Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

Deformation of an amorphous polymer during the fused-filament-fabrication

method for additive manufacturing

Claire McIlroya) and Peter D. Olmstedb)

Department of Physics and Institute for Soft Matter Synthesis and Metrology, Georgetown University, Washington
DC 20057

(Received 4 November 2016; final revision received 30 January 2017; published 1 March 2017)

Abstract
Three-dimensional (3D) printing is rapidly becoming an effective means of prototyping and creating custom consumer goods. The most
common method for printing a polymer melt is fused filament fabrication (FFF) and involves extrusion of a thermoplastic material through a
heated nozzle; the material is then built up layer-by-layer to fabricate a 3D object. Under typical printing conditions, the melt experiences
high strain rates within the FFF nozzle, which are able to significantly stretch and orient the polymer molecules. In this paper, we model the
deformation of an amorphous polymer melt during the extrusion process, where the fluid must make a 90 turn. The melt is described by a
modified version of the Rolie–Poly model, which allows for flow-induced changes in the entanglement density. The complex polymer config-
urations in the cross section of a printed layer are quantified and visualized. The deposition process involving the corner flow geometry dom-
inates the deformation and significantly disentangles the melt. V C 2017 The Society of Rheology.

[http://dx.doi.org/10.1122/1.4976839]

I. INTRODUCTION the nozzle with respect to the previously printed layer (Fig.
1), whereas the width of the layer is determined by a combi-
Fused filament fabrication (FFF), also know as fused
nation of the flow rate, surface tension, and viscoelasticity.
deposition modeling (FDM) [1], is a powerful additive-
The speed of the material flowing through the nozzle is con-
manufacturing tool. The simple-to-use technology allows the
trolled to prevent drawing and buckling, so that the width of
fabrication of complex geometries via build-preparation soft-
the layer is approximately equal to the nozzle diameter.
ware, as well as printed parts with locally controlled proper-
Upon deposition, the melt bonds, cools, and solidifies
ties such as density and porosity [2]. FFF is now considered
with adjoining material so that the structure of the final
indispensable for the rapid manufacturing of concept models,
object consists of a number of partially bonded filaments.
functional prototypes, and customized end-use parts.
Much of the literature to date has focused on how FFF
In FFF, a solid thermoplastic filament is fed into a
parameters, such as build style, raster width, and raster angle,
machine via a pinch-roller mechanism, as shown by the sim-
affect the material properties [7–10]. Analytical models,
plified schematic in Fig. 1. FFF systems contain multiple
based on classical lamination theory combined with the
contractions between the pinch roller and the nozzle exit; for
Tsai-Wu failure criterion, are used to predict the tensile
simplicity Fig. 1 shows a single contraction between the
strength of the bond [8], and adhesives can be used to alter
heated and final sections of the nozzle. The most common
the bonding behavior [11].
printing material investigated is acrylonitrile butadiene sty-
In the absence of a postdeposition cross-linking process,
rene (ABS), an amorphous polymer melt containing rubber
bonding between layers is thermally driven and significantly
(butadiene) nanoparticles [3]. FFF machines can also print
affected by print temperature [12]; higher temperatures can
parts from amorphous polycarbonate [4], semicrystalline
enable better adhesion between printed layers and therefore
poly-lactic acid [5], and other thermoplastic materials [6].
stronger mechanical strength of the final printed part, while
The feedstock is melted and extruded through a nozzle,
temperatures that are too high lead to polymer degradation
with the solid portion of the filament acting as a piston to
and weakening of the product [13]. Recently, carefully cali-
push the melt through. A three-dimensional object is con-
brated infrared imaging has been developed to extract the
structed by printing the extrudate layer-by-layer onto a build
temperature profile at the weld [14], and finite-element anal-
plate. As the material is deposited, the nozzle moves in the
ysis has been used to examine temperature gradients at the
xy-plane to create a prescribed pattern, and the platform
nozzle exit [15]. Laser-assisted heating is proposed to
moves in the z-direction for additional layers to be built. The
improve the thermal-bonding process and consequently the
thickness of the single layer is determined by the height of
strength of the printed part [16].
A number of studies have investigated the thermal-
welding of polymer molecules (e.g., [17,18]), however, these
a)
Author to whom correspondence should be addressed; electronic mail:
studies focus on the diffusive behavior of melts in an equilib-
cm1509@georgetown.edu rium state. During FFF, the polymer experiences large shear
b)
Electronic mail: pdo7@georgetown.edu rates in the nozzle and rapid temperature changes that are
C 2017 by The Society of Rheology, Inc.
V
J. Rheol. 61(2), 379-397 March/April (2017) 0148-6055/2017/61(2)/379/19/$30.00 379
380 C. MCILROY AND P. D. OLMSTED

where Me is the molecular weight between entanglements.


The Rouse and reptation times of a polymer at a given refer-
ence temperature T0 are given by [26]

s0R ¼ s0e Zeq


2
; (2a)
!
3:38 4:17 1:55
s0d ¼ 3s0e Zeq
3
1  pffiffiffiffiffiffiffi þ  pffiffiffiffiffiffiffi3 ; (2b)
Zeq Zeq Zeq

respectively, where s0e is the Rouse time of one entanglement


segment.
Due to the nonisothermal conditions of FFF, the
temperature-dependent rheology must be considered. We
account for this by scaling the relaxation times by a shift fac-
FIG. 1. Simple schematic of typical FFF process, as described in text. In tor a(T), typically measured by rheology, which has the
frame of moving nozzle, the melt exits the nozzle at speed UN and the build
plate moves at speed UL in the y^-direction. The current printed layer is well-known Williams-Landel-Ferry (WLF) form [27]
denoted Lp and the middle of the layer is denoted mp . Welds occur at the  
interface between layers. The layer thickness H is typically less than the noz- C1 ðT  T0 Þ
zle diameter 2R. See Appendix C for typical model parameters. aðT Þ ¼ exp ; (3)
T þ C2  T0
expected to significantly deform the polymer microstructure. for temperature T and constants C1 and C2. At equilibrium,
A nonequilibrium microstructure will affect polymer diffu- the Rouse and reptation times of a melt are given by
sion [19–21] and consequently welding. For example, it is
suggested that polymer alignment in the flow may lead to seq 0
R ðTÞ ¼ sR aðTÞ; (4)
debonding of the layers and create defects in the final printed
object [22]. and
In this paper, we employ a continuum molecularly aware
polymer model [23], which we modify to incorporate flow- seq 0
d ðTÞ ¼ sd aðTÞ: (5)
induced changes in the entanglement density, to describe the
behavior of a typical amorphous polymeric printing material Momentum balance is given by
during the FFF extrusion process. Assuming steady-state and
a uniform temperature profile, the printing flow and polymer Du
q ¼ r  r; (6)
configuration tensor within a cylindrical nozzle and during Dt
the subsequent deposition are calculated using a simple map-
ping to represent the 90o turn and deformation into an for mass density q, fluid velocity u, and the material deriva-
elliptical-shaped layer. We quantify and visualize the poly- tive D=Dt ¼ ð@=@tÞ þ ðu  rÞ. In steady state, we solve
mer microstructure across a printed layer postextrusion and
r  r ¼ 0; (7)
investigate the effect of print speed on disentanglement. The
relaxation of this deformation and the effect on weld proper- for stress tensor r. The total stress in the polymer melt com-
ties will be considered elsewhere. prises solvent and polymer contributions

r ¼ pI þ Ge ðA  IÞ þ 2ls ðK þ KT Þ; (8)


II. MODEL FOR FFF
A. Modified Rolie–Poly model with flow-induced where p is the isotropic pressure and the velocity gradient
disentanglement tensor is denoted Kab ¼ rb ua . The polymer contribution to
the stress is given by the plateau modulus Ge multiplied by
A linear polymer melt is well described by the Doi–
the polymer deformation tensor
Edwards tube model [24]. In this paper, we implement a varia-
tion of this standard theory known as the Rolie–Poly model hRRi
[23] and include a new feature that allows for flow induces A¼ ; (9)
3R2g
changes in the entanglement density [25]. Although the
Rolie–Poly model does not allow for second normal stresses in for end-to-end vector R and radius of gyration Rg. Figure
axisymmetric flows, it provides a simple one-mode constitutive 2(a) shows the polymer microstructure, defined by tensor A,
equation for the stress tensor to describe entangled polymers. as an ellipsoid; a sphere represents an undeformed polymer
At equilibrium, the entanglement number of a melt of at equilibrium ðA ¼ IÞ with radius Rg, whereas an ellipse
molecular weight Mw is defined by signifies stretch and orientation. For times shorter than se,
Mw Rouse modes corresponding to lengths shorter than Me con-
Zeq ¼ ; (1)
Me tribute to a background viscosity defined as [28]
POLYMER FLOWS DURING ADDITIVE MANUFACTURING 381

contribution plus a convective one determined by the rate of


entanglement loss [25]
 
1 1 1 dtrA
¼ þb K:A ; (13)
sd ðT; c_ Þ seq
d T
ð Þ trA dt

where the temperature-dependence of the equilibrium repta-


tion time is given in Eq. (5). The reptation time therefore
implicitly depends on the shear rate c_ . The Rouse time does
not depend on the local shear rate and is given in Eq. (4)
under flow conditions. The steady-state constitutive curve
defined in Eqs. (8), (11), and (12) demonstrates the shear-
thinning behavior typical of FFF-printed materials and is dis-
cussed in Appendix A; increasing the CCR parameter b acts
to suppress excess shear-thinning behavior [29].
In steady state shear, Eq. (12) reduces to

1
¼ ; (14)
1 þ bArs c_ seq
d

where Ars is the shear component of A. Larger shear rates


impose a greater deformation on the polymer microstructure
and the resulting alignment reduces the entanglement frac-
tion. Ianniruberto compared this flow-induced disentangle-
ment theory [25] to molecular dynamics simulations of
simple steady shear flow conducted by Baig et al. [32] for a
wide range of shear rates c_ , finding that b ¼ 0:15 gives the
FIG. 2. (a) Visualization of polymer as a deformed sphere under shear flow. best fit to the entanglement loss data for Zeq ¼ 14. The theory
(b) Velocity profiles w(r) and ellipsoidal representation of polymer deforma- is also compared to the step strain response experiments
tion ðZeq ¼ 37; b ¼ 0:3Þ across the nozzle radius for fast and slow print
cases corresponding to Wi N ¼ 13 and 2 (Table I). reported by Takahashi et al. [33] for a polystyrene melt with
Zeq ¼ 12; in this case good agreement is found for b ¼ 0:25.
Inhomogeneous disentanglement has also been found in dis-
p2 Ge eq sipative particle dynamics simulations by Khomami et al.
ls ¼ s : (10)
12 Zeq R [34,35] of polymer melts with Zeq ¼ 13; 17.
In the following, we show results for b ¼ 0:3. For our
We assume that the polymer deformation tensor A satisfies model parameters, the constitutive curve is monotonic and
the Rolie–Poly equation [23] we avoid shear-banding effects. The effect of increasing b to
unity in our FFF model (as in [23]), which gives the most
DA 1
¼ K  A þ A  KT  ðA  IÞ extreme case of disentanglement, is shown in Sec. V B. We
Dt sd ðT; c_ Þ discuss the effect of choosing smaller b in Appendix A.
rffiffiffiffiffiffiffi ! rffiffiffiffiffiffiffi !
2 3 trA
 1 Aþb ðA  IÞ ; (11) B. FFF parameters and the printing process
sR ðT Þ trA 3
In the following, Eqs. (7), (8), (11), and (12) are solved to
where trA denotes the trace of tensor A. Convective con- determine the melt behavior during steady-state extrusion.
straint release (CCR) is incorporated via the parameter b Extrusion is treated in two stages in a frame fixed with the
[29]. The reptation and Rouse times are denoted sd ðT; c_ Þ and nozzle, where the build plate moves in the y^ direction.
sR ðTÞ, respectively. First, the melt flows through a fixed, vertically orien-
When a melt is subjected to flow, entanglements may be tated nozzle at mass-averaged speed UN. The nozzle is cir-
lost via convection and new entanglements made by reptation. cular in shape so that the flow is axisymmetric. The melt is
We incorporate flow-induced changes in the entanglement then deposited onto the build surface, which moves hori-
fraction  ¼ Z=Zeq to the Rolie–Poly model via the recent zontally at mass-averaged speed UL. During this deposi-
kinetic equation of Ianniruberto and Marrucci [25,30,31] tion, the fluid must speed up and deform to make a 90
  turn. The layer thickness H is typically less that the nozzle
D 1 dtrA 1 diameter, so that the shape of the layer is roughly elliptical
¼ b K : A   þ eq ; (12)
Dt trA dt sd ðT Þ [12]. The print temperature TN is assumed to be uniform
across the nozzle radius and throughout the deposition, and
where entanglement loss can be modified by varying the we assume that the flow is steady (see Appendix C for
CCR parameter b. The reptation rate is given by a thermal details).
382 C. MCILROY AND P. D. OLMSTED

TABLE I. Weissenberg numbers [Eqs. (16)–(18)] calculated during extru- TABLE II. Model parameters for a typical amorphous printing material,
sion for Zeq ¼ 37 and b ¼ 0:3 (similar to polycarbonate printing material), polycarbonate.
print temperature TN ¼ 250  C and two typical speeds UN ¼ 75 and 10 mm/
s. See Appendix C for further details. Polycarbonate properties Notation Value Units

Reptation Wi Fast Slow Rouse WiR Fast Slow Reference temperature T0 260 C
Thermal diffusivity [59] (at 25  C) a 0.14 mm2/s
R
Wi N (average) 13 2 Wi N (average) 0.07 0.009 Molecular weight Mw 60 kDa
WiW (wall) 91 24 WiRW (wall) 1.5 0.4 Entanglement molecular weight [60] Me 1.6 kDa
Plateau modulus [60] Ge 2:6  106 Pa
Entanglement time (at T0) [24] s0e 3:3  107 s
Assuming mass conservation, the speeds are related by WLF parameter C1 3 —
WLF parameter C2 160 —
pRH
pR2 UN ¼ UL ; (15) Equilibrium entanglement number Zeq 37 —
2 Equilibrium reptation time (at TN) Eq. (5) seq
d 0.03 s
Equilibrium rouse time (at TN) Eq. (4) seq
R 5:7  104 s
where R is the nozzle outlet radius. As well as local accelera-
tion due to the corner geometry, the flow must also speed up
to conserve mass while transforming from circular to ellipti- where the equilibrium Rouse time is given in Eq. (4). For the
cal geometry. If UN < HUL =2R, then too little material is fast printing case, the local Weissenberg number is WiRW  1
deposited and drawing occurs; similarly, if UN > HUL =2R near the nozzle wall, which implies stretch of the polymer
then buckling of the printed material occurs. tube during extrusion.
The equilibrium mass-averaged reptation Weissenberg In the following, we show results for Zeq ¼ 37, b ¼ 0:3
number in the nozzle is defined by TN ¼ 250  C and Wi N ¼ 13 and 2, which are typical values
used for FFF of an amorphous polymeric printing material
UN eq
Wi N ¼ s ðTN Þ; (16) [6]. For comparison, we have chosen model parameters for
R d Bisphenol A Polycarbonate. The full set of model parameters
for equilibrium reptation time seqd given in Eq. (5). Values
and the assumptions of this model are discussed in Appendix C;
for a typical print temperature and two typical print speeds the model parameters for polycarbonate are given in Table II,
(fast and slow) are given in Table I. Since Wi N  1, we and typical print speeds and nozzle dimensions (correspond-
expect a significant orientation of the polymer in the nozzle. ing to the simplified schematic in Fig. 1) are given in Tables
The local Weissenberg number at the nozzle wall III and IV, respectively.

WiW ¼ c_ W seq
d ðTN Þ; (17)
III. STEADY-STATE NOZZLE FLOW
where c_ W is the wall shear rate, is an order of magnitude
larger than Wi N . This is due to two effects. For cylindrical A. Calculation of nozzle flow
Poiseuille flow of a Newtonian fluid WiW ¼ 8Wi N because The fluid flows along a direction ^s with arc length coordi-
of the parabolic profile [36]. A shear-thinning fluid has a nate s. The flow direction ^s changes when the material exits
pluglike velocity profile [Fig. 2(b)] with a much higher rela- the nozzle according to
tive shear rate, and hence Weisseberg number, at the wall. 
Similarly, the equilibrium mass-averaged Rouse ^e z ; in the nozzle;
^s  (19)
Weissenberg number in the nozzle is defined as ^e y ; in deposited layer:

R UN eq First, we consider steady-state flow through the circular noz-


Wi N ¼ s ðTN Þ; (18)
R R zle in the vertical ^s -direction. The reasonable steady-state

TABLE III. Model parameters for two typical print speeds corresponding to a “fast” and “slow” case.

Printing parameters Notation Fast case Slow case Units

Mass flow rate Q 9:3  106 1:3  106 kg/s


Mean initial speed (heated nozzle section) U0 3 0.5 mm/s
Mean extrusion speed (final nozzle section) UN 75 10 mm/s
Mean print speed (across deposited layer) UL 100 13 mm/s
Thermal diffusion time (heated nozzle section) Eq. (C2) sa 7 7 s
Residence time (heated nozzle section) Eq. (C3) s0res 2 12 s

Nozzle temperature TN 250 250 C
Deposition time Eq. (28) sdep 0.005 0.03 s
Thermal skin layer in deposit Eq. (C4) Lskin 0.06 0.16 mm
Residence time (final nozzle section) Eq. (C5) sres 0.011 0.08 s
Die swell time ssw 0.0352 0.035 s
Terminal swell distance Eq. (30) zM 2.6 0.35 mm
POLYMER FLOWS DURING ADDITIVE MANUFACTURING 383

TABLE IV. Model parameters for typical nozzle geometry, as shown in typical print speeds corresponding to Wi N ¼ 13 and 2. The
Fig. 1. profiles have a pluglike shape due to shear-thinning behavior
Nozzle dimensions Notation Value Units
and are axisymmetric. The ellipses show how the polymer
chains become more stretched and oriented near the nozzle

Temperature TN 250 C walls due to the increasing shear rate.
Radius (heated nozzle section) R0 1.0 mm The polymer deformation for the two typical print speeds
Length (heated nozzle section) L0 6.0 mm is quantified in Figs. 3(a)–3(d), showing the entanglement
Radius (final nozzle section) R 0.2 mm
fraction , the tube stretch trA  3, the shear orientation Ars,
Length (final nozzle section) L 0.8 mm
and the normal stress difference N ¼ Ass  0:5ðArr þ A// Þ.
Layer thickness H 0.3 mm
For the Rolie–Poly model Arr ¼ Ahh in axisymmetric flow,
so that N is the first normal stress difference in the nozzle.
assumption is discussed in Appendix C. In cylindrical polar As expected, the larger Weissenberg number imposes a
coordinates ðr; /; sÞ, the velocity profile is denoted greater deformation on the polymer, with the chains becom-
ing more stretched and aligned with the flow direction for
u ¼ wðrÞ^s ; (20) the fast-print case. Due to this alignment, the entanglement
fraction decreases dramatically near the wall [Fig. 3(a)].
so that the velocity-gradient tensor is For Wi N ¼ 2,  is reduced to 20% of the equilibrium value
0 1 at the nozzle wall, whereas for Wi N ¼ 13 the melt is nearly
@w fully disentangled at the wall ð ¼ 5%Þ. These profiles pro-
B0 0
@r C vide an initial condition to calculate A during the deposition
K¼B
@0 0
C
0 A: (21)
process.
0 0 0 Figure 4 shows the contours of the described deforma-
tions across the nozzle for the case Wi N ¼ 2; the axisymme-
From Eq. (8), the total shear stress is given by try should be compared with later results (Sec. V) showing
nonaxisymmetric deformation after deposition. In particular,
@w Figs. 4(a) and 4(b) highlight the thin boundary layer of dis-
rrs ¼ Ge Ars þ ls ; (22)
@r entanglement and stretch at the nozzle wall due to the shear-
thinning nature of the flow. The tensor component Ars para-
and satisfies the steady-state momentum balance metrizes the principle shear deformation of the polymer
chain and so Fig. 4(c) demonstrates the shear stress across
@p 1 @
¼ ðrrrs Þ; (23) the nozzle. Since As/ ¼ Ar/ ¼ 0 in axisymmetric flow, the
@s r @r shear deformation is solely responsible for the polymer
for a pressure gradient @p=@s chosen to induce the mean orientation.
This polymer orientation is interpreted as an ellipse with
extrusion velocity
an orientation defined by a polar angle gh and azimuthal
ð angle g/ , i.e.,
w ðr Þ 2
UN ¼ d r: (24)
pR2
^e 1  ^s ¼ cos gh ; (27a)
Finally, the polymer deformation is described by the steady-
state Rolie–Poly equation

1
K  A þ A  KT  ðA  IÞ
sd ðT; c_ Þ
rffiffiffiffiffiffiffi ! rffiffiffiffiffiffiffi !
2 3 trA
 1 Aþb ðA  IÞ ¼ 0; (25)
sR ðT Þ trA 3

where trA ¼ Ass þ A// þ Arr . The reptation time is given by

1 1
¼ eq þ bðK : AÞ; (26)
sd ðT; c_ Þ sd ðT Þ

from Eq. (13) and the entanglement fraction is given in Eq.


(14) for c_ ¼ @w=@r.
FIG. 3. (a)–(d) Polymer deformation properties in the nozzle ðZeq ¼ 37;
B. Polymer deformation in the nozzle b ¼ 0:3Þ: (a) Entanglement fraction profile ðrÞ, (b) tube stretch profile
trAðrÞ  3, (c) shear deformation profile Ars ðrÞ, and (d) normal stress differ-
Figure 2 shows the steady-state velocity profiles calcu- ence profile N(r). Fast (Wi N ¼ 13) and slow ðWi N ¼ 2Þ print cases are
lated from Eqs. (14), (22)–(26) for Zeq ¼ 37, b ¼ 0:3 and two shown.
384 C. MCILROY AND P. D. OLMSTED

theoretically or numerically. Rather than solve the full


problem, we make the following assumptions:
(1) We assume that the temperature is uniform (at TN) dur-
ing deposition. The extrudate exits the nozzle and
reaches the build plate on the time scale

sdep ¼ H=UN : (28)

For our model parameters, the deposition time is typically


of order sdep ¼ 0:005  0:03 s (see Table III). Upon exit,
the material will cool via a combination of convection
and radiation. Thus, a nonuniform temperature profile
with a cool boundary layer near the free surface, where
the layer thickness depends on the print speed, is
expected (see Appendix C). This cooling will conse-
quently delay polymer relaxation due to the diverging
FIG. 4. (a)–(d) Polymer deformation properties in the nozzle for slow-print relaxation time [Eq. (3)]. We neglect the effect of this
case ðZeq ¼ 37; b ¼ 0:3 and Wi N ¼ 2Þ: (a) Entanglement fraction profile
ðr; /Þ, (b) tube stretch trAðr; /Þ  3, (c) principle shear deformation
cooling in our model and assume that the temperature of
Ars ðr; /Þ and (d) normal stress difference Nðr; /Þ shown in the xy-plane. the deposit is uniform. This assumption is roughly com-
The fast case ðWi N ¼ 13Þ induces similar deformation profiles. pensated by assuming that the polymer does not relax
during the deposition stage, as addressed next.
e 1  ^r ¼ sin gh cos g/ ;
^ (27b)
(2) We assume that the deposition occurs sufficiently fast that
we can ignore polymer relaxation. For polycarbonate of
where ^ e 1 is the principle eigenvector (corresponding to the
Zeq ¼ 37, this requires the deposition time to satisfy
largest eigenvalue k1) of the deformation tensor A. Figure 5
shows how the polar angle gh decreases near the nozzle wall,
demonstrating how the polymer becomes more extended and sdep seq
d ¼ 0:03s at TN ;
therefore better aligned with the flow direction in this region. (29)
sdep seq 4
R ¼ 5:7  10 s at TN :
Due to the axisymmetric Poiseulle flow, the azimuthal angle
g/ is zero everywhere and corresponds to ellipses that are Although we estimate sR < sdep ⱗsd (see Table III), a
tilted “inward” toward the center of the nozzle. cooling temperature profile as addressed in item 1 may
arrest relaxation in the skin layer responsible for weld-
ing in a similar way.
IV. STEADY-STATE DEPOSITION FLOW
(3) We assume that the length scale zM for which die swell
A. Assumptions of model deposition flow
develops is greater than layer thickness H; i.e., zM > H.
Due to the small Reynolds number (Re 106 ), the exit- The terminal swell distance downstream of the nozzle
ing flow quickly assumes a uniform plug-flow velocity pro- exit is [37]
file. The filament shape during deposition is a complicated
balance of surface tension, polymer relaxation, and com- zM ¼ ssw UN ; (30)
plex boundary conditions including the free surface.
Although it is known that the material must turn a 90 where ssw is the characteristic time scale for the swell
bend, the actual deposition shape, corresponding flow field, diameter to fully develop. This time scale is associated
and temperature profile have yet to be analyzed either with the relaxation of the first normal stress difference
[38], but little known about this relaxation mechanism.
Experimentally, zM is found to be of the order 2R and
some experiments show ssw to depend on the nozzle shear
rate [36]. For the Rolie–Poly model, the first normal stress
difference N1 ¼ Ass  Arr relaxes on the order of the
reptation time sd, although for WiR > 1 linear relaxation
does not apply. For polycarbonate at print temperature
TN ¼ 250 C, we find that zM =H
1  10 in the typical
print speed range (see Table III). Hence, there is probably
insufficient time for maximum swell ratio to develop dur-
FIG. 5. (a) Polar angle gh [Eq. (27a)] shown in the xy-plane and correspond- ing deposition of the melt onto the build plate. The termi-
ing ellipse located at the outer edge of the nozzle and, (b) gh shown in xyz-
nal swell ratio is discussed in Appendix C.
space with arrows indicating the local polar coordinate axis ðr; /; sÞ for the
ellipse, for Zeq ¼ 37; b ¼ 0:3 and Wi N ¼ 2. The azimuthal angle [Eq. (27b)] (4) We assume a smooth ansatz for the shape of the curved
g/ ¼ 0. filament and demand that the polymeric material
POLYMER FLOWS DURING ADDITIVE MANUFACTURING 385

undergos affine flow of the elements along streamlines. for velocity gradient tensor
The effect of changing the curvature of the corner region 0 1
is discussed in Appendix C. 0 0 0
K ¼ @ vx vy vz A; (35)
B. Deposition flow and polymer deformation wx wy wz

To parametrize the shape and flow field of the curved fila- where the subscripts denote derivatives in the respective
ment, we deform a cylinder around a 90 corner into an directions. In this case, ux ; uy ; uz ¼ 0 since there is no change
elliptic-cylinder, so that the outer edge of the deposition in length in the x^-direction. In this way, the polymer is sim-
traces an ellipse as shown in Fig. 6. Full details of the mesh ply advected with the velocity gradients. Similarly, entangle-
generation are given in Appendix B. The angle h between ments are advected via
the nozzle and the fully deposited layer is in the range
h 2 ½0; p=2 . In the lab frame, the Cartesian velocity profile ðu  rÞ ¼ bðK : AÞ; (36)
u ¼ ð0; v; wÞ of the deformation flow is given by
u ¼ UðhÞ^s ðhÞ; (31) from Eq. (12).
Equations (34) and (36) are solved using a semi-implicit
where finite-difference scheme combined with the velocity profile
from Eq. (33) and the initial polymer tensor A imposed by
^s ¼ sin h^e y þ cos h^e z (32) the nozzle flow. Full details of the calculation are given in
is the flow direction and U is the magnitude of the velocity. Appendix B. For convergence, we require 100 cross sections
with 200  100 mesh points on each plane. This corresponds
The velocity vector field is shown in Fig. 7(a).
to a mesh element at the outer edge of the printed layer hav-
Under the assumption sdep sd ; sR and assuming no sec-
ing volume dx  dy  dz ¼ 6  3  2 lm for a nozzle of
ondary flows, instead of solving the full Navier-Stokes equa-
radius R ¼ 0.2 mm.
tions the velocity profile is calculated from the flux-
Figure 7(b) shows an ellipsoidal visualization of the poly-
conservation condition
mer tensor A during the deposition process for Zeq ¼ 37; b
UðhÞdAðhÞ ¼ UN dAð0Þ; (33) ¼ 0:3 and Wi N ¼ 2. Initially, the polymer ellipsoids are
directed inward toward the nozzle center [as in Fig. 2(b)].
where dA denotes the area of a cross section of the deposit The orientation changes with the flow direction, so that ulti-
at angle h. This is equivalent to imposing local flux conser- mately the polymers are aligned roughly parallel to the
vation on a single mesh element during deposition [see Figs. printed filament layer. Similar orientations of cellulose fibrils
6(b) and 6(c)] and is discussed in detail in Appendix B. are seen in experiments [39]. Figure 7(b) also shows how the
Equation (33) dictates an increase in U to conserve mass ellipses become more stretched along the outer edge of the
during the typical geometric transformation from a circle to deposition due to the increased displacement in this region.
an ellipse [Eq. (15)]. There is a larger displacement ds Next, we consider the final polymer deformation across the
toward the outer edge of the deposit to accommodate the 90 printed cross section (i.e., for h ¼ p=2).
corner [Fig. 7(a)]; this displacement is given by the arc
length ds ¼ r1 dh, where r1 is the radius measured from the
inner corner ð0; R; HÞ and dh is angle between two cross sec- V. RESULTS
tions (see Appendix B for further details). A. Polymer deformation across the printed
Under this assumption (sdep sd ; sR ), the steady-state filament cross section
Rolie–Poly Eq. (11) is reduced to
Figures 8(a)–8(d) show the entanglement fraction , the
ðu  rÞA ¼ K  A þ A  KT ; (34) tube stretch trA  3, the principle shear deformation Ars, and

FIG. 6. (a) Shape of the deposition parametrized by deforming a cylinder into an elliptic-cylinder such that the outside edge of the deposition traces an ellipse.
z -direction and final elliptic cross section (h ¼ p=2) has a velocity UL in the y^-direction. (b) Nozzle
Initial circular cross section (h ¼ 0) has velocity UN in the ^
view ðh ¼ 0Þ in the xy-plane, (c) side view in the zy-plane and (d) layer view ðh ¼ p=2Þ in the xz-plane; r denotes the radial position on a plane and / denotes
the angle around a plane. The points indicate individual mesh points and the shaded area represents the area of a mesh element.
386 C. MCILROY AND P. D. OLMSTED

FIG. 7. (a) Velocity vector field u with shading to indicate the displacement ds (lm). (b) Visualization of polymer chains as deformed spheres during deposi-
tion process. Color indicates three locations; top of layer (green), middle of layer (red) bottom of layer (blue). Views in (c) the xy-plane ðh ¼ 0Þ looking down
the nozzle, (d) the yz-plane, showing a side view of the full deposition and (d) the xz-plane ðh ¼ p=2Þ looking through the printed layer. Model parameters at
Zeq ¼ 37; b ¼ 0:3, and Wi N ¼ 2.

the normal stress difference N profiles, respectively, across Due to the anisotropy, the nonaxisymmetric components
the printed layer for Zeq ¼ 37; b ¼ 0:3, and Wi N ¼ 2; the As/ and Ar/ , corresponding to in-plane tilt and azimuthal
deformation profiles are qualitatively similar for Wi N ¼ 13. shear, respectively, become nonzero and contribute to the
In contrast to the nozzle flow (Fig. 3), the deformation is no total orientation of the polymer. We quantify the effect of
longer axisymmetric and there is a distinct gradient in the these nonaxisymmetric components on the polymer orienta-
polymer microstructure from the top to the bottom of the tion by considering the polar and azimuthal angles, gh and
layer. g/ [Eq. (27)], after deposition. Figure 9(a) shows that the
The structure at the top ðz ¼ HÞ and bottom (z ¼ 0) of polar angle gh decreases (compared to Fig. 5), demonstrating
the layer is of particular interest as welding between how the polymers become more aligned with the flow direc-
adjacent layers in the ^z -direction occurs at these sites. tion during deposition. The effect of the velocity gradients in
The stretch of the free surface due to the curved geometry the center of melt is also demonstrated by this decrease in
induces a large deformation along the outer edge of the gh . The corresponding alignment leads to disentanglement at
deposition, so that the polymer microstructure at z ¼ 0 is r ¼ 0, so unlike flow in the nozzle, the entire cross section of
highly stretched and oriented [Figs. 8(b) and 8(c), see the melt becomes disentangled during deposition. The non-
Appendix C for the effect of changing the curvature]. By zero azimuthal angle g/ [Fig. 9(b)] signifies how the tilt of
far, the largest effects occur during this deposition process, the ellipses becomes nonaxisymmetric after deposition, with
with the stretch increasing significantly (by a factor of 3 ellipses directed “upward” away from build plate across the
under the assumption sdep sR ) in the bottom half of the entire layer (Fig. 7). Ellipses at the top and bottom of the
layer. layer have a similar upward tilt, exhibiting a very different
Alignment of the polymers in the flow direction, together orientation to the inward axisymmetric tilt we see in the
with the velocity gradient profile, disentangles the polymer nozzle.
melt and  is reduced to less than 10% of the equilibrium
entanglement fraction at z ¼ 0 [Fig. 8(a)]. Although the B. Effect of CCR parameter on disentanglement
stretch at z ¼ H is comparatively smaller, the melt also For comparison with Figs. 4 and 8 with b ¼ 0:3, Fig. 10
becomes disentangled in this region (compared to  in the shows the deformation imposed by the nozzle flow and dur-
nozzle before deposition) due to large velocity gradients. ing deposition for CCR parameter b ¼ 1. Since the melt is
Velocity gradients exist primarily due to the material turning less shear-thinning in the case b ¼ 1, there is a larger
90 ; there is also a secondary contribution due to the trans- boundary layer of disentanglement in the nozzle (compared
formation of the deposit from a circular to an elliptical to Fig. 4) and the disentanglement induced by the deposi-
shape. tion process is much more extreme (compared to Fig. 8).
POLYMER FLOWS DURING ADDITIVE MANUFACTURING 387

FIG. 9. (a) Polar angle gh [Eq. (27a)] and (b) azimuthal angle g/ [Eq. (27b)]
shown in the xz-plane; (c) schematic of corresponding ellipses shaded
according to y-coordinate (red is the front of ellipse) to illustrate angles gh
(where k1 is the principle eigenvalue) and g/ (azimuthal rotation from ^ r-
axis). Arrows indicate the local polar coordinate axis ðr; /; sÞ.

For WiW > 1,  N is reduced to less than 20% of the equi-


librium entanglement fraction and reducing the CCR param-
eter slightly inhibits disentanglement at the nozzle wall [Fig.
FIG. 8. (a)–(d) Polymer deformation properties after deposition (final plane
h ¼ p=2) for Zeq ¼ 37; b ¼ 0:3 and Wi N ¼ 2: (a) Entanglement fraction pro-
11(b)]. We find nearly 100% disentanglement at WiW ¼ 100.
file ðr; /Þ, (b) tube stretch trAðr; /Þ  3, (c) principle shear deformation For comparison, a 50% entanglement loss is found for
Ars ðr; /Þ, and (d) local normal stress difference Nðr; /Þ shown in the xz- Wi ¼ 100 in the molecular simulations of simple shear flow
plane. The fast case induces similar deformation profiles. (e) Quantitative by Baig et al. [32]. This is well represented by the theory of
comparison of the stretch (red line) and disentanglement (blue line) along
the z-axis (ðx; yÞ ¼ ð0; RÞ) induced initially in the nozzle (zmax ¼ 2R) and
Ianniruberto and Marrucci for Zeq ¼ 14; b ¼ 0:15 [30],
after deposition across the printed layer (zmax ¼ H), as a function of distance which employs the Doi–Edwards tensor rather than the
from weld site zw. Rolie–Poly model. For our model parameters, b ¼ 0:15
gives a nonmonotonic constitutive curve (see Appendix A).
The deformation of A imposed during deposition is equiva- Figure 11(c) shows the entanglement fraction  L at weld
lent for b ¼ 1 and 0.3 due to the assumption sdep sd ; sR site z ¼ 0 after deposition. Again this process is independent
[Eq. (34)]. of Zeq. For the case b ¼ 1, the idealized deposition process
imposed by the model removes almost all entanglements
C. Effect of shear rate on disentanglement from the melt at the weld site. For smaller values of the CCR
parameter b, the disentanglement process is significantly less
Here, we consider how the predicted disentanglement
severe for moderate Weissenberg numbers, and the melt
varies with the equilibrium entanglement number Zeq (equiv-
only becomes fully disentangled for WiW 100.
alent to changing Mw) and the local Weissenberg number
By spatially advecting the entanglements through the
calculated at the nozzle wall WiW [Eq. (17)].
deposition and assuming sdep sd , Eq. (36) leads to
Figure 11(a) shows the entanglement fraction at the noz-
zle wall (prior to deposition),  N, for b ¼ 1 and three molecu-
@
lar weights. From Eq. (14),  N is given by U  bðK : AÞ; (38)
@s
1
N ¼ ; (37) for flow direction ^s . Since the deformation is dominated by
1 þ bArs WiW
the extension induced by stretching the fluid elements around
and agrees quantitatively with the calculated degree of disen- the corner, we assume
tanglement at the nozzle wall for Ars ¼ 0.5, although Ars is
not independent of WiW. The disentanglement fraction does @U
K:A
Ass ; (39)
not depend on Zeq. @s
388 C. MCILROY AND P. D. OLMSTED

FIG. 10. Polymer deformation properties in the nozzle and after deposi- FIG. 11. Degree of disentanglement  for a range of Weissenberg numbers
tion for Zeq ¼ 37, b ¼ 1 and Wi N ¼ 2: (a) and (e) Entanglement fraction WiW [Eq. (17)]: (a)  N at the nozzle wall before deposition for b ¼ 1 and
profile ðr; /Þ, (b) and (f) tube stretch trAðr; /Þ  3, (c) and (g) princi- equilibrium entanglement numbers Zeq ¼ 22 (*), 27 (x) and 37 (þ), (b)  N
ple shear deformation Ars ðr; /Þ, and (d) and (h) local normal stress dif- before deposition for a range of CCR parameters b and (c)  L at the weld
ference Nðr; /Þ; (i) Quantitative comparison of the stretch (red line) site z ¼ 0 after deposition for b ¼ 1; 0:6, and 0.3. Theory lines given in Eqs.
and disentanglement (blue line) along the z-axis (ðx; yÞ ¼ ð0; RÞ) (37) and (41). The black square marks the slow case Wi N ¼ 2 and the black
induced initially in the nozzle (zmax ¼ 2R) and after deposition circle marks the fast case Wi N ¼ 13.
across the printed layer (zmax ¼ H), as a function of distance from weld
site zw.
 bAss
which gives UL
L  N ; (41)
UN
1 @ bAss @U
 : (40) where  N is given in Eq. (37). Thus, disentanglement
 @s U @s
depends on the geometry, which determines the ratio UL =UN
Integrating yields [Eq. (15)], the total stretch imposed and the CCR parameter
POLYMER FLOWS DURING ADDITIVE MANUFACTURING 389

b. Equation (41) fits the data well for b ¼ 1:0; 0:6, and 0.3, bottom half of the layer. During deposition, the polymer ten-
and Ass ¼ 9 [Fig. 11(c)], although Ass is not independent of sor A is deformed further from equilibrium, nonaxisymmet-
WiW. ric configurations become nonzero and the structure of the
weld region is highly stretched, oriented, and partially disen-
tangled. The degree of disentanglement at the weld site
VI. DISCUSSION (z ¼ 0) depends on both the shear rate in the nozzle and the
A. Model summary and limitations CCR parameter b. Faster printing imposes a greater deforma-
tion and disentangles the melt further during the extrusion
We have developed a model of the fused-filament-fabrica-
process. For WiW > 100, the weld site becomes fully disen-
tion process and tested the effect of changing print speed,
tangled for all b.
entanglement number Zeq, and CCR parameter b on the
Understanding the polymer behavior during extrusion in
degree of polymer deformation and disentanglement during
terms of the material properties, print speed, and nozzle
extrusion. We have used Bisphenol A Polycarbonate as an
geometry is a key to characterizing the strength of the weld
example of a typical amorphous polymer used for FFF. We
between printed filaments. After deposition, the printed melt
model the nozzle flow as axisymmetric, steady-state pipe
will rapidly cool toward the glass transition. The way in
flow. The nozzle flow can stretch and orient the polymer
which the deformation relaxes as a function of temperature
near the nozzle wall, which consequently disentangles the
governs the diffusive behavior at the weld and is therefore
melt via CCR.
the key to understanding the ultimate welding characteristics
Since the material must melt before being deposited, prac-
such as weld thickness, structure, and entanglement. The
tically the upper speed limit for printing is restricted by ther-
effect of this polymer deformation on welding behavior will
mal diffusion in the nozzle. In the model, we assume a
be discussed elsewhere.
uniform temperature profile across the nozzle radius. For
polycarbonate, it is estimated to take 7 s to achieve TN
B. Outlook
across the nozzle radius via thermal diffusivity (see
Appendix C). By comparing to the residence time in the The molecular CCR mechanism is a key to understanding
heated nozzle section, this leads to an upper flow rate limit polymer behavior in highly nonlinear flows such as the FFF
of 3  106 kg/s for our model, although faster rates are technique for additive manufacturing. CCR was first added
often used (e.g., 9  106 k/s for UL ¼ 100 mm/s). Arguably to the original Doi–Edwards tube model by Ianniruberto and
only the outer side of the filament must be melted to ensure Marrucci [29,41]. The recent GLaMM model [42] refines the
welding. Moreover, fluorescence-based measurements dur- tube theory further to include the effects of CCR on the chain
ing polymer extrusion report temperature gradients of up to stretch. CCR in the tube model has now been revisited to
5  C/mm between the center of the nozzle and the wall due account for flow-induced changes in the entanglement den-
to shear heating effects [40]. A more detailed model is sity [25,30].
required to capture the effects of an inhomogeneous temper- In this paper, we have modified the Rolie–Poly model
ature profile in the nozzle. [23] to incorporate Ianniruberto’s flow-induced disentangle-
After exiting the nozzle, the extrudate deforms to make a ment theory and capture inhomogeneous disentanglement at
90 turn and is deposited into a elliptical-shaped layer. the continuum level of the orientation tensor. Although the
Rather than calculate the full fluid mechanics, we have cal- Rolie–Poly model handles stretch in a slightly different way
culated the steady-state flow by assuming flux conservation to the molecular GLaMM model, the advantage of the
and a uniform temperature profile. However, we estimate Rolie–Poly model is the simple one-mode constitutive equa-
that a cool boundary layer with thickness 0:1 mm will tion that can be applied to arbitrary inhomogeneous flows.
develop during deposition (see Appendix C); a more detailed We have shown that disentanglement during FFF is sensitive
model is required to capture these complex cooling dynamics to the chosen CCR parameter b, particularly during the depo-
upon exiting the nozzle. We also neglect polymer relaxation sition stage due to the large stretch induced by the corner-
during deposition. The assumption that reptation is slow flow geometry.
compared to the deposition time (sd > sdep ), yields the valid- Flow-induced disentanglement is observed in numerous
ity condition simulations, including Brownian simulations [43], molecu-
lar dynamics simulations [32], and dissipative particle
H dynamics simulations [34]. To date b serves as a fitting
Wi N > ; (42)
R parameter between simulations and the tube theory, and
different values are required depending on the flow type
which for our model parameters leads to Wi N > 1:5. For and entanglement number [25]. Thus, testing FFF-induced
smaller Weissenberg numbers where polymer relaxation disentanglement using a range of constitutive models is a
must be considered, the flow may not be in steady state dur- must.
ing deposition. Various other CCR theories have arisen based on molecu-
Due to the corner flow geometry and the transition from a lar simulations. For example, Wang and Larson [44] incor-
circular to an elliptical shape, the polymer deformation is porate kink-jump motions of the tube segments to capture
affected primarily by the deposition flow rather than the noz- the constraint release effect and find a broad distribution of
zle flow, with polymer stretch becoming significant in the constraint lifetimes. In contrast to tube models, slip-link
390 C. MCILROY AND P. D. OLMSTED

models construct an effective field to represent entangle-


ments so that the chain satisfies random-walk statistics at all
length scales and constraint release is determined by a slip-
link friction [45,46]. However, with little experimental evi-
dence, flow-induced disentanglement continues to be a
debated topic.
Finally, the model presented here is restricted to predict-
ing the behavior of linear amorphous polymer melts. FFF
systems can handle a wide range of rheologically complex
materials, including semicrystalline melts [5] and filled
melts, containing nanoscale particles, such as ABS [3]. Since
semicrystalline polymers tend to flow more readily com-
pared to amorphous melts above the glass transition tempera-
ture, FFF systems incorporate fans that rapidly cool the
extruded material [47]. Thus, it is expected that semicrystal-
line printed parts will exhibit a greater degree of anisotropy
than parts made from amorphous materials.
Although the most-commonly printed polymer is ABS, an
amorphous melt containing rubber nanoparticles that provide
toughness even at low temperatures, this material has been
rarely characterized rheologically [48]. The addition of fibers
to an amorphous melt can also enhance both thermal and
mechanical properties. Yet how these fillers behave during
the printing process and how they modify viscoelasticity
remain open questions.

ACKNOWLEDGMENTS
FIG. 12. Log-linear plot of constitutive curve given in Eq. (8) and the corre-
The authors thank Jonathan Seppala and Kalman Migler sponding disentanglement fraction  for Zeq ¼ 37 and increasing
for advice and an enjoyable collaboration, as well as the Weissenberg number Wi; (a) effect of allowing disentanglement in Eq. (8)
via Eq. (12) and (b) effect of changing CCR parameter b.
National Institute for Standards and Technology (NIST),
Georgetown University, and the Ives Foundation for funding.
APPENDIX B: FULL DEPOSITION CALCULATION

APPENDIX A: THE CONSTITUTIVE CURVE 1. Parametrization


The steady-state constitutive curve defined in Eqs. (8), To parametrize the shape of the curved filament, we deform
(11), and (12) is plotted in Fig. 12(a) for a Zeq ¼ 37 melt that a cylinder into an elliptic-cylinder by rotating successive planes,
remains fully entangled ð ¼ 1Þ and that is allowed to disen- as shown in Fig. 6(a). We divide the space into a three-
tangle ð < 1Þ, where the value of  varies with shear rate. dimensional mesh, where each plane is specified in terms of the
The CCR parameter is set to b ¼ 0 and 1. The curve indicates spherical coordinate system ðr; h; /Þ, where r defines the radial
the shear-thinning nature of the Rolie–Poly model. Feed position from the center of the plane and / is the azimuthal
stocks for FFF processes are typically shear thinning and are angle around the plane. The angle between the nozzle and the
often assumed to follow a power-law viscosity model [15, fully deposited layer is denoted by h and is in the range
49–51]. These treatments are not molecularly aware and can- h 2 ½0; p=2 . The initial plane (located at the nozzle exit) is a
not capture normal stress effects of complex flow fields. The circle of radius R centered at ð0; 0; HÞ with h ¼ 0 [Fig. 6(b)].
Rolie–Poly model, on the other hand, includes key aspects of The final plane at h ¼ p=2 is an ellipse centered at ð0; R;
the molecular melt structure. H=2Þ with major radius R and minor radius H=2 [Fig. 6(d)].
For b ¼ 0, there is no CCR so that  can only equal unity The mesh points R ¼ ðx; y; zÞ are expressed as Cartesian
in a steady flow and the constitutive curve is nonmonotonic. functions of the spherical coordinate system ðr; h; /Þ. That
For b ¼ 1, disentanglement can occur and the entanglement is, in the frame moving with the nozzle,
fraction  becomes less than unity for sufficiently large shear
rates. This disentanglement mechanism acts to suppress Rðr; h; /Þ ¼ xðr; h; /Þ^e x þ yðr; h; /Þ^e y þ zðr; h; /Þ^e z : (B1)
excess shear-thinning behavior in a similar way to increasing
the CCR parameter b [29], as demonstrated in Fig. 12(b). Coordinates for each plane are calculated by applying a
Smaller b flattens the constitutive curve, enhancing shear- deformation to the initial plane, parametrized by initial polar
thinning behavior by reducing the rate of CCR. For coordinates ðr 0 ; /0 Þ, rotated by angle h about the stagnation
b ¼ 0:15, the constitutive curve becomes nonmonotonic for point at the nozzle exit. That is,
Zeq ¼ 37, in which case shear-banding instabilities would be
expected in the nozzle [52]. Rðr; h; /Þ ¼ T ðhÞ  Rðr 0 ; 0; /0 Þ; (B2)
POLYMER FLOWS DURING ADDITIVE MANUFACTURING 391

where Axy ¼ 0; (B9d)


T ðhÞ  KðhÞ  Mð^
x ; hÞ: (B3) Axz ¼ cos /0 Ars ; (B9e)
That is, a rotation about the x^-axis for angle h defined by
Ayz ¼ sin /0 Ars ; (B9f)
0 1
1 0 0
x ; hÞ ¼ @ 0
Mð^ cos h sin h A; (B4) since there is zero second normal stress (Arr ¼ Ahh ) in the
0 sin h cos h Rolie–Poly model under axisymmetric flow.

and a deformation factor given by 3. Mesh spacing

0 1 The initial geometry (polar angle h ¼ 0) is given by a cir-


kx ðhÞ 0 0 cular plane defined in Cartesian coordinates by
KðhÞ ¼ @ 0 ky ðhÞ 0 A: (B5)
0 0 kz ðhÞ; Rðr 0 ; 0; /0 Þ ¼ r 0 cos /0 ^e x þ r0 sin /0 ^e y ; (B10)

for initial polar coordinates ðr0 ; /0 Þ. This initial plane is


This is a general mapping for any function K. In our case,
divided up into a numerical mesh [Fig. 6(b)] with mesh spac-
the deformation is defined by
ing given by
kx ¼ ky ¼ 1; (B6a) 0
^ ;
ds ¼ dr0 ^r 0 þ r0 d/0 / (B11)
H
kz ¼ ; (B6b) where
2R
r0 ¼ jR  Ro j; (B12)
so that the outer corner of the deposition shape traces an
ellipse. Equation (B6) defines the typical geometry transfor- for plane center Ro ¼ 0. For this initial circular geometry,
mation imposed by the FFF process; since the nozzle head is the radial and azimuthal spacing, dr0 and d/0 , are uniform
placed at height H above the build plate (or previously and given by
printed layer), which is less than the nozzle diameter 2R,
printed layers are elliptically shaped and mass is conserved 2R 2p
dr0 ¼ and d/0 ¼ ; (B13)
by balancing UN and UL [Eq. (15)]. Mmax Pmax
where Mmax and Pmax are the number of radial and azimuthal
2. Initial condition
mesh points, respectively.
The initial velocity is assumed to be uniform at the nozzle Due to the nature of the mapping, it is natural to continue
exit and is given by u ¼ ð0; 0; UN Þ. The initial polymer con- adopting the parametrization ðr; h; /Þ for subsequent planes.
figuration A induced by flow through the nozzle is calculated However, the initial polar coordinates ðr0 ; /0 Þ defined for
in Sec. III and in cylindrical polar coordinates ðr0 ; /0 ; sÞ. the circular plane are not equivalent for subsequent planes.
This is then converted to the Cartesian frame (x, y, z) to cal- In general, the coordinates
culate the deposition flow. In the following i; j; k; … label
Cartesian and a; b; c; … label spherical polar coordinates. r  rðr 0 ; h; /0 Þ and /  /ðr 0 ; h; /0 Þ; (B14)
The polymer tensor Aab is converted to Cartesian coordi-
nates (x, y, z) via the rotation depend on the shape of the plane at h, which is determined
by deformation that Eq. (B2) imposes onto a circular plane.
Aij ¼ Xia Aab Xbj ; (B7) Note that r and / are not explicitly required to calculate R,
but act as counters to locate adjacent mesh points within a
where the rotation matrix is given by plane.
In the case discussed in this paper, the deformation
0 1
cos /0 sin /0 0 imposes elliptical geometry. Thus, for h > 0,
B C
Xia ¼ B
@ sin /
0
cos /0 0CA ; (B8)
r 6¼ r 0 and / 6¼ /0 ; (B15)
0 0 1
ia
and the mesh spacing is not uniform across and around each
so that the initial polymer configuration for the deposition plane
calculation is given by
dr 6¼ dr 0 and d/ 6¼ d/0 : (B16)
Axx ¼ Arr ; (B9a)
Since the mesh spacing must now reflect the spacing
Ayy ¼ Arr ; (B9b) between mesh points that are mapped using Eq. (B2), we
must make the following distinctions.
Azz ¼ Ass ; (B9c) First, the mesh spacing is given by
392 C. MCILROY AND P. D. OLMSTED

ds6 ¼ dr ^r þ r1 dh^h þ r2 d/ /;
^ (B17) where

where 6 signifies the forward and backward directions for Rs  ð0; R; 0Þ and Ro  T ðhÞ  Rð0; 0; 0Þ; (B19)
each coordinate. Second, the arc lengths in Eq. (B17) are cal-
culated from the average of two successive arcs by defining are the stagnation point and center of each plane. Finally, the
radial spacing dr is given by
1
r16  ðjR  Rs j þ jRðr; h6dh; /Þ  Rs jÞ; (B18a)
2 dr 6 ¼ jR  Rðr6dr; h; /Þj; (B20)
1
r26  ðjR  Ro j þ jRðr; h; /6d/Þ  Ro jÞ; (B18b) the azimuthal angle d/ is calculated via the law of cosines
2

jR  Ro j2 þ jRðr; h; /6d/Þ  Ro j2  jR  Rðr; h; /6d/Þj2


cos d/6 ¼ : (B21)
2jR  Ro jjRðr; h; /6d/Þ  Ro j

uðr; h; /Þ ¼ Uðr; h; /Þ ^s ðhÞ;


and the polar angle is chosen to vary uniformly according to (B23)
¼ vðr; h; /Þ^e y þ wðr; h; /Þ^e z ;
p
dh ¼ ; (B22)
2Nmax where
where Nmax is the total number of planes. The mesh spacing pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
U¼ v2 þ w2 and ^s ¼ sin h^e y þ cos h^e z (B24)
defined in Eq. (B17) is shown in Fig. 13.
are the magnitude of the velocity and the local unit normal
4. Flow field vector (normal to the plane), respectively. On exiting the
The velocity profile is also written as Cartesian functions nozzle, h ¼ 0 and u ¼ ð0; 0; UN Þ, whereas at the end of depo-
of the spherical coordinate system ðr; h; /Þ, such that sition h ¼ p=2 and the velocity profile of the layer is
u ¼ ð0; UL ; 0Þ.
Instead of solving the full Navier-Stokes equations, we
assume sdep sd ; sR . Thus, assuming no secondary flows,
the velocity profile is calculated from the local flux-
conservation condition

Uðr; h; /Þ daðr; h; /Þ ¼ UN daðr0 ; 0; /0 Þ; (B25)

where da denotes the area of a single mesh element imposed


by the prescribed shape (B2), as shown in Figs. 6(b) and
6(c). Thus, the horizontal and vertical velocity components
are given by
 
da r0 ; 0; /0
vðr; h; /Þ ¼ UN sin h; (B26a)
daðr; h; /Þ
 
da r 0 ; 0; /0
wðr; h; /Þ ¼ UN cos h; (B26b)
daðr; h; /Þ

respectively.

5. Polymer deformation
Since the flow starts in steady state and polymer relaxa-
FIG. 13. Schematic of numerical method showing the initial circular plane tion is ignored, the deposition flow remains in steady state.
and two successive planes that are elliptic in shape. Cartesian mesh points
are parametrized by spherical polar coordinates ðr; h; /Þ. Mesh spacing ds Thus, to advect the polymer with velocity gradients during
a
is given in Eq. (B17). Note that the radial spacing dr and the azimuthal angle deposition we solve
spacing d/ are not uniform around and across each plane for h > 0 due to
the elliptical geometry. See text for details. ðu  rÞA ¼ K  A þ A  KT ; (B27)
POLYMER FLOWS DURING ADDITIVE MANUFACTURING 393

in the Cartesian frame. In Einstein notation, Eq. (B27) is 1 @Ajk Ajk ðr; h þ dh; /Þ  Ajk
C2 ¼ ; (B37)
written as r1þ @h Dt

ðui @i ÞAjk ¼ Kjl Alk þ Ajl Kkl ; where


¼ @l uj Alk þ Ajl @l uk ; (B28) r1þ dh
Dt ¼ (B38)
C2
where derivatives in the Cartesian frame are denoted is the advection time scale that captures a greater displacement
  ds ¼ r1þ dh at the outside edge of the deposition [Fig. 7(a)].
@ @ @
@i ¼ ; ; : (B29) Then, the semi-implicit finite-difference scheme is
@x @y @z i defined by the generalized matrix system
Since the Cartesian mesh points are defined as functions of ½Ajk  DtðM1 1
al @a uj Alk þ Ajl Mal @a uk Þ ðr;hþdh;/Þ
the spherical polar coordinate system ðr; h; /Þ, derivatives in 
the ^r ; ^ ^ directions, that is
h, and / ¼ Ajk þ DtðM1 1
al @a uj Alk þ Ajl Mal @a uk Þ
  X
@ 1 @ 1 @  Dt Ca @a Ajk : (B39)
@a ¼ ; ; ; (B30)
@r r1 @h r2 @/ a a¼1;3 ðr;h;/Þ

are easily computed from the mesh, as shown in Fig. 13. to be solved for Ajk ðr; /Þ on the plane at angle h þ dh using
Due to the nonuniform nature of the mesh, velocity gra- information from the previous plane at angle h. Note that
dients in the ^ ^ and ^h directions are given by the average
r ; /, Mal and uj are known for all planes a priori based on the
of a forward and backward first-order finite-difference transformation given in Eq. (B2) and the flux-conservation
approximation. For example, velocity gradients in the ^r - condition Eq. (B26), respectively. The final results are given
direction are given by in the flow coordinate system via
  Aab ¼ Mai Aij Mjb : (B40)
1 ui ðr þ dr þ ; h; /Þ  ui
@r ui ¼
2 drþ
  APPENDIX C: APPROXIMATIONS AND
1 ui  ui ðr  dr  ; h; /Þ ASSUMPTIONS IN THE MODEL
þ ; (B31)
2 dr 
Here, we discuss further details of the validity of the FFF
and similarly for the ^h and /-directions.
^ One-sided finite-dif- model. The model parameters for polycarbonate are given in
ference approximations are used at the boundaries. Table II, and typical print speeds and nozzle dimensions
Derivatives are converted to the Cartesian frame via (corresponding to the simplified schematic in Fig. 1) are
given in Tables III and IV, respectively. The assumptions
@a ¼ Mai @i : (B32) made are as follows:
(1) We ignore viscous heating and assume that there is zero
where temperature gradient across the nozzle radius. A thick fil-
0 1 ament of solid polycarbonate is fed into the FFF nozzle
cos / cos h sin / sin h sin / at a mass flow rate Q. As the solid filament enters the
Mai ¼ @ 0 sin h cos h A : (B33) nozzle, a heated element increases the temperature of the
sin / cos h cos / sin h cos / ai material to TN so that it becomes molten and flows at
average speed U0 such that
Thus, Eq. (B28) becomes
Q ¼ qpR20 U0 ; (C1)
ui M1
ai @a Ajk ¼ M1
al @a uj Alk þ Ajl M1
al @a uk ; (B34)
for density q and nozzle radius R0. The time scale for
heat diffusion is given by
and tensor Ajk is advected in the ^h-direction through angle h
via
sa ¼ R20 =a; (C2)
X
C2 @2 Ajk ¼ ðM1
aj @a ul Alk þ Ajl M1
ak @a ul Þ  Ca @a Ajk ;
for thermal diffusivity a. If the residence time in the
a¼1;3
heated section of the nozzle satisfies
(B35)
s0res ¼ L0 =U0  sa
7s; (C3)
where
for length L0, then the temperature is expected to be uni-
Ca ¼ ui M1
ai : (B36) form across nozzle radius. Finite-element analyses of ther-
mal diffusion in the heated nozzle section (which ignore
We make the forward finite-difference approximation viscous heating effects) show that uniform temperature
394 C. MCILROY AND P. D. OLMSTED

profiles ðDT 1  CÞ are rapidly achieved [49–51]. For swell ratio DM =2R, where DM denotes the maximum
our model parameters s0res  sa for typical print speeds steady-state diameter of the melt after exiting the nozzle,
(Table III), so effects of a nonuniform temperature profile is estimated from the first normal stress difference and
within the nozzle may need to be considered. the shear stress at the wall [53]. Using the values calcu-
(2) We assume that the temperature throughout the deposition is lated in Sec. III, the die-swell ratio is found to be [53]
uniform. A simple calculation from thermal diffusivity sug-
gests a cool boundary layer near the free surface of thickness   !1=6
DM 1 Ass  Arr
¼ 1þ þ 0:13
1:2; (C6)
pffiffiffiffiffiffiffiffiffiffi 2R 2 2Ars
Lskin ¼ sdep a; (C4) W

which depends on the print speed through the deposition for Wi N ¼ 2 and 13. Reported values for FFF-like pro-
time; for a typically fast print speed Lskin ¼ 0.06 mm cesses range from  1.05 to 1.3 [6], and this swelling phe-
(Table III). We choose to neglect the complicated heat nomenon is found affecting the alignment of extruded
transfer process here. fiber suspensions [54]. Ceramic particles [55] and carbon
(3) We assume that the flow is steady state in the nozzle and fibers [56] may be used to reduce the swelling effect.
during deposition. The molten material exits the final (5) We prescribe the shape of the deposition and neglect any
nozzle section at average speed UN (usually after passing spreading of the deposition on the build plate. Models
through two contractions). If the residence time in the that address the spreading of a printed layer [57,58] have
final nozzle section satisfies yet to be applied to polymer melts. The deposition shape
described in Sec. IV assumes a smooth corner region
sres ¼ L=UN  seq
d ; (C5) [Fig. 6(a)]. However, the prescribed shape can affect the
deformation imposed by the deposition process. As a
for length L, then the flow can be assumed to be steady. comparison, we have calculated the effect of having a
For polycarbonate rheology, sres  seq
d for typical print-
sharp corner, whose deformation is defined by
ing speeds (Table III), thus a more detailed calculation
of the flow may be required to capture start-up effects in kx ¼ 1; (C7a)
the nozzle. 8
1
>
>
(4) We assume that the time scale for die swell to fully < ; for h < h ;
develop is larger than the deposition time scale. Upon ky ¼ kz ¼ cos h (C7b)
exiting the nozzle, since the melt is no longer con- >
> tan h
: ; for h > h ;
strained, the polymer conformations relax and elastically sin h
stored energy is released leading to die swell. The die-
where h ¼ tan1 ðR=HÞ denotes the angle at which the
corner is reached [Fig. 14(a)]. In this way, the total depos-
ited volume equals that of the cylinder that would be
deposited during vertical extrusion with no die swell.
Figures 14(b) and 14(c) show that qualitatively the stretch
and disentanglement profiles across the layer are similar
to the smooth corner case [Fig. 8(b)]. A quantitative com-
parison is shown in Fig. 15. A square outer-corner region
induces more stretch and disentangles the melt further; at

FIG. 14. (a) Numerical mesh with a square corner at ðy; zÞ ¼ ðR; 0Þ and
resulting polymer deformation across the printed layer: (b) entanglement FIG. 15. Quantitative comparison of the stretch trðAÞ (red lines) and disen-
ðr; /Þ and (c) Tube stretch profile trAðr; /Þ  3 for Zeq ¼ 37, b ¼ 0:3, and tanglement  (blue lines) along the z-axis (ðx; yÞ ¼ ð0; RÞ) for a deposition
Wi N ¼ 2. The blue dotted lines show the equivalent cylindrical mesh with shape with a smooth corner [Eq. (B6), Fig. 6(a)] and a square corner [Eq.
the same volume. (C7), Fig. 14]. Model parameters are Zeq ¼ 37, b ¼ 0:3, and Wi N ¼ 2.
POLYMER FLOWS DURING ADDITIVE MANUFACTURING 395

z ¼ 0 the stretch increases from trA ¼ 10:13 to 10.87 a Thermal diffusivity


(approximately 93% smaller with a smooth corner) and  b CCR parameter
decreases from L ¼ 0:0075 to 0.0068 (approximately c_ Shear rate
10% larger with a smooth corner).
c_ W Shear rate at nozzle wall
Dt Advection time
NOMENCLATURE gh Polar angle
g/ Azimuthal angle
A Polymer deformation tensor
h Polar angle between nozzle and layer
Ars Principle shear deformation
k1 Principle eigenvalue
a(T) WLF shift factor
K Deformation factor tensor
C1 WLF parameter
ls Rouse viscosity
C2 WLF parameter
 Entanglement fraction
dA Area of deposition cross section
L Entanglement fraction at weld site
da Area of mesh element
N Entanglement fraction at nozzle wall
ds Displacement
q Mass density
^
e1 Principle eigenvector
r Total stress
^e x ; ^ey; ^ez Orthonormal Cartesian coordinate basis
s0d Reptation time of chain at T0
Ge Plateau modulus
s0e Rouse time of one entanglement segment at T0
H Layer thickness
s0R Rouse time of polymer chain at T0
K Velocity gradient tensor
sdep Deposition time
L Nozzle length
sres Residence time in nozzle
Lskin Thermal skin layer in deposit
ssw Die swell time scale
L0 Length of heated nozzle section
X Rotation Matrix
M Rotation matrix
M Transformation matrix
References
Me Entanglement molecular weight
Mw Molecular weight [1] Chua, C. K.: K. F. Leong, Rapid Prototyping: Principles and
N Normal stress difference Applications (World Scientific, NJ, 2003), Vol. 1.
p Pressure [2] Li, L., Q. Sun, C. Bellehumeur, and P. Gu, “Composite modeling and
Q Mass flow rate analysis for fabrication of FDM prototypes with locally controlled
R Nozzle outlet radius properties,” J. Manuf. Processes 4(2), 129–141 (2002).
R Polymer end-to-end vector [3] Ziemian, C., M. Sharma, and S. Ziemian, Anisotropic Mechanical
R Mesh points Properties of ABS Parts Fabricated by Fused Deposition Modelling
Re Reynolds’ number (Intech Open Access, Rijeka, 2012).
Rg Polymer radius of gyration [4] Hill, N., and M. Haghi, “Deposition direction-dependent failure crite-
R0 Radius of heated nozzle section ria for fused deposition modeling polycarbonate,” Rapid Prototyping J.
r; / Transformed polar coordinates 20(3), 221–227 (2014).
^r ; ^
h; /^ Orthonormal spherical coordinate basis [5] Drummer, D., S. Cifuentes-Cuellar, and D. Rietzel, “Suitability of
r 0 ; /0 Initial polar coordinates PLA/TCP for fused deposition modeling,” Rapid Prototyping J. 18(6),
^s Flow direction 500–507 (2012).
T Temperature [6] Turner, B. N., R. Strong, and S. A. Gold, “A review of melt extrusion
t Time additive manufacturing processes: I. Process design and modeling,”
Rapid Prototyping J. 20(3),192–204 (2014).
trA  3 Stretch deformation
[7] Ahn, S.-H., M. Montero, D. Odell, S. Roundy, and P. K. Wright,
TN Print temperature
“Anisotropic material properties of fused deposition modeling ABS,”
T0 Reference temperature
Rapid Prototyping J. 8(4), 248–257 (2002).
U Magnitude of velocity vector
[8] Ahn, S. H., C. Baek, S. Lee, and I. S. Ahn, “Anisotropic tensile failure
UL Horizontal average print speed
model of rapid prototyping parts-fused deposition modeling (FDM),”
UN Vertical average print speed
Int. J. Mod. Phys. B 17(08n09), 1510–1516 (2003).
u Velocity vector
[9] Arivazhagan, A., and S. H. Masood, “Dynamic mechanical properties
WiW Equilibrium reptation Weissenberg number at
of ABS material processed by fused deposition modelling,” Int. J. Eng.
nozzle wall
Res. Appl. 2(3), 2009–2014 (2012).
WiRW Equilibrium Rouse Weissenberg number at noz- [10] Lee, C. S., S. G. Kim, H. J. Kim, and S. H. Ahn, “Measurement of
zle wall anisotropic compressive strength of rapid prototyping parts,” J. Mater.
Wi N Mass-averaged equilibrium reptation Process. Technol. 187, 627–630 (2007).
Weissenberg number [11] Esplain, D., K. Arcaute, E. Anchondo, A. Adame, F. Medina, R.
R
Wi N Mass-averaged equilibrium Rouse Weissenberg Winker, T. Hoppe, and R. Wicker, “Analysis of bonding methods for
number FDM-manufactured parts,” in 21st Annual International Solid
Zeq Entanglement number Freeform Fabrication Symposium - An Additive Manufacturing
zM Terminal swell distance Conference (University of Texas, TX, 2010), pp. 37–47.
396 C. MCILROY AND P. D. OLMSTED

[12] Sun, Q., G. M. Rizvi, C. T. Bellehumeur, and P. Gu, “Effect of proc- [33] Takahashi, M., T. Isaki, T. Takigawa, and T. Masuda, “Measurement
essing conditions on the bonding quality of FDM polymer filaments,” of biaxial and uniaxial extensional flow behavior of polymer melts at
Rapid Prototyping J. 14(2), 72–80 (2008). constant strain rates,” J. Rheol. 37(5), 827–846 (1993).
[13] Gibson, I., D. W. Rosen, and B. Stucker, Additive Manufacturing [34] Mohagheghi, M., and B. Khomami, “Molecular processes leading to
Technologies (Springer, New York, 2010). shear banding in well entangled polymeric melts,” ACS Macro Lett.
[14] Seppala, J. E., and K. D. Migler, “Infrared thermography of welding 4(7), 684–688 (2015).
zones produced by polymer extrusion additive manufacturing,” Addit. [35] Nafar Sefiddashti, M. H., B. J. Edwards, and B. Khomami, “Individual
Manuf. 12, 71–76 (2016). chain dynamics of a polyethylene melt undergoing steady shear flow,”
[15] Yardimici, R. S., T. Hattori, S. I. Guceri, and S. C. Danforth, “Thermal J. Rheol. 59(1), 119–153 (2015).
analysis of fused deposition modelling,” in Solid Freeform [36] Bird, R. B., R. C. Armstrong, and O. Hassager, Dynamics of Polymeric
Fabrication Proceedings, edited by D. L. Bourell, J. J. Beaman, H. L. Liquids. Vol. 1: Fluid Mechanics (John Wiley and Sons Inc., New
Marcus, R. H. Crawford, and J. W. Barlow (University of Texas, TX, York, 1987).
1997). [37] Cloitre, M., T. Hall, C. Mata, and D. D. Joseph, “Delayed-die
[16] Du, J., Z. Wei, X. Wang, J. Wang, and Z. Chen, “An improved fused swell and sedimentation of elongated particles in wormlike micel-
deposition modeling process for forming large-size thin-walled parts,” lar solutions,” J. Non-Newtonian Fluid Mech. 79(2), 157–171
J. Mater. Process. Technol. 234, 332–341 (2016). (1998).
[17] Ge, T., F. Pierce, D. Perahia, G. S. Grest, and M. O. Robbins, [38] Allain, C., M. Cloitre, and P. Perrot, “Experimental investigation and
“Molecular dynamics simulations of polymer welding: Strength from scaling law analysis of die swell in semi-dilute polymer solutions,”
interfacial entanglements,” Phys. Rev. Lett. 110(9), 098301 (2013). J. Non-Newtonian Fluid Mech. 73(1), 51–66 (1997).
[18] Wool, R. P., and K. M. O’Connor, “A theory of crack healing in poly- [39] Gladman, A. S., E. A. Matsumoto, R. G. Nuzzo, L. Mahadevan, and J.
mers,” J. Appl. Phys. 52(10), 5953–5963 (1981). A. Lewis, “Biomimetic 4D printing,” Nat. Mater. 15, 413–418 (2016).
[19] Hunt, T. A., and B. D. Todd, “Diffusion of linear polymer melts in [40] Migler, K. B., and A. J. Bur, “Fluorescence based measurement of
shear and extensional flows,” J. Chem. Phys. 131(5), 054904 (2009). temperature profiles during polymer processing,” Polym. Eng. Sci.
[20] Ilg, P., and M. Kr€oger, “Molecularly derived constitutive equation for 38(1), 213–221 (1998).
low-molecular polymer melts from thermodynamically guided simu- [41] Marrucci, G., “Dynamics of entanglements: A nonlinear model consis-
lation,” J. Rheol. (1978-present) 55(1), 69–93 (2011). tent with the Cox-Merz rule,” J. Non-Newtonian Fluid Mech. 62(2),
[21] Uneyama, T., K. Horio, and H. Watanabe, “Anisotropic mobility 279–289 (1996).
model for polymers under shear and its linear response functions,” [42] Graham, R. S., A. E. Likhtman, T. C. B. McLeish, and S. T. Milner,
Phys. Rev. E 83(6), 061802 (2011). “Microscopic theory of linear, entangled polymer chains under rapid
[22] Sood, A. K., R. K. Ohdar, and S. S. Mahapatra, “Experimental investi- deformation including chain stretch and convective constraint release,”
gation and empirical modelling of FDM process for compressive J. Rheol. 47(5), 1171–1200 (2003).
strength improvement,” J. Adv. Res. 3(1), 81–90 (2012). [43] Yaoita, T., T. Isaki, Y. Masubuchi, H. Watanabe, G. Ianniruberto, F.
[23] Likhtman, A. E., and R. S. Graham, “Simple constitutive equation for Greco, and G. Marrucci, “Statics, linear, and nonlinear dynamics of
linear polymer melts derived from molecular theory: Rolie–Poly equa- entangled polystyrene melts simulated through the primitive chain net-
tion,” J. Non-Newtonian Fluid Mech. 114(1), 1–12 (2003). work model,” J. Chem. Phys. 128(15), 154901 (2008).
[24] Doi, M., and S. F. Edwards, The Theory of Polymer Dynamics (Oxford [44] Wang, Z., and R. G. Larson, “Constraint release in entangled binary
University, Oxford, 1988). blends of linear polymers: A molecular dynamics study,”
[25] Ianniruberto, G., “Quantitative appraisal of a new CCR model for Macromolecules 41(13), 4945–4960 (2008).
entangled linear polymers,” J. Rheol. 59(1), 211–235 (2015). [45] Likhtman, A. E., “Single-chain slip-link model of entangled polymers:
[26] Likhtman, A. E., and T. C. B. McLeish, “Quantitative theory for linear Simultaneous description of neutron spin-echo, rheology, and dif-
dynamics of linear entangled polymers,” Macromolecules 35(16), fusion,” Macromolecules 38(14), 6128–6139 (2005).
6332–6343 (2002). [46] Schieber, J. D., and M. Andreev, “Entangled polymer dynamics in
[27] Williams, M. L., R. F. Landel, and J. D. Ferry, “The temperature equilibrium and flow modeled through slip links,” Annu. Rev. Chem.
dependence of relaxation mechanisms in amorphous polymers and Biomol. Eng. 5, 367–381 (2014).
other glass-forming liquids,” J. Am. Chem. Soc. 77(14), 3701–3707 [47] Goyanes, A., U. Det-Amornrat, J. Wang, A. W. Basit, and S. Gaisford,
(1955). “3d scanning and 3d printing as innovative technologies for fabricating
[28] Graham, R. S., “Molecular modelling of entangled polymers under personalized topical drug delivery systems,” J. Controlled Release
flow,” Ph.D. thesis, University of Leeds, 2002. 234, 41–48 (2016).
[29] Ianniruberto, G., and G. Marrucci, “On compatibility of the Cox-Merz [48] Aoki, Y., A. Hatano, T. Tanaka, and H. Watanabe, “Nonlinear stress
rule with the model of Doi and Edwards,” J. Non-Newtonian Fluid relaxation of ABS polymers in the molten state,” Macromolecules
Mech. 65(2), 241–246 (1996). 34(9), 3100–3107 (2001).
[30] Ianniruberto, G., and G. Marrucci, “Convective constraint release [49] Bellini, A., S. Guceri, and M. Bertoldi, “Liquefier dynamics in fused
(CCR) revisited,” J. Rheol. 58(1), 89–102 (2014). deposition,” J. Manuf. Sci. Eng. 126(2), 237–246 (2004).
[31] Ianniruberto, G., and G. Marrucci, “Erratum:Convective constraint [50] Mostafa, N., H. M. Syed, S. Igor, and G. Andrew, “A study of melt
release (CCR) revisited[J. Rheol. 58, 89–102 (2014)],” J. Rheol. 58(4), flow analysis of an ABS-iron composite in fused deposition modelling
1083–1083 (2014). process,” Tsinghua Sci. Technol. 14, 29–37 (2009).
[32] Baig, C., V. G. Mavrantzas, and M. Kroger, “Flow effects on melt [51] Ramanath, H. S., C. K. Chua, K. F. Leong, and K. D. Shah, “Melt flow
structure and entanglement network of linear polymers: Results from behaviour of poly-e-caprolactone in fused deposition modelling,”
a nonequilibrium molecular dynamics simulation study of a polyeth- J. Mater. Sci.: Mater. Med. 19(7), 2541–2550 (2008).
ylene melt in steady shear,” Macromolecules 43(16), 6886–6902 [52] Olmsted, P. D., “Perspectives on shear banding in complex fluids,”
(2010). Rheol. Acta 47(3), 283–300 (2008).
POLYMER FLOWS DURING ADDITIVE MANUFACTURING 397

[53] Tanner, R. I., “A theory of die-swell,” J. Polym. Sci. A-2: Polym. [57] Crockett, R. S., “The liquid-to-solid transition in stereodeposition
Phys. 8(12), 2067–2078 (1970). techniques,” Ph.D. thesis, The University of Arizona, 1997.
[54] Heller, B. P., D. E. Smith, and D. A. Jack, “Effects of extrudate swell [58] Crockett, R. S., and R. S. Calvert, “The liquid-to-solid transition in
and nozzle geometry on fiber orientation in fused deposition modeling stereodeposition techniques,” in Solid Freeform Fabrication
nozzle flow,” Addit. Manuf. 12, 252–264 (2016). Proceedings, edited by D. L. Bourell, J. J. Beaman, H. L. Marcus, R.
[55] Bellini, A., “Fused deposition modelling of ceramins: A comprehen- H. Crawford, and J. W. Barlow (University of Texas, TX, 1996), pp.
sive experimental, analytical and computational study of material 257–264.
behaviour, fabrication process and equipment design,” Ph.D. thesis, [59] Zhang, X., W. Hendro, M. Fujii, T. Tomimura, and N. Imaishi,
Drexel University, 2002. “Measurements of the thermal conductivity and thermal diffusivity of
[56] Shofner, M. L., F. J. Rodr iguez-Mac Ias, R. Vaidyanathan, and E. V. polymer melts with the short-hot-wire method,” Int. J. Thermophys.
Barrera, “Single wall nanotube and vapor grown carbon fiber rein- 23(4),1077–1090 (2002).
forced polymers processed by extrusion freeform fabrication,” [60] Mark, J. E., Physical Properties of Polymers Handbook (Springer,
Composites, Part A 34(12), 1207–1217 (2003). New York, 1996).

You might also like