You are on page 1of 20

Modeling of In Situ Strength Development

for the Thermoplastic Composite


Tow Placement Process

JOHN TIERNEY1 AND J. W. GILLESPIE, JR2,*


1
Center for Composite Materials and Department of Materials Science and
Engineering, University of Delaware, Newark, DE 19716-3144, USA
2
Department of Civil and Environmental Engineering, University of Delaware
Newark, DE 19716-3144, USA

(Received September 30, 2004)


(Accepted August 22, 2005)

ABSTRACT: The composite tow placement process is a non-autoclave consolida-


tion method that involves in situ consolidation and bonding of thermoplastic tape on
a preconsolidated substrate. It can be used as a method for manufacturing large
composite structures with a high degree of process control and dimensional accuracy.
The bond strength that develops between layers within the laminate ultimately
governs the final strength of the part. In this work, a model is presented for
predicting through-thickness heat transfer and bond strength development based on
intimate contact and healing at the ply interface. Experiments are carried out to
validate these results for a wide variety of process conditions, and they show that
model based predictive control can be used as a method for process optimization.
The numerical results show that bond strength development is significantly affected
by the process set points (e.g., head velocity heat input and roller pressures), and that
the resulting strength can vary significantly within the part based on these set points.

KEY WORDS: composite, thermoplastic, tow placement process, ATP, strength,


bonding.

INTRODUCTION

N THE THERMOPLASTIC composite tape placement process, the part is manufactured by


I applying adequate heat and pressure to a thermoplastic prepreg tape, so that welding
is achieved at the interface between the incoming tape and preconsoldiated substrate, and
consolidation is obtained within the composite tow [1]. By utilizing an industrial robot
and/or a gantry, it is possible to manufacture a variety of shapes ranging from simple flat
plates, to panels with mild curvatures to complex three-dimensional (3-D) surfaces [2–5].

*Author to whom correspondence should be addressed. E-mail: gillespie@ccm.udel.edu


Figures 1, 4 and 8–13 appear in color online: http://jcm.sagepub.com

Journal of COMPOSITE MATERIALS, Vol. 40, No. 16/2006 1487


0021-9983/06/16 1487–20 $10.00/0 DOI: 10.1177/0021998306060162
ß 2006 SAGE Publications
1488 J. TIERNEY AND J. W. GILLESPIE, JR.

One of the most important measures of quality in any part is its strength. The primary
strength component in automated tow placement (ATP) processing is the bond strength
that develops between each ply upon placement. It is this strength that governs the ability
of the laminate to transfer interlaminar stresses between individual plies. These stresses are
a function of laminate stacking sequence, and as such can be controlled by proper material
selection and stacking sequence. One of the assumptions made, however, is that a perfect
bond is obtained between each layer during processing. This may not be the case, as the
material undergoes a complex thermal and pressure history, both of which affect the bond
strength development. The objective of this study is to develop a model that describes the
strength development in ATP composite laminates and to validate this theory with
experimental results.

THE THERMOPLASTIC COMPOSITE TOW-PLACEMENT PROCESS

The ATP system used in this study is shown in Figure 1. It makes use of two hot gas
nitrogen torches to heat the material and two rollers to provide the pressures required for
consolidation. The purpose of the first torch and roller is to preheat the composite surface
and the incoming tow together. The material is thus ‘tacked’ to the surface with this roller.
This tacking approach also aids in improving the efficiency of the cut and refeed
mechanism [6,7]. The second torch (main heater) provides supplemental through thickness
heating to facilitate consolidation and bonding of the tow and substrate under the
consolidation roller. These rollers provide the necessary forces to achieve complete
intimate contact across the tow interface, and the boundary pressure for compressing

Figure 1. Schematic of the automatic tow placement process used in this study.
In Situ Strength Development for Thermoplastic Composite 1489

internal voids present in the incoming material. The consolidated tows then cool under
ambient air temperature conditions. The forces applied to both rollers are controlled
independently using a series of pneumatic actuators. Positioning of the tows with
this method also enables precise positioning of the material on the laminate structure,
which is often not possible with hand lay-up. Also the principal process variables
are shown in the figure which are known to influence the thermal profile and
resulting material properties: process velocity vx, nozzle heights h1 and h2 which are
independently controlled, and the roller forces F1 and F2, which are also independently
controlled.
Other process parameters which influence final material properties are gas temperatures,
tool temperatures, and head design. These settings are fixed in this analysis as, (a) the gas
temperatures are maintained at the maximum possible so that improved thermal efficiency
is achieved; (b) the tool is not heated as this can dramatically increase the overall
production costs, and (c) the head design is fixed, as this study is not based on process
optimization, but model validation based on the current head design.
In the ATP process, the laminate is placed on a substrate using an in situ (or on-line)
consolidation process. This process involves the concurrent use of both heat and pressure
such that the material fuses or bonds at the interface. The two mechanisms that contribute
to the strength development at the polymer–polymer interface in this process are: (a) the
development of complete mechanical, or intimate, contact, and (b) the diffusion of
polymer chains across those contacting surfaces known as healing [8–12]. The first
mechanism is quantified by the degree of intimate contact, Dic, which describes the
fraction of bond area that is in intimate contact. The second is quantified by the degree
of healing, Dh, which describes the degree of polymer diffusion across a contacting
interface. The degree of intimate contact is directly affected by temperature dependent
viscosity and external pressures, and the healing process is directly related to process
temperature. As both mechanisms occur simultaneously during any fusion bonding
process, a coupled bonding model is required to determine the final bond strength, Db [13].
This model tracks the degree of healing for each incremental increase in interfacial
area, and is applicable to all fusion bonding processes. The degree of consolidation
or void fraction within the bond can also be determined from the pressure and
temperature distribution within the interfaces. A summary of these models is shown in
Figure 2.
The results from these models provide a framework for process optimization through a
model based predictive approach. It has been found that, exploration, of these processing
parameters is time consuming and may not yield optimum conditions for this process. As
a result, a series of neural network simulations with an optimized control system aimed at
improving quality was developed and used as a method towards the ultimate development
of a virtual manufacturing environment for the ATP based thermoplastic composites.
This method is also used to monitor the complete processing history at any region of
the part, and as such is used for continuous improvement in the final quality through
multiple passes [14].

HEAT TRANSFER MODEL

A number of thermal models have been developed for the ATP process, such as a simple
steady state solution [1], and the 2-D transient solution [15]. In this study, a 1-D steady
1490 J. TIERNEY AND J. W. GILLESPIE, JR.

Temperature solution
from ply interface

Consolidation/deconsolidation
Polymer healing, Dh* pressure field P(x), viscosity µ,
%compaction,

Intimate contact, Dic*

Final bond strength, Db* Final volume fraction


of voids

Figure 2. Schematic of models used for describing bond strength development and consolidation for fusion
bonded composites.

state gas impingement model has been developed to predict the thermal response of
the composite in the vicinity of the tow placement head. The governing equation is as
follows:

@T 2 T x
¼ 2 ¼ ð1Þ
@ z V

where  is dependent on the material’s thermal properties in the through-thickness


direction. The solution details are omitted here for the sake of brevity but can be found
in numerous references [16,17].
One of the important aspects of this process is the sequential placement of the incoming
tow. Repeated passes can effectively anneal and improve the tow interface properties and
help to increase final bond strength and reduce the bulk void content in the material, due
to repeated compaction at the roller region [16]. The effect of multiple passes on the
response of the composite material can be obtained by applying the following solution
technique to the heat transfer model:
1. The steady state solution is obtained for the placement of layer n. The temperature
solution for the surface layer is then applied to the material models.
2. A new layer is added, and the model recomputes the temperature field for a lamina of
n þ 1 layers. The temperature data for the second layer beneath the surface is extracted
and applied to the process models shown in Figure 2.
3. This procedure is continued until the maximum ply temperature drops below the glass
transition temperature, where high polymer viscosities inhibit the growth and/or
reduction of voids within the laminate and constrain further development of
intimate contact.
In Situ Strength Development for Thermoplastic Composite 1491

Table 1. Critical process parameters of the ATP head.


Model input Value
Initial thickness of tow 0.1778 mm
Width of tow 6.35 mm
Radius of roller (1) 15.8 mm
Radius of roller (2) 19.0 mm
Distance between rollers 80 mm from tacking roller
Length of region under preheat nozzle 75 mm
Length of region after consolidation 200 mm
Ti, Initial composite temperature 100 C
Tm, Initial mandrel temperature 25 C
Ta, Ambient air temperature 25 C
_ Gas flow rate
m, 50 L/min
Troller Roller temperature 100 C
Tgas N2 Gas temperature 750–900 C
x1, Location of preheater torch 75 mm from nip point location
x2, Location of main torch 35 mm from nip point

Table 2. Material properties used for APC-2/PEEK.


Model input Value
 Density 1584 kg/m3
kx Longitudinal conductivity 3.5 W/m K
ky Transverse conductivity 0.34 W/m K
Tg Glass transition temperature 143 C
Tm Melting temperature 334 C
cp Specific heat capacity 1370 J/kg K

The model geometry selected for this analysis corresponds to the actual specifications
used in the process. The initial temperatures and boundary conditions applied to this
steady state model are given in Table 1. The initial temperature of the composite material
is set at 100 C, which was found to be the average surface temperature during continuous
steady state conditions. Cold startup conditions are examined by setting the initial
mandrel temperature to room temperature values (25 C) with the roller temperatures
maintained at 100 C.
The material examined in the validation of the thermal model is comprised of a 6.35 mm
wide APC-2 PEEK (Polyetheretherketone) preimpregnated tape. The density of the
material is assumed constant, though small density gradients develop if the material
undergoes crystallization. The material properties used in the model correspond to the
transverse through-thickness properties of the composite prepreg material and are given
in Table 2.
Figure 3 is a plot of through-thickness temperature gradients taken at the interface of
each ply verses location on the composite surface. In this case, the velocity is set at
25 mm/s, and both torch heights are set at 4 mm. The two peaks correspond to the torch
locations, and the two sharp dips correspond to the roller locations. The melt temperature
and glass transition temperatures are also shown in this figure. One can see that the melted
region has a depth of two layers and the rubbery region (temperatures between the melt
1492 J. TIERNEY AND J. W. GILLESPIE, JR.
500
Composite–mandrel interface
Substrate surface
400 Composite surface
Tm
Temperature (°C)
300

200 Tg

100

0 50 100 150 200 250


Length of domain (mm)

Figure 3. Through-thickness temperature profile for an AS4/PEEK laminate (T1 ¼ T2 ¼ 900 C, v ¼ 25 mm/s,
H1 ¼ H2 ¼ 4 mm, F1 ¼ F2 ¼ 200 N) [12].

temperature and glass transition temperature) exists up to seven layers beneath the surface,
which demonstrates the need to consider multiple passes on material quality development
for this process. This model was validated for a wide variety of processing conditions with
results from this work given in the literature [16].

PRESSURE MODEL

A compressible squeeze flow model of a Newtonian fluid in a 2-D geometry, developed


by Ranganathan et al. [18] and Pitchumani et al. [19] is used to predict the pressure field
under the rollers. This pressure field is an important input for the coupled bonding model
(via intimate contact effects are discussed below) as well as tow consolidation.
Consolidation under the rollers is modeled as a squeeze flow continuum, in which the
rheological properties are dependent on the temperature, fiber volume fraction, and void
content. The consolidation model also defines the roller/substrate contact area
required for the heat transfer analysis. A macroscopic flow model is used to predict
material flow during consolidation of the fiber/resin system, where the fiber–resin-
void mixture is modeled as an equivalent homogeneous fluid, with the rheological
properties of the continuum dependent on the temperature, fiber volume fraction, and
the void content.
Figure 4 shows a close-up of the consolidation region under the compaction roller. In
this step, a tow of a given height, hi and width wi, enters the region under the roller where
it undergoes a thickness reduction due to the applied consolidation force. As a result, its
height reduces to hf while its width increases to wf. As the tow dimension in the x-direction
is much larger than the y- and z-directions, flow in the x-direction may be neglected.
The fluid motion under the compaction roller is governed by the continuity and
momentum equations in Cartesian coordinates. As the thickness of the prepreg tow is
much smaller than its width, the dominant velocity gradient terms in the momentum
equations are the shear strain rate through the gap, @vx =@z, and its derivative with respect
to the thickness direction, z.
In Situ Strength Development for Thermoplastic Composite 1493

Figure 4. Close-up of the consolidation region.

The following general expression of the macroscopic governing equation for


determining the pressure distribution under the roller is [15]:
 Z h Z  
@ @  dP z  1 @h
h þ  vy ðz ¼ 0Þ d þ C1 ð yÞ d dz þ  ¼ 0 ð2Þ
@ @y 0 dy 0   @

where  is a dummy variable of integration, C1( y) is a constant of integration from


the velocity profile, and * is the density of the fiber–resin-void mixture normalized
with respect to the density of the mixture in the absence of voids. The viscosity, , of
the resin–fiber system is found using an Arrhenius-type equation of the following
form [20],
 
B
 ¼ A exp ð3Þ
T

where A and B are empirically determined constants. For the APC-2/PEEK system,
A ¼ 132.95 N s/m2 and B ¼ 2969 K [21].
Since Equation (2) involves the second derivative of pressure with respect to y, two
pressure boundary conditions are required to solve this problem. On the surface ply, the
tow is unconstrained along its width, the appropriate pressure conditions are simply
atmospheric pressure and P( y ¼  w/2) ¼ patm. The terms within the square brackets in
Equation (2) denote the expression for vy(z). The unknowns vy(0) and C1(y) in the
expression are determined using the velocity boundary conditions at z ¼ 0 (tow-substrate
interface) and z ¼ h (tow-roller contact). The boundary conditions may be written in
a generalized form as given below:
   
@vy Kb @vy Kb
¼ vy ð0Þ; ¼ v y ð hÞ ð4Þ
@z z ¼ 0  @z z ¼ h 

where Kb/ is a parameter that determines the type of velocity boundary condition at the
tow-substrate interface. A value of Kb/ ¼ 1 implies a no-slip condition, i.e., vy(z ¼ 0) ¼ 0,
1494 J. TIERNEY AND J. W. GILLESPIE, JR.

Figure 5. Schematic of idealized surface elements being deformed to the dotted line with the application of
pressure.

whereas Kb/ ¼ 0 corresponds to a perfect slip condition. Similarly, the value of Kb/
decides the velocity boundary condition at the tow-roller interface. In the present case, a
no-slip condition exists at the tow-substrate interface, i.e., Kb/ ¼ 1, while a condition of
partial slip tending towards complete slip exists at the tow-roller interface, i.e., Kb/ ¼ 0.
The full solution details for this pressure model have been omitted here but can be found in
the literature [15–19].
For layers beneath the surface, a boundary condition of no width change is applied to
the model as the tow material is constrained by adjacent consolidated material. The change
in resulting tow dimensions under pressure is then just a function of void fraction within
the material.

INTIMATE CONTACT MODEL

The degree of intimate contact is defined as the fraction of the total area in contact at
any given time. It is directly related to the applied temperature and pressures and the
residence time at conditions, where the viscosity is low enough for the material to deform
and reduce surface roughness.
Dara and Loos [20] developed a model to describe intimate contact by considering
an irregular surface as a distribution of rectangular elements of varying size.
Their approach incorporates three submodels: a statistical distribution describing the
prepreg geometrical non-uniformity of the tow heights, a mechanics model simulating
the viscoelastic response of the composite to compressive loading, and an assumed
constitutive relationship for viscosity. This resulting intimate contact model was complex
in its formulation, and inconvenient to use in a practical sense. Lee and Springer [22]
followed the development of Dara and Loos with a simplified model where they
considered the surface a periodic array of identically sized elements. The application
of pressure causes the elements to deform and spread along the interface as shown in
Figure 5.
This model showed the dependence of applied pressure and temperature on the
development of intimate contact. Mantell and Springer [21] extended this model to
In Situ Strength Development for Thermoplastic Composite 1495

account for time-varying properties and process conditions. Based on their idealizations,
the degree of intimate contact Dic can be expressed geometrically as:

bðtÞ
Dic ¼ ð5Þ
wo þ bo

where wo and bo are given in Figure 5, and b(t) is the width after pressure has been applied
for a time t. By assuming that the volume of each element remains constant, and applying
the law of conservation of mass to a control volume of each element, the following
expression for the degree of intimate contact can be obtained
 Z tp 1=5
P
Dic ¼ Dico 1 þ C1 dt ð6Þ
0 mf

where

1
Dico ¼ ð7Þ
1 þ wo =bo

and
  
wo ao 2
C1 ¼ 5 1 þ ð8Þ
bo bo

In the above equations, Dico represents the amount of area that is in initial contact
before any force is applied, and C1 is a function of the size of the rectangular elements. The
quantity mf is the fiber–matrix viscosity, tp is the duration of the pressure cycle, and ao,
bo, and wo are the geometric parameters defining the idealized asperity elements as shown
in Figure 5. If the scale of the deformation is greater than the distance of a fiber from the
surface, the following expression for mf can be used [21]

1
mf ¼ pffiffiffiffiffiffiffi  ð9Þ
1  f=F

where f, is the fiber volume fraction, F is the packing factor, and  is the resin viscosity.
In this study, it is assumed that a resin rich layer is close to the bonding surface,
which results in f ¼ 0 and mf ¼  from Equation (3) [23]. By defining the roughness
parameter as:

Rc ¼ Dico C10:2 ð10Þ

and by discretizing the time integral in Equation (6), the degree of intimate contact
becomes:

"  #1=5
p =t 
tX
ðPiþ1 þ Pi Þ=2
Dic ¼ Rc 1 þ t ð11Þ
i¼0
ððTiþ1 þ Ti Þ=2Þ
1496 J. TIERNEY AND J. W. GILLESPIE, JR.

This roughness constant basically represents a ‘lumped’ coefficient based on


experimentally determined information of the surface asperities. The constant for AS4/
PEEK is approximated as 0.29 based on experimental measurements of an AS4/PEEK
thermoplastic composite system [19]. Thus, by knowing the pressure and temperature
dependent viscosity at each discrete point under the rollers, it is possible to determine the
degree of intimate contact.

HEALING MODEL

When two amorphous thermoplastic parts are brought into contact above the glass
transition temperature of the resin (or above the melt temperature for semicrystalline
material systems), interdiffusion of the polymer chains across the interface, referred to
as healing, contributes to the strength development of the bond area. The model used to
describe healing in this process is based on reptation theory of chain mobility [24,25]. In
this model, the melt material is considered as an entanglement of polymer chains that
restrict each other in movement. This entanglement with neighboring chains imposes a
topological constraint upon the chain restricting lateral motion within an idealized tube.
Chains are allowed to move backward and forward along the curvilinear length of the
tube, and only the chain ends can exit through the ends of the tube. Because the chain ends
are free to move in a random direction from the original tube, the memory of this tube
is gradually lost.
Wool and O’Connor [10] extended the work of de Gennes [24] governing the molecular
motion within the bulk to the motion of the molecular chains across the interface. In doing
so, he assumed that the molecular chain entanglements at the interface affect the chain
motion in the same manner as the entanglements in the bulk. Reptation theory has shown
that the average interpenetration distance of the chain ends across the interface, ,
increases with time and follows the relation:

 t 1=4
¼ ð12Þ
M

where M is the molecular weight. Kim and Wool [26] demonstrated how the strength
development is proportional to the average interpenetration distances of the polymers
across the interface. From this work, the dependence between the strength development
over time and the molecular weight for monodisperse systems is found to be:

 ð13Þ

Substituting in Equation (12) gives:

 t 1=4
¼ ð14Þ
M

Wool showed that when the time approaches the reptation time, the diffusion and
randomization stages are complete and the full potential strength of the bond, 1 , has
In Situ Strength Development for Thermoplastic Composite 1497

been reached [26]. The ultimate strength of the material after full healing can then be
represented by:

 t 1=4
r
1 ¼ ð15Þ
M

where tr is the reptation time at a given reference temperature, and also represents the
characteristic time for full healing to occur. Thus, the degree of healing, Dh can be defined
by the ratio of the autohesive strength to the cohesive strength as follows:


Dh ¼ ð16Þ
1

Thus, by combining Equations (14) and (16), we get the following expression:
 1=4
 t
Dh ¼ / ð17Þ
1 tr

To apply this model to the ATP process the solution is discretized to account for the
nonisothermal nature of this process. Bastien and Gillespie [9] developed such a solution
for thermoplastic amorphous polymer systems which was extended to PEEK-based
thermoplastic systems by Agarwal [8]. Their model involved discretizing Equation (17) and
summing these discretizations over each isothermal step as follows

 1=4 X " 1=4  1=4 #


h =
1 j  j1
Dh ¼   1=4 ð18Þ
tr j¼1 aT Tj

where aT represents a shift factor applied to the reptation time tr (0.11 s), at a reference
temperature Tref of 400 C, and  h is the elapsed time for the healing process. Sonmez and
Hahn [27] also modified the isothermal equation using an approach similar to that of
Bastien and Gillespie to capture nonisothermal healing during thermoplastic composite
tape placement.
Wool pointed out that Equation (17) is valid for low molecular weight thermoplastics in
the range Mc<M<8Mc, where Mc is the critical entanglement molecular weight [28,29].
For typical engineering thermoplastics, the molecular weight range is higher than 8Mc,
so that the interpenetration depth and minor chain length do not have to reach their
maximum values. This results in a welding time tw, which is less than tr, resulting in a more
general expression for the degree of healing as follows:
 1=4
 t
Dh ¼ / ð19Þ
1 tw

From this equation, one can see that the linear relationship between strength and time
to the one quarter power is followed until the cohesive strength of the material is attained,
at which point, the strength is independent of time, and the interface and bulk resin are
indistinguishable.
1498 J. TIERNEY AND J. W. GILLESPIE, JR.

More recently, Yang and Pitchumani [30,31] developed an integral formulation for
nonisothermal healing from first principles:
Z t 1=4
1
Dh ðtÞ ¼ dt ð20Þ
0 tw ðTÞ

The time required for welding tw of AS4/PEEK was found by Yang and Pitchumani
through a series of isothermal lap shear experiments which compare the measured strength
values for different time–temperature cases with the maximum strength value for this
material. An Arrhenius relationship was then used to determine the welding time with
respect to a reference temperature Tref as follows:
  
Ea 1 1
tw ¼ A exp  ð21Þ
R T Tref

where A ¼ 0.11 s, Ea ¼ 57.3 kJ/mol, Tref ¼ 400 C, and R is the universal gas constant
(8.314 kJ/kmol K). In the case of PEEK, it is assumed that healing occurs above the melt
temperature (338 C), where the polymer chains are free to migrate and the material is
amorphous. The degree of healing for each interface is found for each pass and summated
to give the final degree of healing based on the entire processing history. Interface healing
development for the ATP process using the models of Bastien and Gillespie, as well as
Yang and Pitchumani, are used to determine the final strength development from the
coupled bonding model presented in the following section.

COUPLED BONDING MODEL

Intimate contact is necessary to initiate healing across the interface. As a result, a


coupled bonding model is used, which takes into account this time-dependent strength
development [13]. The degree of healing for each spatial increase in intimate contact is
integrated throughout the process history to give the final degree of bonding. Thus, using
a convolution integral, the net bond strength developed at the interface in dimensionless
form can be found from the following:
Z p  
d Dic
Db ðh Þ ¼ Dic ð0ÞDh ðh Þ þ Dh d 0 ð22Þ
0 d t0

Here  0 is the time available for the incremental area to heal, and Dic(0) is the initial
degree of intimate contact at  ¼ 0. This equation is presented schematically in Figure 6.
At any time within the process, the material will have a certain degree of intimate contact
with some bond strength associated with that area. With any increase in area of contact
under appropriate thermal conditions, the healing for this newly formed interface will
commence. The overall bond strength is then simply a summation of each interfacial area
times its degree of healing.
In the tow placement process, the residence time of the tows under the consolidation
roller,  ic, which also represents the available time for intimate contact to develop,
is small relative to the time available for healing to occur,  h. Thus, the development of
In Situ Strength Development for Thermoplastic Composite 1499

Intimate contact (τ) ∆Dic

Healing
∆Dh

Initial bonding

Figure 6. Schematic of how bond strength develops as a function of intimate contact and healing.

some intimate contact under the roller is considered to develop instantaneously. By using
a Heaviside step function, the degree of bonding then becomes:

Z p
Db ðh Þ ¼ Dic ð0ÞDh ðh Þ þ Dh ð   0 ÞHð   0 ÞDic ð Þ d ð23Þ
0

This equation is a simplified yet accurate evaluation of the degree of interfacial bonding
for this process. The accuracy of Equation (23) was examined in a separate study, where
the bond strength predictions were found to be in good agreement with experimental data
obtained from short beam shear (SBS) testing of in situ consolidated panels [13].

MODEL RESULTS

Figure 7 shows the pressure distribution in the plane of travel under the main
consolidation roller with variation of applied force. The pressures are low upon initial
entry under the roller, as the melted resin system has a low viscosity, thus allowing for
significant displacement in the transverse direction. As the resin cools due to further
contact with the roller, the viscosity increases, which allows for load reaction by
the material. As a result, the maximum pressure is found to occur towards the end of
roller contact which also corresponds to the region where the roller surface is parallel
to the plane of travel. Increasing the force by 67% is found to increase the maximum
pressure by only 40%, but is found to broaden the load distribution under the roller.
A uniform pressure distribution is applied to the remaining layers beneath the surface
as the pressure profile is more uniform at the interface between these layers below the
surface ply.
Multiple passes can have a significant effect on the overall bulk properties of the
material. This is particularly true in the strength development of ATP composites.
As the thickness of the melt and glass transition region extends well beyond the thickness
of the first tow, the interfacial area will increase under the rollers and then be in a position
to heal downstream of these zones as long as temperatures remain sufficiently high.
Figure 8 shows model prediction of intimate contact development for a single interface
1500 J. TIERNEY AND J. W. GILLESPIE, JR.
200
180 F=150N
160
F=200N
140

Pressure (MPa)
F=250N
120
100
80
60
40
20
0
0.00E+00 1.14E-04 2.31E-04 3.45E-04 4.60E-04
Location under roller (m)

Figure 7. Pressure distribution under the main consolidation roller with variation in applied force (v ¼ 40 mm/s
H1 ¼ H2 ¼ 4 mm, Trollers ¼ 100 C, Rr ¼ 19 mm).

0.8
Degree of intimate contact (Dic)

v = 20mm/s
v = 40mm/s
v = 60mm/s
0.6
v = 80mm/s

0.4

0.2

0
0 First pass Second pass Third pass

Figure 8. Development of intimate contact with subsequent passes with variation in velocity. (H1 ¼ H2 ¼ 3 mm,
Trollers ¼ 100 C, F1 ¼ F2 ¼ 200 N).

undergoing repeated passes. In this case, the process velocity was varied from
20 to 80 mm/s. The force applied to each roller was 200 N, and the roller temperatures
were set at 100 C. Both torches were fixed at a height of 3 mm. The time domain was
extrapolated under both rollers to observe the rate at which intimate contact develops.
In the case of a roller contact point, the closing speed and temperature is greatest at the
roller entry point, which results in the highest growth of interfacial area.
Higher intimate contact is seen to develop at lower velocities in the first layer due to
higher surface temperatures and lower viscosity for longer dwell times. For any
appreciable bond strength at these forces, the best range of process velocity is seen to
In Situ Strength Development for Thermoplastic Composite 1501
0.6

0.5
Degree of intimate contact (Dic)

0.4
h1 = h2 = 3mm
h1 = h2 = 6mm
0.3
h1 = h2 = 9mm

0.2 h1 = h2 = 12mm

0.1

0
0 1 2
Number of passes

Figure 9. Degree of intimate contact development of a single ply with variation of torch height (v ¼ 40 mm/s,
Trollers ¼ 100 C, F1 ¼ F2 ¼ 200 N).

exist between 20 and 40 mm/s. At higher velocities intimate contact development is very
low, which result in parts with inferior quality. Another option is to significantly increase
the forces applied to these rollers. The maximum available force with the hardware
configuration is currently 350 N.
Figure 9 is a similar plot of intimate contact development where both torch heights are
varied from 3 to 12 mm in 3 mm increments. The process velocity is set to 40 mm/s and the
roller forces are maintained at 200 N each. As expected, the lower torch heights result
in higher temperatures within the laminate, and further intimate contact development.
The interesting point to note is that, while more heat will aid in reducing the viscosity on
the surface and allow for good intimate contact, it also allows for void growth away
from the rollers [16]. As a result, it is more desirable to apply more force to the
rollers rather than more heat to increase intimate contact and reduce void growth in the
surface plies.
Figure 10 is a plot of the final degree of healing through the thickness of a 16-ply lamina
with variation of torch heights from both the Bastien and Gillespie (B/G) and Yang and
Pitchumani (Y/P) models. Both forces are set to 200 N, and the process velocity is set at
40 mm/s. As expected, the tool acts as a heat sink in reducing the final degree of healing,
and the healing on the surface plies is a little lower in magnitude due to lack of repeated
processing (i.e., multiple passes). In all cases the results for the Bastien and Gillespie are
conservative and predict slightly lower levels of degree of healing than the integral
formulation. For highly nonisothermal thermal histories, the model of Bastien and
Gillespie was found to underestimate healing during heat-up and overestimate healing
during cool-down phase of the ATP process. In summary, the model of Yang and
Pitchumani was found to be a more rigorous model for the nonisothermal healing process.
Note that all models give equivalent results when temperatures approach isothermal
conditions.
1502 J. TIERNEY AND J. W. GILLESPIE, JR.
3mm/3mm (B/G)
6mm/6mm (B/G)
1 9mm/9mm (B/G)
3mm/3mm (Y/P)
0.9 6mm/6mm (Y/P)
9mm/9mm (Y/P)
0.8
Degree of healing (Dh)

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Layer

Figure 10. Final degree of healing through the thickness of a 16-ply lamina. (F1 ¼ F2 ¼ 200 N, Trollers ¼ 100 C,
v ¼ 40 mm/s).

30 mm/s (B/G)
0.6 40 mm/s (B/G)
50 mm/s (B/G)
60 mm/s (B/G)
Degree of bonding (Db)

0.5 30 mm/s (Y/P)


40 mm/s (Y/P)
50 mm/s (Y/P)
0.4 60 mm/s (Y/P)

0.3

0.2

0.1

0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Layer number

Figure 11. Through thickness final bond strength of a 16-ply lamina with variation of process velocities.
(F1 ¼ F2 ¼ 350 N, T1 ¼ T2 ¼ 850 C, Trollers ¼ 100 C, H1 ¼ H1 ¼ 3 mm).

Figure 11 is a plot of the final coupled bond strength (from Equation (23))
with variation in process velocity for a 16-ply lamina with both rollers applying
350 N, and both torch heights are maintained at 3 mm. These settings represent the
best possible process conditions applied to the prepreg material being placed, namely
near maximum force and maximum heat input. With these conditions, the maximum bond
strength was still only 74%, which means that the hardware configuration, force, and heat
input are inadequate for manufacturing panels with optimum properties. These results,
In Situ Strength Development for Thermoplastic Composite 1503

however, do provide the requirements needed in order to improve the overall design of
the ATP head.

EXPERIMENTAL PROCEDURES

Short beam shear (SBS) experiments were carried out according to ASTM standard
designation: D 2344-84. The SBS specimens were cut to dimensions of 6.4 mm (0.25 in.)
width and a length of six times the specimen midpoint thickness. The support span of the
machine was set at four times the specimen midpoint thickness, and the specimens were
aligned so that the midpoint was centered. The tests were performed on an Instron
universal testing machine at room temperature with a 454 kg (1000 lb) load cell with a
crosshead speed of 1.3 mm/min (0.05 in./min). Twelve samples were tested for each panel
so as to obtain a satisfactory average and coefficient of variation. Six of these samples were
reconsolidated in an autoclave to determine the baseline strength and whether any
degradation occurred during the ATP process. The first maximum load was recorded as
the failure load. The shear strength is calculated from the following:

0:75Pf
Smax ¼ ð24Þ
bd

where P is the failure load, b is the midpoint width, and d is the midpoint thickness. The
predicted strength values are found by using the autoclave interfacial shear strength of
AS4/PEEK as 1 found from SBS experiments, in conjunction with Equation (17). The
centerline value of degree of bonding is used, as this point corresponds to the location of
maximum interlaminar shear stress in the material during testing. The centerline degree
of bonding is predicted for a wide variety of processing conditions and compared with
experimental results.

EXPERIMENTAL RESULTS

Figure 12 is a plot of experimental results for in situ and reconsolidated shear strengths
with variation in nozzle heights. The baseline values (i.e., reconsolidated in autoclave)
remain fairly constant for these conditions, and the average of these values is taken as the
baseline theoretical strength in the model. Also shown in this figure are centerline strength
predictions (i.e., mid-thickness of the laminate) from the coupled bonding model. Fairly
good agreement is obtained between model predictions and experimental results for the
in situ consolidation. The strengths drop away with increasing torch heights as expected
due to lack of adequate heating of the material. In addition, results from the two
nonisothermal healing models are presented. It is found that the integral formulation from
Yang and Pitchumani provided a better correlation with experimental results.
Figure 13 is a similar plot where the process velocity is increased from 30 to 60 mm/s in
10 mm/s increments. The baseline values do show some drop off at the lower process
speeds and this may be attributed to the fact that the torch heights in this study were set to
3 mm vertical height, thus yielding some small degree of degradation. In summary, good
agreement is obtained between model results and experimental values, with the integral
formulation yielding a better fit to the data.
1504 J. TIERNEY AND J. W. GILLESPIE, JR.

Reconsolidated
In situ
100 Bastien and Gillespie
Yang and Pitchumani
80
Strength (MPa)

60

40

20

0
3 6 9 12
Both torch heights (mm)

Figure 12. Resulting strength with variation in nozzle height. Also shown is the autoclave reconsolidated
strength from the same panels. (v ¼ 40 mm/s, F1 ¼ F2 ¼ 270 N, Trollers ¼ 100 C).

100
90
80
70 ATP
Reconsolidated
Strength (MPa)

60 Yang and Pitchumani


Bastien and Gillepsie
50
40
30
20
10
0
30.0 40.0 50.0 60.0

Speed (mm/s)

Figure 13. Resulting strength with variation process speed. Also shown are autoclave reconsolidated
strength results from the same panels. (v ¼ 40 mm/s, F1 ¼ F2 ¼ 270 N, Trollers ¼ 100 C, H1 ¼ H2 ¼ 3 mm).

CONCLUSIONS

In this study, a strength model was developed and validated with experimental results
for the ATP process. An intimate contact model, which models the polymer–polymer
interface as a series of periodic rectangular asperities, was applied to the ATP process
to determine the development of interfacial area as a function of processing conditions.
In Situ Strength Development for Thermoplastic Composite 1505

A healing model based on reptation theory was applied to this process to determine the
degree of polymer chain diffusion across the interface. A coupled bonding model, which
takes into account the interdependence between both the intimate contact development
under the rollers and the healing along the length of travel of the process, was then used to
determine the final strengths of automated tow placement (ATP) laminates under varying
process conditions. A series of short beam shear (SBS) experiments were carried out, and
they showed good agreement with model predictions for varying process velocities and
torch heights.
From model predictions and experimental results, it was found that the bond strength
varied through the thickness of the laminate due to heat sink effects and variation in the
processing history of each interface. It was found that the center of the laminates would
possess the highest bond strength, as this region is not close enough to the substrate such
that heat sink effects would not be present, and would experience the benefit of repeated
processing from the surface. Although the center of the laminate generally experiences the
maximum load when the part undergoes shear, either through bending forces or in-plane
motion, it is important that all the plies are bonded to the maximum degree in order to
maximize part performance. This can be accomplished by carrying out repeated passes of
the ATP head on the surface without any tow being deposited. This ensures that each
interface undergoes the same thermal and pressure cycle such that the resulting interface
bond strength is at a maximum in all plies. As far as the substrate region is concerned,
preheating the tool will help increase the bond strength, but this solution adds
significant costs to manufacturing the part. One potential solution is to also carry out
dry placement (conduct passes without incoming material) on the first few plies with a
short duration between passes, such that the localized temperatures are high enough
to compensate for these heat sink effects and the maximum strength is obtained at these
first few interfaces.

REFERENCES

1. Pitchumani, R., Gillespie, Jr., J.W. and Lamontia, M.A. (1997). Design and Optimization of a
Thermoplastic Tow-placement Process with In situ Consolidation, J. Composite Materials, 31(3):
244–275.
2. Insert Sawicki, A., Schultze, E., Fitzwater, L. and Harris, K. (1995). Structural Qualification of
V-22 EMD Tow-placed Aft Fuselage, Annual Forum Proc., American Helicopter Soc., 2:
1641–1653.
3. Hinkley, J.A., Messier, B.C. and Marchello, J.M. (1997). Effect of Pressure
in Thermoplastic Ribbon Thermal Welding, In: Int. SAMPE Sym. Proc., Vol. 42(2),
pp. 1209–1216.
4. Weihn, M.P. and Hale, R.D. (2002). Low Cost Robotic Fabrication Methods for Tow Placement,
In: Int. SAMPE Sym. Proc., Vol. 47, pp. 1842–1852.
5. Manson, J.E., Wakeman, M.D., Hagstrand, P.O., Bonjour, F. and Bourban, P.E. (2002).
Robotic Tow Placement for Local Reinforcement of Glass Mat Thermoplastics (GMTs),
Composites Part A, 33(9): 1199–1208.
6. Felderhoff, Klaus D. and Steiner, K.V. (1993). New Compact Robotic Head for Thermoplastic
Fiber Placement, In: Proc. 38th Int. SAMPE Sym., Vol. 38(1), pp. 138–151.
7. Gillespie, J.W., Howie, I. Don, R.C. and Holmes, S.T. (1997). Adjustable Hot Gas Torch Nozzle,
U.S. Patent No. 5,626,472, issued May 6, 1997.
8. Agarwal, V. (1991). The Role of Molecular Mobility in the Consolidation and Bonding of
Thermoplastic Composites, PhD Thesis, University of Delaware, Newark, Delaware.
1506 J. TIERNEY AND J. W. GILLESPIE, JR.

9. Bastien, L.J. and Gillespie, Jr., J.W. (1991). A Nonisothermal Healing Model for Strength and
Toughness of Fusion Bonded Joints for Amorphous Thermoplastics, Polymer Engineering and
Science, 31(24): 1720–1730.
10. Wool, R.P. and O’Connor, K.M. (1981). A Theory of Crack Healing in Polymers, J. Applied
Physics, 52(10): 5953.
11. Wool, R.P. (1983). Molecular Aspects of Tack, Rubber Chemistry and Technology, 57: 307.
12. Wool, R.P. (1993). Polymer Entanglements, Macromolecules, 26: 1564.
13. Butler, C.A., Pitchumani, R., Gillespie, Jr., J.W. and Wedgewood, A.G. (1994). Coupled
Effects of Healing and Intimate Contact on the Strength of Fusion-Bonded Thermoplastics,
In: Proc. 10th Annual ASM/ESD Advanced Composites Conference, Dearborn, MI, pp. 595–604.
14. Heider, D., Tierney, J.J., Piovoso, M.J. and Gillespie, Jr., J.W. (1999). Artificial Neural
Network Modeling of the Automated Thermoplastic Composite Tow-placement System,
J. Materials Processing and Manufacturing Science, 7(4): 360–379.
15. Pitchumani, R., Don, R.C. and Gillespie, Jr., J.W. (1994). Simulation of the Transients in
Thermoplastic Fiber Placement In: Proc. 9th International SAMPE Symposium, pp. 1521–1535.
16. Tierney, J. and Gillespie, J.W. (2003). Modeling of Heat Transfer and Void Dynamics for the
Thermoplastic Composite Tow Placement Process, J. Composite Materials, 37(19): 1745–1768.
17. Tierney, J.J., Heider, D. and Gillespie, Jr., J.W. (1997). Welding of Thermoplastic Composites
Using the Automated Tow Placement Process: Modeling and Control, In: Proc. ANTEC 97,
Society of Plastics Engineers, pp. 1165–1170.
18. Ranganathan, S., Advani, S.G. and Lamontia, A.A. (1993). Model for Consolidation and Void
Reduction during Thermoplastic Tow Placement, In: 25th International SAMPE Technical
Conference, October 1993, pp. 620–631.
19. Pitchumani, R., Don, R.C., Gillespie, Jr., J.W. and Ranganathan, S. (1994). Analysis of On-Line
Consolidation during the Thermoplastic Tow-placement Process, In: Alam, M.K. and
Pitchumani, R. (eds), Heat and Mass Transfer in Composites Processing, ASME Press.
20. Dara, P.H. and Loos, A.C. (1985). Thermoplastic Matrix Composite Processing Model, Virginia
Polytechnic Institute, CCMS-85-10.
21. Mantell, S.C. and Springer, G.S. (1992). Manufacturing Process Models for Thermoplastic
Composites, J. Composite Materials, 26(16): 2348–2377.
22. Lee, W.I. and Springer, G.S. (1987). A Model for the Manufacturing Process of Thermoplastic
Matrix Composites, J. Composite Materials, 21: 1017.
23. Pipes, R.B., Hearle, J.W.S., Beaussart, A.J., Sastry, A.M. and Okine, R.K. (1991). A
Constitutive Relation for the Viscous Flow of an Orientated Fiber Assembly, J. Composite
Materials, 25: 1204.
24. De Gennes, P. (1971). Reptation of a Polymer Chain in the Presence of Fixed Obstacles,
J. Chemical Physics, 55(2): 572.
25. Doi, M. and Edwards, S.F. (1978). Dynamics of Concentrated Polymer Systems, Faraday
Transactions, 74(1): 1789.
26. Kim, Y.H. and Wool, R.P. (1983). Theory of Healing at a Polymer–Polymer Interface,
Macromolecules, 16(7): 1115–1728.
27. Sonmez, F.O. and Hahn, H.T. (1997). Analysis of the On-line Consolidation Process in
Thermoplastic Composite Tape Placement, J. Thermoplastic Composite Materials, 10: 543–572.
28. Wool, R.P., Yuan, B.L. and McGarel, O.J. (1989). Welding of Polymer Interfaces, Polym. Eng.
Sci., 29(19): 1340–1367.
29. Wool, R.P. (1995). Polymer Interfaces, Structure and Strength, Hanser Publishers, Munich,
Germany, Vienna and New York, pp. 74–81.
30. Yang, F. and Pitchumani, R. (2002). Healing of Thermoplastic Polymers at the Interface under
Nonisothermal Conditions, Macromolecules, 35: 3213–3224.
31. Yang, F. and Pitchumani, R. (2003). Nonisothermal Healing and Interlaminar Bond Strength
Evolution During Thermoplastic Matrix Composites Processing, J. Polymer Composites,
24(2): 263–278.

You might also like