ArticlePDF Available

Self assembly of ABC triblock copolymer thin films on a brush-coated substrate

Authors:

Abstract and Figures

Self assemblies of ABC triblock copolymer thin films on a densely brush-coated substrate were investigated by using the self-consistent field theory. The middle block B and the coated polymer form one phase and the alternating phase A and phase C occur when the film is very thin either for the neutral or selective hard surface (which is opposite to the brush-coated substrate). The lamellar phase is stable on the hard surface when it is neutral and interestingly, the short block tends to stay on this hard surface. The rippled structure forms when the cylindrical phase exists near the surface between grafted polymers and ABC block copolymers. Due to the existence of the hydrophilic brush-coated surface serving as a soft surface of the film, the energy fluctuation existing in the film confined by two hard surfaces disappears. The results are helpful for designing the nanopattern of the film and realizing the functional thin film, such as adding the functional short block A to the BC diblock copolymer.
Content may be subject to copyright.
Chinese Journal of Polymer Science Vol. 27, No. 4, (2009), 583
592 Chinese Journal of
Polymer Science
©2009 World Scientific
SELF ASSEMBLY OF ABC TRIBLOCK COPOLYMER THIN FILMS ON A BRUSH-
COATED SUBSTRATE*
Zhi-bin Jiang, Rong Wang** and Gi Xue**
Department of Polymer Science and Engineering, State Key Laboratory of Coordination Chemistry, Nanjing National
Laboratory of Microstructures, School of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210093, China
Abstract Self assemblies of ABC triblock copolymer thin films on a densely brush-coated substrate were investigated by
using the self-consistent field theory. The middle block B and the coated polymer form one phase and the alternating phase
A and phase C occur when the film is very thin either for the neutral or selective hard surface (which is opposite to the brush-
coated substrate). The lamellar phase is stable on the hard surface when it is neutral and interestingly, the short block tends to
stay on this hard surface. The rippled structure forms when the cylindrical phase exists near the surface between grafted
polymers and ABC block copolymers. Due to the existence of the hydrophilic brush-coated surface serving as a soft surface
of the film, the energy fluctuation existing in the film confined by two hard surfaces disappears. The results are helpful for
designing the nanopattern of the film and realizing the functional thin film, such as adding the functional short block A to the
BC diblock copolymer.
Keywords: Functional self-assembly; Thin film; Simulation; Self-consistent field theory; ABC triblock copolymer.
INTRODUCTION
Block copolymers can microphase separate to form a variety of ordered nanoscale morphologies[1, 2]. Self
assembly and phase separation in diblock copolymers have been well studied and characterized both
theoretically and experimentally in the last few decades[310]. Recently, there has been a great interest in studying
the thin film of block copolymers experimentally and theoretically. The function of the thin film is very
important for its application besides the perfect nanopattern of the film. In thin films, incompatibility between
different polymer components and the thin film thickness creates frustration, the effect of which can be probed
by surface-induced effect. The presence of a surface or interface can strongly influence the microdomain
morphologies and the kinetics of microdomain ordering. In general, phase separation is accompanied by a
minimization of the interfacial area of contact between dissimilar components, resulting in a reduction of the
enthalpy. This is accompanied by a reduction of the conformational entropy of the polymer chains. The complex
and rich phase behaviors depend not only on molecular parameters such as the interaction energies between
distinct blocks and the architectures of block copolymers, but also on external variables such as electric
fields[11, 12], temperature gradients[13], chemically patterned substrates[1421] and interfacial interactions[2225].
The phase behavior of diblock copolymer melts confined in a parallel slit or in a thin film has been
extensively studied[3, 2629]. Comparing with two-component systems (i.e. AB diblock copolymers), the phase
behavior of ABC triblock terpolymers is much more complex, and a much larger number of simulation
* This work was supported by the National Natural Science Foundations of China (Nos. 20504013, 20674035, 20874046 and
50533020), the National Basic Research Program of China (No. 2007CB825101), the Nanjing University Talent
Development Foundation (No. 0205004107), and the Natural Science Foundation of Nanjing University (No. 0205005216).
** Corresponding author: Rong Wang (汪蓉), E-mail: rong_wang73@hotmail.com
Gi Xue (薛奇), E-mail: xuegi@nju.edu.cn
Received September 8, 2008; Revised October 27, 2008; Accepted November 11, 2008
Z.B. Jiang et al.
584
parameters is needed to describe thin films: the film thickness (H), the volume fractions of the components fA, fB
and fC (1 – fA fB), three mutual interaction parameters between the components
χ
AB,
χ
AC,
χ
BC and interaction
parameters between the interfaces and the three components (
χ
AS,
χ
BS, and
χ
CS, and S representing the surface).
Pickett and Balazs have used the self-consistent field theory to probe the preferential orientation of lamellae
formed by an ABC triblock terpolymer confined between two walls attracting the middle block[30]. They found
that an orientation of the lamellae perpendicular to the plane of the film orientation is highly favored, indicating
that triblock terpolymers possess distinct advantages over diblocks in technological applications. Monte Carlo
simulations by Feng and Ruckenstein for ABC melts in thin films show that the microdomain morphology can
be very complicated and is affected by the composition, the interactions[31] and even the geometry of the
confinement. Chen and Fredrickson have applied the self-consistent field theory and strong segregation limit
studies (SSL) to investigate the films of linear ABC triblock terpolymer melts[32] for the particular case where A
and C blocks are in equal size, the interaction parameters are identical, and both walls have identical chemical
properties. Self-consistent field theory (SCFT)[33] and dynamic density functional field theory[34] are also used to
study symmetric ABC triblock copolymer thin films. The above results show that the morphology of block
copolymer thin films depends on the film thickness, the surface-polymer interaction, and the incompatibility
between different blocks of the copolymer, which is usually characterized by a Flory-Huggins parameter. We
can tailor surface fields, film thicknesses and even the geometry of the confinement to manipulate the
microdomain structure, shape and orientation.
In practice, besides using pure (silicon) substrates, the hydrophobic or hydrophilic substrates are prepared
by chemically grafting self-assembled monolayer onto (silicon) substrate[35], which form the so-called “polymer
brush” to provide a simple way of modifying surface properties, such as adhesion, lubrication and wetting
behavior. Ma and his group members[36, 37] observed the structure transformation of the AB diblock copolymer
thin film by tailoring the grafting density of the coated surface or the concentration of the copolymer. In this
paper, we use a combinatorial screening method based on the real space implementation of the SCFT, originally
proposed by Drolet and Fredrickson for block copolymer melts[38, 39], to search the equilibrium microphases of
ABC linear triblock copolymers with short end block confined between the polymer brush-coated surface and
the flat impenetrable surface in 2D due to a computationally cost task of solving SCF equations in 3D. Polymer
films on brush-coated substrate are mimicked as a polymer melt confined between the brush-coated surface and
a hard surface. The simulations are performed on a square grid Lx × Lz. The walls are presented as lines at z = 0
and Lz . Hence, the film thickness is Lz.
THE MODEL
We assume the ABC triblock copolymer is confined between the polymer-grafted substrate and the flat
impenetrable surface (hard surface) with a distance Lz along the z-axis. There are nc ABC triblock
copolymer chains with polymerization N and ng polymer chains with polymerization P (here, we take P =
N) grafting on the substrate. Each copolymer chain consists of N segments with compositions (average
volume fractions) fA and fB (fC = 1 – fA fB), respectively. The monomers of the ABC triblock copolymer
and the grafted polymers are assumed to be flexible with a statistical length a and the mixture is
incompressible with each polymer segment occupying a fixed volume 1
0
ρ
. The substrate coated by the
polymers is horizontally placed on the xy plane at z = 0 and the hard free surface is placed at z = Lz. The
volume of the system is V = LxLz, where Lx is the lateral length of the substrate and the flat impenetrable
surface along the x-axis and Lz is the film thickness. The grafting density is defined as
σ
= nga/Lx. The
average volume fractions of the grafted chains and copolymers are
ϕ
g = ngN/
ρ
0V and
ϕ
c = ncN/
ρ
0V,
respectively.
In the SCFT one considers the statistics of a single copolymer chain in a set of effective external
fields wi, where i represents block species A, B, C or grafted polymers. These external fields, which
represent the actual interactions between different components, are conjugated to the segment density
Self Assembly of ABC Triblock Copolymer Thin Films on a Brush-Coated Substrate 585
fields,
φ
i, of different species i. Hence, the free energy (in unit of kBT) of the system is given by
]
2
1
[d/
)]([d/)/ln()/ln(
s
,S
gggccc
∑∑
+
++=
jii
iijiij
ii
iii
NNV
wVVQVQF
rr
r
r
δφχφφχ
φξφϕϕϕϕ
1
11
(1)
where
χ
ij is the Flory-Huggins interaction parameter between species i and j,
ξ
is the Lagrange multiplier
(as a pressure),
χ
iS is the interaction parameter between the species i and the free hard surface S. rs is the
position of the free hard surface. ),(d cc 1rrqQ
= is the partition function of a single copolymer chain in
the effective external fields wA, wB and wC, and ),(d gg 1rrqQ
= is the partition function of a grafted
polymer chain in the external field wg. The fundamental quantity to be calculated in mean-field studies is
the polymer segment probability distribution function, q(r, s), representing the probability of finding
segment s at position r. It satisfies a modified diffusion equation using a flexible Gaussian chain model
),(),(
6
),( 2
2swqsq
Na
sq
srrr =
(2)
where w is wA when 0 < s < fA, wB when fA < s < fA + fB, wC when fA + fB < s < 1 for ABC triblock
copolymer and wg for the grafted polymer. The initial condition of Eq. (2) satisfies qc(r, 0) = 1 for ABC
triblock copolymer. Because the two ends of the block copolymer are different, a second distribution
function ),(
csq r
+ is needed which satisfies Eq. (2) but with the right-hand side multiplied by 1 and the
initial condition 1)1,(
c=
+rq. The initial condition of qg(r, s) for grafted polymer is qg(r = rz, 0) = 1 and
qg(r rz, 0) = 0, where rz presents the position of the substrate at z = 0, and that of ),(
gsq r
+
is 1)1,(
c=
+rq. The periodic boundary condition is used for qc(r, s), ),( sqcr
+, qg(r, s), and ),(
gsq r
+ along
x-direction when ],0[ z
Lz . qc(r, s), ),( sqcr
+, qg(r, s), and ),( sqgr
+ are equal to zero when z < 0 or
z > Lz.
Minimization of the free energy with respect to density, pressure, and fields, δF
φ
= δF/δ
ξ
=
δF/δw
= 0, leads to the following equations.
++=
A
,ASAA s
)()()(
i
ii NNw rr
rrr
δχξφχ
(3)
++=
B
,BSBB s
)()()(
i
ii NNw rr
rrr
δχξφχ
(4)
++=
C
,CSCC s
)()()(
i
ii NNw rr
rrr
δχξφχ
(5)
++=
g
,gSgg s
)()()(
i
ii NNw rr
rrr
δχξφχ
(6)
φ
Α(r) +
φ
Β(r) +
φ
C(r) +
φ
g(r) = 1 (7)
+
=A
0cc
c
c
A),(),(d)( fsqssq
Q
Vrrr
ϕ
φ
(8)
Z.B. Jiang et al.
586
++
=BA
A
),(),(d)( cc
c
c
B
ff
fsqssq
Q
Vrrr
ϕ
φ
(9)
+
+
=1
BA
),(),(d)( cc
c
c
Cff sqssq
Q
Vrrr
ϕ
φ
(10)
+
=1
0gg
g
g
g),(),(d)( sqssq
Q
Vrrr
ϕ
φ
(11)
Here, we solve Eqs. (3)−(11) directly in real space by using a combinatorial screening algorithm proposed
by Drolet and Fredrickson[38, 39]. The algorithm consists of randomly generating the initial values of the fields
wi(r). By using a Crank-Nicholson scheme and an alternating-direct implicit (ADI) method[40], the diffusion
equations are then integrated to obtain q and q+, for 0 < s < 1. Next, the right-hand sides of Eqs. (8)−(11) are
evaluated to obtain new values for the volume fractions of blocks A, B, C, and grafted polymers.
The simulations are performed on a two-dimensional square grid Lx × Lz. The walls are presented as lines at
z = 0 (substrate) and Lz (hard surface) and Lx = 100 (in unit of a) along x-direction in all simulations. The
film thickness is Lz. And we can vary Lz to change the thickness of the thin film. The simulation is carried
out until the phase patterns are stable and the free energy difference between two iterations is smaller than 105,
i.e., ΔF < 105. It should be noted that the resulting microphases largely depend on the initial conditions.
Therefore, all the simulations are repeated at least 10 times using different random states to guarantee the
structure is not occasionally observed. We only address the thin films of ABC triblock copolymer on the
densely polymer-grafted substrate and the grafted polymers are assumed to be identical with the middle block
B. The grafting density of the grafted chains is set to
σ
= 0.2 to insure that the polymer brush is in the
dry brush regime (
σ
N1/2 > 1) and the pattern of the ABC triblock copolymer film is perfect. The
polymerization of ABC triblock copolymer is N = 100 and that of the grafted chains is same with the
copolymers, i.e., P = N = 100. The lamellar phase with compositions fA = 0.18, fB = 0.39 and fC = 0.43 in
melts[41] is studied, which is similar with the composition in experiments[35]. The interaction parameters
between different components are
χ
ijN = 35. In this work, we focus on two cases: the hard surface (opposite
to the substrate) is (1) neutral and (2) good for the middle block B.
RESULTS AND DISCUSSION
Morphology of the ABC Triblock Copolymer
Figure 1 presents the morphologies of the ABC block copolymer by increasing the film thickness with the
neutral hard surface. The morphology of the coated polymer is shown with the black color under the
morphology of ABC triblock copolymer. The microphase patterns, displayed in the form of density, are the gray,
white and black, assigned to A, B and C, respectively. Due to the higher grafting density of the coated polymer,
the brush height is h =
σ
Pa[42], therefore, the effective thickness of the ABC block copolymer is only about
hLL zz =
eff . When the film thickness is very small, such as )22.1(25 g
eff RLL zz == , the end blocks A and C
form the alternating cylinders in the phase consisting of the middle block B and the coated polymer B. The
vertical lamellar phase occurs when the film thickness is )18.3(33 g
eff RLL zz == . With the film thickness
increasing to )67.3(35 g
eff RLL zz == in Fig. 1(c), the short end block A stays at the upper hard interface, so,
the lamellar phase forms near the hard surface. The block B forms lamellae on the coated surface due to the
thinner film thickness. At the time, the block A which is adjacent to block B cannot separate from the block
copolymer, so they form one phase. Then only the lamellar phase occurs in this case. With further increasing the
film thickness to )16.4(37 g
eff RLL zz == as shown in Fig. 1(d), the short block A also forms lamellae parallel
Self Assembly of ABC Triblock Copolymer Thin Films on a Brush-Coated Substrate 587
to the hard surface. But at the polymer coated surface, the dispersed A cylindrical phase occurs owing to the
coated polymer and the block B forming the majority. The domain size of the A cylinder phase near the coated
surface becomes large with the film thickness increasing until the lamellar phase near the coated surface occurs,
such as from Lz = 37 to 48. Another period will form when further increasing the film thickness, for example, Lz
= 50 to 60 (not shown). The lamellar phase will always form near the hard surface but the size of the cylinder
phase A near the coated surface enlarges as the film thickness increases.
(a)
(b)
(c)
(d)
(e)
(f)
Fig. 1 Morphologies of the ABC triblock copolymer confined between the polymer coated surface and the
neutral free hard surface for fA = 0.18, fB = 0.39, fC = 0.43 at
χ
ABN =
χ
ACN =
χ
BCN = 35,
χ
ASN =
χ
BSN =
χ
CSN = 0 with different film thicknesses
a) )22.1(25 g
eff RLL zz == ; b) )18.3(33 g
eff RLL zz == ; c) )67.3(35 g
eff RLL zz == ;
d) )16.4(37 g
eff RLL zz == ; e) )9.4(40 g
eff RLL zz == ; f) )86.6(48 g
eff RLL zz == ;
Gray, white and black phases correspond to blocks A, B and C, respectively; The morphology of the
polymer coated substrate is shown in black under the morphology of the ABC triblock copolymer.
When the upper surface is good for B, such as
χ
BSN = 0 and
χ
ASN =
χ
CSN = 20, the self assembly of the
ABC triblock copolymer changes a lot although the two surfaces (one is the polymer coated surface, the other is
the free hard surface) are good for the block B. Figure 2 shows the morphologies of the ABC block copolymer
with different film thicknesses. The morphology of the coated polymer is shown with the black color under the
morphology of ABC triblock copolymer. The morphologies are more complex and show much difference
compared with the ABC block copolymer thin films in a slit[30], where the perpendicular lamellar phase easily
forms. The block B of the copolymers and the coated polymers B form one phase when the film thickness is
x
z
Z.B. Jiang et al.
588
(a)
(b)
(c)
(d)
(e)
(f)
(g)
(h)
(i)
Fig. 2 Morphologies of the ABC triblock copolymer confined between the polymer coated surface and the selective
free hard surface for fA = 0.18, fB = 0.39, fC = 0.43 at
χ
ABN =
χ
ACN =
χ
BCN = 35,
χ
BSN = 0,
χ
ASN =
χ
CSN = 20 with
different film thicknesses
a) )96.1(28 g
eff RLL zz == ; b) )45.2(30 g
eff RLL zz == ; c) )18.3(33 g
eff RLL zz == ;
d) )67.3(35 g
eff RLL zz == ; e) )41.4(38 g
eff RLL zz == ; f) )86.6(48 g
eff RLL zz == ;
g) ).( g
eff RLL zz 35750 == ; h) )08.8(53 g
eff RLL zz == ; i) )80.9(60 g
eff RLL zz == ;
Gray, white and black phases correspond to blocks A, B and C, respectively; The morphology of the polymer coated
substrate is shown in black under the morphology of the ABC triblock copolymer.
Self Assembly of ABC Triblock Copolymer Thin Films on a Brush-Coated Substrate 589
small )96.1(28 g
eff RLL zz == as shown in Fig. 2(a). The end blocks A and C form alternating phases along the
film, which is consistent with the experimental results[35]. With the film thickness increasing to
)45.2(30 g
eff RLL zz == in Fig. 2(b), the short block A will dissolve into the adjacent block B, and the lamellar
phase forms. When )18.3(33 g
eff RLL zz == in Fig. 2(c), the block B stays on the free surface and the block A
forms the cylinder phase immersed in the B phase near the free surface. The block B in the copolymers and the
coated polymers B form one phase on the other brush coated surface. So the cylindrical phase A forms between
the two surfaces and distributes on the lamellar phase C. With the film thickness further increasing, the
cylindrical phase A on the upper hard surface forms and the domain size of the cylindrical phase A increases a
little, but the domain of the phase A on the other brush coated surface clearly increases, which is shown in
Figs. 2(d) and 2(e). When the film thickness comes to )86.6(48 g
eff RLL zz == , another period begins, see
Figs. 2(f)−2(i). The cylindrical phase A on the hard surface and on the polymer coated surface enlarges with the
film thickness increasing. We can vary the film thickness of the film to obtain the different domain size of the
cylindrical A phase on the two sides of the film. Although the two surfaces (one is the hard surface and the other
is the polymer brush coated surface) of the film is good for the middle block B of the ABC triblock copolymer,
the vertical lamellae phase is not always stable compared with the ABC triblock copolymer film confined
between the two hard surfaces which are good for the middle block B[30].
In general, the vertical lamellar phase is stable when the effective thickness is very small for the case of the
neutral hard surface; the short end block will accumulate near the neutral hard surface when the effective film
thickness is greater than (34)Rg. But the middle block B accumulates on the selective surface for block B,
which leads to the block A forming the cylindrical phase immersed in the B phase near the free hard surface.
The phase A forms the cylinder at the polymer coated surface, and the domain size will vary with the film
thickness. The lamellar phase parallel to the film always forms in the middle. The number of the periods
increases with the film thickness increasing.
When the cylinder phase A near the brush coated surface occurs, we can note the rippled structure of ABC
block copolymer near the polymer grafted surface due to the soft surface of grafted polymer, see Figs. 1(d), 1(e),
2(d), 2(e) and 2(h). The corresponding structure of the grafted polymer is also rippled. In order to clearly see the
rippled distribution of the grafted polymers, we give the distribution sum of the grafted polymers along the z-
direction as a function of x, which is shown in Fig. 3. From the figure, we can see that the distribution of the
grafted polymers will change a lot near the interface between the block copolymers and the grafted polymers to
Fig. 3 Density sum of the grafted polymers along z-direction as a function of x for fA = 0.18,
fB = 0.39 and fC = 0.43 at
χ
ABN =
χ
ACN =
χ
BCN = 35
a) Neutral free hard surface,
χ
ASN =
χ
BSN =
χ
CSN = 0; b) Good for block B,
χ
BSN = 0,
χ
ASN =
χ
CSN = 20; The density sum is the integral in the z-direction of the volume fraction of the
polymer brush for a fixed x.
Z.B. Jiang et al.
590
minimize the interfacial energy. The structure of the grafted polymers can be tailored with the film thickness
varying. It can be the flat or rippled structure. The rippled structure in the brush forms when the cylindrical
phase A is near the interface between the block copolymers and the grafted polymers, which is also observed in
the AB diblock copolymer confined between the polymer brush-coated surfaces[36, 37]. The results show that the
grafted polymers serve as a soft surface for self assembly of the ABC triblock copolymer. The frustration that
occurs at the non-tailored surface disappears or decreases in this case.
Free Energy Analysis
In the following, we provide insight into the entropic and the enthalpic free energies of the system. Figure 4
shows the entropic free energy TS (spheres), the enthalpic free energy U (squares) and the total free energy F
(triangles) (F = U
TS) of the system as a function of the reduced film thickness g
eff /RLz. With the film
thickness decreasing, the entropic free energy and the enthalpic free energy increase. But there are shallow
sharps when one period transfers to another period, such as g
eff )32( RLz , 7.5Rg, 11Rg for the neutral free
surface and g
eff RLz2, 6Rg, 10Rg for the selective free surface for the middle block B. Due to the grafted
polymer serving as a hydrophilic part (identical with the block B) of the block copolymer, the large energy
fluctuation is suppressed in our system as compared with the block copolymer confined in the film only by hard
surfaces.
But if the free hard surface is selective for the end block A or C, the lamellar phase forms on the hard
surface. The cylindrical phase A is still stable on the brush-coated surface.
Fig. 4 Enthalpic free energy U (squares), entropic free energy TS (spheres), and total free energy F
(triangles) of the system for fA = 0.18, fB = 0.39 and fC = 0.43 at
χ
ABN =
χ
ACN =
χ
BCN = 35
a)
χ
ASN =
χ
BSN =
χ
CSN = 0; b)
χ
BSN = 0,
χ
ASN =
χ
CSN = 20
CONCLUSIONS
The self assemblies of ABC triblock copolymer thin films on a densely brush-coated substrate were investigated
by using the self-consistent field theory. The coated polymers are identical with the middle block B of the ABC
triblock copolymer. The middle block B and the coated polymer form one phase together and the alternating
phase A and phase C occur when the film is very thin either for the neutral or selective upper hard surface. The
lamellar phase parallel to the surfaces is stable on the neutral hard surface. Interestingly, the short block (A in
the paper) always stays on the free hard surface when the effective film thickness larger than (34)Rg. But when
the hard surface is good for the middle block B, the phase A forms cylinders staying on the interface between
the phase B and phase C near the free surface. The cylindrical phase A easily forms on the polymer coated
surface for either neutral or selective hard surfaces. The cylindrical phase can be controlled, and the domain size
can be tailored by varying the width of the slit. The rippled structure forms when the cylindrical phase exists
Self Assembly of ABC Triblock Copolymer Thin Films on a Brush-Coated Substrate 591
near the interface between grafted polymers and block copolymers. Due to the existence of the hydrophilic
polymer coated surface, the large energy fluctuation existing in the film confined by two hard surfaces
disappears. These observations demonstrate that enthalpic interactions, in addition to entropic considerations,
can play a major role in forming the above complex morphologies.
The results are helpful for designing the functional nanopattern of the film. For example, the short
functional block A can be added to the diblock BC block copolymer, and the functional surface of the short
block A on the film can easily forms in the case for the neutral hard surface. When the hard surface is good for
the middle block B, the functional beads coming from the short block A occur on the surfaces and their size can
be controlled by tailoring the film thickness. Therefore, it is very useful for designing and realizing the
functional thin films besides the structural ones by adding the functional short block A to the BC diblock
copolymer.
REFERENCES
1 Hamley, I.W., "The Physics of Block Copolymers", Oxford University Press, New York, 1998
2 Balsamo, V., Collins, S. and Hamley, I.W., Polymer, 2002, 43(15): 4207
3 Matsen, M.W., J. Chem. Phys., 1998, 108(2): 785
4 Jiang, Y., Huang, R. and Liang, H.J., J. Chem. Phys., 2005, 123(12): 124906
5 Srinivas, G., Discher, D.E. and Klein, M.L., Nat. Mater., 2004, 3(9): 638
6 Srinivas, G., Shelley, J.C., Nielsen, S.O., Discher, D.E. and Klein, M.L., J. Phys. Chem. B, 2004, 108(24): 8153
7 Glass, R., Moller, M. and Spatz, J.P., Nanotechnology, 2003, 14(10): 1153
8 Sun, R.G., Wang, Y.Z., Wang, D.K., Zheng, Q.B., Kyllo, E.M., Gustafson, T.L., Wang, F.S. and Epstein, A.J., Synthetic
Met., 2000, 111: 595
9 Wurlitzer, A., Politsch, E., Huebner, S., Kruger, P., Weygand, M., Kjaer, K., Hommes, P., Nuyken, O., Cevc, G. and
Losche, M., Macromolecules, 2001, 34(5): 1334
10 Muthukumar, M., Curr. Opin. Colloid & Interface Sci., 1998, 3(1): 48
11 Morkved, T.L., Lu, M., Urbas, A.M., Ehrichs, E.E., Jaeger, H.M., Mansky, P. and Russell, T.P., Science, 1996, 273: 931
12 Boker, A., Muller, A.H.E. and Krausch, G., Macromolecules, 2001, 34(21): 7477
13 van der Meer, P.R., Meijer, G.C.M., Vellekoop, M.J., Kerkvliet, H.M.M. and van den Boom, T.J.J., Sensor Actuat. A-
Phys., 1998, 71(1-2): 27
14 Wolff, M., Scholz, U., Hock, R., Magerl, A., Leiner, V. and Zabel, H., Phys. Rev. Lett., 2004, 92(25): 255501
15 Ionov, L., Minko, S., Stamm, M., Gohy, J.F., Jerome, R. and Scholl, A., J. Am. Chem. Soc., 2003, 125(27): 8302
16 Tsori, Y. and Andelman, D., Macromolecules, 2001, 34(8): 2719
17 Harrison, C., Adamson, D.H., Cheng, Z.D., Sebastian, J.M., Sethuraman, S., Huse, D.A., Register, R.A. and
Chaikin, P.M., Science, 2000, 290(5496): 1558
18 Pereira, G.G. and Williams, D.R.M., Langmuir, 1999, 15(6): 2125
19 Pereira, G.G. and Williams, D.R.M., Macromolecules, 1998, 31(17): 5904
20 Pereira, G.G. and Williams, D.R.M., Phys. Rev. Lett., 1998, 80(13): 2849
21 Ludwigs, S., Schmidt, K., Stafford, C.M., Amis, E.J., Fasolka, M.J., Karim, A., Magerle, R. and Krausch, G.,
Macromolecules, 2005, 38(5): 1850
22 Luzinov, I., Minko, S., and Tsukruk, V.V., Prog. Polym. Sci., 2004, 29(7): 635
23 Peters, R.D., Yang, X.M., and Nealey, P.F., Macromolecules, 2002, 35(5): 1822
24 Walton, D.G., Soo, P.P., Mayes, A.M., Allgor, S.J.S., Fujii, J.T., Griffith, L.G., Ankner, J.F., Kaiser, H., Johansson, J.,
Smith, G.D., Barker, J.G., and Satija, S.K., Macromolecules, 1997, 30(22): 6947
25 Mansky, P., Russell, T.P., Hawker, C.J., Mays, J., Cook, D.C., and Satija, S.K., Phys. Rev. Lett., 1997, 79(2): 237
26 Morkved, T.L. and Jaeger, H.M., Europhys. Lett., 1997, 40(6): 643
27 Geisinger, T., Muller, M. and Binder, K., J. Chem. Phys., 1999, 111(11): 5251
Z.B. Jiang et al.
592
28 Huinink, H.P., Brokken-Zijp, J.C.M., van Dijk, M.A. and Sevink, G.J.A., J. Chem. Phys., 2000, 112(5): 2452
29 Rasmussen, K.O., J. Polym. Sci. Part B-Polym. Phys., 2004, 42(20): 3695
30 Pickett, G.T. and Balazs, A.C., Macromol. Theor. & Simul., 1998, 7(2): 249
31 Feng, J. and Ruckenstein, E., Polymer, 2002, 43(21): 5775
32 Chen, H.Y. and Fredrickson, G.H., J. Chem. Phys., 2002, 116(3): 1137
33 Ludwigs, S., Krausch, G., Magerle, R., Zvelindovsky, A.V. and Sevink, G.J.A., Macromolecules, 2005, 38(5): 1859
34 Knoll, A., Lyakhova, K.S., Horvat, A., Krausch, G., Sevink, G.J.A., Zvelindovsky, A.V. and Magerle, R., Nat. Mater.,
2004, 3(12): 886
35 Fukunaga, K., Hashimoto, T., Elbs, H. and Krausch, G., Macromolecules, 2002, 35(11): 4406
36 Ren, C.L. and Ma, Y.Q., Phys. Rev. E, 2005, 72(5): 051804
37 Ren, C.L., Chen, K. and Ma, Y.Q., J. Chem. Phys., 2005, 122(15): 154904
38 Drolet, F. and Fredrickson, G.H., Phys. Rev. Lett., 1999, 83: 4317
39 Drolet, F. and Fredrickson, G.H., Macromolecules, 2001, 34: 5317
40 Press, W.H., Flannery, B.P., Teukolsky, S.A. and Vettering, W.T., "Numerical Recipes", Cambridge University Press,
Cambridge, 1989
41 Tang, W.H., Macromolecules, 2000, 33(4): 1370
42 Aubouy, M., Fredrickson, G.H., Pincus, P. and Raphael, E., Macromolecules, 1995, 28: 2979
... The phase behavior of diblock copolymers melts confined in a parallel slit or in a thin film has been extensively studied [15][16][17][18][19][20][21][22][23][24][25]. When the number of distinct blocks increases from two, i.e., ABC triblock copolymer, the complexity and variety of self-assembled structures are increased dramatically [1,[26][27][28][29][30][31][32][33][34][35][36][37][38][39]. If a surface or interface exists, the microdomain morphologies and the kinetics of microdomain ordering can change significantly. ...
... The previous work mainly concentrated on phases of several compositions of ABC triblock copolymer by varying the film thickness or the interfacial interaction. As we know, the polymer brush-coated surface is good from the energy view [30,31]. It is equivalent to changing the surface-polymer interaction as polymer brush acts as a soft surface [30,31,61,62]. ...
... As we know, the polymer brush-coated surface is good from the energy view [30,31]. It is equivalent to changing the surface-polymer interaction as polymer brush acts as a soft surface [30,31,61,62]. Experimentally, random copolymers were used to control the wetting behavior of block copolymer [63,64]. ...
Article
Full-text available
The morphology and the phase diagram of ABC triblock copolymer thin film directed by polymer brushes are investigated by the self-consistent field theory in three dimensions. The polymer brushes coated on the substrate can be used as a good soft template to tailor the morphology of the block copolymer thin films compared with those on the hard substrates. The polymer brush is identical with the middle block B. By continuously changing the composition of the block copolymer, the phase diagrams are constructed for three cases with the fixed film thickness and the brush density: identical interaction parameters, frustrated and non-frustrated cases. Some ordered complex morphologies are observed: parallel lamellar phase with hexagonally packed pores at surfaces (LAM3 ll -HFs), perpendicular lamellar phase with cylinders at the interface (LAM⊥-CI), and perpendicular hexagonally packed cylinders phase with rings at the interface (C2⊥-RI). A desired direction (perpendicular or parallel to the coated surfaces) of lamellar phases or cylindrical phases can be obtained by varying the composition and the interactions between different blocks. The phase diagram of ABC triblock copolymer thin film wetted between the polymer brush-coated surfaces is very useful in designing the directed pattern of ABC triblock copolymer thin film.
... To decrease the tension, the size of the micelles increased and their morphology changed to a shape decreasing the free energy of the whole system (Ian & Liu, 2013a). Although the shrinking of the contact area between the solvent and the core caused by the collapse of Pt BA corresponded with the enthalpy decrease, it was also accompanied by a loss of entropy (Jiang et al., 2009). When the enthalpy change did not compensate the loss of entropy, the micelles were close to each other or adhered together to form micellar strings decreasing the free energy. ...
Article
Linear ABC triblock copolymer PtBA154-b-PS300-b-P2VP240 was successfully synthesized by RAFT polymerization. Block copolymer micelles were prepared by the two-step hierarchical selfassembly process. Size exclusion chromatography and 1H NMR were used to characterize the structure of samples. Morphologies and size of micelles were determined by transmission electron microscope. The results showed that the densely dispersed spherical micelles of PtBA154-b-PS300-b-P2VP240 were obtained in the first step of the hierarchical self-assembly process. In the second step, core-compartmentalized micelle strings with different lengths and distribution densities were obtained when the primary self-assembled solution was dialyzed in distilled water with pH ≈ 3. When distilled water with pH ≈ 3 was added drop-wise to this solution, uniformly dispersed spherical core-compartmentalized micelles of PtBA154-b-PS300-b-P2VP240 were prepared. Thus, hierarchical self-assembly structure of linear ABC triblock copolymer was obtained successfully and the preparation of uniformly dispersed spherical micelles of triblock copolymers was realized simply by changing the secondary self-assembly methods.
... It can greatly affect conformation, [5][6][7][8][9][10][11] glass transition [12][13][14][15] of polymers, even the self-assembly of block copolymers. [16][17][18][19][20][21] The conformational properties of polymer glasses on planar surfaces are widely investigated experimentally and theoretically. [10][11][12] Even the curvature of the surface/interface is an important factor to change the phase behavior of polymers. ...
Article
Full-text available
Structural properties of polymers confined in nanocylinders are investigated by Monte Carlo simulation, which is successfully used to consider the conformational property of constrained polymers. The conformational properties of the polymers close to the walls exhibit different features. The density profiles of polymers are enhanced near the wall of the nanocylinder, which shows that the packing densities differ near the wall and far from the wall. The highest densities near the wall of the nanocylinder decrease with increasing radius of the nanocylinder. Furthermore, the density excess is not only near the wall of the nanocylinder, but also shifts to the center of the nanocylinder at lower temperatures. The radius of gyration and the bond length of polymers in the nanocylinder show that the polymer chains tend to extend along the axis of the nanocylinder in highly confined nanocylinder and contract at lower temperature. Our results are very helpful in understanding the packing induced physical behaviors of polymers in nanocylinders, such as glass transition, crystallization, etc.
Article
The confined environment plays a very important role in the phase separation of copolymers, which can change bulk phase behaviors of copolymers. The different confinement conditions can induce the formations of various interesting and novel morphologies, which can be used in a variety of nanotechnology applications such as high-density medium storage, nanolithography and photonic crystals. The grafting of polymers to confined surfaces is an efficient means for tailoring surface properties. In this work, we investigate the effect on architecture of the AB diblock copolymer confined between mixed brush-grafted surfaces by using self-consistent field theory. The brush contains two types of homopolymers. We study the effects of the fraction of A block, grafted period and the volume fraction of the polymer brush, the distance between two surfaces and the interaction strength between two blocks on the morphology. 1) With the increase of the fraction of A block (fA), the phase morphology changes from the A-block hexagonal cylinder to the parallel lamellae, to the curving lamellae, and then to the B-block hexagonal cylinder. The period of hexagonal cylinder and curving lamellae is equal to the grafted period of the polymer brush due to the influence of the polymer brush. 2) The grafted period of polymer brush is a very important factor for the morphology of diblock copolymer. When fA=0.3, we change the grafted period of the polymer brush. We obtain the phase transition from the hexagonal cylinder to the alternating phase of tetragonal and hexagonal cylinder, then to the alternating phase of tetragonal and octagonal cylinder. When fA=0.4, the structure changes from the hexagonal cylinder to the order phase of the waving lamellae and cylinder with the increase of the grafted period of the polymer brush. Compared with the single homopolymer brush system, the mixed brush enlarges the range of ordered phase and reduces the range of disordered phase. Block copolymers are prone to forming cylinder in mixed brush system and tending to form lamellae in single homopolymer brush system. 3) When fA=0.3, we obtain the phase transition from the hexagonal cylinder to the one-layered cylinder phase by increasing the volume fraction of the polymer brush. This transition is different from that of the single homopolymer brush system. Interestingly, when fA=0.45, the structure of AB block copolymer changes from the parallel lamellae to the perpendicular lamellae with the increase of the volume fraction of the polymer brush. The entropic energy plays an important role in this transition process. Similarly, we also observe the phase transition from the parallel lamellae to the perpendicular lamellae by decrease the distance between two surfaces. 4) We construct the phase diagram for a range of the fraction of A block and the interaction strength. The results provide an effective approach to obtaining the desired microstructures for fabricating nanomaterials.
Article
Full-text available
An extensive and systematical calculation was performed to explore hierarchical cylindrical structures and the order-to-order transitions of AB diblock copolymers (fA = 0.3) on a saw-toothed substrate using self-consistent mean-field theory. We obtained fifteen relatively simple morphologies, including the existing morphologies observed experimentally and from simulations, and five more complicated structures, by varying the lateral periodicity of the substrate, the film thickness of diblock copolymers, the interaction between A-block and the substrate and the tilt angles (or the shape) of the substrate. These structures show that the orientation and number of layers of cylinders can be tailored. Even lamellae and spherical microdomains were observed. Most interestingly, hierarchical structures are also observed, such as the morphology of Cab// within the upper cylinder perpendicular to the bottom cylinder, SCb// morphology that the upper is cylinder but the bottom is sphere. In addition, we discussed these complex hierarchical structures in detail and have analyzed the order-to-order transitions between the cylindrical morphologies with distinct orientations and layers.
Article
Linear diblock (SB) styrene-butadiene block copolymer, linear triblock [(SB)2] and star-shaped four-arm triblock [(SB) 4] poly(styrene-block-butadiene-block-styrene) copolymers were respectively dissolved in tetrahydrofuran (THF) and cyclohexane (CH), and then their intrinsic viscosities were measured. The obtained viscosity radii (R-eta) of the polymers in those solvents were calculated by a modified Einstein equation. The radii of gyration (R-g) of the polymers dissolved in each solvent under different temperatures were obtained from static laser light scattering measurements. Due to the similarity between R-n and hydrodynamic radius (R-h), the characteristic parameter rho(= R-g/R-h), which can be used to describe the morphology of a polymer chain in solution, can be calculated by rho = R-g/R-eta. The p value of (SB)4 in THF is 1.51,but is 0.76 in CH, showing flexible chain and hard sphere morphology, respectively. The rho values of (SB)2 are 0.97 and 0.81 in THF and CH, which are located in the region of hollow sphere morphology. However, the rho values of SB are greater than 1, showing flexible chain morphology in both solvents. Therefore, the molecular architecture of a polymer has great influence on the morphology.
Article
Self-assembly of supramolecular AB diblock copolymer/C homopolymer mixtures in selective solvents was examined by real-space self-consistent field theory. The non-covalent bond is formed between the free end of hydrophobic B block of amphiphilic AB diblock copolymer and one end of hydrophobic C homopolymer,, and thereby the mixtures consist of unbonded AB diblock copolymers, unbonded C homopolymers and supramolecular ABC triblock copolymers. The constitutions of the mixtures are dependent on the non-covalent bonding strength and the initial AB/C mixing ratio, which are essential to the morphologies of the self-assembled micelles. The AB diblock copolymer/C homopolymer mixtures are self-assembled into a variety of micelles such as core-shell-corona micelles and wormlike micelles in selective solvents. With increasing the bonding strength, the volume fraction of supramolecular ABC triblock copolymers increases, and that of unbonded AB diblock copolymers and C homopolymers decreases. In this process, transitions from core-shell-corona micelles to wormlike micelles were observed. In the core-shell-corona micelles, the bonded C homopolymers are distributed at the hydrophobic/hydrophilic interface, while the unbonded C homopolymers are dispersed in solvents due to their weak hydrophobic property. In the wormlike micelles, most bonded C homopolymers localize between two cores in one micelle, while the unbonded C homopolymers transit to the center of C domains. With increasing the volume fraction of homopolymers, the volume fraction of supramolecular ABC triblock copolymers increases to a maximum and then decreases. The morphological transitions of these micelles were also observed as a function of the volume fraction of homopolymers. The work provided an insight into the effect of non-covalent bonds on the morphology of supramolecular micelles.
Article
Polymeric self-consistent field theory is used to investigate microstructures and interphase properties of diblock copolymers grafted onto solid surfaces in a homopolymer melt. The calculations show that the grafted diblock copolymers can self-assemble into hemispherical microstructures at low grafting densities of the diblock copolymers. The morphology transforms into hemicylinder-like and sandwich-like lamellar microstructures with an increase in the chain-grafting density. The effective thickness of the grafted block layer and the interphase width between the homopolymer melt and the grafted copolymers strongly depend on the physicochemical parameters of the system, such as the composition of the grafted copolymer, the chemical incompatibility between the different components, the length ratio of grafted copolymer to homopolymer, and the grafting density of the diblock copolymers. In addition, the above computational results of microphase-separated structures and interphase properties are qualitatively compared with our previous experimental observations. The comparison indicates that our theoretical results not only reproduce the general feature of the experimental observations, but also elucidate the internal structural information and complement the findings in the region of high grafting densities of diblock copolymers.
Article
The morphology transition of binary mixtures of polystyrene-block-poly(butadiene)-block-poly(2-vinylpyridine)(SBV) triblock and polystyrene (PS) homopolymer thin films was investigated as a function of the volume fraction of added homopolymer and the annealing time in benzene vapor. It was found that the weight ratio of PS in the blends influenced the transition process. When PS content was >5%, the order-order transition (OOT) of core-shell cylinders (C) sphere in “diblock Gyroid” (sdG) sphere in lamella (sL) sphere (S) was observed, which was similar to ABC triblock copolymer except for the increased surface area of the PS phase. When PS content reached to 10–30%, the OOT in the sequence of C sL S was observed. The disappearance of the Gyroid phase is due to the change of the effective volume fraction. Further increasing the PS content, C phase also disappeared and sL S was expected to take place. © 2014 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys, 2014
Article
The phase behavior of coil-comb copolymers A–(Bm+1Cm) (side chain number m = 1–5) is investigated by real-space self-consistent field theory (SCFT). Depending on the copolymer composition and architecture, eight two-dimensional ordered phases are observed, including two-color lamellae (LAM2), three-color lamellae (LAM3), hexagonal cylinders (HEX), core–shell hexagonal phase (CSH), hexagon outside hexagonal phase (HEX2), two interpenetrating tetragonal phase (TET2), lamellae with beads inside (LAM + BD), and lamellae with core–shell beads (LAM3 + CSB). When the volume fractions are comparable, i.e., fA ≈ fB ≈ fC, LAM_3 phase is found to be stable for m = 1 while the hexagonal phases (core–shell hexagonal phase CSH or hexagon outside hexagonal phase HEX2) are stable if m > 1. The phase region of the hexagonal phases HEX, CSH or HEX2 enlarges with increasing m. For short coil length, such as fA = 0.1, the phase diagram is complex, especially when m = 1. For longer coil length, the lamellae become the dominant phase. The phase transition from lamellar phase to hexagonal phase is observed with the increase of the side chain length when the side chain number m is large, which is in agreement with the experimental results. Our results give a good way to tailor the phase behavior of block copolymer and are very useful to further study the hierarchical structure of the coil-comb block copolymer.
Article
Thin film miscible blends of poly(methyl methacrylate) (PMMA) and a branched random copolymer of methyl methacrylate and methoxy poly(ethylene glycol) monomethacrylate, P(MMA-r-MnG), were investigated by neutron reflectivity. The branched copolymer, which has a higher surface tension than PMMA, was nevertheless found to segregate to and completely cover both the surface and silicon substrate following annealing in 2000 Å thick films with ≥2 wt % P(MMA-r-MnG). This is in contrast to linear polyethylene oxide, which was depleted at both film interfaces when blended with PMMA and annealed. The reflectivity results were confirmed by contact angle studies, which indicate that the surfaces of P(MMA-r-MnG)/PMMA blends behave like that of pure P(MMA-r-MnG), resulting in a hydrophilic surface that is stable against dissolution in water-based environments. The branched hydrophilic additive is further shown to render PMMA resistant to protein adsorption and cell adhesion.
Article
We study the self-assembly in thin films of a lamella-forming polystyrene-block-poly(2-vinylpyridine)-block-poly(tert-butyl methacrylate) triblock copolymer by means of scanning force microscopy (SFM) and cross-sectional transmission electron microscopy (TEM). We follow the time development of the microdomain structure in the films during solvent vapor treatment. After preparation on a polar substrate preferentially attracting the middle block of poly(2-vinylpyridine), the films exhibit a spongelike microdomain morphology. With subsequent solvent vapor treatment, metastable lamellae with the spacing much smaller than the equilibrium value are first formed from free surface and grow toward the substrate and eventually establishes a multilayered structure throughout the film. As a result of this process, defects associated with the lamellar structure are annihilated from the free surface and accumulated close to the solid substrate. On an extended solvent vapor treatment, the lamellae thicken toward the equilibrium value. This lamella thickening is accompanied by autophobic dewetting of the thin film.
Article
This paper deals with the thermal and temperature-control aspects of a gas sensor based on surface-acoustic-wave (SAW) devices in a dual delay-line configuration. Application of smart temperature sensors combined with a micro-controller in a temperature-control system provides for a high-performance low-cost A/D conversion. An anti-windup control mechanism has been implemented to solve the problem of actuator saturation.
Article
The microstructure of six types of acrylic-based and hydrated cellulose-based carbon fibres of strengths from 1650 to 6120 MPa and elastic moduli from 97 to 228 GPa were studied by scanning electron microscopy (SEM). A ‘microcomposite’ structure of carbon fibres studied consisting of quasi-amorphous (‘matrix’) and orientated fibrillar carbon was revealed. This led to a new model of the fibre structure. The analysis of results of testing different carbon fibres defines the elastic modulus of ‘matrix’ carbon, and shows plastic drawing of fibrils. The model describes the properties of fibres and predicts ways to improve the fibre properties.
Article
We have used a dynamic density functional theory (DDFT) for polymeric systems, to simulate the formation of micro phases in a melt of an asymmetric block copolymer, AnBm(fA=1/3), both in the bulk and in a thin film. In the DDFT model a polymer is represented as a chain of springs and beads. A spring mimics the stretching behavior of a chain fragment and the spring constant is calculated with the Gaussian chain approximation. Simulations were always started from a homogeneous system. We have mainly investigated the final morphology, adopted by the system. First, we have studied the bulk behavior. The diblock copolymer forms a hexagonal packed array of A-rich cylinders, embedded in a B-rich matrix. Film calculations have been done by confining a polymer melt in a slit. Both the slit width and surface-polymer interactions were varied. With the outcomes a phase diagram for confined films has been constructed. Various phases are predicted: parallel cylinders (C&par;), perpendicular cylinders (C⊥), parallel lamellae (L&par;), and parallel perforated lamellae (CL&par;). When the film surfaces are preferentially wet by either the A or the B block, parallel oriented microdomains are preferred. A perpendicular cylindrical phase is stable when neither the A nor B block preferentially wets the surfaces. The predicted phase diagram is in accordance with experimental data in the literature and explains the experimentally observed differences between films of asymmetric block copolymers with only two parameters: the film thickness and the energetic preference of the surface for one of the polymer blocks. We have also observed, that confinement speeds up the process of long range ordering of the microdomains.
Article
Using a two-dimensional self-consistent field calculation, we determine the equilibrium morphology of thin films of ABC triblock copolymers confined between hard, smooth plates. The B segment is chosen to be the central block and all the blocks are incompatible. The chains microphase-segregate into a lamellar phase, with the stripes either perpendicular or parallel to the walls. When all the monomer-surface interactions are identical, the perpendicular orientation has the lowest free energy. When a repulsion is introduced between the surface and the A and C monomers. The surface interactions further stabilize the perpendicular orientation. At strong surface interactions, the morphology of the perpendicular structure is controlled by the overall thickness of the molten layer. In comparing diblocks to triblocks as candidates for forming laterally patterned films, our work indicates that triblocks possess distinct advantages over diblocks. First, no special effort needs to be taken to establish neutral surfaces. Second, the film does not have to be confined between two substrates. Thus, triblocks can be used to fabricate patterned polymer surfaces, which can be used for novel optical or electronic applications.
Article
We study a symmetric melt AB diblock copolymer thin film which is placed on a periodically striped CD surface. Because of different surface energies each stripe has a different preference for each of the blocks. There is thus a competition between the preferred bulk lamellar spacing, Lb, and the width of the stripes, lambda. This system is similar to the Frenkel-Kontorowa model of solid state physics. We consider the case Lb>=lambda, and assume perpendicular alignment. The periodic striped surface will induce lamellae with unequal spacings. We also show that the striped potential can induce inverted bilayers, i.e., an AB,AB pattern rather than the more usual AB,BA.
Article
The response of disordered P({ital d}-S-{ital b}-MMA) diblock copolymers to variable strength surface fields has been studied by neutron reflectivity. Surface interactions were controlled by end grafting P(S-{ital r}-MMA) random copolymers with various styrene contents onto Si substrates. The degree interfacial segregation of the block copolymer was proportional to the surface potential. A first-order transition in the degree of segregation was observed as the brush composition was varied. Conditions were found which yielded neutral boundary conditions {ital simultaneously} at the vacuum and substrate interfaces. {copyright} {ital 1997} {ital The American Physical Society}
Article
Monodisperse melts of ABC linear triblock copolymer are examined using self-consistent field theory (SCFT). Our study is restricted to symmetric triblocks, where the A and C blocks are equal in size and the A/B and B/C interactions are identical. Furthermore, we focus on the regime where B forms the majority domain. This system has been studied earlier using density functional theory (DFT), strong-segregation theory (SST), and Monte Carlo (MC), and it corresponds closely to a series of isoprene–styrene–vinylpyridine triblocks examined by Mogi and co-workers. In agreement with these previous studies, we find stable lamellar, complex, cylindrical, and spherical phases. In the spherical phase, the minority A and C domains alternate on a body-centered cubic lattice. In order to produce alternating A and C domains in the cylinder phase, the melt chooses a tetragonal packing rather than the usual hexagonal one. This amplifies the packing frustration in the cylinder phase, which results in a large complex phase region. Contrary to the previous evidence that the complex phase is double-diamond, we predict the gyroid morphology. The earlier theoretical results are easily rationalized and the experimental data are, in fact, more consistent with gyroid. © 1998 American Institute of Physics.
Article
Recent experiments have shown it is possible to produce striped surfaces with periodicities of the same order as those for self-organized diblock copolymer lamellae. Here we study theoretically the behavior of a thin film of AB diblocks on a CD striped surface. The competition between the preferred bulk lamellar spacing and the spacing favored by the stripes leads to several novel phases. We study this system by analytical and numerical means and in one limit by mapping it onto the Frenkel−Kontorowa model of solid state physics. The main conclusion is that the periodically striped surface can induce lamellae of unequal spacing. We also show that the surface can induce inverted bilayers where an AB is followed directly by another AB rather than a BA.