Sie sind auf Seite 1von 10

Glass Transition and Viscosity prediction of Thin Polymer Films from

a Leveling experiment. A Newtonian Fluid Approach


a Paul F. Nealey,b,c and Juan P. Hernandez-Ortza,b
Daniel A. Olaya-Munoz,
Received Xth XXXXXXXXXX 20XX, Accepted Xth XXXXXXXXX 20XX
First published on the web Xth XXXXXXXXXX 200X
DOI: 10.1039/b000000x

At present, the study of the behavior of polymeric films at nanoscale has become a very useful tool to understand how the
principal properties of polymers change as the structure decreases in size. The dependence of properties of the polymers with
the size of the arrangement, is important in the development of devices like nanostructured semiconductors, based on block
copolymers structures. Properties like Glass Transition Temperature , Loss and Storage energy moduli deviate significantly of
the bulk values, and the characterization of them let us manipulate the behavior of these devices. Through the experimentation,
was suggested that the elastic modulus of polymeric nanostructures is dimension dependent, decreasing as the beam width is
decreased below 50 nm. An initial analysis of thin films is developed using a simple model for the leveling of a Newtonian
Fluid, which allows the calculation of viscosity as a function of temperature. Then, a Numerical Solution (through the Radial
Basis Function Method) is implemented in order to find an approximation of the theoretical viscosity previously found. Thus,
the calculation of the viscosity in function, not only of temperature but also of the size of the arrangement, can be performed,
and the previous approximation is able to provide the Glass Transition Temperature as a function of the period of the film. The
behavior of the films is then simulated through RBF-Method;the heat rate-dependence of the leveling process is analized in a 80
nm Structure, and the behavior of a non-symmetric film is simulated for the same period.

1 Introduction
The material properties of polymeric structures exhibit size
dependence when one or more of the dimensions of the system are below 50-100 nm. However, in the most common applications of polymer films, the thickness of the systems does
not approach to this length scale, and the behavior explained
above doesnt affect the properties. On the other hand, in the
micro and nano-electronics industry, the changes in the behavior of thin films are associated with the decreasing in size of
the polymeric nanostructures, and it can deviate significantly
from the bulk one 14 . Thereby, the need of finding these properties becomes evident, but the measurement of them is actually a big trouble, because of the length scale does not let
make experiments easily in order to extract the properties.
Several experimental and theoretical studies have been developed in order to analyse the behavior of thin films on pata Departamento de Materiales, Universidad Nacional de Colombia, Sede
Medellin, Calle 75 79A-51, Bloque M17, Medellin, Colombia
b Institute for Molecular Engineering, Physical Science Division, The University of Chicago, Chicago, Illinois, 60637, United States.
c Materials Science Division, Argonne National Laboratory, 9700 South Cass
Avenue, Argonne, Illinois, 60439, United States
E-mail: daaolayamu@unal.edu.co
E-mail: nealey@uchicago.edu
E-mail: jphernandezo@wisc.edu

terned surfaces 58 , due to the self-assembly of materials on


chemically patterned surfaces is a feasible approximation to
fabricate dense arrangements with a size below 30 nm, with
applications in the microelectronic industry 9 and magnetic
data storage 10 . Some of these studies have shown that the
Glass Transition Temperature Tg of polymer thin films can
change with thickness 1,11,12 . Mattsson, Forrest and Borjesson 13 report that the Tg of free-standing polystyrene (PS) thin
films decreases monotonically below thickness of 100 nm 14 .
Also, Torres, Nealey and de Pablo have showed in molecular dynamic simulation that the diffusivity of polymers segments is highly heterogeneous in thin polymer films and that
it is strongly correlated with deviations of Tg from the bulk 15 .
These changes presented in the Thermophysical properties of
polymer nanostructures affect the sensitivity of these arrangements, generating unexpected deviations in the morphology of
the structure, related with the thermal and mechanical stability of the structures, and the presence of defects. 16 This behavior has been observed in devices with polymer thin films,
like Block Copolymers (BCP), used in semiconductor industry, where it is essential to create BCP thin films that exhibit,
not only long-range order, but also a structure free of defects
over macroscopic areas 17,18 . Diverse theoretical studies have
been developed for thin films 1926 , in order to predict certain
properties, like Viscosity and Navier Slip Lenght 27, by measuring relaxation time of the film, through an analytical model;

110 | 1

or with the purpose of study the non-linear effects presented in


the leveling process through Non-Newtonian models 28 ; how
the behavior of thin films is affected by changes presented
due to the presence of surfactants 29, and the measurement
of the diffusivity of thin polymer films by a curvature-driven
flow 30,31 .
For concreteness, a first problem based on the leveling of a
polymer thin film that is being subjected to a temperature gradient is considered. In this manner, the viscosity in function of
temperature will be extracted, using a theoretical model developed by Orchard 32 and based on the experimental leveling developed by Delcambre 33 . Then, from the experimental data, a
numerical simulation is performed with Radial Basis Function
Method (RBF) 34 , in order to find a viscosity profile which is
engaged with the theoretical model. After this validation and
employing the calculated viscosity, a simulation of the membrane is developed, in order to predict how the leveling takes
place in this structure and make an approach of the complete
performance during the heating experiment. The Newtonian
Approximation gives valid information about Tg , showing that
It is dimension-dependent and decreases with the decrease of
the size of the membrane.

A study of the emergence of capillary forces that causes


the collapse of the beams, as the width of the pattern is decreases (See Fig. 1), due to the height instabilities in the arrangement, was implemented. It was found that the addition
of TCPP affects the collapse properties of the defined PMMA
nanostructures. The patterns were stabilized by introducing
TCPP, and it was discovered that the pattern containing 5%
TCPP by mass, remains entirely stable. The enhancement in
mechanical stability achieved by TCPP Antiplasticization was
benchmarked through the measurement of the Critical Aspect
Ratio of Collapse (CARC) of the arrangement. As the Aspect
Ratio, shown in the Figure 1, increases beyond a critical value
or CARC, the collapse of the nanostructures due to capillary
forces was observed.

2 Experimental Section
The effects of antiplasticization on the mechanical properties
of dense arrays of poly(methyl methacrylate) (PMMA) beams
less than 75 nm in width were investigated by Delcambre 33,
through observation of the deformation and collapse of these
beams during lithographic processing. In that work, Delcambre observed the deformation and collapse of dense, lithographically fabricated grating nanostructures to quantify the
effects of an antiplasticizing additive 35, tris(2-chloropropyl)
phospate (TCPP), on the mechanical properties of PMMA
structures.
Spacing (S)

Beam Width
(BW)

Length (L)
Pitch = BW + S
Aspect Ratio
AS = H/BW

Height
(H)

Silicon Wafer
Figure 1. Schematic of Dense Arrays of Linear Nanostructures. The
Height (H), Beam Width (BW), Spacing (S), Pitch, Length (L) and Aspect
Ratio (AR) are defined as shown. 33

2|

110

Figure 2. Schematic of the Experimental SAXS Layoyt. Polymer grating samples were mounted on a heating stage in a transmission geometry.
Synchrotron X-Rays penetrated the sample and contacted the detector. 35,36

The shapes of antiplasticized PMMA grating nanostructured 80 nm to 400 nm in width were monitored upon heating in situ Small Angle X-Ray Scattering (SAXS) and ex situ
Atomic Force Microscopy (AFM), in order to quantify the
shape evolution, or leveling, of the nanoimprinted patterns as
a function of temperature and TCPP antiplasticizer concentration. Delcambre found that the nanostructures rapidly flowed
into flat films near a critical temperature, T f low , that decreases
monotonically with decreasing initial structure width, caused
by an increase in applied stresses with decreasing period due
to surface tension and a dimension-dependence of Tg . Also, it
was confirmed, through the topographical profiles determined
with SAXS and AFM, that leveling process is driven by surface tension at the air-polymer surface, and that the leveling
rate was less dependent on line width in TCPP-antiplasticized
structures than in structures composed of pure PMMA.

It was also observed that Tg of grating nanostructured of


PMMA decreases with decreasing line width dimension, according with the results of Mundra, et.al. 37 , and a relation
with the decrease in T f low with decreasing linewidth, was
found by Delcambre 35. However, a dependence of Tg , not
only of the dimension of the arrangements, but also of TCPP
concentration, was found, through a relation between the dependence of T f low on additive concentration and the monotonic depression presented in Tg .

3 Theoretical Model
The focus of this work is the study of thin polymer films subjected to a temperature gradient. In the experimental section
of the research 33,36 , a PMMA sample was used to develop the
leveling in order to extract some important properties. The
system consists in a silicone substrate and a polymeric membrane; a sinusoidal shape is assumed for the last one. The
dimensions of the membrane are around 100 nm (see Fig. 3).
The arrangement has a residual height Hr and a length of repetition interval or Period P.

be neglected (Re 0), and the boundary conditions are determined by means of a stress balance at the interfaces 3840 :
P + 2 u = 0
u = 0,

u = 0

at the Substrate

[ n] = ( n)n,

at the Free Surface

(1)
(2)
(3)

where represents the Stress Tensor ( = P + ), u


represents the velocity vector, P is the pressure of the system,
n is the normal vector at the interfaces, t is the tangential vector at the interfaces, is the Surface Tension, and n represents the mean curvature of the free surface. A symmetry
condition is imposed in the model, due to the periodicity of
the film, and is represented as a boundary condition at the side
boundaries as 28 :
ux =

uy
=0
x

(4)

n=s(n)n

air

polymer

-P + 2u=0
u=0
Figure 3. SEM Transversal Image. Silicon Substrate. Height of the
pattern (H) is 100 nm and the period (P) is 160 nm. The residual layer is 20
nm approximately. 36

u=0
Figure 4. Representation of the Geometry of the Membrane. The

This film relaxes into a flat surface, due to the action of a


thermal heating, generating an increase in the size of the residual layer. Thereby, it is possible to extract some properties,
like viscosity and Tg , as a function of the calculated velocities.
But this is possible, just if it is assumed that the viscosity is
only function of Temperature and Pressure i.e., the polymeric
film is assumed to have a Newtonian Fluid Behavior (See Fig.
4).
In order to analyse the behavior of the thin films subjected to
a temperature gradient, the governing equations of the system
are established. The evolution of the system is represented
through the Stokes Equations, due to the Inertial Forces can

geometry of the system is assumed to be sinusoidal, that relaxes into a flat


surface. The stress balance at the air-polymer interface gives the boundary
condition for the velocity at the free surface.

For incompressible fluids, the pressure is not thermodynamic variable. It is a Lagrange multiplier that constrains the
velocity field in order to remain divergence-free. The boundary conditions for pressure were developed for the incompressible Navier-Stokes Equations 41 by Gresho et. al. The
leveling problem is represented by the Stokes Flow, which
leads to the Poisson Equation for the Pressure field:
2 P = 0

(5)
110 | 3

(r2 + c2 )m/2

Due to the film is not subjected to a pressure gradient, at the


free surface, the pressure boundary condition is:
n P = 0

100

80

80

60

60

40

40

20

20

20
60
100

40

20

20

40

20
60 60
100

80

80

60

60

40

40

20

20

40

20

20

40

20
60 60

(10)

gaussian

40

20

20

40

60

Thin-plate splines have been considered optimal for interpolation of multivariate functions but they only converge linearly. 44,45 In this work, the generalized thin-plate spline function is used to approximate the variables, due to the simplicity
in the formulation of the systems of equations, and it avoids
inclusion of the shape parameter c that is related with multiquadratic and gaussian functions. The implementation of the
RBF into the leveling problem started with the making of a
uniform grid and approaching the velocity and pressure functions, found in the Stokes Equations, as follows:

[
j

N
40

20

20

40

ux u

(11)

uy u

(12)

P P

(13)

Then, the governing equations of the system, i.e. Stokes


Equations, can be represented as follows:

20
60

er/c

(9)

(6)

Through an analytical model developed by Orchard 32,


which calculates and predicts the behavior of a Newtonian
Thin Film, is possible to solve the governing equations of the
system, assuming a smooth free surface, a constant surface
tension and a viscosity only as a function of temperature. In
this way, the analytical model was performed, under the conditions of the structure (See Fig. 4) and a constant viscosity
and surface tension were chosen. This analysis proves that the
model is able to reproduce the conditions of the leveling of a
polymeric film, in an approximate way, as shown in the Figure
5.
100

generalized multiquadratic,

60

P
2 u 2 u
+
) j] = 0
j + (
x
x2
y2

(14)

P
2 u 2 u
+
) j] = 0
j + (
y
x2
y2

(15)

Figure 5. Newtonian Fluid Leveling. Simulated nanostructure with conN

stant viscosity 7 105 Pa s and surface tension 0.0382 N/m. These values are

characteristic of bulk PMMA

u
u
j +
j] = 0
x
y

(16)

The boundary conditions for velocity are expressed as follows:

3.1 Radial Basis Function (RBF) Approach

The idea behind the RBF approach is finding a solution to the


leveling problem, through the approximation of the velocity
and pressure of the arrangement as a linear combination of
a complete set of N functions. The coefficients of the linear combination are obtained by solving a linear system that
arises when the function is forced to adopt specific values at N
known locations 34 . Using the meshless RBF, a function f (x)
is approximated by

u j = 0
N

[ p j +(2

(17)

nx ny + (2

u
u
u
j )nx nx +2 (
j +
j)
x
y
x

u
j )ny ny ] = 2H
y

f (x) j (kx x j k)

at the Substrate

at the Free Surface


(18)

(7)

where N is the number of total collocation points, j s are unknown coefficients and (kx x j k) is the Radial Basis Function. Commonly used RBFs include: 42,43
2a

r log r
4|

generalized thin-plate spline,


110

(8)

[(2
j

u
u
u
j )nx ny + (
j +
j )(n2y n2x )
x
y
x

(2

u
j )nx ny ] = 0
y

at the Free Surface (19)

where nx and ny are the vectorial components of the Normal


Vector to the interfaces 2 and 3 ; and H represents the Mean
Curvature of the Free Surface, which is calculated as shown
below. 38, 46




d2y


dx2

2H =
(20)
dy 2 3/2
[1 + ( dx ) ]
where y(x) represents the free surface function.
The symmetry condition can be described as:
N

u j =
j

N
u
P
j =
j = 0
x
j x

RBF is able to predict the theoretical behavior of the viscosity, previously found with the Orchards Model.
2

10

Orch.

TG

10

10

10

(21)

10

10

4 Results and Discussion


4

10

330

340

350

360

370

4.1 Analytical Viscosity Profile

390

400

410

420

Figure 6. vs. T for a Nanostructured PMMA Grating. Calculated


using the Analytical Orchard Leveling Model.
0

0.1

0.2

u [nm/s]

The theoretical velocities of the system can be calculated using the analytical model of Orchard. By this way, it is possible
to use the results of the experimental leveling developed by
Delcambre explained in the experimental section of this paper, where the height of the structure was measured through
the in situ monitoring with the Small Angle X-Ray Scattering, as it was decreasing in function of temperature, in order
to obtain a characteristic curve for the viscosity as a function
of temperature, by minimizing the residual error between the
experimental velocity calculated from the measured height of
the experiments and the theoretical velocity found based on
the Orchard Model. This calculation was developed through a
least square minimization between the experimental and theoretical velocities, using the MINPACK library 47, with the purpose of find values of viscosity, which allowed minimize that
residual error.The analysis explained above is able to provide a
continuous viscosity profile that passes through the glass transition, observed at the inflexion point.

380

0.3

0.4

0.5

u Orch.
u

RBF.

0.6

0.7
330

340

350

360

370

380

390

400

410

420

Figure 7. Experimental and Theoretical Velocities vs. Temperature for


a Nanostructured PMMA Grating. Calculated using the Analytical Orchard
Leveling Model.
2

10

Orch.

4.2 Viscosity Profile: RBF Approach

RBF.
1

10

10

Through the RBF approach, a viscosity profile is obtained, by


means of the minimization of the experimental velocity profile found by Delcambre 33, with the numerical velocity found
with RBF Method. The form of the free surface is assumed
to be sinusoidal in the simulation. A regular grid was implemented in the fitting process, in order to achieve a better
interpolation error; the resolution of the grid was chosen as a
function of the interpolated pressure laplacian. The parameters of the TPS function used to approximate the velocity and
pressure fields were chosen as au = 3 and a p = 2. The least
square minimization between the theoretical and experimental velocities was also developed using the MINPACK library.
The results show that the numerical solution developed with

10

10

10

10

330

340

350

360

370

380

390

400

410

420

Figure 8. vs. T for a Nanostructured PMMA Grating. The graph


shows the predicted viscosity profile using the RBF approach and the analytical solution predicted by the Orchards Model.

110 | 5

4.3 Simulation: 120 nm Arrangement


The simulation of the leveling of a 120 nm film is developed
through the RBF Method, using the boundary conditions and
the Theoretical model described above. The time evolution of
the free surface is governed by a kinematic equation expressing that the surface must move with the fluid 48 :

h
h
= uy ux
(22)
t
x
where h is the height function that represents the free surface, and ux and uy are the fluid velocity components. With
the purpose of avoid temporal instabilities in the simulation,
an adaptive time-step is implemented as a function of the maximum velocity in the film, as follows:
tadapt. =

k
umax.

(23)

where k is an scaling constant chosen to be 0.00005.


In order to ensure the smoothness of the free surface of the
mesh, three smoothing methods were implemented in the simulation. The first method is based on the connection of the
surface nodes with its neighbors through an elastic force 49.
At equilibrium, the net force over a specified node (xi ) is zero,
and its displacement can be calculated using the equilibrium
equation:
Neigh.

ki j (x j xi ) = 0

(24)

where ki j is a spring constant related to the inverse of the


distance between xi and each x j . The displacement of the surface is iterated until the displacement converges with a relative
tolerance 1010. The second method that is applied is the
Mean Curvature Flow (MCF) Correction 50, due to the instabilities at the free surface are reflected in changes of concavity
of the Mean Curvature. The approximation of the MCF is
given by:
(25)
xit+1 = xi t + H (xit )n(xi t )
where H is the Mean Curvature of the Free Surface, xit+1
is the position of the node in the time t + 1 and xit is the actual position of the node, and n is the normal vector. These
smoothing methods were implemented in the leveling simulation, but a distortion in the free surface was achieved and the
velocity field found with the interpolation method presented
instabilities, leading to a distortion in the grid.
Then, the third method implemented in the numerical solution is based on a Laplacian Smoothing for the position of
the boundary nodes. The Mean Curvature H is calculated using the equation (20) and the first and second derivatives are
calculated as follows 51 :
6|

110

yix =
yixx =

1
(y 8yi1 + 8yi+1 yi+2 )
12h i2

(26)

1
(yi2 + 16yi1 30yi + 16yi+1 yi+2 ) (27)
12h2

Then, when the velocity field is calculated, the surface is


moved using the kinematic equation and a Laplacian Smooth2
ing is applied to the second derivative xy of a central node,
based on the second derivative of the neighboring nodes as
follows:

2 yn+1
2 yn
i
= (1 ) 2i +
2
x
x
Ni

2 ynj
2
j=1 x
Ni

(28)

where is a relaxation parameter, here taken to be 0.05,


and Ni is the number of neighbors, set as 4. Through this
method, a smooth mean curvature was achieved during all the
simulation, and any instabilities in the grid were observed,
leading to a continuous and homogeneous interpolated velocity and pressure fields.
For the update of the internal nodes of the grid, three methods were implemented. The first consists in a spring-based
smoothing method 52,53 , where the nodes at the free surfaces
move according to the kinematic equation, and all the internal nodes are connected with springs to its neighbors, and
the movement of them is given by a relaxation of the elastic
force between the boundary and the internal nodes, as shown
in the equation (24). The second method is an analogy to
the Diffusion-Based Smoothing, used in the CFD Modules,
based on the Elastic Mesh Model based on that proposed by
Lynch 54 . The linear elasticity model takes the form of a Poisson Problem:
[C(x)] = 0

(29)

where x is a interior vector of displacements to be found


and C is a 4th -order elasticity tensor, that was taken to be
the identity tensor. The boundary conditions for this Poisson
Problem are given by the displacements found at the boundaries of the film previously, and this system of equations is
also solved through the RBF Method.
The Spring-Based and the Elastic Mesh Method were implemented in the leveling simulation and the solution for small
deformations of the mesh allows an appropriate interpolation
of the solution, but when the grid is deformed in the simulation near the Glass Transition, the resolution of the mesh is
not uniform in all the domain and the interpolation leads to an
error in the solution for the pressure and velocity fields.
Then, for the update of the internal nodes, a remeshing technique is implemented in the numerical solution, in order to

10

ensure a constant resolution of the mesh in all the simulation.


This resolution is chosen as a function of the minimal value
for the pressure laplacian found with the RBF Approximation,
and remains constant, which results in a good interpolation
through all the steps of the leveling simulation.

10

10

0.000494

t = 240 s.

t = 0 s.

0.0108

120

120

0.00862

0.000395

10

100

100

80

80

0.00646

0.000296

60

60

0.00431

0.000198

10

10

40

40

0.00215

9.88e05
20

20

0
60

40

20

20

40

60

3.81e12

60

40

20

20

40

60

6.1e11

12

10
t = 360 s.

120

0.33

t = 390 s.

120

0.594

330

340

350

360

370

380

390

400

410

420

100

100

0.475

0.264

80

80
0.198

0.356

60

Figure 11. Pressure Laplacian. The Pressure Poisson Equation is calcu-

60

0.132

0.238

40

lated with the BCs established in the leveling problem in an indirect way. For
Stokes Flow, it is equal to zero, and the RBF approximation leads to a good

40

0.0661

0.119

20

20

0
60

40

20

20

40

60

t = 410 s.

120

5.22e09

60

0.626

40

20

20

40

60

t = 488 s.

120

approach of the solution of the Poisson Equation for the Pressure Field.

6.14e09

0.0521

4.4 Heat Rate Dependence: 80 nm Profile

100

100

0.0416

0.501

80

80

0.0312

0.375

In this section will be performed a simulation showing the dependence of the leveling of a 80 nm film, with the heat rate of
the arrangement (5 K/min, 10 K/min, 20 K/min).

60

60

0.0208

0.25
40

40

0.0104

0.125
20

20

0
60

40

20

20

40

60

3.5e09

60

40

20

20

40

60

1.88e10

4.5 Glass Transition Temperature


Figure 9. Simulation of the 120 nm Arrangement. The extracted viscosity using the RBF Method is implemented in this simulation. The performance of the film during the experiment is simulated from the initial temperature, until the film becomes into a flat surface.
1

Based on the solution found for the Viscosity profile with


the RBF approximation, simulations of the behavior of membranes with a different period were performed, in order to
analyse how the decay of the arrangement unfolds and predict, not only the Newtonian Viscosity of the pattern, but also
the Glass Transition Temperature Tg .

0.9

1
80nm
120nm
160nm

0.8
0.9

0.7
0.8

0.6

0.7

0.5
0.6

0.4
0.5

0.3
0.4

0.2
0.3

0.1
0
330

0.2

340

350

360

370

380

390

400

410

420

0.1

0
330

340

350

360

370

380

390

400

410

420

Figure 10. Normalized Height. The amplitude of the film is calculated


as a function of the Temperature in the simulation of the 120 nm Film. A
non-linear phenomena is observed at the crest and trough of the film, leading
to a no-conservation of the sinusoidal shape of the free surface, due to the
magnitude of the velocity at these points is not the same.

Figure 12. n vs. T for 3 different Nanostructured PMMA Patterns


. The Normal Height n of 3 different membranes with period of 80 nm,
120 nm, and 160 nm respectively, is shown in function of temperature, in the
experiment developed by Delcambre 35,36 .

110 | 7

0.1

01

02

03

Two different arrangements were selected for the analysis,


with a period of 80 nm and 160 nm, respectively. RBF simulations were implemented with the purpose of extract the viscosity profile of each membrane, and the viscosity gradient T
was calculated in order to extract the Tg data, as a function of
the membrane period. As shown in the Figure 8, a dependent
viscosity of the size of the membrane is presented in function
of temperature. The results of simulation indicate that the viscosity decreases as the structure decreases in size, and exhibits
a non-linear behavior in function of the period of the arrangement. The Figure 9 presents a viscosity gradient T for each
film size, and shows a turning point in the curve, that is related
with the point where the polymer realizes a transition from a
solid-like state to a liquid-like state or Glass Transition.

04

05

06

07

08

09

1
330

340

350

360

370

380

390

400

410

420

Figure 14. T vs. T for Films with a period of 80 nm, 120 nm and
160 nm .The gradient of Viscosity allows to calculate the Glass Transition

10

Temperature, by choosing the turning point showed in the curve.

80nm
120nm

10

125

TG Bulk : 121 C

160nm
120

10

115

10

110

105

T [C]

10

100

10

95

10

90

85

10

330

340

350

360

370

380

390

400

410

420
80

50

100

150

200

250

300

350

400

Period [nm]

Figure 15. Tg vs. Membrane Period.The Glass Transition Temperature


is presented as a function of the period of the pattern.
Figure 13. vs. T for a Nanostructured PMMA Grating with 3 Different Period. The graph shows the predicted viscosity profile using the RBF
approach, showing the dependence of the Viscosity with the Period of the
Pattern.

The temperature where this phenomena occurs is known as


Glass Transition Temperature Tg and can be calculated locating the turning point in the curve. For the 80 nm membrane,
Tg is equal to 111C. The corresponding Tg for the 120 nm
membrane is 116 C, and for the 160 nm membrane is approximately equal to 120 C, as shown in the Figure 12. The Tg
data for the 400 nm membrane was taken from Delcambres
research 35, according to the calculation of the viscosity profile for this structure, and It was reported that Tg for this pattern was approximately equal to the bulk value for the glass
transition temperature (121 C).
8|

110

Conclusions

It has been developed a solution for a Leveling of Newtonian


Fluids, based on the analytical solution performed by Orchard.
This solution found by simulation, shows that the theoretical
model works, and is able to provide a continuous viscosity
profile, that presents a first approximation to the calculation of
Tg , if the polymer is assumed to be a Newtonian Fluid. Numerical solutions using RBF Method were performed in order
to find a viscosity profile, for different sizes of the PMMA
Membranes and was found that the viscosity is dimensiondependent and decreases as structure decreases in size. From
these calculations, Tg was extracted, according to the analysis of the viscosity gradient, in function of the period of the
membrane, and It shows a decreasing monotonic behavior as
the size of the structure decreases. Further work shall consist in the addition of a linear viscoelasticity model, because
of the length scale of the system is associated with small deformations. Thereby, the viscosity will be function not only

of temperature and pressure, but also of shear rate and shear


stress, leading to a non-Newtonian fluid problem.

Acknowledgements
The authors acknowledge to S.P. Delcambre for the experimental results used in this work. D.A.O-M is grateful to D.
Calle, A. Martiliano, H.A. Castano, for helpful discussions
and support.

Appendix A: Analytical Model for a Newtonian


Leveling (Orchards Model)

where is the angle made by the surface with the x-axis.


The surface gradient is assumed to be small and the surface
tension gradient is zero, due to an air-polymer interface is
present. Then:

(36)
y =
R
xy = 0
(37)
The condition of no-shear stress at the free boundary now
yields:
An Kn + Dn = 0
(38)
With the normal stress balance, we obtain:
Bn =

With the purpose of calculate the theoretical velocities of the


system, the Orchards model is implemented 32, using a stream
function , defined as follows:
ux =

uy =

P + 2 u = 0

u = 0

An = Bn

(31)

If the pressure term is eliminated of the equation (22) using the divergence of the velocity field, yields the biharmonic
equation:
2 2 = 0
(32)
Suppose that the form of the free surface is arbitrary, and
that it extends in the x-direction. Then, the solution of (23)
may be sought in the form of Complex Fourier Series:
= eiKK n x [An Cosh(Kn y) + Bn Sinh(Kn y)
+ Cn y Cosh(Kn y) + Dn y Sinh(Kn y)] (33)
The coefficients An , Bn , Cn and Dn are calculated using
both, the free and fixed boundary conditions. At the surface,
there is no shear stress component tangential to the liquid-air
interface, but a stress of magnitude /R ( : Surface Tension
and R: Radius of Curvature of the Free Surface) exists normal to the interface. The stress components at the surface are
related by:

( )Cos( ) = yCos( ) + xy Sin( )


R

(34)

( )Sin( ) = y Sin( ) + xyCos( )


R

(35)

(39)

where an is the amplitude of the Fourier components.


Using the fixed boundary conditions i.e., (ux = uy = 0) at y
= Hr , we obtain:

(30)

where ux is the x-component of the velocity and uy is the ycomponent. The Reynolds number of the system is zero, because there is not presence of inertial forces, due to the characteristic length of the system is very small. Taking into consideration this condition is obtained the Stokes Equation and
the divergence of velocity field:

an
2

Tanh( ) Sech2( )
1 + 2Sech2( )
1
1 + 2Sech2( )

(41)

Tanh( ) Sech2( )
1 + 2Sech2( )

(42)

Cn = Kn Bn
Dn = Kn Bn

(40)

and using the definition of velocity at free surface:


an = a0 exp(

Kn f ( )

dt)

(43)

where = Kn Hr .

References
1 R. S. Tate, D. S. Fryer, S. Pasqualini, M. F. Montague, J. J. de Pablo and
P. F. Nealey, The Journal of Chemical Physics, 2001, 115, 99829990.
2 M. P. Stoykovich, K. Yoshimoto and P. F. Nealey, Applied Physics A:
Materials Science and Processing, 2008, 90, 277283.
3 R. a. Riggleman, K. Yoshimoto, J. F. Douglas and J. J. De Pablo, Physical
Review Letters, 2006, 97, 14.
4 T. R. Bohme and J. J. De Pablo, Journal of Chemical Physics, 2002, 116,
99399951.
5 C.-c. Liu, A. Ramrez-Hernandez, E. Han, G. S. W. Craig, Y. Tada,
H. Yoshida, H. Kang, S. Ji, P. Gopalan, J. J. D. Pablo and P. F. Nealey,
Macromolecules, 2013, 46, 14151424.
6 G. Liu, F. Detcheverry, A. Ramrez-Hernandez, H. Yoshida, Y. Tada, J. J.
de Pablo and P. F. Nealey, Macromolecules, 2012, 45, 39863992.
7 S. O. Kim, H. H. Solak, M. P. Stoykovich, N. J. Ferrier, J. J. De Pablo and
P. F. Nealey, Nature, 2003, 424, 411414.
8 A. Chakrabarti and H. A. O. Chen, Journal of Polymer Science: Part B:
Polymer Physics, 1998, 36, 31273136.
9 M. P. Stoykovich, H. Kang, K. C. Daoulas, G. Liu, C.-C. Liu, J. J.
de Pablo, M. Muller and P. F. Nealey, ACS nano, 2007, 1, 16875.
10 R. a. Segalman, Materials Science and Engineering: R: Reports, 2005,
48, 191226.

110 | 9

11 D. S. Fryer, P. F. Nealey and J. J. de Pablo, Macromolecules, 2000, 33,


64396447.
12 D. S. Fryer, R. D. Peters, E. J. Kim, J. E. Tomaszewski, J. J. de Pablo,
P. F. Nealey, C. C. White and W.-l. Wu, Macromolecules, 2001, 34, 5627
5634.
13 J. Mattsson, J. Forrest and L. Borjesson, Physical review. E, Statistical
physics, plasmas, fluids, and related interdisciplinary topics, 2000, 62,
5187200.
14 J. Forrest and J. Mattsson, Physical review. E, Statistical physics, plasmas,
fluids, and related interdisciplinary topics, 2000, 61, R536.
15 J. a. Torres, P. F. Nealey and J. J. De Pablo, Physical Review Letters, 2000,
85, 32213224.
16 M. P. Stoykovich, M. Muller, S. O. Kim, H. H. Solak, E. W. Edwards, J. J.
de Pablo and P. F. Nealey, Science (New York, N.Y.), 2005, 308, 14426.
17 U. Nagpal, M. Muller, P. F. Nealey and J. J. de Pablo, ACS Macro Letters,
2012, 1, 418422.
18 R. Ruiz, H. Kang, F. a. Detcheverry, E. Dobisz, D. S. Kercher, T. R. Albrecht, J. J. de Pablo and P. F. Nealey, Science (New York, N.Y.), 2008,
321, 9369.
19 D. W. Bousfield, Journal of Non-Newtonian Fluid Mechanics, 1991, 40,
4754.
20 F. Joos, AiChE Journal, 1996, 42, 623637.
21 N. T. Malamataris and T. C. Papanastasiou, Industrial & Engineering
Chemestry Research, 1991, 30, 22112219.
22 E. Rognin, S. Landis and L. Davoust, Langmuir, 2014, 30, 69636969.
23 S. Livescu, R. V. Roy and L. W. Schwartz, Journal of Non-Newtonian
Fluid Mechanics, 2011, 166, 395403.
24 A. Oron and S. G. Bankoff, Reviews of Modern Physics, 1997, 69, 931
980.
25 R. V. Craster and O. K. Matar, Reviews of Modern Physics, 2009, 81,
11311198.
26 L. E. Stillwagon and R. G. Larson, Journal of Applied Physics, 1988, 63,
52515258.
27 P. Gilormini and H. Teyss`edre, 2013, 20130457, 123.
28 R. Keunings and D. Bousfield, Journal of Non-Newtonian Fluid Mechanics, 1987, 22, 219233.
29 L. W. Schwartz, D. E. Weidner and R. R. Eley, Langmuir : the ACS journal of surfaces and colloids, 1996, 11, 36903693.
30 I. Karapanagiotis and W. W. Gerberich, Macromolecules, 2005, 38, 3420
3425.
31 I. Karapanagiotis and W. W. Gerberich, Surface Science, 2006, 600,
11781184.
32 S. Orchard, Applied Scientific Research, Section A, 1963, 11, 451464.
33 S. P. Delcambre, R. a. Riggleman, J. J. de Pablo and P. F. Nealey, Soft
Matter, 2010, 6, 2475.
34 T. Osswald and J. Hernandez-Ortz, Polymer Processing. Modeling and
Simulation, Carl Hanser-Verlag, 1st edn, 2006.
35 S. P. Delcambre, PhD thesis, Department of Chemical and Biological Engineering, University of Wisconsin-Madison, 2010.
36 K. Nyga rd, S. P. Delcambre, D. K. Satapathy, O. Bunk, P. F. Nealey and
J. F. van der Veen, Macromolecules, 2012, 45, 57985805.
37 M. K. Mundra, S. K. Donthu, V. P. Dravid and J. M. Torkelson, Nano
Letters, 2007, 7, 713718.
38 W. Deen, Analysis of Transport Phenomena, Oxford University Press, 2nd
edn, 2011.
39 L. G. Leal, Advanced Transport Phenomena. Fluid Mechanics and Convective Transport Processes., Cambridge University Press, 1st edn, 2007.
40 V. Ajaev, Interfacial Fluid Mechanics. A Mathematical Modeling Approach, Springer, 2012.
41 P. M. Gresho and R. L. Sani, International Journal for Numerical Methods in Fluids, 1987, 7, 11111145.
42 F. Hildebrand, Introduction to Numerical Analysis, Dover Publications

10 |

110

Inc., 2nd edn, 2003.


43 G. Demirkaya, C. W. Soh and O. Ilegbusi, Applied Mathematical Modelling, 2008, 32, 1848 1858.
44 C. Chen and M. Goldberg, Discrete Projection Methods for Integral
Equations, WIT Press., 1996.
45 M. Powell, Numerische Mathematik, 1994, 68, 107128.
46 S. K. M. G. Bronshtein, I.N. and H. Muehlig, Handbook of Mathematics,
Springer, 2005.
47 H. K. Garbow, B.S. and J. More, Documentation for Minpack Subroutines, Argonne National Laboratory, 1980.
48 C. Hirt and B. Nichols, Journal of Computational Physics, 1981, 39, 201
225.
49 M. Dai and D. P. Schmidt, Journal of Computational Physics, 2005, 208,
228252.
50 Y. Ohtake, A. Belyaev and I. Bogaevski, CAD Computer Aided Design,
2001, 33, 789800.
51 R. Peterson, PhD thesis.
52 B. Perot and R. Nallapati, Journal of Computational Physics, 2003, 184,
192214.
53 S. Quan and D. P. Schmidt, Journal of Computational Physics, 2007, 221,
761780.
54 D. R. Lynch, Journal of Computational Physics, 1982, 47, 387411.

Das könnte Ihnen auch gefallen