90% of the tensile modulus

is realized at sub-millimeter fiber lengths. Less than half of the tensile modulus (~43%)

is seen in the randomly oriented model at 90% of the theoretical tensile modulus, in

contrast to the uniaxially aligned fiber model.

Cottrell impact model. In 1964, Cottrell developed a model to predict the notched

impact strength of discontinuous uniaxially aligned fiber composites [13]. It is important

to identify the failure mechanisms present during impact that account for energy

absorption. Deformation and fracture of the matrix takes place in front of the crack tip.

Concurrently, the matrix transfers load to the fibers by shear. If the applied load exceeds

the fiber-matrix interfacial shear strength, debonding may occur.

Transfer of load may still occur to a debonded fiber via frictional forces along the

interface. Fibers may fracture if the stress level exceeds the fiber strength, or fracture

may occur prematurely from local flaws present along the fiber length, and inherent in

the fiber itself. Fibers that have debonded will still dissipate energy as they are pulled out

from the matrix. All of these mechanisms are incorporated into the Cottrell impact model

[13]. Like the Kelly-Tyson model, the Cottrell model incorporates equations for

subcritical and supercritical fiber lengths, as defined by equation (1). When L>Lc, the

predicted impact energy dissipation is given in equation (15):

(15) ( )⎥⎥⎦

⎢⎢⎣

⎡+⎥

⎤⎢⎣

⎡+−=

DLV

DLUV

UVU ffdfmfc 6

12τ

Page 30: Bartus 2003_Thesis

15

where the three terms encompass matrix fracture, fiber/matrix debonding, and fiber pull-

out energies. When L>Lc, the predicted impact energy dissipation is calculated from

equation (16):

where the four terms account for matrix fracture, fiber fracture and debonding, and pull-

out limited to the critical fiber length. The energy dissipated by the matrix is small,

because the presence of fibers inhibits large deformations [6].

The interfacial shear friction, τf, will not normally equal the interfacial shear

strength, τ. The calculated interfacial friction (τf = 0.910 MPa) is an order of magnitude

lower than the interfacial shear strength (τ = 3.8 MPa) reported by Thomason and Vlug

for polypropylene and E-glass [6]. The coefficient of friction between the fiber and

interface, τf = μdσr, can be determined from the radial stresses, due to thermal mismatch

between the fiber and matrix at processing temperature and test temperature from

equation (17). The variables are described in Table 1.

The interfacial friction tends to be significant in thermoplastic composites, with a

high degree of thermal mismatch between the fiber/matrix and a large temperature

change due to elevated processing temperatures. The interfacial friction plays an

important role in the fiber pull-out energy and is thought to be the predominate

mechanism of energy dissipation [6].

( ) ( )⎥⎥⎦

⎢⎢⎣

⎡+

⎥⎥⎦

⎢⎢⎣

⎡+⎥

⎤⎢⎣

⎡ −+−=

DLLV

DLULV

LULLV

UVU fcfdcffcfmfc 6

132 τ

(16)

( )( )( ) ( ) mmfff

mftsfmR EEV

EETTυυ

αασ

++++

−−=

121(17)

Page 31: Bartus 2003_Thesis

16

The Cottrell model for notched impact energy dissipation versus log fiber length

is shown in Figure 5. Due to the brittle fracture of glass fibers, the contribution of energy

dissipation in considered negligible (e.g. Uf = 0). The large peak in the predicted impact

strength is associated with the critical fiber length. When L < Lc, the failure mode is

predominately fiber pull-out, whereas fiber failure dominates for L > Lc. There is very

little experimental support for the peak at the critical fiber length in the Cottrell model.

Cooper [18] reported some evidence of this phenomenon for ductile fibers in a brittle

matrix (copper/epoxy), but Thomason and Vlug found no indication of a decrease in

impact strength for fibers above critical length for PP/E-glass composites [6]. Wald and

Figure 5 Cottrell impact model for notched impact energy versus log fiber length of uniaxially aligned E-Glass fibers (40 wt. %) in a polypropylene matrix. The peak corresponds to the critical fiber length, Lc = 3.35 mm.

Page 32: Bartus 2003_Thesis

17

Schriever [19] also found evidence of high impact strength in GMTs, which typically

contain fiber lengths of 25-50 mm.

The experimental data from Thomason and Vlug on PP/E-glass suggested that

fiber strain energy might play an important role in the energy absorption process,

accounting for the discrepancy in long-fiber-reinforced composite impact strength [6].

The total energy involved in fracture of a single fiber is given by equation (18)

where L is the gauge length of the test specimen and the remaining variables are given in

Table 1. The strain energy stored in a fiber is dissipated in the form of heat and acoustic

energy upon fracture [6]. In the case of fibers below the critical length, the strain energy

is released in a similar manner as they debond from the matrix. This also contributes to

the explanation of an increase in impact strength of PP/E-glass composites with a

decrease in temperature [6]. As the test temperature decreases, the interfacial friction

increases from equation (17), increasing the impact strength. In addition, an increase in

strength of glass fibers has been reported with decreasing temperature, σf/dT ~5 MPa0C-1,

which would contribute to the fiber strain-energy absorption in equation (18) [6].

If fiber strain energy contributes to impact energy absorption, equation (18) can

be incorporated into equation (16) to yield the impact strength. The Cottrell impact

strength versus log fiber length is plotted with the original and modified models in Figure

6.

f

ff E

LU

2

2σ= (18)

Page 33: Bartus 2003_Thesis

18

A comprehensive experimental database on the effect of aspect ratio on composite

strength, stiffness, and impact properties is not currently available [3]. Moreover, the

work done on characterizing the effect of aspect ratio on the aforementioned is typically

done with fiber-friendly processing methods when, in actuality, a significant amount of

fiber length degradation and bending occur during processing [21]. Thomason and Vlug

have shown that a direct relationship exists between Charpy impact energy dissipation

and tensile strength, which indicates that the parameters governing laminate strength also

govern resistance to impact [6]. Assuming a uniform or, more accurately, an average

fiber length and fiber concentration for a given discontinuous fiber-reinforced material,

the composite strength is dictated by only the fiber orientation. If one accepts that impact

Figure 6 Cottrell impact models (with and without consideration of fiber strain energy) for notched impact energy versus log fiber length of uniaxially aligned E-Glass fibers (40 wt. %) in a polypropylene matrix. The fiber diameter is 14 μm.

Page 34: Bartus 2003_Thesis

19

strength is dictated by the fiber orientation, the composite failure will occur along planes

parallel to areas of high preferential fiber orientation. Therefore, knowledge of the

orientation state should allow one to qualitatively predict areas that may be susceptible to

damage under impact.

Page 35: Bartus 2003_Thesis

20

CADPRESS-TP BACKGROUND

Cadpress-Thermoplastic (or Express) was developed jointly by M-Base

Engineering and Software, in conjunction with their academic partner, the Institut fur

Kunststoffverarbeitung (IKV) in Aachen, Germany, and The Madison Group with their

academic partner, the Processing Research Center at the University of Wisconsin,

Madison [22]. Currently, CADPRESS-TP is probably the only commercial software

suitable for simulating compression-molding of discrete long-fiber-reinforced

thermoplastics. The fundamental background of Cadpress-TP (Cadpress) and its

relationship to other work on process modeling of thermoplastic composites are given in

this chapter.

Cadpress performs two discrete, albeit dependent, simulations that reproduce the

process-induced material properties, shrinkage, and warpage of complex discontinuous

fiber-reinforced thermoplastic matrices when compression molded using finite element

methods [22]. The first part of the program simulates the flow behavior of the melt

during the filling stage of compression molding. It is during the filling stage that flow-

induced fiber orientation develops, upon which the final mechanical and

thermomechanical properties are highly dependent [22].

The user has the option of an isothermal or non-isothermal flow calculation, but

the non-isothermal simulation is inherently more accurate, due to the non-Newtonian

nature of polymer solutions, which undergo shear thinning. The temperature dependent

melt viscosity does not share a linear relationship with shear rate. In the case of shear

Page 36: Bartus 2003_Thesis

21

thinning, the apparent viscosity decreases with increasing shear rate [23]. Background

will not be given for the isothermal flow calculation for brevity. The flow simulation

follows the control volume approach based on a static, finite element mesh. The flow

front progression is defined by a fill factor assigned to each element for three stages of

filling: fi = 1, for fully filled; 0 < fj < 1, for partial filling; and fk = 0, for no filling.

In the case of thermoplastic melts, the material cools rapidly as it encounters the

mold walls, increasing the viscosity locally until the melt no longer flows. The non-

isothermal calculation takes into account the local variations in viscosity with equations

19 and 20.

The shear rate, γ& , in equation 19 corresponds to the velocity gradient through the

flow channel height, accounting for the shear rate dependency of the viscosity. The three

parameters in equation 19 e.g., P1, P2 and P3, are the zero shear or null viscosity, the

infinite shear viscosity, and the power law index, respectively. The temperature

dependency of the viscosity is addressed by the temperature shift coefficient, aT, equation

20 [22].

The flow simulation can be described using the generalized Hele-Shaw flow

model for incompressible, inelastic, non-Newtonian fluid under non-isothermal

conditions [22, 24-32]. The temperature and shear-rate dependent viscosity are then used

to determine the non-isothermal, non-Newtonian flow conductivity, which is given in

( ) 3

2

1

1 pT

T

PaaP

γη

&+=

( ) ( )( )

( )( )S

oS

SBo

SBT TTC

TTTTC

TTa−+

−−

−+−

=6.10186.8

6.10186.8log

(19)

(20)

Page 37: Bartus 2003_Thesis

22

equation 21. The flow conductivity, equation 21, is a function of the flow channel height,

h.

Due to the highly temperature-dependent nature of compression molding,

calculation of the temperature distribution is imperative. Energy transport is considered

in equations 22, 23, and 24, which are the convection, conduction, and diffusion terms,

respectively. The energy transport equations are solved in conjunction with the

governing flow equations. After filling, the diffusion and convection terms drop out of

the energy equation, leaving only the conduction term. The solution to the differential

equation is solved using a one-dimensional implicit finite difference form of the equation

(22).

Upon ejection, a second energy balance must be applied to account for heat

transfer to the surroundings as the part cools to ambient temperature, in which case the

diffusion and convection terms must be reintroduced. The warpage phenomenon,

common in compression and injection molding, requires modifying the boundary

conditions for the energy equations. The coefficient of thermal diffusion, λ, must then be

modified to reflect the heat transfer from the part to the surroundings [22].

η12

3hS = (21)

⎟⎟⎠

⎞⎜⎜⎝

⎛∂∂

+∂∂

−yT

xTc yxp ννρ (22)

zzy

yzx

xz ∂

∂−

∂∂

−ν

τντ (23)

2

2

zT

tTc p ∂

∂=

∂∂ λρ (24)

= 0

= 0

Page 38: Bartus 2003_Thesis

23

The most critical calculation in the simulation is the prediction of fiber

orientation, from which the thermomechanical and final mechanical properties of the

component are determined. The fiber orientation calculation starts at the very beginning

of the simulation and is continued until the mold cavity is filled. Jeffery provided the

first model to attempt to predict fiber orientation [23]. The Jeffery model does not

consider fiber-fiber interactions and can only be applied in the dilute fiber volume

regime, φf <<(d/l) 2, in which the fiber-fiber interactions are infrequent and purely

hydrodynamic [24].

The Jeffery model for predicting fiber orientation was modified by Folgar and

Tucker to account for fiber interaction with a damping term, known as the fiber

interaction coefficient, C1. The rotary diffusivity can than be expressed as Dr = C1γ&

[25]. The same model has been used by several other authors for the prediction of fiber

orientation in injection, injection/compression, and compression-molded short-fiber-

reinforced thermoplastics [22, 24-32].

The fiber interaction coefficient can be expressed in terms of the root-mean-

square angle change caused by the fiber-fiber interactions or can be determined

empirically in the absence of any knowledge of the fiber angle change caused by

mechanical contact [24]. The fiber interaction coefficient depends on the number of fiber

contacts, N, built up in the flow fields, so it not only accounts for the fiber volume

content but also the fiber aspect ratio [8]. Fibers have a tendency to resist alignment in

the flow where they are in contact with one another. It is, however, inherently unstable,

requiring the help of an additional numerical procedure.

Page 39: Bartus 2003_Thesis

24

Several simplifying assumptions must also be made to efficiently conduct the

rather intensive calculations [22]. The first of which is that the fibers are considered rigid

bodies with uniform length and diameter. Also, the fiber-matrix melt is considered

incompressible, and the viscosity of the matrix is so high that inertial and buoyancy

effects are negligible. Another assumption is that there are no externally applied forces

or moments on the fibers. In addition, the interaction between two fibers is assumed to

take place when the fibers’ centers of gravity move past one another within a distance

that is smaller than the length of the fibers. Long fibers may deviate from the rigid body

behavior by bending. This is handled by treating the filament as a series of single, linked,

inflexible filaments [22].

Since it is computationally inefficient to consider each fiber interaction

separately, the Folgar-Tucker model uses a statistical approximation of the entire domain.

The Gaussian probability distribution, ψ, of the fiber orientation must satisfy the

continuity equation, which accounts for all of the fibers rotating in and out of an arbitrary

control volume, equation (25). Equation (26) gives the rotational speed of a single fiber

with the addition of a damping term. The variable, γ& , is the scalar magnitude of the

strain rate tensor, describing the frequency of fiber-fiber interaction. Equation (27) is

used in equation (28) to calculate the fluid velocity for variable layers, based on the shear

velocity present at a given layer.

Page 40: Bartus 2003_Thesis

25

The differential equation for the calculation of fiber rotation is not solvable by

analytical methods, requiring a numeric solution. The governing equations are first

discretized and rewritten in implicit form [22]. The fiber angles are then discretized into

25 angle classes, from 0° to 180°, with respect to the local coordinate system and are

displayed as a fiber orientation distribution, fiber frequency versus angle class [22]. This

is done for each of the five discrete layers from the midplane, assuming symmetry about

the midplane. Dividing the geometry into layers allows for the consideration in the

different velocity profiles with respect to the flow channel height. This is necessary since

the flow channel height is transient, because of mold closing and the cooling and eventual

freezing of the material. The systems of equations are solved using a Gaussian-based

matrix-solving algorithm [22].

In the flow front, the fibers rotate while being transported across element

boundaries, implying that a different fiber orientation as well as volumetric content may

be present at a given time step. An average flow rate for each element is calculated in

order to satisfy the continuity requirements. The anisotropic elastic material properties

(25) ⎥⎥⎦

⎢⎢⎣

+

+−−

∂∂

−∂∂

=∂

),,2

,2

,

2

2

cossincos

sincossin(

yyxy

yxxxIC

t φνφφν

φνφνφψ

φφψγψ

&

yyxyyxxxICt ,,

2,

2, cossincossincossin1 φνφφνφνφνφ

φψ

ψγφ

++−−∂∂

−=∂∂

& (26)

1++= μνμ νγν Sh& (28)

yyxyyxxx ,22

,,,2 2)(2 ννννγ +++=& (27)

Page 41: Bartus 2003_Thesis

26

are calculated by first applying the micromechanical Halpin-Tsai empirical expressions,

given as equations 29-34. The expressions account for the aspect ratio of discontinuous

fiber reinforcement. The variables for equations 29–34 are described in Table 1.

However, the Halpin-Tsai expressions can only be used in the unidirectional case.

Therefore, the 25 different fiber orientation angles are thought of as 25 discrete layers

and treated using classical laminate theory. The laminates are combined using the

principle of superposition, also known as Continuum Theory, derived by Puck and

Halpin, to give the composite anisotropic elastic material properties for each element

[22].

Physical and mechanical material property data are very limited in the relatively

new realm of LFT composites. The material behavior, both in processing and in the final

product, differs significantly from its SFRT analogy. Significant improvements in tensile

strength and stiffness, as well as impact resistance, have been documented with

whereGG M ,5.015.01

12 Φ−Φ+

1=ξ (34)

1

22112 E

Eνν = (33)

2=ξwhereEE M ,1

12 Φ−

Φ+=

κξκ (30)

whereEE M ,1

11 Φ−

Φ+=

κξκ

F

F

dl2=ξ (29)

ξκ

+

−=

M

F

M

F

EEEE 1

(31)

)1(21 Φ−+Φ= ννν F (32)

Page 42: Bartus 2003_Thesis

27

increasing aspect ratio [2, 5, 6, 10-12, 14-16, 19, 33]. However, a major drawback is

seen in the processing of such materials. The high melt viscosity of LFTs with large fiber

aspect ratios and high volumetric fiber content require extrusion/compression molding,

whereas injection molding can still be used for shorter reinforcements with low fiber

loading.

Compression molding can be modeled effectively using Cadpress-TP

compression-molding software, which simulates the processes from the time the charge is

placed in the mold to the time when the part has cooled to the user-determined limiting

temperature outside the mold. From this important processing information, mechanical

property data, and shrinkage and warpage data are calculated. This offers the immense

advantage of decreased development time. Furthermore, the part design, processing

conditions, and mechanical performance can be optimized without having to produce an

actual component. Moreover, the manufacturing process can be controlled/optimized to

generate favorable processing conditions and fiber orientation states to obtain the best

possible components from such a process [27,28].

Difficulties, however, arise in obtaining accurate material data for the

compression-molding simulations. A vast array of data is required, much of which is

costly and/or difficult to obtain. The software can be utilized to determine the sensitivity

to changes in processing conditions and material parameters, and to resolve which of the

parameters are most likely to effect processing and component performance through a

simulation matrix.

Page 43: Bartus 2003_Thesis

28

EXPERIMENTAL PROCEDURE: PROCESS MODELING

Simulation Matrix

The properties of a material are related to its structure, and processing controls

the structure. This has a particular relevance to the compression-molding process, in

which the material is made while producing the component [29]. The goal in the

experimental Cadpress-TP simulation matrix is to determine the effect of material

properties and processing parameters on the manufacturing and properties of compression

molded long-fiber thermoplastics. There are three required inputs for Cadpress-TP, the

discretized model, the material properties, and the processing parameters. The processing

parameters, along with their default values, are given in Table 2.

Table 2 Processing parameters Parameter Default value Units Mold closing velocity 10 mm s-1 Mold temperature (top) 60 oC Mold temperature (bottom) 60 oC Idle time (to place charge) 0 s Ambient temperature 20 oC Cooling time 40 s Heat transfer coefficient (air) 5.00E-06 W mm-2 K-1 Heat transfer coefficient (mold) 0.002 W mm-2 K-1 Maximum press force 1000 kN Boundary conditions Geometry dependent Charge location Geometry dependent Charge preorientation 0.318, all 25 layers

Page 44: Bartus 2003_Thesis

29

The component chosen for the study is a tab plaque flow tool with a flat insert

located at Southern Research Institute Composites Manufacturing Center, Birmingham,

AL. In order to create the finite element mesh, a solid model was first constructed in

Pro/ENGINEER (Pro/E). The model was simplified by removing small radii and drafts.

Next, the flow tool model was discretized using both HyperMesh 5.1 and ANSYS 5.7

with several different mesh densities and was imported into Cadpress-TP. Altair

HyperMesh is a powerful finite element pre- and post-processor that enables generation

of finite element and finite difference models for engineering simulation and analysis.

The Cadpress software requires an elastic triangular three-node element with six

degrees of freedom and midplane nodes (ANSYS: SHELL 63). Optimizing the mesh

density is essential because CADPRESS-TP is computationally intensive. In the final

discretized model, a mesh density of 0.0356 elements mm-2 (5446 elements with an area

of 1.53 X 105 mm2) was used.

Process Variables

Mold temperature, Simulations 1–3. A simulation matrix was constructed in

order to investigate the effect of various processing conditions and material properties.

The selected parameters were looked at independently, holding all other variables

constant. The required processing parameter inputs are shown in Table 2. The default

values are approximately those of GMT, PP/40 wt. % E-glass. The process parameters

selected for study, namely mold temperature and charge location, can be readily

controlled in practice.

Page 45: Bartus 2003_Thesis

30

A temperature gradient between the male and female mold halves is often

maintained in order to control the shear edge through thermal expansion. For the

material parameter simulations, the mold temperatures were set at 70oC and 60oC for the

male and female mold halves, respectively. In the mold temperature effect simulations,

the temperatures of the top and bottom molds were made equal. Intuitively, increasing

the mold temperature will aid in mold filling but may actually increase the cycle time

through an increase in cooling time. Quantitative analysis of this type is difficult to

ascertain without the aid of modeling. Simulation 2 was the control run, with a default

mold temperature of 60°C, while in Simulations 1 and 3, the mold temperatures were

30°C and 90°C, respectively. The effect of mold temperature on flow fronts, the time

and pressure required for consolidation, and shrinkage and warpage in the consolidated

part were considered.

Charge location, Simulations 4-8. The charge location is the area in which the

fiber/matrix melt is placed on the compression-molding tool just prior to closing the

mold. The charge location can be easily modified, both in simulation and in practice, and

can have a significant effect on the mechanical and thermomechanical properties. In

addition, knowledge of the flow progression can aid in the prediction of knit lines (weld

lines), which have adverse affects on mechanical properties. The flow progression also

allows for prediction of areas for air entrapment. The different charge locations

examined are shown in Figure 7.

Page 46: Bartus 2003_Thesis

31

Figure 7 Charge locations. Charge location (a) was used in Simulations 1-3, (b) in Simulation 4, (c) in 5, (d) in 7, (e) in 6, (f) in 8.

(a) (c) (b)

(f) (e) (d)

Page 47: Bartus 2003_Thesis

32

The preorientation of the fibers in the charge, conversely, are dependent on the

plastication process. Typically, a single screw extruder is utilized; the charge

preorientation is mainly dependent on the type of plasticator. Therefore, a random

preorientation will be assumed for all 25 layers of the charge.

Boundary conditions. The boundary conditions are required to prevent free body

rotation in the shrinkage and warpage calculation. Default values of the remaining

variables are believed to have little influence on the process or are difficult to control in

practice; for example, the thermal conductivity of the mold is dependent on the mold

material.

Material Parameters. Tables 3, 4, and 5 provide the material property inputs and

their default values for the fiber, matrix, and the composite melt, as required for the

simulations.

Table 3 Fiber material parameters Parameter Default value Units Fiber aspect ratio 2000 Elastic modulus 7.30E+04 N mm-2 Coefficient of linear expansion 3.00E-06 oK-1 Poisson’s ratio 0.22 Thermal conductivity 8.50E-04 W mm-1 oK-1 Specific heat capacity 0.84 J g-1 oC-1 Density 2.52 g cm-3

Page 48: Bartus 2003_Thesis

33

Table 4 Composite material parameters Parameter Default value Unit No-flow temperature 150 oC Volumetric fiber content 0.16 Thermal diffusivity 0.08 mm2 s-1 Thermal conductivity 0.17 W m-1 oK-1 Carreau-Parameter 1 (Null viscosity) 4149 Pa s

Carreau-Parameter 2 (Infinite shear viscosity) 1 s

Carreau-Parameter 3 (Power law index) 0.599

Fiber interaction coefficient 0.14

Table 5 Matrix material parameters Parameter Default value Units Poisson’s ratio 0.35 1.89E-03 oK-1 -1.54E-06 oK-2 -3.05E-08 oK-3 Elastic modulus 1.68E+03 N mm-2 -2.21E+01 N mm-2 oC-1 7.98E-02 N mm-2 oC-2 -4.15E-05 N mm-2 oC-3 Coefficient of linear expansion 3.41E+04 cm3 bar g-1 5.48E+04 cm3 bar g-1 oK-1 2.38E+08 bar 3.07E+04 bar 1.17E-07 cm3 g-1 0.11 oK-1 2.98E-03 bar-1 PvT data 3.37E+04 cm3 bar g-1 1.12 cm3 bar g-1 oK-1 1.38E+03 bar 2.91E+04 bar Crystallization temperature 1.34E+02 oC 0.02 oC bar-1

Page 49: Bartus 2003_Thesis

34

Many of the parameters used in the flow simulation and fiber orientation

calculations are not solvable by analytical methods or are based on other calculations or

are solvable only by numeric methods [4].

Quantitative analysis is beneficial in understanding which properties are the most

critical in obtaining accurate results, as properties vary greatly depending on the material

purveyor. Other properties, such as fiber aspect ratio, can be chosen in the production of

the LFT pellets. Knowledge of material property effects could save a considerable

amount of time and money while optimizing the performance from a given material and

aid in obtaining accurate modeling results.

The material parameters chosen for the study are those thought to suit this

criterion and are as follows: the fiber interaction coefficient, melt viscosity, and the PvT

coefficients. These parameters were expected to have the greatest effect in material

parameter simulations.

Fiber interaction coefficient, Simulations 9 and 10. The fiber interaction

coefficient has not been determined accurately via analytical methods and is usually

determined empirically [24]. The effect of the interaction coefficient was evaluated by

varying it + 25% from the default value of 0.140. The effect on the degree of fiber

orientation was evaluated graphically as well as quantitatively for selected elements.

Melt viscosity, Simulations 11–17. The melt viscosity is particularly difficult to

determine experimentally. This is due not only to its non-Newtonian, non-isothermal

nature, but also because standard methods of measuring it produce in situ fiber length

Page 50: Bartus 2003_Thesis

35

degradation, thereby decreasing the viscosity [8]. Equation 19 gives the viscosity as a

function of temperature and shear rate. Noting only the shear-rate dependency of the

viscosity, it becomes a function of three parameters: the zero shear or null viscosity, the

infinite shear viscosity, and the power law index. The null viscosity was varied + 25% of

the default value given in Table 5. The infinite shear viscosity and power law index were

varied + 10% of the default values given in Table 5. The time and pressure for

consolidation and flow front effects were investigated in the study of melt viscosity.

PvT parameters, Simulations 18 and 19. Due to the complicated nature of

thermoplastic composites in the filling stage of compression-molding, one must consider

the influence of the temperature and pressure of a material during flow and in the

calculation of shrinkage and warpage in order to determine if the material is frozen. This

requires knowledge of the glass transition and crystallization temperatures with respect to

their pressure dependency. The coefficient of thermal expansion is also dependent on

transient temperature and pressure. This information is obtained from a PvT diagram.

Unfortunately, PvT data is very limited and rather difficult to obtain, particularly when

considering the wide number of matrix materials available. This made the influence of

the PvT coefficients on processing of special interest.

All four PvT coefficients were varied + 25% of their default values given in Table

4. The results were examined in terms of flow front effects, pressure, and the time

required to consolidate, as well as nodal pressure.

Page 51: Bartus 2003_Thesis

36

Simulation verification, Simulation 20. A qualitative relationship between

fracture paths of impacted specimens and the predicted fiber orientation was used to

verify a simulation, representing the tab plaque component produced, which is discussed

in material processing. The results are discussed in detail in the BOI results and

discussion. In the process simulation, all the default variables were used, with the

exception of mold temperature (60oC and 70oC for the male and female mold halves,

respectively); charge shape and location, which is shown in Figure 8; and the melt

temperature (230oC).

Figure 8 Charge location for the simulation verification study and for simulations 9-20. The charge shape and location were chosen as best to represent the processing conditions outlined in material processing.

Charge location

Page 52: Bartus 2003_Thesis

38

RESULTS AND DISCUSSION: PROCESS MODELING

A parametric study of processing and material property effects was conducted

based on Cadpress-TP compression-molding simulations of LFT in order to determine

the sensitivity of said parameters using the flow tool geometry, which was shown in

Figure 8. The graphic representations of selected results are given in Figures 8-46 at the

end of the chapter.

Process Parameter Results, Simulations 1-8

Mold temperature effects, Simulations 1-3. In the study of mold temperature

effects, Simulation 2 was the control run, with a default mold temperature of 60°C; in

Simulations 1 and 3, the mold temperatures were 30°C and 90°C, respectively. All other

variables given in Tables 3, 4, and 5 were held constant. The charge location shown in

Figure 8(a) was used in all three runs.

In Simulation 1, the force required for consolidation increased by approximately

100 kN (4% increase) in contrast to the control run, while in Simulation 3, the force

decreased by 200 kN (8% decrease). The greater decrease in force for an equivalent

increase in temperature (30oC increase) is attributed to the non-isothermal temperature

shift coefficient used in the calculation of the melt viscosity. The time required to fill the

mold cavity remained roughly the same. The mold temperature effected the deformation

after ejection from the mold. The least amount of deformation occurred in the 60°C mold

temperature (1.653 mm), and the greatest deformation in the low temperature (30°C)

Page 53: Bartus 2003_Thesis

39

simulation (3.317 mm). Only a slight increase in deformation was noted in the high mold

temperature simulation (1.824 mm). No significant effects were seen in the flow fronts

or fiber orientations. For the geometry used in Simulations 1-3, a 30oC increase in

temperature yielded the greatest effect on processing (two-fold decrease in consolidation

force), in contrast to an equivalent decrease in temperature from the control run. The

component warpage was minimized at the control run mold temperature of 60oC.

Effect of charge location, Simulations 4-8. Five different charge locations were

investigated in order to determine their effect on the force and time required to

consolidate, and to determine locations for knit lines. The different charge locations

investigated are shown in Figure 8(b)-(f). The maximum force required for consolidation

remained approximately the same, with slight variations in the force versus time profiles.

These small variations are not likely to have a drastic influence in processing. However,

significant differences in flow fronts were noted. The formation of knit lines can be seen

in Figures 13 and 14 for charge locations (d) and (f). The extent of knit lines in Figure

13 was unexpected, as opposed to multiple charge configurations in which knit lines are

inevitable. Knit lines are undesirable, since they have been shown to exhibit adverse

affects on the mechanical properties and aesthetics of the final component.

Material Processing Effects, Simulations 9 – 20.

The flow progression in Simulations 9-20 was essentially the same in all cases.

Typical flow fronts at 50% and 90% filling for the respective simulations are shown in

Figures 15-16.

Page 54: Bartus 2003_Thesis

40

Fiber interaction coefficient, Simulations 9 and 10. The fiber interaction

coefficient is difficult to ascertain, making its influence on the flow simulation important.

For the geometry and material properties used in the study, very little effect from the

fiber interaction coefficient was noted. The numeric values used in the study was given

in Table 6, along with the maximum nodal pressure in Table 7. The software is unlikely

to be able to resolve minute differences in the nodal pressure. More intriguing, however,

was the lack of an effect on the fiber orientation, shown graphically in Figures 30(a)-(c)

and quantitatively in Figures 32-40. The elements selected are shown in Figure 31.

Virtually no distinction can be made between the simulations. One would expect a

decrease in fiber orientation with increasing fiber interaction coefficient, which inhibits

the ability of a fiber to rotate within the melt. The lack of difference may be due to the

rather simple geometry investigated and the large ratio of the charge area to the mold area

(roughly 26%).

Table 6 Fiber interaction coefficient parameters investigated Simulation Fiber interaction coefficient Control 0.140 Interaction low 0.105 Interaction high 0.175

Table 7 Fiber interaction coefficient study results Simulation Maximum nodal pressure (MPa) Control 12.29 Interaction low 12.29 Interaction high 12.28

Page 55: Bartus 2003_Thesis

41

Effect of melt viscosity, Simulations 11 – 17. Three parameters are required to

model the non-Newtonian nature of polymers in the flow simulation: the zero shear or

null viscosity, the infinite shear viscosity, and the power law index. Each was

investigated independently using the values given in Table 8. The values obtained in

Table 9 are from the force versus time plots in Figures 17-28. The results for each

simulation are shown in Table 9 along with Figures 41-46, which are given at the end of

the chapter.

In all the simulations, the consolidation force, estimated from the force versus

time plots in Figures 20-28, remained a constant 1.17 MN. The maximum nodal pressure

Table 8 Melt viscosity parameters investigated

Simulation Zero shear viscosity (Pa s) Infinite shear viscosity (s) Power law index

Control 4149 1.00 0.599 Zero shear low 3112 1.00 0.599 Zero shear high 5186 1.00 0.599 Infinite shear low 4149 0.90 0.599 Infinite shear high 4149 1.10 0.599 Power index low 4149 1.00 0.539 Power index high 4149 1.00 0.659

Table 9 Melt viscosity parameter study results

Simulation Maximum nodal pressure (MPa)

Consolidation force (MN) Consolidation time (s)

Control 12.29 1.17 1.49 Zero shear low 12.37 1.17 1.19 Zero shear high 12.26 1.17 2.10 Infinite shear low 12.26 1.17 1.67 Infinite shear high 12.31 1.17 1.35 Power index low 12.17 1.17 2.35 Power index high 12.80 1.17 1.15

Page 56: Bartus 2003_Thesis

42

varied slightly, from 12.17–12.80 Mpa; however, the time for the melt to consolidate

varied considerably. The most significant effect was seen in the melt viscosity study

based on the power law index variations. The time to consolidate increased with a

decreasing power law index as shown in Table 9, in which the two extremities varied by

1.2 s. This may seem insignificant at first, but in a 500,000-part run, typical of an

original equipment manufacture (OEM), it would result in an increased manufacturing

time of approximately 167 h. Figures 26, 27, and 28 showing the force versus time

consolidation plots, indicate that the force required for consolidation increases faster as

the power law index decreases, which is expected from the non-Newtonian viscosity

equation (19).

The infinite shear viscosity effect seemed the least significant in the study. This

could be due to the relatively low shear rate used, based on a linear mold closing velocity

of 10 mm s-1, which would also indicate why the zero shear or null viscosity had a

slightly greater significance. In the zero shear viscosity study, the slope of the force

versus time consolidation plot decreased with increasing zero shear viscosity, shown in

Figures 20, 21, and 22. In the high zero shear simulation (5186 Pa s), the slope of the

force versus time plot is almost vertical as shown in Figure 22.

The fiber orientation was relatively unaffected in the zero shear viscosity study.

A slightly higher degree of orientation can be seen in the low zero shear study, Figure 44.

This may be due to an increase in mobility for the fibers to orient themselves in the flow

resulting from a decrease in melt viscosity. There was virtually no difference in the

graphical fiber orientations seen in the infinite shear viscosity study in Figure 45, again

most likely attributed to the low shear rate used in the simulations. The power law index

Page 57: Bartus 2003_Thesis

43

inquiry exhibited some variation in the graphical fiber orientation shown in Figure 46.

The highest degree of relative fiber orientation was seen in the high power law index

simulation, followed by the control run and then the low power law index simulation. In

all cases, as the melt viscosity increased, the fiber orientation decreased.

PvT parameter effects, Simulations 18 and 19. The PvT parameters had the least

significant effect in the material parameter study. The parameters investigated are given

in Table 10, with the respective results in Table 11. The results indicate no prominent

effects in the processing or nodal pressure. No effect from the PvT parameters study was

seen in the fiber orientation either. This may be because the PvT parameters are used

primarily in the warpage calculation, which was not investigated in this case. The PvT

parameters can be used to determine whether a given layer is frozen. However, the

Table 10 PvT parameters investigated

Simulation Coeff. 1 (cm3 bar g-1)

Coeff. 2 (cm3 bar g-1 oK-1) Coeff. 3 (bar) Coeff. 4 (bar)

Control 33709 1.12 1379 29122 PvT low 25282 0.84 1035 21842 PvT high 42136 1.40 1724 36402

Table 11 PvT parameter study results

Simulation Maximum nodal pressure (MPa)

Consolidation force (MN)

Consolidation time (s)

Control 12.29 1.18 1.5 PvT low 12.29 1.18 1.5 PvT high 12.29 1.18 1.5

Page 58: Bartus 2003_Thesis

44

crystallization temperature can be used by itself without taking into consideration its

pressure dependency. The Cadpress Theory manual does not discuss this in detail.

Control run, Simulation 20. The results from the control run were used as a base

line for all of the material parameter simulations. The process and material parameters

used are given in Tables 2-5, and results are given in Figures 15-17, 20, 23, 26, 30(a), 32,

35, 38, 41, 44(a), 45(a), and 46(a). The fiber orientation results were then used to verify

the control run simulation in results and discussion: blunt object impact, through a

qualitative comparison between the predicted fiber orientation and the failure modes of

LFT subjected to BOI.

Page 59: Bartus 2003_Thesis

45

Figure 10 Flow fronts at 50% filling for Simulation 4.

Figure 12 Flow fronts at 50% filling for Simulation 6.

Figure 11 Flow fronts at 50% filling for Simulation 5.

Figure 9 Typical flow front at 50% filling for Simulations 1–3.

Page 60: Bartus 2003_Thesis

46

Figure 13 Flow fronts for 80% filling showing the formation of knit lines in Simulation 7.

Figure 14 Flow fronts for 95% filling showing the formation of knit lines for two charges placed in Simulation 8

Page 61: Bartus 2003_Thesis

47

Figure 15 Typical flow front for Simulations 9-20 showing 50% filling.

Figure 16 Typical flow front for Simulations 9–20 showing 90% filling.

Page 62: Bartus 2003_Thesis

48

Figure 17 Force versus time for the control simulation, fiber interaction coeff. = 0.140.

Figure 19 Force versus time diagram for the fiber interaction coefficient low (0.105) simulation.

Figure 18 Force versus time diagram for the fiber interaction coefficient high (0.175) simulation.

Page 63: Bartus 2003_Thesis

49

Figure 22 Force versus time for the high null viscosity (5186 Pa s) simulation.

Figure 21 Force versus time for the low null viscosity (3112 Pa s) simulation.

Figure 20 Force versus time for the control simulation, null viscosity = 4149 Pa s.

Page 64: Bartus 2003_Thesis

50

Figure 25 Force versus time for the high infinite shear viscosity (1.10 s) simulation.

Figure 24 Force versus time for the low infinite shear viscosity (0.90 s) simulation.

Figure 23 Force versus time for the control simulation, infinite shear viscosity = 1.0 s.

Page 65: Bartus 2003_Thesis

51

Figure 28 Force versus time for the high power law index (0.569) simulation.

Figure 27 Force versus time for the low power law index (0.539) simulation.

Figure 26 Force versus time for the control simulation, power law index = 0.599.

Page 66: Bartus 2003_Thesis

52

b

a

Figure 29(a) and (b) Force versus time plots for low and high PvT parameter simulations, respectively.

Page 67: Bartus 2003_Thesis

53

Figure 30 Graphical representation of the fiber orientation in the (a) control run (0.14), (b) low fiber interaction coefficient (0.105) simulation, and (c) the high fiber interaction coefficient (0.175) simulation.

b

a

c

Graphical representation of an element indicating a close to random fiber orientation

Graphical representation of an element in which the main fiber orientation is indicated by the arrow

Page 68: Bartus 2003_Thesis

54

Figure 31 Illustration of selected element locations for the fiber orientation distribution function comparison: (a) element 39, (b) element 29, and (c) element 937. The locations were chosen as intermediate locations of interest.

a c b

Page 69: Bartus 2003_Thesis

55

Figure 32 Fiber distribution function of element 39 in the control simulation, fiber interaction coeff. = 0.140.

Figure 33 Fiber distribution function of element 39 in the low fiber interaction coefficient (0.105) simulation.

Figure 34 Fiber distribution function of element 39 in the high fiber interaction coefficient (0.175) simulation.

Page 70: Bartus 2003_Thesis

56

Figure 35 Fiber distribution function of element 29 in the control simulation, fiber orientation coeff. = 0.140.

Figure 36 Fiber distribution function of element 29 in the low fiber interaction coefficient (0.105) simulation.

Figure 37 Fiber distribution function of element 29 in the high fiber interaction coefficient (0.175) simulation.

Page 71: Bartus 2003_Thesis

57

Figure 38 Fiber distribution function of element 937 in the control simulation, fiber orientation 0.140.

Figure 39 Fiber distribution function of element 937 in low fiber interaction coefficient (0.105) simulation.

Figure 40 Fiber distribution function of element 937 in high fiber interaction coefficient (0.175) simulation.

Page 72: Bartus 2003_Thesis

58

Figure 41 Maximum nodal pressure (Pa) for the control simulation, where the maximum nodal pressure is 12.29 MPa, null viscosity = 4149 Pa s.

Figure 43 Maximum nodal pressure (Pa) for the high null viscosity (5186 Pa s) simulation, where the maximum nodal pressure is 12.26 MPa.

Figure 42 Maximum nodal pressure (Pa) for the low null viscosity (3112 Pa s) simulation, where the maximum nodal pressure is 12.37 MPa.

Page 73: Bartus 2003_Thesis

59

a

b

c

Figure 44 Graphical representation of the fiber orientation in the (a) control run (null viscosity = 4149 Pa s), (b) low zero shear viscosity (3112 Pa s) simulation, and (c) the high zero shear viscosity (5186 Pa s) simulation.

Graphical representation of an element indicating a close to random fiber orientation

Graphical representation of an element in which the main fiber orientation is indicated by the arrow

Page 74: Bartus 2003_Thesis

60

c

Figure 45 Graphical representation of the fiber orientation in the (a) control run (infinite shear viscosity = 1.0 s), (b) low infinite shear viscosity (0.90 s) simulation, and (c) the high infinite shear viscosity (1.10 s) simulation.

b

a

Graphical representation of an element indicating a close to random fiber orientation

Graphical representation of an element in which the main fiber orientation is indicated by the arrow

Page 75: Bartus 2003_Thesis

61

a

b

c

Figure 46 Graphical representation of the fiber orientation in the (a) control run (power law index = 0.599), (b) low power law index (0.539) simulation, and (c) the high power law index (0.659) simulation.

Graphical representation of an element indicating a close to random fiber orientation

Graphical representation of an element in which the main fiber orientation is indicated by the arrow

Page 76: Bartus 2003_Thesis

62

SUMMARY AND CONCLUSIONS: PROCESS MODELING

Eight simulations were used to investigate the effect of mold temperature (three)

and charge location (five) in the processing of LFT composites for the tab-plaque flow

tool geometry.

Mold temperature had a significant affect on processing. A 30oC increase in mold

temperature (mold temperature = 90oC) decreased the force required for consolidation by

8%, whereas a 30oC decrease in mold temperature (mold temperature = 30oC) increased

the force required for consolidation by only 4%. This is attributed to the non-isothermal

dependency of the viscosity, which is accounted for by the temperature shift coefficient.

A 60oC mold temperature simulation yielded the least warpage in the consolidated

component, followed closely by the 90oC mold temperature simulation (~10% increase).

In the 30oC mold temperature simulation, warpage increased by approximately 90%. The

warpage increase with decreasing mold temperature is thought to arise from an increase

in the high thermal gradient between the polymer melt and the mold.

The charge location study indicated that flow progression could be modeled and

optimized in order to avoid weld (knit) lines. The weld lines in this simple geometry

could be avoided by using a slight variation in charge placement.

Eleven simulations were conducted to understand material property effects on the

process simulations of the tab-plaque flow tool geometry. Three simulations were

conducted to determine the effect of the four parameter PvT coefficients, three

Page 77: Bartus 2003_Thesis

63

simulations investigated the fiber interaction coefficient, and six simulations were used to

determine the effects of the non-Newtonian, non-isothermal viscosity equation.

Negligible effects were seen in varying the fiber interaction coefficient and the

PvT parameters studies. The PvT parameters are thought to play a more significant role

in the warpage calculation, which was not studied in detail. The low dependency on the

fiber interaction coefficient may be attributed to the relatively small distance required for

flow progression. This decreases the time for the fibers to reorient in the flow. Also, a

random pre-orientation of the charge was assumed. This could influence the tendency for

fibers to orient depending on the direction of the fiber pre-orientation relative to the flow

progression

Melt viscosity played a substantial role on the processing effects and, to a lesser

degree, the fiber orientation. The power law index had the most significant effect in the

melt viscosity study. Therefore, it is the most important value to determine

experimentally. As the power law index decreased (increasing the viscosity), the time for

the melt to consolidate increased and fiber orientation decreased. This may be due to an

increase in the melt viscosity. The zero shear viscosity also had a significant effect on

processing, and to a lesser degree, the fiber orientation. The same dependency on melt

viscosity seen in the power law index study was observed in the zero shear viscosity

study. As melt viscosity increases, the time required for consolidation increases and fiber

orientation decreases. The infinite shear viscosity study resulted in the least significant

effects. This may be attributed to the low shear imposed on the melt in the simulation.

The shear rate is dependent on the mold-closing velocity. The mold-closing velocity

Page 78: Bartus 2003_Thesis

64

used in all simulations was 10 mm s-1. Forty to fifty mm s-1 may have been a more

appropriate choice for the mold-closing velocity.

The simulation matrix illustrated how a process might be optimized before a

component is ever produced. Practical goals in obtaining material parameters were

established for the geometry studied. It is important to note that geometries of increasing

complexity may have different dependencies on the material properties that were

investigated. In addition, the synergistic effects between properties were not taken into

account, since only one parameter was varied in a given simulation.

For the simulation matrix in this study, varying the PvT parameters + 25% from

their default value had very little effect on the results. The same was noted in the case of

the fiber interaction coefficient, which was varied + 10%. For the material and

processing parameters used with the geometry considered, the simulations were relatively

insensitive to the PvT parameters and the fiber interaction coefficient. However, the

three-parameter melt viscosity (null viscosity + 25%, infinite shear viscosity + 10%, and

power law index + 10%) had a considerable effect on the processing results and also an

effect on the mechanical properties in the consolidated component, making an accurate

characterization of the melt viscosity of greatest importance in the material parameter

study. In all cases in the melt viscosity parameter study, the most significant effect was

seen in the fiber orientation and the time for the component to consolidate. As the

parameters were varied in such a way that the viscosity would increase, the degree of

fiber orientation decreased and the time for the component to consolidate increased.

Page 79: Bartus 2003_Thesis

65

LITERATURE REVIEW: BLUNT OBJECT IMPACT

Categorization of impact

A considerable amount of work has gone into studying transverse impact on

composite structures [34-41]. The authors of this work [34-41] have focused primarily

on continuous fiber composites with thermoset matrices. The work done in thermoplastic

matrices has been mainly been limited to PEEK, an aerospace material. Most of the

literature available pertains either to Charpy and Izod tests or low or high velocity

impact.

Charpy and Izod, which are pendulum-type test configurations, are constrained in

terms of specimen size, impact direction, and boundary conditions [5]. Moreover,

Charpy and Izod testing, especially in the notched configuration, force the specimen to

fail at a predetermined area rather than along the weakest plane. This could skew results

because of the highly anisotropic nature of LFTs. The velocity range, impactor

geometry, and impactor mass should reflect the type of threats a component may

encounter in service. In this case, a more representative test would be in the intermediate

velocity range with a blunt object impactor. This would simulate the effect of stones and

other debris a vehicle might encounter at highway speeds. In addition, the importance of

this test methodology lies not only in the characterization of debris hits, but also energy

dissipation and failure mechanisms under high loading rate.

It is important to define low, intermediate, and high velocities, as the use of these

terms tend to vary among authors [34]. Some authors contend that high velocity refers to

Page 80: Bartus 2003_Thesis

66

conditions resulting in complete perforation of the target. However, this phenomenon

can be readily observed in drop tower testing in which the maximum obtainable velocity

is typically less than 10 m s-1. For the remainder of this thesis, the term low velocity will

be reserved for velocities less than 10 m s-1 with large mass impactors and large target

deformations. The definition of high velocity impact will follow that of Abrate, in which

the ratio between the impactor velocity and the transverse compressive wave velocity is

greater than the maximum strain to failure in that direction [34]. Damage induced by

high velocity impact is typically introduced by the first few compression waves though

the thickness when the global plate motion has not been established [34].

Intermediate velocity will be considered to fall between the low and high velocity

regimes. The main difference is that, depending on the projectile mass, large deformation

may occur in the intermediate velocity range, particularly in the case of massive

projectiles, but may differ from low velocity impacts with regard to loading rate and

momentum effects.

Impactor mass and geometry. The impactor size, shape, mass, material, and angle

of incidence all have a strong influence on the response of the specimen [34]. A blunt

object will be defined as a large diameter projectile that will emulate the effects of debris

to characterize energy absorption upon impact. Jenq et al. [40] found that the momentum

transfer to graphite/epoxy targets by flat impactors was about four times greater than that

of a sharp or conically tipped impactor when the penetrator was fired at high velocity.

Cantwell and Morton [41] studied the effect of projectile mass on energy absorption in

composites while maintaining the impactor size and shape. It was found that varying the

Page 81: Bartus 2003_Thesis

67

projectile mass had a significant effect on the resulting damage. Cantwell and Morton

[41] suggest that lighter projectiles are more damaging to the overall load-bearing

capacity of the composite because the incident energy is dissipated over a very small area

immediately adjacent to the point of impact. This may or may not be the case with

intermediate velocity impact, depending on the target material, boundary conditions,

projectile shape, and projectile mass.

Impact Energy

The energy absorbed in the system, Eabsorbed, can be described as having two

components: the energy absorbed in creating damage in the specimen, Edamage, and the

energy absorbed by the system through vibration, heat, elastic response of the specimen,

and elastic behavior of the projectile or supports, ESL, equation (37):

where the Edamage term includes the energies associated with the specimen indentation,

matrix damage, fiber breakage, fiber debonding, and fiber pull-out. The energy absorbed

in the system can be equated to the energy put into the system, e.g. the kinetic energy

(KE) of the impactor shown in equation (38):

where m is the projectile mass, and Vo and Vf are the initial and residual projectile

velocities. The energy spent in the elastic response of the projectile and supports has

been shown to be small and will not be considered, simplifying the equation (38) to yield

equation (39) [35].

Eabsorbed = KE = ½ m Vo 2 - ½ m Vf 2 +ESL (38)

Eabsorbed = Edamage + ESL (37)

Page 82: Bartus 2003_Thesis

68

The highest impact energy is absorbed at the ballistic limit of the material in

which the projectile imbeds itself in the specimen. However, embedding a projectile in

thin specimens is unlikely, due to an inadequate area between the target and projectile for

fictional forces to act upon. Since the residual velocity of a non-penetrating projectile is

small, it will not be included in the energy balance. Therefore, the Vf term is not part of

the energy balance unless the projectile perforates the specimen and has a detectible exit

velocity. Otherwise, the initial velocity is assumed equal to the critical velocity, Vo = Vc.

This is similar to the V50 ballistic limit, which is defined as the projectile having a 50%

probability of completely penetrating the specimen at the critical velocity with a

deviation in velocity less than 40 m s-1 [34]. Given the low critical velocity of blunt

object impact, a more conservative deviation is appropriate. Since a standard test method

does not exist for blunt object impact, the critical velocity will be considered on the basis

of a standard deviation in velocity no greater than 10 m s-1, in which 50% of the

projectiles do not perforate the specimen.

Eabsorbed = KE = ½ m Vo 2 - ½ m Vf 2 (39)

Page 83: Bartus 2003_Thesis

69

EXPERIMENTAL PROCEDURE: BLUNT OBJECT IMPACT

Impact Test Apparatus

An understanding of LFT behavior while it undergoes high energy, high strain

rate impact is essential in promoting its use in automotive and other markets. A gas gun

was used to propel a variety of projectiles as shown in Table 12, with a velocity range of

approximately 45 m s-1 to 140 m s-1 to simulate the effects of blunt object impacts. The

wide range of projectile mass helped in determining velocity effects, while the blunt

projectile shape aided in characterization of perforation mechanisms.

Since an impact apparatus was unavailable, the design and construction had to be

undertaken before testing could proceed. The objective in the design phase was to

encompass velocities from approximately 10 m s-1 to 400 m s-1 for projectile masses

ranging from 2 g up to 500 g with a maximum projectile diameter of 38 mm. The test

Table 12 Blunt object impact projectile types Projectile Weight (g) Shape Material Dimensions (mm) Sabot 160 Flat Aluminum 37.71 (φ) x 50 Sabot 100 Flat Aluminum 37.71 (φ) x 50** Sabot 50 Flat UHMWPE 37.71 (φ) x 48 Sabot 50 Conical UHMWPE 37.71 (φ) x 55, 60° shoulder Sabot 25 Flat UHMWPE 37.71 (φ) x 37** Sabot 25 Conical UHMWPE 37.71 (φ) x 41, 60° shoulder** ** Denotes sabots in which material was removed from the center to reduce mass

Page 84: Bartus 2003_Thesis

70

apparatus required a high degree of flexibility so that a wide range of tests can be

performed.

Pneumatic propulsion or gas guns have been successfully implemented for such

tests in the past and are more prevalent than propellant propulsion guns firing standard

cartridges [34, 37, 38]. A driving force for this trend is the flexibility in directly varying

the projectile velocity by working fluid pressure, as opposed to propellant propulsion, or

center fire guns, where the powder charge is metered to alter the projectile velocity. In

addition, propellant guns fire projectiles that should be classified as high velocity.

Another drawback with chemical propulsion is that projectile types are limited in terms

of size, shape, mass, and materials. The largest round available in standard center fire

cartages is the 0.50 caliber Browning machine gun (BMG), which would not accurately

emulate the effect of blunt object impacts. Clearly, the logical choice for a test apparatus

is a gas gun, consisting of a high-pressure fluid source, regulator, pressure transducer,

pressure vessel, valve, barrel, velocity sensors, capture chamber, specimen fixture, and

projectile arrest.

Firing valve and pressure vessel. The firing valve is a critical component in the

test apparatus since it controls the working fluid flow. The crucial criteria in the valve

design are fast actuation, ability to operate at working pressures in excess of 1.70 MPa,

high flow rate, safe operation, and repeatability. Several candidates are commercially

available in the form of pneumatic or solenoid-actuated ball and butterfly valves.

However, a poppet valve similar to the type used in air rifles meets the aforesaid criteria

and can be manufactured in-house cost effectively. The disassembled poppet valve is

Page 85: Bartus 2003_Thesis

71

shown in Figure 47. The valve outside diameter is 45 mm and is sealed to the valve

guide by two rubber O-rings. The valve operates at a pressure range of 27.5 kPa and 1.72

MPa. The valve is guided by the center carrier and is held in place by a sear pin. The

valve body and center carrier guide the sear pin.

The pressure vessel limits the upper pressure range to approximately 1.70 MPa.

The pressure vessel was welded at North Dakota State University, Fargo, ND, and was

non-destructively tested with X-ray radiography at Midwest Industrial X-ray Inc., Fargo,

ND. Using a conservative yield strength for mild steel of 210 MPa, the factor of safety is

a

a

b

c

d

e

Figure 47 Disassembled poppet valve: (a) two halves of the valve body, (b) valve center carrier, (c) valve, (d) barrel union, and (e) front valve face.

Page 86: Bartus 2003_Thesis

72

five, based on an operating pressure of 1.70 MPa and the assumption of a thin-walled

pressure vessel, using equation (40) given by Shigley and Michke:

where σt,max is the maximum principle tangential stress, Pmax is the maximum working

pressure, di is the inside diameter of the pressure vessel, and t is the wall thickness [42].

Firing mechanism. Pulling the sear pin, which retains the valve and guide, opens

the valve. A four-bar linkage is used to gain mechanical advantage and is actuated by a

solenoid. Nitrogen, the working fluid in the pressure vessel, forces the valve open. The

solenoid is operated via remote control containing an ‘arm/safe’ two position toggle

switch and a ‘fire’ push button switch for safety. A mechanical safety is located on the

four-bar linkage, preventing the sear pin from being pulled prematurely. The firing

sequence requires that the mechanical safety be removed, and then, once the operator is

in a safe location, the ‘arm’ switch is thrown. A red LED indicates that the remote

trigger is hot. Finally, the fire switch is depressed, actuating the solenoid and opening the

valve.

Gas gun carriage and barrel. The valve, pressure vessel, and firing mechanism

are mounted on a carriage, allowing them to translate such that the barrel can be loaded

from the breach. The valve is coupled to the barrel with a union. The barrel is mild steel

seamless tubing, 4.57 m long with a 37.49 mm inside diameter and 51.31 mm outside

ttdP i

t 2)(max

max,+

=σ (40)

Page 87: Bartus 2003_Thesis

73

diameter with 50.8 μm true indicated roundness (TIR). A muzzle brake with sixteen

15.875-mm diameter holes + 15 μm was implemented 61 cm from the barrel end to

exhaust the working fluid before it enters the capture chamber and to reduce recoil. The

gas assembly and carriage are shown in Figure 48.

Pressure data acquisition. Since the velocity is particularly sensitive at lower

pressures, accurate and repeatable pressure readings are critical. A Rosemont® 2088

Smart pressure transducer was used for this task. The operating pressure range is from 0

- 1.034 x 106 Pa gauge (0 - 150 psig) with a resolution of 0.690 Pa gauge (1 x 10-4 psig).

Figure 48 Gas gun assembly on the carriage showing the pressure vessel, pressure transducer, firing valve, and firing mechanism.

Page 88: Bartus 2003_Thesis

74

The transducer can be readily interfaced with a personal computer serial port using the

VIATOR® RS 232 Interface. The transducer also has a digital readout for taking manual

measurements, thereby negating the need for a PC. It was calibrated for a pressure range

of 0 - 1034 kPa (0 - 150 psig) in accordance with the user manual.

Velocity data acquisition. Projectile velocity is a critical measurement in impact

testing. The energy of an impact is a function of the velocity squared, so precise

measurement is imperative. The most common way to measure projectile velocity is via

chronographs, which consist of two photoelectric sensors separated by a known distance

with a timer (chronograph) to record the event. When the first photoelectric sensor

detects an object, the timer is tripped and stopped when the second photoelectric sensor

detects an object. Knowing the time for a projectile to travel a given distance allows for

the calculation of velocity.

Two CED Millennium chronographs were used for velocity acquisition. An

independent German test lab reported that the CED Millennium chronographs were

within 0.2% (99.8%) of the true laboratory recordings [43]. The first chronograph was

placed just in front of the sample holder to measure the incident velocity and the other

chronograph just behind the sample holder to measure residual velocity. Both are

mounted on the outside of the capture chamber to prevent damage from the projectile or

impact debris. A light source is placed on the other side of the chamber with 25.7 mm (1

in) polycarbonate windows separating them as shown in Figure 51. Light diffusers were

placed in front of the windows between the light sources and chronographs to alleviate

variations in lighting. Four 150-W flood lamps were used for the light source.

Page 89: Bartus 2003_Thesis

75

Calibration Curves. Plotting the working fluid pressure versus the velocity for a

given projectile and fitting a curve to the plot establishes a calibration curve. A typical

calibration curve is shown in Figure 49 for a 100 g sabot. A curve, in this case a second

order polynomial, is fitted to the data, allowing calculation of the required pressure to

obtain a given velocity. The calibration curves have also shown that, if a consistent

projectile diameter is maintained, the velocity will become a function of only the

projectile mass and pressure. Hence, the velocity of a new projectile can then be

predicted based on previous testing.

Velocity2 = -0.004*(P)2 + 1.3*(P) - 3.2

R2 = 0.9986

Velocity1 = -0.005*(P)2 + 1.5*(P) - 9.0

R2 = 0.9989

30

40

50

60

70

80

90

100

110

30 50 70 90 110 130Pressure (kPa)

Vel

ocity

(m s-1

)

Chronograph 2 Chronograph 1

Figure 49 Typical calibration curves for pressure versus velocity for the incident (1) and residual (2) velocity chronographs shown for a 100 g sabot. A second-order polynomial was fitted to the data with an R2 value approaching one where y is the velocity (m s-1) and x is the pressure (kPa). The y error bars indicate a 5% deviation

Page 90: Bartus 2003_Thesis

76

Capture chamber. A capture chamber is a control volume with a large access

door, which provides for velocity data acquisition, the sample holder, and the projectile

arrest. The components were designed and assembled using Pro-E. The overall

dimensions are 33 cm by 33 cm by 183 cm. The Pro-E assembly is shown in Figure 50,

and the finished capture chamber is shown in Figure 51.

An important requirement of the chamber is to safely arrest the projectile after

impacting the specimen. All of the structural materials used in the capture chamber

constructions of 6.35 mm (1/4 in.) thick mild steel with the exception of a 12.7 (1/2 in.)

kinetic deflector, which serves as a bullet trap. LSDYNA 3-D was used to verify whether

the chamber wall could safely contain a projectile. A flat 12.7 mm (1/2 in.), 16 g,

fragment-simulating projectile was used to impact of the center of a fully clamped 30.48

mm by 30.48 cm (12 in. by 12 in.), 6.35 mm (1/4 in.) thick mild steel plate. The impact

was assumed normal to the plate to determine the minimum impact velocity required for

perforation of the chamber walls in a worst-case scenario. No plastic deformation was

assumed in the projectile. The first velocity examined was 400 m s-1, and penetration

occurred. The velocity was decreased until the ballistic limit was determined at

approximately 330 m s-1. Since a normal impact on the walls is not possible, the capture

chamber provides adequate protection as impact resistance increases substantially as the

impact angle of incidence increases. This concept was used in the design of the kinetic

deflector, which was used as the projectile arrest as shown in Figure 50. The design

insured that, regardless of where a projectile made contact with the arrest, it would be

safely contained. Thick, 12.7 mm cold-rolled 1018 steel was used for the kinetic defector

and was welded as shown in the schematic in Figure 52.

Page 91: Bartus 2003_Thesis

77

Figure 50 Pro/E drawing of the capture chamber with the access door removed. The windows are backlit for the photoelectric sensors in the chronographs

Kinetic Deflector Barrel Sample Holder Mounts

Figure 51 Image of the capture chamber showing the (a) light bank, (b) barrel, (c) access door, (d) toggle clamps, (e) polycarbonate data acquisition windows, and the (f) incident and (g) residual velocity chronographs.

a

b

cf

d

e

g

Page 92: Bartus 2003_Thesis

78

A heavy-duty, 1.52 m long hinge allows full access to the capture chamber door.

Six 1780 N (400 lbf) toggle clamps, placed on three sides of the chamber, secure the door

before firing as shown in Figures 51(c) and 51(d).

Sample holder and boundary conditions. The boundary condition for testing was

simply supported on four sides, with the supports 160 mm apart. The simply supported

boundary condition allowed for rotation and was chosen to emulate the support

conditions LFT would see in service. The supports allowed 25.4 mm of deflection.

Sample preparation. A minimum of five samples were tested for each

configuration in an attempt to accurately determine the critical velocity. Three equal size

Projectile Path

Figure 52 Pro/E drawing of the kinetic deflector showing the path of the projectile.

Page 93: Bartus 2003_Thesis

79

samples (470.4 cm2) were cut from each of the LFT plaques produced for blunt object

impact testing, and the tabs were removed. The samples are shown schematically (with

the tabs still in place) in Figure 53.

The average weight and standard deviation for the three samples was 206.32 g

(S.D.= 1.358 g), 219.8 g (S.D.= 1.822 g), and 225.27 g (S.D.= 1.056 g) for the bottom,

center, and top samples, respectively. This yielded an average areal density of 4.39 g cm-

2, 4.67 g cm-2, and 4.79 g cm-2 for the bottom, center, and top samples. The areal density

is of interest because impact energy absorption is often described in terms of it. The

Figure 53(a) and (b) (a) Schematic of the top and side views of the tab plaque (not shown to scale) and the representative sections that were cut from it and (b) Pro/E isoperimetric drawing of the tab plaque.

600 mm

Top

Center

Bottom

240 mm

a b

Page 94: Bartus 2003_Thesis

80

variations in areal density are due to a thickness disparity caused by the lack of automatic

leveling on the compression molder.

Blunt object impact test matrix. A minimum of five samples was examined for

each of the 18 configurations in the test matrix outlined in Table 13.

The goal in the impact-testing phase was to establish whether a relationship

existed between the critical energy dissipation and impact mass, e.g. effect of projectile

velocity, the projectile geometry, and areal density.

The residual velocity due to rebound was measured using a Vision Research, Inc.

high-speed camera (Model: Phantom V5.0), at 14,000 frames second-1 with 40 μs

Table 13 Blunt object impact test matrix Projectile mass (g) Projectile geometry Areal density of specimen (g cm-2) 25 Flat 4.39 25 Flat 4.67 25 Flat 4.79 50 Flat 4.39 50 Flat 4.67 50 Flat 4.79 100 Flat 4.39 100 Flat 4.67 100 Flat 4.79 160 Flat 4.39 160 Flat 4.67 160 Flat 4.79 25 Conical 4.39 25 Conical 4.67 25 Conical 4.79 50 Conical 4.39 50 Conical 4.67 50 Conical 4.79

Page 95: Bartus 2003_Thesis

81

exposures. The maximum rebound velocity of a 100 g non-perforating sabot was

calculated to be approximately 8.4 m s-1. Since the residual velocity from a rebounding

projectile was small, it will not be included in the energy balance.

Page 96: Bartus 2003_Thesis

82

MATERIAL PROCESSING

Celstran®PP-GF40-03 Processing and Material Properties

Test plaques, with the geometry shown in Figure 53 were fabricated at Southern

Research Institute, Birmingham, Alabama, using Celstran® PP-GF40-03, 40 wt. % long

glass fiber-reinforced, chemically coupled, heat stabilized polypropylene manufactured

by Ticona. The color number is AF 3001 (natural), stock number GL0175, and lot

number W7650. The starting fiber length was 25.4 mm. The plasticator and 3.56 MN

(400(US)-ton) compression molder used were manufactured by C.A. Lawton Co. The

idle time was 5 seconds with a 30 second cooling time. This gave a cycle time of

approximately 60 seconds. The system pressure in the plasticator was 1.25 MPa, the

backpressure 2.07 MPa, the screw speed 9-10 rpm, and the melt temperature at the knife

was 232°C.

The mold temperatures, checked every 40 cycles, remained constant at 54°C and

76°C for the top and bottom mold halves respectively. The plaque temperature at the

time of ejection varied from 87.5°C to 90.0°C and from 45°C to 51°C five minutes after

ejection for an ambient temperature of 31-32°C. Twenty-four trial plaques formed before

making a 160 plaque run to ensure the system was at steady state. The plaque ID is as

follows: (year, month, date, lot number, plaque number) e.g. (020919-1-#). The first lot

of plaques was manufactured on September 19, 2002. The mechanical properties are

given in Table 14 from the Ticona website [44].

Page 97: Bartus 2003_Thesis

83

Table 14 Material properties of Celstran® PP-GF40-03 adopted from the Ticona website [44]. Property Method Value Units Density ISO 1183 1,210 kg m-3

Tensile modulus (1mm/min) ISO 527-2/1A 7,900 MPa Tensile stress at break ISO 527-2/1A 100 MPa Tensile strain at break ISO 527-2/1A 2.0 % Flexural modulus ISO 178 8,000 MPa Flexural strength ISO 178 175 MPa Charpy notched impact strength at 23°C ISO 179/1eA 20 kJ m-2

Page 98: Bartus 2003_Thesis

84

RESULTS AND DISCUSSION: BLUNT OBJECT IMPACT

Blunt Object Impact Results

LFT panels (PP/40% E-glass) were impacted in the intermediate velocity range of

42.8 m s-1–140.4 m s-1, with projectile masses varying between 23.7 g–160.7 g, and with

both conical and flat configurations. This range in projectile velocity and geometry will

simulate the effect of blunt objects, such as rocks and debris, traveling at highway speeds

and greater. This work also provides insight into high-strain-rate energy dissipation and

failure mechanisms, such as a structural composite may undergo in the event on an

automobile accident.

A minimum of five samples were tested for each eighteen configurations in an

attempt to determine the critical impact velocity (or ballistic limit). Three test specimens

of LFT PP/40 wt. % E-glass, shown in Figure 53(a), were considered, representing fiber

orientation effects, in addition to unintentional variations in areal density. The results

will be discussed in terms of projectile mass, projectile geometry, and areal density,

respectively.

Effect of projectile mass. In the first phase of testing, the effect of projectile mass

was examined by determining the critical impact velocity for three test specimens, top

(areal density = 4.789 g cm-2), center (areal density = 4.674 g cm-2), and bottom (areal

density = 4.386 g cm-2). The goal in this phase of testing was to establish if impact

velocity had an effect on energy dissipation and failure modes. Impact mass variation is

Page 99: Bartus 2003_Thesis

85

an established technique in varying strain rate and often used in instrumented low

velocity drop-weight testing [44]. The critical impact velocity versus projectile mass

results are given in Figures 54, 55, and 56 for the top, center, and bottom specimens, and

the statistical significance is shown in Tables 23, 24, and 25.

Impact velocity alone does not describe the impact properties of a material, but it

can affect the energy dissipated in an impact. It is shown only to illustrate the variations

in impact velocity with projectile mass. In all cases examined, the critical velocity

decreased exponentially as a function of projectile mass. Material failure requires

energy, and the remaining results will be represented as such.

The impact energy versus projectile mass is given in the subsequent Figures 57,

58, and 59 for same samples, calculated from the data shown in Figures 54, 55, and 56.

For the range investigated, no clear relationships were observed between impact energy

and projectile mass. This is shown in Figure 60, where a linear regression was fit to the

mean of all the impact energy versus projectile mass data. The mean critical energy

dissipation data and the numeric results and analysis for the top, center, and bottom are

given in Tables 15, 16, and 17 as a function of projectile mass and shape. Variations in

energy dissipated for the three different samples are attributed to deviations in areal

density and fiber orientation effects. The effect of orientation will be discussed in detail

at the end of the chapter.

Page 100: Bartus 2003_Thesis

86

Table 15 Numerical results and analysis for the impact data on the top specimen Projectile Mean (J) S.D. (J) 95% C.I. (J)25 g Flat 190.2 6.5 5.7 25 g Conical 127.7 5.9 5.2 50 g Flat 178.4 21.1 18.5 50 g Conical 134.3 12.9 11.3 100 g Flat 153.8 12.5 11.0 160 g Flat 177.0 15.9 11.8 Flat projectiles 175.0 19.0 11.8 Conical projectiles 131.0 10.1 4.2

Table 16 Numerical results and analysis for the impact data on the center specimen Projectile Mean (J) S.D. (J) 95% C.I. (J)25 g Flat 201.9 30.5 26.8 25 g Conical 118.6 15.5 15.2 50 g Flat 174.9 26.4 19.6 50 g Conical 126.1 16.4 14.4 100 g Flat 141.6 5.7 5.0 160 g Flat 179.9 4.5 3.6 Flat projectiles 174.9 28.2 18.4 Conical projectiles 122.8 15.5 6.3

Table 17 Numerical results and analysis for the impact data on the bottom specimen Projectile Mean (J) S.D. (J) 95% C.I. (J)25 g Flat 157.0 14.0 12.3 25 g Conical 114.0 11.1 10.9 50 g Flat 146.4 20.5 13.4 50 g Conical 110.5 13.4 11.7 100 g Flat 143.0 7.9 6.9 160 g Flat 160.5 7.9 7.0 Flat projectiles 151.3 15.8 6.1 Conical projectiles 112.0 7.7 7.7

Page 101: Bartus 2003_Thesis

87

Numeric Results and Analysis of the Impact Data

Numeric analysis was conducted on the velocity data. Since the energy

dissipation is a function of the velocity squared, the standard deviation is greater for the

energy data as opposed to the statistic significance of the experimental velocity data, as

shown in Tables 15, 16, 17, 23, 24, and 25. The largest standard deviation in the velocity

data was 9.8 m s-1, which corresponds to a standard deviation in the energy data of 30.5 J

for the center specimen impacted with the 25 g flat projectile. This standard deviation is

actually greater than that of all the flat and conical projectiles for a given sample, e.g. the

deviation of the flat projectiles on the center specimen was 28.2 J (Table 16). Moreover,

the deviation for the flat projectiles on the center specimen was approximately 33%

greater than the next highest deviation, corresponding to the average deviation for the flat

projectiles on the top specimen, as shown in Table 15. Figure 61 illustrates the average

impact energy versus areal density for the flat impactors. Overall, the 25 g projectile

exhibited the highest energy dissipated upon impact, followed by the 160 g, 50 g, and 100

g projectiles, respectively. In a similar plot, Figure 62, for the conical projectiles,

virtually no influence of mass is observed.

Damage Characterization

The primary theory behind the deviation stems from observations of three series

of high-speed images capturing the impact event of a 100 g, flat-tipped projectile. Upon

closer examination of the impact event, a slight tilting of the projectile was noticed. This

is illustrated in Figure 63, which captures the projectile tilting just after impact. The

damage progression is also shown at the end of the chapter via high-speed imagery in

Page 102: Bartus 2003_Thesis

88

Figures 64 and 65. Another point of interest noted from the high-speed images is that the

impact phenomenon exhibited characteristics of both low and high velocity.

The first impact event captured was below the critical velocity, resulting in no

visible damage. In the other two image series, where the impact occurred at close to the

critical velocity for the 100 g flat projectile, incipient damage was observed before global

plate motion was established. This is a characteristic of high velocity impact. Then, as

momentum effects took over, a large global displacement was observed (approximately

25 mm), an attribute of low velocity impact. The specimen actually met the back of the

sample support because it was only simply supported on two sides in order to allow for

observation of the event.

Damage began at the first point of contact on the target, which resulted in a shear

plug or punch-though failure. This is thought to be a result of an increase in contact

stresses at the point of impact due to a decrease in the contact area of the projectile. All

samples impacted with the 100 g projectile were impacted under the exact test

configuration, and all data show a marked decrease in energy dissipation. The gas gun

was realigned before further testing was carried out, and the impression left by

subsequent impactors did not indicate projectile tilting. A typical impactor impression

for an oblique impact is shown in Figure 66(a), and a normal impact impression is shown

in Figure 66(b).

A marked increase in contact stresses is also thought to be responsible for a

decrease in energy dissipation for the conical projectiles, discussed in the next section. In

cases where the projectile impacted normal to the specimen, damage is thought to initiate

at weak planes parallel to areas of high fiber orientation or in areas where fiber clumping

Page 103: Bartus 2003_Thesis

89

occurs. Other factors that may have contributed to data deviation are impactor material

properties and inconsistencies in the impact response of test specimen. Abrate noted that

the elastic material properties of the impactor could affect the perforation energy, e.g. that

higher modulus materials result in higher contact stresses due to a decrease in elastic

deformation during the onset of damage [34]. This is, of course, assuming that there is

no plastic deformation of the impactor. Since two different materials were used in the

mass affect study, UHMWPE and aluminum, a decrease in energy dissipation would be

expected in the case of the higher modulus aluminum projectile. The standard deviation

is too large to make any conclusions as to whether the impactor material type played a

significant role in energy dissipation.

Variations in data were most likely caused by irregularities in the impact response

of test specimens. Although great care was taken in consistently impacting the same area

of all the specimens, sample shifting is impossible to control in the case of a simply

supported sample. In addition, with the current sample holder, there is no way to

accurately insure even clamping pressure, especially with variations in sample thickness

and process-induced warping. Both factors may have led to variations in the impact

response of the target. After impacting a specimen, the line of support at the four support

areas leaves an imprint. Variations in the imprint were noted, although attempts to

minimize this were unsuccessful and may have influenced the energy dissipation.

Effect of projectile geometry. In the preliminary testing of the flat impactors, it

was surmised that inadvertent projectile tilting affected the energy dissipation of the

sample. It is thought that this marked decrease in energy stems from an increase in

Page 104: Bartus 2003_Thesis

90

contact stresses at the point of impact. As one would expect, a sharp or conically tipped

projectile would also result in a substantial decrease in perforation energy, in contrast to a

normally impacting flat projectile. Projectile mass seemed to play even less of a role in

the study of conically shaped projectiles, whose results exhibited a greater statistical

significance than the flat-shaped projectile study. If contact stress plays a dominant role

in sample perforation, it would follow that less deviation would occur in conically shaped

projectiles in which projectile tilting would not adversely affect the contact area.

Both projectile geometries inflicted planar cracking of the specimens upon

impact. Damage from flat projectiles initiated along the periphery of the impact area and

propagated radially along two or three planes, away from the area of initial damage.

However, specimens impacted with conical projectiles exhibited a higher degree of crack

branching. Damage from conical projectiles began at the point of initial contact,

subjecting the specimen to high local contact stresses. It is possible that crack branching

results from the increased fiber strain energy at the point of contact for samples subjected

to impact with conical projectiles. When the fibers fracture, the release of strain energy

is greater than the energy required to create a single crack plane. The crack must

bifurcate in order to dissipate excess energy.

Neglecting projectile mass affects and plotting projectile geometry as a function

of energy dissipation and areal density, an increasingly linear trend develops. Figure 63

illustrates this proclivity in which a linear regression was fit through the mean of the data

for the flat and conical projectile geometries, resulting in R-squared values of 0.93 and

0.97 respectively. The mean numeric data is reported next to the respective data set,

Figure 67. On average, a decrease in critical energy dissipation for the bottom, center,

Page 105: Bartus 2003_Thesis

91

and top specimens was approximately 26% (areal density = 4.39 g cm-2), 30% (areal

density = 4.67 g cm-2), and 25% (areal density = 4.79 g cm-2), respectively, when

subjected to impact by conically tipped projectiles.

Areal density effects. It is common practice to report impact data as a function of

areal density, or weight area-1, for continuous fiber-reinforced composites. This was also

done in the case of the LFT specimens, shown in Figures 61, 62, and 67. Variations in

the energy dissipation of each of the samples were first attributed to the variations in

areal density, especially considering the expected linear increase in energy dissipation.

However, in the post mortem examination of the impacted specimens, similarities

between the fracture patterns of the three different samples were qualitatively established,

suggesting that a common failure mechanism might be at play.

Fiber orientation effects. Thomason and Vlug [6] surmised that a linear

relationship exists between tensile and impact strength. As the ultimate tensile strength

increases, the impact strength increases. Since tensile strength is a strong function of

fiber orientation, as shown in Figure 4, impact strength would be expected to follow

similar trends. No known models of impact strength as a function of discontinuous fiber

orientation exist, and current models for uniaxially aligned fibers, Figures 5 and 6, have

not been well documented [6]. Intuitively, fracture seems most likely to follow the

weakest plane, e.g. the plane perpendicular to the main fiber orientation. This type of

failure was seen in sections taken from the specimens representing typical failure modes,

shown in Figure 68 using scanning electron microscopy (SEM). The SEM images were

Page 106: Bartus 2003_Thesis

92

taken normal to the fracture surface, as in Figures 69-78. A simplified illustration of

what may be occurring is shown in Figure 79, which shows a theoretical transverse view

of fiber orientation and fracture path. The impact occurs at the top of the illustration and

follows the plane with a higher degree of fiber orientation. As the crack propagates,

energy is dissipated though fiber debonding, fiber pull-out, fiber breakage, and matrix

fracture. The crack will continue to propagate until the energy for fracture is spent or the

crack is blunted by fibers oriented perpendicular to the fracture path.

Micrograph analysis

SEM analysis of the impacted LFT sample fracture surface revealed that the

predominant modes of failure where fiber debonding/pull-out, fiber fracture, fiber-matrix

pull away and matrix fracture, as in Figures 69-78. It is difficult to discern the

difference between fiber fracture and fiber pull-out in discontinuous fiber composites.

Upon close examination, part of the fiber/matrix interfacial surface can be seen where the

fiber fractured and pulled out, as in Figure 71. This type of fracture also appears in

Figures 70-72, 74, and 76-78. Fiber pull-out is evident in every micrograph.

Fiber pull-out occurs when the interfacial stresses at the fiber-matrix interface

exceed the interfacial strength, causing the fiber to debond from the matrix. The impact

load is carried by the interfacial friction force imposed on the fiber from the matrix and

through matrix cracking. The magnitude of the interfacial friction is a function of the

debonded fiber surface roughness, the fiber length, and the radial stresses on the fiber

imposed by thermal contraction. The thermal contract results from thermal mismatch

between the fiber and matrix at processing temperature and test temperature.

Page 107: Bartus 2003_Thesis

93

As the test temperature decreases, e.g. increasing the thermal gradient, the radial

stresses increase. The surface roughness of the fiber is mainly dependent on chemical

coupling and fiber wetting at the fiber-matrix interface. Fibers sized for the matrix

exhibit increased chemical coupling though an increase in fiber wetability. The

debonded surface of a properly sized fiber is typically rough in appearance when viewed

at sufficient magnification.

If the fiber is above critical length, fiber breakage must occur for the fiber to pull-

out. This phenomenon, known as fiber fragmentation, occurs when the interfacial

stresses exceed the fiber strength. This is likely to have occurred in the impacted LFT

samples. However, it is not easy to characterize the phenomenon without low fiber

loading in a transparent matrix, clearly not the case with the samples examined.

Three possible modes of failure appear in the matrix: brittle fracture, crazing-

tearing (stress whitening), and separation of fibrils, as in Figures 71, 72, and 74-76. A

smooth fracture surface indicates brittle fracture, as in Figures 71 and 72. This was the

most common failure mode seen in the SEM micrograph analysis. The matrix fails in a

brittle manner because the fiber reinforcement inhibits large matrix deformations.

Crazing-tearing, indicated by the white areas in the micrographs shown in Figures 70, 71,

and 73, is also common in thermoplastic matrices subjected to strain. The ductile pulling

of fibrils, as in Figures 76 and 77, also indicates a brittle failure. No attempt was made to

quantify the matrix failure modes. The contribution of energy dissipation from matrix is

typically small in contrast to the reinforcement.

Page 108: Bartus 2003_Thesis

94

Correlation Between Predicted Fiber Orientation and Impact Failure Mode

The failure mode of discontinuous fiber-reinforced composites subjected to

impact differs from the failure mode of laminated composites, both in appearance and

damage mechanisms. Laminated composites subjected to transverse impact usually

exhibit damage variation though the thickness, e.g. delamination. The delamination

usually grows in area through the thickness (Christmas tree pattern or conical pattern),

toward the tensile face of the impact specimen [34]. This damage mechanism contributes

a significant amount of impact energy dissipation. Conversely, discontinuous fiber-

reinforced composites do not have a laminated architecture and dissipate energy through

planar cracking, as in Figures 66, 68, and 80-90. As shown before, the planar cracking

appears to advance along areas with preferential fiber orientation. Since it is not practical

to quantify the fiber orientation along the entire fracture surface of the specimens, an

attempt was made to correlate the fracture path with the fiber orientation predicted in the

Cadpress control run simulation.

The control run simulation used material and process parameters closest to those

used to produce the LFT tab plaque specimens. The parameters are outlined in Tables 2-

5. One noteworthy assumption was a random pre-orientation of the charge. Matrix burn

off revealed some preferential orientation in the charge roughly parallel to the extrusion

axis. However, a quantitative analysis of the fiber pre-orientation in the charge is

difficult, so it was assumed to be random.

Assuming that the control run simulation provided a reasonably accurate

representation of the fiber orientation, one would expect to see a correlation between

fiber orientation and the fracture path. Macro photos were taken of representative

Page 109: Bartus 2003_Thesis

95

specimens exemplifying typical fracture patterns. The fracture path was traced using

free-form sketch and grouped to a frame around periphery of the specimen. The aspect

ratio of the trace was maintained and superimposed over the respective area from the

fiber orientation predication. This is shown in Figures 80-91 for both flat and conical

impactors. The graphical representation of fiber orientation is described in Figures 30,

and 44-46.

The fracture pattern correlates well with the planes of high or preferential fiber

orientation predicted in the control run simulation. In addition, it appears that cracks may

have a tendency to advance along areas where an abrupt change in local fiber orientation

occurs. This is shown in Figures 80-91; however, it was not confirmed experimentally.

The abrupt changes in local fiber orientation may also be a mechanism for crack blunting.

Understanding the correlation between LFT failure mechanisms and the predicted

fiber orientation can aid engineers in component and process design. Because of the

strong off-axis effect from fiber orientation, tensile strength and modulus are optimized

(in one direction) when the fibers are uniaxially aligned. If this was the case, the

component may actually be susceptible to impact damage along the uniaxially aligned

fibers. A random orientation would likely provide the highest impact strength by creating

a torturous path for crack propagation. More over, LFTs with a random fiber orientation

would also exhibit greater damage tolerance because of crack blunting.

Page 110: Bartus 2003_Thesis

96

Table 18 Numeric results and analysis of the velocity data for the top specimen Projectile Mean (m s-1) S.D. (m s-1) 95% C.I. (m s-1) 25 g Flat 124.4 2.5 2.4 25 g Conical 103.5 2.4 2.1 50 g Flat 84.2 4.5 4.0 50 g Conical 76.4 3.9 3.4 100 g Flat 56.0 2.3 2.0 160 g Flat 46.9 2.1 1.6

0

30

60

90

120

150

0 20 40 60 80 100 120 140 160 180Projectile Mass (g)

Vel

coci

ty (m

s-1)

25 g flat projectile 25 g conical projectile

50 g flat projectile 50 g conical projectile

100 g flat projectile 160 g flat projectile

Figure 54 Critical velocity versus projectile mass for the top specimen (areal density = 4.79 g cm-2) , showing the velocity as an exponentially decreasing function of projectile mass.

Page 111: Bartus 2003_Thesis

97

Table 19 Numeric results and analysis of the velocity data for the center specimen Projectile Mean (m s-1) S.D. (m s-1) 95% C.I. (m s-1) 25 g Flat 127.2 9.8 8.6 25 g Conical 97.8 6.6 7.5 50 g Flat 83.3 6.3 4.7 50 g Conical 72.9 4.9 4.3 100 g Flat 53.8 1.1 0.9 160 g Flat 47.3 0.6 0.5

0

30

60

90

120

150

0 20 40 60 80 100 120 140 160 180Projectile Mass (g)

Vel

coci

ty (m

s-1)

25 g flat projectile 25 g conical projectile

50 g flat projectile 50 g conical projectile

100 g flat projectile 160 g flat projectile

Figure 55 Critical velocity versus projectile mass for the center specimen (areal density = 4.67 g cm-2), showing the velocity as an exponentially decreasing function of projectile mass.

Page 112: Bartus 2003_Thesis

98

Table 20 Numeric results and analysis of the velocity data for the bottom specimen Projectile Mean (m s-1) S.D. (m s-1) 95% C.I. (m s-1) 25 g Flat 111.1 6.2 5.4 25 g Conical 97.9 4.9 4.8 50 g Flat 76.8 5.3 3.6 50 g Conical 68.2 4.1 3.6 100 g Flat 54.0 1.5 1.3 160 g Flat 44.7 1.1 1.0

0

30

60

90

120

150

0 50 100 150 200Projectile Mass (g)

Vel

ocity

(m s-1

)25 g Flat 25 g Conical

50 g Flat 50 g Conical

100 g Flat 160 g Flat

Figure 56 Critical velocity versus projectile mass for the bottom specimen (areal density = 4.39 g cm-2), showing the velocity as an exponentially decreasing function of projectile mass.

Page 113: Bartus 2003_Thesis

99

25

50

75

100

125

150

175

200

225

250

0 20 40 60 80 100 120 140 160 180Projectile Mass (g)

Ene

rgy

(J)

Figure 57 Effect of projectile mass, critical energy versus projectile mass for the top specimen.

Flat projectiles

Conical projectiles

Page 114: Bartus 2003_Thesis

100

Figure 58 Effect of projectile mass, critical energy versus projectile mass, for the center specimen.

25

50

75

100

125

150

175

200

225

250

0 20 40 60 80 100 120 140 160 180Projectile Mass (g)

Ene

rgy

(J)

Flat projectiles

Conical projectiles

Page 115: Bartus 2003_Thesis

101

25

50

75

100

125

150

175

200

225

250

0 20 40 60 80 100 120 140 160 180Projectile Mass (g)

Ene

rgy

(J)

Figure 59 Effect of projectile mass, critical energy versus projectile mass, for the bottom specimen.

Flat projectiles

Conical projectiles

Page 116: Bartus 2003_Thesis

102

25

50

75

100

125

150

175

200

225

250

0 20 40 60 80 100 120 140 160 180Projectile Mass (g)

Ene

rgy

(J)

bottom, flat impactor (average)bottom, conical impactor (average)center, flat impactor (average)center, conical impactor (aveage)top, flat impactor (average)top, conical impactor (average)

Figure 60 Energy (J) versus projectile mass (g) for the bottom, center, and top specimens with a linear regression analysis, fitted through the mean of each data set, indicative of the independent relationship between projectile mass (impact velocity) and the energy dissipation upon impact. The trend in decreasing energy dissipation in the flat and conical data sets is most likely due to a decrease in areal density and fiber orientation effects.

areal density

areal density

Page 117: Bartus 2003_Thesis

103

25

50

75

100

125

150

175

200

225

250

4.3 4.4 4.5 4.6 4.7 4.8 4.9Areal Density (g cm-2)

Ene

rgy

(J)

50 g flat tip projectile

25 g flat tip projetile

100 g flat tip projectile

160 g flat tip projectile

Figure 61 Energy (J) versus areal density (g cm-2), plotted as a linear regression fit through the mean of the data for the 25 g, 50 g, 100 g, and 160 g flat projectiles, illustrating that no significant relationship exits between the projectile mass and energy dissipation in the mass range examined.

Page 118: Bartus 2003_Thesis

104

25

50

75

100

125

150

175

200

225

250

4.3 4.4 4.5 4.6 4.7 4.8 4.9Areal Density (g cm-2)

Ene

rgy

(J)

50 g conical tip projectile

25 g conical tip projectile

Figure 62 Energy (J) versus areal density (g cm-2), plotted as a linear regression fit through the mean of the data for the 25 g and 50 g conical projectiles, illustrating that no significant relationship exits between the projectile mass and energy dissipation in the mass range examined.

Page 119: Bartus 2003_Thesis

105

Figure 63 High-speed image taken at 14,000 frames s-1 showing a 100 g flat- tipped projectile exemplifying projectile tilting just after impacting the sample.

Punch through

50 mm

Page 120: Bartus 2003_Thesis

106

Figure 64 High-speed images of a BOI illustrating the onset of damage, K.E. 142.3J. High-speed image taken at 14,000 frames s-1 showing a 100 g flat-tipped projectile just prior to impact, exemplifying projectile tilting.

a

)

b

)

Figure 65 High-speed images taken at 14,000 frames s-1 showing a 100 g flat projectile after the initial impact and while rebounding, K.E. 142.3J.

Page 121: Bartus 2003_Thesis

107

Figure 66 Impactor impressions left on the target for a projectile tilted (a) and normal (b).

(a)

(b)

Page 122: Bartus 2003_Thesis

108

Figure 67 Energy versus Areal Density for flat and conically tipped impactors showing a linearly increasing trend via linear regression fit through the mean of all the impact data. The results of 106 impacted samples are shown, which represent increasing areal density for the bottom, center, and top specimens, respectively, which illustrate the influence of impactor geometry. The linear regression for the flat and conically tipped projectiles is given in the top right of the graph with the respective R2 values.

175.0 J174.9 J

151.3 J

131.0 J

122.8 J112.0 J

Energy (Flat) = 63.2*(Areal Density) - 124.6R2 = 0.93

Energy (Conical) = 45.2*(Areal Density) - 86.5R2 = 0.97

25

50

75

100

125

150

175

200

225

250

4.3 4.4 4.5 4.6 4.7 4.8 4.9Areal Density (g cm-2)

Ene

rgy

(J)

bottom, flat impactor center, flat impactor

top, flat impactor bottom, conical impactor

center, concial impactor top, conical impactor

Conical impactors

Flat impactors

Page 123: Bartus 2003_Thesis

109

a

e

c d

b

f

Figure 68 Impacted samples showing the location of sections taken for SEM analysis. The respective samples are (a) 34T, (b) 38T, (c) 34B, (d) 39B(1) and 39B(2), (e) 33C, and (f) 1C. At least two representative samples from the impacted LFT panels were sectioned from each of the three locations- top, center, and bottom. The locations analyzed are indicated by the rectangular box.

Page 124: Bartus 2003_Thesis

110

Figure 69 SEM of sample 34B, taken normal to the fracture surface, showing the path of fracture following the main fiber orientation angle. The dominant failure modes are fiber pull-out, fiber breakage, and matrix fracture.

Impact direction

Page 125: Bartus 2003_Thesis

111

Figure 70 SEM image of sample 34B showing fiber pull-out and fiber breakage. The rough fiber surface indicates strong interfacial fiber-matrix bonding. A mixture of crazing-tearing and brittle fracture is evident in the matrix.

Fiber pull-out

Fiber breakage

Page 126: Bartus 2003_Thesis

112

Figure 71 SEM image of sample 39C showing fiber pull-out. Matrix failure appears to have a combination of crazing-tearing (stress whitening) and brittle fracture (smooth appearance).

Crazing-tearing

Brittle fracture

Page 127: Bartus 2003_Thesis

113

Figure 72 SEM of sample 1C showing fiber pull-out (fiber sliding) and fiber matrix pull away.

Impact direction

Fiber pull-out

Fiber-matrix pull away

Page 128: Bartus 2003_Thesis

114

Figure 73 Micrograph of sample 1C showing fiber pull-out lengths in excess of approximately 3 mm. Also, note again, a high degree of fiber orientation

Impact direction

Page 129: Bartus 2003_Thesis

115

Figure 74 SEM image of sample 1C showing fiber pull-out.

Page 130: Bartus 2003_Thesis

116

Figure 75 SEM of sample 34T showing a high degree of orientation and ductile pulling of fibrils. Fiber clumping also appears to be present.

Fibrils

Page 131: Bartus 2003_Thesis

117

Figure 76 Micrograph of sample 34T illustrating a brittle-matrix fracture with ductile pulling of fibrils.

Page 132: Bartus 2003_Thesis

118

Figure 77 SEM image of sample 38T normal to the fracture surface showing fiber pull-out and breakage.

Page 133: Bartus 2003_Thesis

119

Figure 78 SEM of sample 39C showing variations in fiber orientation through the thickness of the section, taken normal to the fracture plane.

Page 134: Bartus 2003_Thesis

120

Figure 79 Illustration of a possible failure mechanism of the samples tested showing a transverse view of a test plaque with the fracture path following a preferential plane of fiber orientation.

Fracture path

Fibers blunting crack propagation

Impact area

Page 135: Bartus 2003_Thesis

121

(a)

(b)

Figure 80 (a) Bottom specimen (020920-1-12B), impacted with a 50 g flat projectile, exhibiting a typical fracture pattern, (b) Graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (B) with the fracture pattern superimposed over the results.

Page 136: Bartus 2003_Thesis

122

Figure 81 (a) Bottom specimen (020920-1-63B), impacted with a 25 g flat projectile, exhibiting a typical fracture pattern, (b) Graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (B) with the fracture pattern superimposed over the results.

(b)

(a)

Page 137: Bartus 2003_Thesis

123

Figure 82 (a) Center specimen (020920-1-16C), impacted with a 25 g flat projectile, exhibiting a typical fracture pattern, (b) Graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (C) with the fracture pattern superimposed over the results.

(a)

(b)

Page 138: Bartus 2003_Thesis

124

Figure 83 (a) Center specimen (020919-1-35C), impacted with a 160 g flat projectile, exhibiting a typical fracture pattern, (b) Graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (C) with the fracture pattern superimposed over the results.

(b)

(a)

Page 139: Bartus 2003_Thesis

125

Figure 84 (a) Center specimen (020919-1-32C), impacted with a 160 g flat projectile, exhibiting a typical fracture pattern, (b) Graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (C) with the fracture pattern superimposed over the results.

Page 140: Bartus 2003_Thesis

126

Figure 85 (a) Top specimen (020919-1-35T), impacted with a 50 g flat projectile, exhibiting a typical fracture pattern, (b) Graphical representation of fiber orientation predicted in Cadpress showing the representative area of the specimen (b) with the fracture pattern superimposed over the results.

(a)

(b)

Page 141: Bartus 2003_Thesis

127

Figure 86 (a) Top specimen (020919-1-107T), impacted with a 160 g flat projectile, exhibiting a typical fracture pattern, (b) Graphical representation of fiber orientation predicted in Cadpress showing the representative area of the specimen (b) with the fracture pattern superimposed over the results.

(b)

(a)

Page 142: Bartus 2003_Thesis

128

Figure 87 (a) Bottom specimen (020920-1-27B), impacted with a 50 g conical projectile, exhibiting a typical fracture pattern, (b) Graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (B) with the fracture pattern superimposed over the results.

(b)

(a)

Page 143: Bartus 2003_Thesis

129

Figure 88 (a) Bottom specimen (020920-1-22B), impacted with a 50 g conical projectile, exhibiting a typical fracture pattern, (b) Graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (B) with the fracture pattern superimposed over the results.

(a)

(b)

Page 144: Bartus 2003_Thesis

130

Figure 89 (a) Center specimen (020919-1-61C), impacted with a 25 g flat projectile, exhibiting a typical fracture pattern, (b) Graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (C) with the fracture pattern superimposed over the results.

(b)

(a)

Page 145: Bartus 2003_Thesis

131

Figure 90 (a) Bottom specimen (020920-1-87T), impacted with a 50 g conical projectile, exhibiting a typical fracture pattern, (b) Graphical representation of fiber orientation predicted in Cadpress showing the representative area of specimen (T) with the fracture pattern superimposed over the results.

(b)

(a)

Page 146: Bartus 2003_Thesis

132

Figure 91 (a) Superimposed view of the fracture patterns over the entire tab plaque for the flat projectiles from Figures 75(a), 79(a), and 80(a). (b) Superimposed view of the fracture patterns over the entire tab plaque for the conical projectiles from Figures 83(a), 84(a), and 85(a).

(a)

(b)

Increasingly random fiber orientation

Higher degree of orientation

Page 147: Bartus 2003_Thesis

133

CONCLUSION: BLUNT OBJECT IMPACT

The effects of LFTs subjected to BOI were investigated as a function of projectile

mass and geometry. The average critical energy dissipated for PP/40 wt. % E-glass

(average areal density = 4.61 g cm-2) was 167 J and 121 J for the flat and conical

projectiles, respectively. The predominant energy dissipation mechanisms are fiber

fracture, fiber debonding, fiber pull-out, and matrix fracture.

For panels impacted by the flat-tipped projectile with normal incidence, e.g. full

contact of the projectile face with the sample, the limiting damage occurred by punch-

through (shear plug). The damage initiated around the periphery of the impactor as a

result of high transverse shear stresses. Assuming a uniform shear stress distribution,

failure occurred in areas with preferential fiber orientation tangent to the impact area.

Away from the impactor, the dominant failure mode occurred by simultaneous tearing

(planar cracking) across planes of preferential fiber orientation.

In the 100 g flat impactor study, the edge of the projectile made the first contact

resulting from a slight tilting upon or prior to contact. This consequently resulted in an

initial notch arising from increased contact stresses, followed by punch-through and

subsequent tearing. The increase in contact stresses decreased the critical energy

required for perforation. This is also thought to be the mechanism for a decrease in

energy dissipation for the conically tipped projectiles.

Energy dissipation (critical energy) for samples subjected to conical projectile

impact was approximately 27% less than for samples impacted by flat projectiles.

Page 148: Bartus 2003_Thesis

134

However, samples impacted by conical projectiles exhibited a higher degree of damage,

e.g. more fracture surface as a result of increased crack branching. This may be a result

of increased fiber-strain energy at the point of impact for samples impacted with the

conical projectiles. When the fibers fracture, the release of strain energy is greater than

the energy required to create a single crack plane, resulting in crack bifurcation in order

to dissipate the excess energy.

Impactor velocity did not result in any appreciable effects on the critical energy

dissipation in the specimens. This indicates that the material was not sensitive to the

loading rate in the range investigated. Neglecting projectile mass effects, and plotting

projectile geometry as a function of critical energy dissipation and specimen areal

density, a linearly increasing trend develops. In addition to areal density effects, fiber

orientation may also play a critical role in energy dissipation. The energy dissipation of

impacted LFTs appears to decrease as areal density decreases and as fiber orientation

increases. The degree to which this occurs was analyzed qualitatively in the control run

simulation.

The fracture pattern correlated well with the planes of high or preferential fiber

orientation predicted in the Cadpress control run simulation. The cracks have a tendency

to advance along areas where an abrupt change in local fiber orientation occurs. Abrupt

changes in local fiber orientation may also be a mechanism for crack blunting. This

indicates that components exhibiting a high degree of anisotropy, or abrupt changes in

fiber orientation, will be susceptible to impact damage. A random orientation would

likely provide the maximum impact strength by creating a torturous path for crack

propagation. More over, LFTs with a random fiber orientation would also exhibit greater

damage tolerance, due to crack blunting.

Page 149: Bartus 2003_Thesis

135

LIST OF REFERENCES

1 Hartness, T., Husman, G., Koenig, J., and Dyksterhouse, J., The characterization of low cost fiber-reinforced thermoplastic composites produced by the drift process. Composites, 2001, 32, 1155-1160.

2 Marsh, G., Composites on the road to big time? Reinforced Plastics, 2003, 47, 33-36, 38-47.

3 Thomason, J.L. and Vlug, M.A., Influence of fibre length and concentration on the properties of glass fibre-reinforced polypropylene: 1. Tensile and flexural modulus. Composites, 1996, 27A, 477-484.

4 Saito, M., Kukula, S., and Kataoka, Y., Practical Use of the Statistically Modified Laminate Model for Injection Moldings. Part 1: Method and Verification. Polym. Compos., 1998, 19, 497-505.

5 Lee, S.M., Cheon, J.S., and Im, Y.T., Experimental and numerical study of the impact behavior of SMC plates. Compos. Structures, 1999, 47, 551-561.

6 Thomason, J.L. and Vlug, M.A., Influence of fibre length and concentration on the properties of glass fibre-reinforced polypropylene: 4. Impact properties. Composites, 1997, 28A, 277-288.

7 Miller, K. and Ramani, K., Process-Induced Residual Stresses in Compression Molded UHMWPE. Polym. Eng. Sci., 1999, 39, 110-118.

8 Bush, S.F., Torres, F.G., and Methven, J.M., Rheological characterization of discrete long glass fibre (LGF) reinforced thermoplastics. Composites, 2000, 31, 1421-1431.

9 Chawla, K.K., Composite Materials Science and Engineering, 2nd Ed. Springer-Verlag New York, Inc., 1998.

10 Curtis, P.T., Bader, M.G., and Bailey, J.E., The stiffness and strength of a polyamide thermoplastic reinforced with glass and carbon fibers. J. Mater. Sci., 1978, 13, 377-390.

11 Kelly, A. and Tyson, W.R., Tensile properties of fiber-reinforced metals: Copper/Tungsten and Copper/Molybdenum. J. Mech. Phys. Solids, 1965, 13, 329-350.

Page 150: Bartus 2003_Thesis

136

12 Cox, H.L., The elasticity and strength of paper and other fibrous materials. Brit. J. Appl. Phys., 1952, 3, 72-79.

13 Thomason, J.L., Vlug, M.A., Schipper, G., and Krikor H. G. L. T., Influence of fibre length and concentration on the properties of glass fibre-reinforced polypropylene: Part 3. Strength and strain at failure. Composites, 1996, 27A, 1075-1084.

14 Chou, T.W., Microstructural Design of Fibrous Composites. Cambridge University Press, Cambridge, 1992.

15 Krenchel, H., Fibre Reinforcement. Akademisk Forlag. Copenhagen, 1964.

16 Toll, S. Packing Mechanics of Fiber Reinforcements. Polym. Eng. Sci., 1998, 38, 1337-1350.

17 Eichhorn, S.J. and Young, R.J., The Young’s modulus of microcrystalline cellulose. Cellulose, 2001, 18, 197-207.

18 Cooper, G.A., The fracture toughness of composites reinforced with weakened fibers. J. Mater. Sci., 1970, 5, 645-654.

19 Wald, H.H. and Schriever, H., Glass-mat reinforced thermoplastics. Kunststoffe Plast Europe, 1985, 75, 497-503.

20 Turkovich, R.V. and Erwin, L., Fiber Fracture in Reinforced Thermoplastic Processing. Polym. Eng. Sci., 1983, 23, 743-749.

21 Cadpress Theory Manual, Volume 2. The Madison Group: Polymer Processing Research Corp.

22 Munson, B.R., Young, D.F., and Okishi, T.H., Fundamentals of Fluid Mechanics, 3rd Ed. John Wiley & Sons, Inc., 17, 1998.

23 Gupta, M. and Wang, K.K., Effect of processing conditions on fiber orientation in injection-molded composites. SPE ANTEC Technical Papers, 1993, 39, 2290-2296.

24 Gupta, M. and Wang, K.K., Fiber Orientation and Mechanical Properties of SFRIM Composites. Polym. Compos., 1993, 15, 367-373.

25 Chung, S.T. and Kwon, T.H., Coupled Analysis of Injection Molding Filling and Fiber Orientation, Including In-Plane Velocity Gradient Effect. Polym. Compos., 1996, 17, 859-872.

26 Advani, S.G. and Talreja, R., A Continuum Approach to Determination of Elastic Properties of Short Fibre Composites. Mech. Compos. Mater., 1993, 29, 171-183.

Page 151: Bartus 2003_Thesis

137

27 Advani, S.G. and Tucker, C.L., The Use of Tensors to Describe and Predict Fiber Orientation in Short Fiber Composites. J. Rheology, 1987, 8, 751-784.

28 Jackson, W.C., Advani, S.G., and Tucker, C.L., Predicting the Orientation of Short Fiber in Thin Compression moldings. J. Compos. Mater., 1986, 20, 539-557.

29 Altan, M.C., Advani, S.G., Guçeri, S.I., and Pipes, R.B., On the Description of the Orientation State for Fiber Suspensions in Homogeneous Flows. J. Rheology, 1989, 7, 1129-1155.

30 Gibson, A. G. and Toll, S., Mechanics of the squeeze flow of planar fibre suspensions. J. Non-Newtonian Fluid Mech., 1999, 82, 1-24.

31 Park, C.H., Lee, W.I., Yoo, Y.E., and Kim, E.G., A study on fiber orientation in the compression molding of fiber-reinforced polymer composite material. J. Mater. Processing Technol., 2001, 111, 233-239.

32 Thomason, J.L. and Groenewoud, W.M., Influence of fibre length and concentration on the properties of glass fibre-reinforced polypropylene: 2. Thermal properties. Composites, 1996, 27A, 555-565.

33 Abrate, S., Impact on Composite Structures. Cambridge University Press, 1998.

34 Delfrosse, D. and Poursartip, A., Energy-based approach to impact damage in CRFP laminates. Composites, 1997, 28A, 647-655.

35 Dutta, P.K., Farrell, D., Taylor, S., Tadayon, A., and Hui, D., CRREL Special Report 96-29, 1996.

36 Okafor, A.C., Otieno, A.W., Dutta, A., and Rao, V.S., Detection and characterization of high-velocity impact damage in advanced composite plates using multi-sensing techniques. Compo. Structures, 2001, 54, 289-297.

37 Iremonger, M.J., and Went, A.C., Ballistic impact of fibre composite armours by fragment-simulating projectiles. Composites, 1996, 27A, 575-581.

38 Tarim, N., Findik, F., and Uzun, H., Ballistic impact performance of composite structures. Compos. Structures, 2002, 56, 13-20.

39 Jenq, S.T., Wang, S.B., and Sheu, L.T., A Model for Predicting the Residual Strength of GFRP Laminates Subject to Ballistic Impact. J. Reinforced Plast. Compos., 1992, 11, 1127 – 1141.

40 Cantwell, W.J. and Morton, J., The Influence of Varying Projectile Mass on the Impact Response of CFRP. Compos. Structures, 1989, 11, 101-114.

41 Shigley, J.E., and Mischke, C.R., Mechanical Engineering Design, 5th Edition. MeGraw-Hill, Inc., 1989.

Page 152: Bartus 2003_Thesis

138

42 http://www.shootingsoftware.com/chronographs.htm, December 5, 2002.

43 http://www.ticona-us.com/datasheet/DataSheet.cfm?ID=827&STANDARD=ISO, August 9, 2002.

44 Hsiao, H.M. and Daniel, I.M., Stain rate behavior of composite materials. Composites, 1998, 29B, 521-531.