Anda di halaman 1dari 10

Article

pubs.acs.org/Macromolecules

Solution Self-Assembly of Blends of Crystalline-Coil


Polyferrocenylsilane-block-polyisoprene with Crystallizable
Polyferrocenylsilane Homopolymer
Graeme Cambridge, M. Jose Gonzalez-Alvarez, Gerald Guerin, Ian Manners,*,
and Mitchell A. Winnik*,

Department of Chemistry, University of Toronto, 80 St. George St., Toronto, ON, Canada M5S 3H6
School of Chemistry, University of Bristol, Bristol BS8 1TS, United Kingdom

S Supporting Information
*

ABSTRACT: The self-assembly of block copolymers in


solution leads to micellar structures with various morphologies.
One way to modify the morphology of these micelles is to
blend the block copolymer with a homopolymer corresponding to the core-forming block. Although the self-assembly of
blends of amorphous homopolymers and block polymers has
been extensively studied, there are few examples of solution
self-assembly of blends of a core-crystalline block copolymer
with a semicrystalline homopolymer. Here we describe a
systematic study of the assembly in decane of blends of a
polyferrocenylsilane-block-polyisoprene sample (PFS48-b-PI264)
with two dierent PFS homopolymer samples (PFS50 and PFS20). We examine the structures formed as a function of blend
composition and compare them to the structures formed from the individual components. PFS48-b-PI264 itself forms long
cylindrical micelles, while the two homopolymer samples form stacks of lamellar crystals. Self-assembly of block copolymer
mixtures leads to structures with an elongated planar core and ber-like protrusions from the ends. The details of the structure
vary in an interesting and systematic way as the ratio of homopolymer/block copolymer is increased, with important dierences
seen for the PFS50 and PFS20 homopolymer samples. This study demonstrates that cocrystallization plays a crucial role in
determining the structures formed from these mixtures.

INTRODUCTION
In the search for new materials with interesting properties and
novel morphologies, scientists have sometimes turned to blends
of block copolymers with homopolymers. The earliest examples
involved amorphous materials in the bulk state in which the
homopolymer had the same composition as one of the
components of the block copolymer.13 From these studies
emerged the important observation that if the homopolymer
was shorter in length than its counterpart in the block
copolymer, it was miscible in the block copolymer melt and
became incorporated into the microphase domains formed by
this component. Under these wet brush conditions, the
homopolymer added to the volume fraction of its component
in the system and often led to a change in morphology of the
system. In contrast, if the homopolymer was signicantly longer
than its counterpart in the block copolymer, it remained phase
separated in the blend. Because it had little inuence on the
morphology of the block copolymer component, this
morphology was referred to as dry brush.
Additional factors can aect the self-assembly of a block
copolymer in solution in the presence of a homopolymer
corresponding to the insoluble block.46 The rst example was
reported by Zhang and Eisenberg.4 They rst studied the self 2015 American Chemical Society

assembly of poly(styrene-b-acrylic acid) (PS-b-PAA) for a wide


range of PS-rich polymer compositions when water was added
to solutions of these polymers in dimethylformamide (DMF),
showing that as the content of the insoluble PS block increased,
the morphology of the micelles changed progressively from
crew-cut spherical micelles to cylinders to bilayers. These
experiments were then repeated in the presence of PS
homopolymer, dissolved in the copolymer/DMF solutions
before the addition of water. In the presence of homopolymer,
the size of the spherical particles increased and became more
polydisperse. In the case of the bilayers or cylinders, the
morphology changed from nonspherical to spherical micelles in
the presence of PS. The authors ascribe this change in
morphology to the reduced degree of stretching of the PS
blocks into the micelle core and also to the entropy gained
from mixing the homopolystyrene with the PS blocks in the
core.4
When one or more of the components of the block
copolymer or the homopolymer is a rigid-rod polymer or can
Received: November 11, 2014
Revised: December 19, 2014
Published: January 21, 2015
707

DOI: 10.1021/ma502279b
Macromolecules 2015, 48, 707716

Article

Macromolecules

In systems with crystalline-coil block copolymers and a


semicrystalline homopolymer, the introduction of a crystallizable block brings even more complex phase behavior.11 For
example, crystalline-coil block copolymers normally selfassemble in solution to form low curvature structures. These
are typically lamellae or elongated bers with cross-sectional
dimensions shorter than the fully extended length of the coreforming block, suggesting that the polymer chains fold as they
crystallize in the core. For some systems, uniform homopolymer or block copolymer single crystals can be obtained by a
self-seeding technique.
Chengs group12 used self-seeding in a polystyrene-selective
solvent to generate uniform lozenge-shaped thin platelet single
crystals of poly(L-lactic acid)-block-polystyrene (PLLA-b-PS)
with PLLA as the crystalline block and square-shaped platelet
single crystals of PEG-b-PS with PEG as the crystalline block.
These crystallites were used as seeds and added to supersaturated solutions of the corresponding PLLA or PEG
homopolymers. The homopolymer nucleated on the faceted
growth fronts and grew as a uniform thin crystal around each
block copolymer nucleus. In AFM images of the composite
structures, one could see the uniformity of the nucleated
homopolymer and the enhanced height of the central core due
to the presence of the PS corona.
Next, we turn our attention to PFS (polyferrocenyldimethylsilane) diblock copolymers. These block copolymers form
long cylindrical micelles in solvents selective for the corona
chains when the corona-forming chains are longer than the PFS
block. Examples include PFS-b-PI and PFS-b-PDMS (PDMS =
polydimethylsiloxane) in aliphatic hydrocarbon solvents such as
hexane or decane. Micelle growth is driven by the crystallization
of the PFS block. Spontaneous nucleation appears to be a rare
event, so that when additional unimer in a common good
solvent (such as tetrahydrofuran, THF) is added to a micelle
solution in hexane or decane, the newly added block copolymer
grows epitaxially o the ends of the existing micelles.13,14 Fiberlike PFS block copolymer micelles can also be nucleated from
the rough surface of a semicrystalline PFS homopolymer lm.15
There are two examples in the literature of hybrids formed
between micelles of PFS diblock copolymers and semicrystalline PFS homopolymer, both involving sequential addition of
the components. Yussof et al.16 reported that when a small
amount of a solution of PFS28 homopolymer in THF was
added to short (ca. 200 nm) monodisperse cylindrical micelles
of PFS54-b-PI324 in hexane, the homopolymer added to the ends
of the micelles, often linking the cylinders into elongated onedimensional structures. Under the same conditions, a longer
homopolymer, PFS65, crystallized preferentially at the ends of
the micelles. These crystals were larger and acted as a glue for
multiple seed cylinders, resulting in aggregated structures that
formed micelle networks.
More recently, Qiu et al.17 examined a dierent sequential
addition protocol, adding a small amount of PFS43 as a
concentrated solution in THF to hexane followed by an aliquot
of PFS53-b-PI637 block copolymer in THF. Hexane is a poor
solvent for PFS, and the homopolymer eventually precipitates.
On its own, PFS53-b-PI637 forms long ber-like micelles under
these conditions. When the block copolymer solution was
added quickly after adding the homopolymer, platelet micelles
were formed. On the other hand, if the block copolymer
addition was delayed by 24 h, then multiarm star-shaped
micelles were obtained, each characterized by four to eight
uniform cylindrical micelles emanating from a core that likely

crystallize during self-assembly, then the structures formed in


solution often depend sensitively on the protocol for sample
preparation. In some instances, one adds a solution of the block
copolymer to a suspension of the insoluble homopolymer in a
selective solvent for one of the block copolymer components.
In other instances, the homopolymer is added to a colloidal
solution of the self-assembled block copolymer. Alternatively, a
mixture of both components is subjected to self-assembly
conditions. For these coassemblies, one needs to know how the
homopolymer is incorporated into the block copolymer
structure. In addition, one would like to understand how the
sample preparation protocol controls the nature of the
structures formed. In the paragraphs below, we review some
interesting examples of these types of block copolymer
homopolymer assemblies in solution to set the context for our
work on the crystallization-driven solution coassembly of
polyferrocenyldimethylsilane-block-polyisoprene (PFS-b-PI)
block copolymer in the presence of PFS homopolymer.
We rst consider blends of rodcoil block copolymers with
rod-like homopolymers. Wang and co-workers recently
reported how mixtures of poly(-benzyl-L-glutamate) (PBLG)
and a PBLG-b-PEG diblock copolymer (PEG = poly(ethylene
glycol)) co-operatively self-assemble in water to form hybrid
helical rods and rings.5 These aggregates were prepared by a
selective precipitation method. First, PBLG130-b-PEG45 and
PBLG2200 homopolymers (the subscripts refer to the mean
degrees of polymerization) were dissolved in a mixed solvent of
THFDMF. Then water, a selective solvent for PEG, was
added, followed by dialysis against water to remove the organic
solvents. Hybrid helical superstructures were obtained with a
PBLG axis coated by PEG chains.
Another example involved solution self-assembly of mixtures
of regioregular poly(3-hexylthiophene) (P3HT 200 ) and
P3HT20-b-PEG108. Parks group7 showed that the type of
coassembly obtained depended sensitively on the solvent
quality. The authors rst formed long nanobers of P3HT in
anisole and then added this solution to a thin dry lm of
P3HT20-b-PEG108. After 20 min, the mixture was diluted with
either methanol or water, both selective solvents for the PEG
chains. After stirring overnight, the anisole was removed under
a low ow of nitrogen. The coassembly of the P3HT20-bPEG108 and P3HT200 nanobers in methanol led to bundled
nanobers encapsulated in P3HT20-b-PEG108. When water was
used instead of methanol, shish-kebab structures were obtained.
This interest in homopolymer and diblock copolymer blends
in solution has recently drawn the attention of the theoretical
chemists.810 Fengs group used two-dimensional self-consistent eld calculations to examine the self-assembly of
homopolymerdiblock copolymer blends in a selective
solvent.10 The groups of Lin and Wang studied the selfassembly of a binary system containing rodcoil (or coilcoil)
block copolymers and rod-like (or coil) homopolymers by
Brownian dynamics.8 The simulation results suggested that the
interpolymer association forces, the rigidity of the polymers,
and the mixture ratio of the two polymers all inuence the selfassembly behavior. By adjusting these parameters in their
calculations, the morphologies of the self-assembled objects
could be systematically varied from a beads-on-a-wire structure
to superhelices and plain bers. The simulations were
supported by experimental results using an array of block
copolymers (PBLG-b-PEG and PS-b-PEG) and the corresponding homopolymers (PBLG and PS).
708

DOI: 10.1021/ma502279b
Macromolecules 2015, 48, 707716

Article

Macromolecules

Addition of PFS48-b-PI264 to Lamellar Crystals of PFS. We


added 30 L of an 8 mg/mL solution in THF of PFS48-b-PI264 to a
suspension of lamellar crystals of PFS50 (and of PFS20) in decane (5
mL, 0.1 mg/mL), prepared as indicated above.
Blends of PFS48-b-PI264 with PFS Homopolymer. Two sets of
blends were prepared with PFS48-b-PI264 plus PFS20 or PFS50. Table 1

consists of a PFS homopolymer nanocrystal. Since PFS


homopolymer is insoluble in hexane, these multiarm structures
likely represent a kinetic trapping of the PFS homopolymer
before it precipitates. One of the main lessons of this work is
that the detailed nature of the colloidal objects formed in
solution depends sensitively on self-assembly protocols.
In a traditional blend experiment, both components are
present initially. For a blend of a coil-crystalline block
copolymer with the corresponding crystalline homopolymer,
one imagines that nuclei formed in solution will inuence the
self-assembly of both components. A fascinating example of this
kind of coself-assembly was reported recently by the van de
Ven group, who examined mixtures of a short polycaprolactone
(PCL10) homopolymer in the presence of a poly(ethylene
glycol)-block-polycaprolactone diblock copolymer (either
PCL18-b-PEG45 or PCL24-b-PEG45).18 The block copolymer
and PCL were both dissolved in dioxane, and then water (bad
solvent for PCL) was added. Lamellar structures were observed,
and the analysis suggested that these rafts were formed by a
sequence of directional adhesion steps consisting of the linear
aggregation of spheres to yield rods followed by the planar
alignment of rods to form the lamellae. The width and length of
the lamella depended on the homo-PCL content. The authors
also grew block copolymer assemblies without PCL in nhexanol. After sedimentation, the structures could be transferred into water. Here a similar lamella morphology was
obtained, indicating that the PCL homopolymer was not
necessary to obtain a lamellar morphology in n-hexanol.
However, only by coassembly of the block copolymer with
homo-PCL could the lamellar crystals be formed in water.
Here we describe a systematic study of the coassembly in
decane of blends of a PFS-b-PI block copolymer sample
(PFS48-b-PI264) with two dierent PFS homopolymers (PFS50
and PFS20). Well-dened mixtures were prepared in the dry
state. Then decane was added, and the sample was heated to
dissolve the mixture. We used transmission electron microscopy (TEM) to examine the structures that formed upon
cooling and show that the types of structures obtained depend
upon the blend composition. In many cases, the blend
structures had uniform features that we were able to quantify
through analysis of their TEM images. From these results, we
propose a model to explain how the dierent PFS
homopolymers cocrystallize with the PFS48-b-PI264 block
copolymer.

Table 1. Blends Prepared of PFS Homopolymer and PFS48b-PI264


PFS20 blends with PFS48-b-PI264a
mol %b PFS20

volume PFS20 (L, 50 g/mL)

10
25
50
75

10
25
51
76
PFS50 blends with PFS48-b-PI264

mol %b PFS50

volume PFS50 (L, 50 g/mL)

10
25
50
75

26
55
131
196

0.1 mL of PFS48-b-PI264 (1.0 mg/mL) in THF was added to each


blend, followed by the amount of PFS homopolymer indicated in the
right-hand column. Then the THF was evaporated. bThe mol % of
PFS homopolymer in the blend with PFS48-b-PI264.
describes the quantities of each component used for the blends. Stock
solutions of PFS20 and PFS50 in THF were prepared by dissolving
polymer (0.5 mg) in THF (10 mL). A stock solution of PFS48-b-PI264
was made by dissolving PFS48-b-PI264 (1.0 mg) in THF (1 mL). All the
blends samples have the same quantity of PFS48-b-PI264 block
copolymer (0.1 mg).
As an example of sample preparation, consider the 10:90 (by mole)
blend of PFS48-b-PI264/PFS20. PFS48-b-PI264 (0.1 mL of the 1.0 mg/mL
stock solution) and PFS20 (10 L of 50 g/mL solution) were added
to a 7 mL vial. The vial was swirled by hand to mix the two solutions.
The solvent was removed by a owing stream of air, and the mixture
was further dried overnight in a vacuum oven at 25 C. Decane (2
mL) was added to the vial, and then the mixture was heated to 100 C
in an oil bath. After 1 h, the sample was removed from the oil bath and
allowed to cool to room temperature in air. The sample was then aged
at room temperature for 3 days before grids were prepared for TEM
imaging.
Instrumentation. Transmission electron microscopy (TEM)
measurements were performed using a Hitachi H-7000 CTEM
operating at 75 kV. Samples for TEM analysis were prepared by
placing a drop of sample onto a carbon coated copper grid and then
wicking away the excess solvent with a KimWipe. Images were
analyzed using Image-J, published by the National Institutes of Health.
To ensure proper statistics, a minimum of 200 micelles were measured
by hand by drawing contour lines over each micelle in the image. In
the study of the dimension of the micelles we analyzed the length of
the bers and the areas of the lamellae obtained. Further details about
how we performed these measurements are given in the Supporting
Information (Figure S1).
Scanning electron microscopy (SEM) images were taken using a
Hitachi S-5200 SEM instrument. Energy-dispersive X-ray spectroscopy
(EDX) analysis to verify the presence of Pt catalyst coordinated to the
PI was performed using the EDX attachment (INCA, Oxford
Instruments) on the Hitachi S-5200 SEM instrument. The accelerating
voltage for EDX measurements was 15 kV, and the current was 20 A.
Staining the PI Corona of PFS48-b-PI264 in Blends with PFS.
To enhance the contrast in the images and to help determine the
location of PFS48-b-PI264 in the self-assembled structures, we used
Karstedts catalyst as a selective stain for PI.21 A drop of Karstedts
catalyst was added to a suspension of a blend of PFS and PFS48-b-PI264

EXPERIMENTAL SECTION

Synthesis of PFS and PFS48-b-PI264. PFS50 and PFS20 were


synthesized by anionic polymerization by procedures analogous to that
described previously. 19 PFS48-b-PI264 used in this study was
synthesized and characterized as described elsewhere.13,20 The
dispersity determined for all the homopolymers and block copolymers
was less than 1.1.
Homopolymer Self-Assembly. PFS50 (0.5 mg) and decane (5
mL) in a 20 mL scintillation vial were heated to 130 C. The
temperature of the bath was reduced after 1 h to 90 C. The vial was at
this temperature for another hour, after which the heating bath was
turned o. Then the vial in the oil bath was allowed to cool slowly to
room temperature. The homopolymer crystals were sonicated briey
(30 s) in a bath sonicator to break up aggregates to aid in imaging. The
same protocol was followed to prepare the homopolymer PFS20.
Self-Assembly of PFS48-b-PI264. Micelles were formed by heating
PFS48-b-PI264 to 100 C in decane (c = 0.5 mg/mL) for 1 h. After 1 h
the sample was removed from the bath and allowed to cool at room
temperature in air. TEM images were taken after 3 days of aging.
709

DOI: 10.1021/ma502279b
Macromolecules 2015, 48, 707716

Article

Macromolecules
after self-assembly. The solution was swirled and then left to react in
the dark for a period of 24 h. A drop of the solution was then placed
on a TEM grid, and the solvent was wicked away using a KimWipe.
The grids were left to dry in the open air.

yellow, and a small amount of precipitate was observed at the


bottom of the vial. The amount of precipitate here was lower
than for PFS50. As shown in Figure 1C, after solvent
evaporation, PFS20 appeared in TEM images as large elongated
crystal aggregates, greater than 10 m in length. The crystals
had rectangular features but tapered toward the ends. Upon
closer inspection, the aggregates appeared to consist of layers of
stacks of narrow tapering lamellae (more examples are shown
in Figure S2C). No isolated lamellae were observed.
Micelles of PFS48-b-PI264 (c = 0.5 mg/mL) were prepared by
suspending a polymer sample in decane, heating the mixture to
100 C for 1 h to obtain a clear solution, cooling the solution to
room temperature, and aging it for 3 days.13 In the unstained
TEM image (Figure 1D), contrast comes primarily from the
iron-rich core of the micelles. The width of this core was
uniform and ca. 8 nm across. Lengths varied throughout the
sample but most micelles were longer than 1 m.
As mentioned above, semicrystalline lms or suspensions of
PFS homopolymer can nucleate the growth of ber-like PFS
block copolymer micelles. We conrmed this behavior for the
addition of PFS48-b-PI264 unimer in THF (30 L, 8 mg/mL) to
a suspension of PFS50 platelets and their aggregates in decane
(5 mL, 0.1 mg/mL). As shown in the TEM images in Figure
S3A,B, stacks of PFS50 lamellae become surrounded by a dense
eld of ber-like micelles. Less aggregated homopolymer
crystals in the sample (Figure S3C,D) show a preference for
the attachment of the bers to the tips of the lamellar crystals.
Self-Assembly of Blends of PFS50 and PFS48-b-PI264. As
described in the Experimental Section, blends of PFS48-b-PI264/
PFS50 at dierent mole ratios were prepared in THF to
promote mixing, transferred to small vials, and dried. Decane
was added then. The samples were heated to 100 C for 1 h to
dissolve the polymers, cooled in air to room temperature, and
aged 3 days. Figure 2 presents selected TEM images for
structures formed from dierent ratios of PFS50 homopolymer
to block copolymer. Overall, the structures of the aggregates at
all compositions look similar: one observes a rectangular center
with long narrow bers at both short edges. The rectangular
structures are narrow and resemble the isolated lamellar crystals
found in the supernatant following crystallization of PFS50
homopolymer. However, for the blends, the composite
aggregates remain colloidally stable in solution. Upon drying,
they show no evidence for stacking in TEM images. One
prominent feature of the aggregates that varied with blend
composition was the mean width of the central lamellae, which
increased with increasing PFS50 content, from 94 nm at 10%
PFS50 to 152 nm at 75% PFS50 (Table S1).
Within this morphology, two dierent types of structures can
be distinguished. One consists of a long but broad lamella that
separates into multiple bers at the end. We refer to this
morphology as a blunt lamella. An example is indicated by the
red oval in the lower left of the TEM image for the 50:50 blend
(Figure 2C), and another example is shown at higher
magnication in Figure 2E for the 75:25 PFS48-b-PI264/PFS50
blend. The other type of structure is characterized by a long but
much thinner lamella that tapers into one or two bers at the
end. An example of this tapered lamella is indicated by the red
oval in the lower right of Figure 2C for the 50:50 blend. The tip
of one of these tapered structures for the 75:25 PFS48-b-PI264/
PFS50 blend is shown at higher magnication in Figure 2F.
Additional images are presented also in Figure S4.
The TEM images in Figure 2 also show that both the bers
and lamellae have similar contrast. This result might suggest

RESULTS
Self-Assembly of the Individual Components. Before
presenting results of the coassembly of PFS48-b-PI264 with the
two homopolymer samples PFS50 and PFS20, we rst examine
the behavior of the individual components under our selfassembly conditions.
When a sample of PFS50 in decane was heated to 130 C for
1 h, the polymer appeared to dissolve. The sample was then
held at 90 C for 1 h, where noticeable crystallization took
place. After cooling the sample to room temperature, we
obtained a yellow precipitate suspended in a pale yellow
supernatant. TEM images taken from this sample are presented
in Figure 1A,B and Figure S2. From the supernatant, we found

Figure 1. TEM images of the building blocks: (A, B) Two images of


rectangular crystals formed by PFS50 homopolymer in decane. (C)
Elongated crystals formed by PFS20 in decane. Homopolymer samples
(0.1 mg/mL) were heated to ca. 130 C to dissolve them; crystals or
micelles formed upon cooling. (D) Cylindrical micelles formed by
PFS48-b-PI264 diblock copolymer in decane. The block copolymer (0.5
mg/mL) was heated to 100 C. Micelles formed upon cooling.

isolated narrow rectangular lamellae (Figure 1A) as well as


stacked lamellae in other sections of the grid (Figure 1B). The
main portion of the sample, the yellow precipitate, consisted of
large arborescent structures formed by multiples lamellae
(Figure S2A,B). These lamellae appeared to be nucleated from
a central point and grew outward to form branches. The
arborescent structures were large in size compared to the
isolated narrow crystals, with dimensions up to tens of microns.
Attempts to obtain information about the crystal structure by
selected area electron diraction in the TEM failed because the
thin crystals were damaged by the electron beam before any
diraction data could be obtained.
Samples of PFS20 crystals in decane were prepared in the
same way. Upon cooling, the decane solution was tinged
710

DOI: 10.1021/ma502279b
Macromolecules 2015, 48, 707716

Article

Macromolecules

Figure 2. TEM images of blends of PFS48-b-PI264/PFS50. Mole ratios


of PFS48-b-PI264/PFS50 are (A) 10:90, (B) 25:75, (C) 50:50, and (D)
75:25. (E) Higher magnication image of the blunt end of a structure
formed by a 75:25 blend. (F) Higher magnication image of the blunt
end of a tapered structure formed in the same 75:25 blend.

Figure 3. TEM images of blends of PFS48-b-PI264/PFS20. Mole ratios


of PFS48-b-PI264/PFS20 are (A) 10:90, (B) 25:75, (C) 50:50, and (D)
75:25. (E, F) Higher magnication images of dierent parts of one
structure formed by a 10:90 blend. (E) shows the central lamellar core,
and (F) shows the ber-like end of the structure.

that they are similar in thickness. Evidence for the very thin
nature of these structures is provided by the high-magnication
TEM image in Figure S5 of a 50:50 blend of PFS48-b-PI264/
PFS50. In this gure we found a folded lamella with a triangle of
increased contrast at the turn. Other areas of increased contrast
can be seen in this image as well as in Figure 2E,F, where
micelle bers cross on top of other structures. These results
demonstrate that the lamella is thin enough to be able to fold
on itself and that the lamellae and the bers are thin enough to
be partially transparent to the incident electrons.
Self-Assembly of Blends of PFS20 and PFS48-b-PI264.
Parallel experiments were carried out on blends with PFS20.
Figure 3 presents low-magnication TEM images of the
structures formed for a series of dierent compositions for
blends of PFS48-b-PI264/PFS20. The structures in all of the
images consist of elongated lamellae with long bers protruding
from both ends. Here, however, the mean widths of the center
portions of the lamellae (ca. 85 nm) did not change with
sample composition.
One structure formed by a 10:90 blend of PFS48-b-PI264/
PFS20 is shown at higher magnication in Figure 3E,F. The
bers close to the lamellar core tend to be broad and ill-dened
but become more uniform in width toward the tip. At longer
distances from the lamellar core, the widths of the bers
decrease (Figure 3F). In the core region (Figure 3E), it appears
that at least some of the lamellae may be formed from aligned
ber-like ribbons that become attached side-by-side. Additional
images of the PFS48-b-PI264/PFS20 blend are presented in
Figure S6.
Location of PFS48-b-PI264 in the Structures Formed. In
the TEM images presented above, it is not possible to identify
the location of the homopolymer and block copolymer

components in the self-assembled structures. In order to


obtain further insights into the location of block copolymer in
the blend structures, we use Karstedts catalyst as a selective
stain of the PI corona in the blends.21 Karstedts catalyst is an
organoplatinum compound generally assumed to exist in
solution as mixture of related Pt(0) complexes with labile
alkene ligation.22 We recently reported that this catalyst is an
eective Pt stain for polyisoprene (PI).21
Figure 4 presents two TEM images of structures formed
from a 75:25 blend of PFS48-b-PI264/PFS50. The sample in

Figure 4. TEM images of 75:25 blend of PFS48-b-PI264/PFS50. (A) No


stained and (B) stained micelles treated with Karstedts catalyst, which
stains PI with Pt. Scale bar is 100 nm.

Figure 4A was not stained, whereas Figure 4B shows the same


sample after addition of Karstedts catalyst. We use gray scale
analysis (Figure S7) to determine the contrast of the objects in
these images. Remarkably, the central lamella and the bers
growing from it showed almost identical contrast. After
staining, the bers became larger in width, increasing from 10
711

DOI: 10.1021/ma502279b
Macromolecules 2015, 48, 707716

Article

Macromolecules

Figure 5. TEM image of two lamella of 75:25 blend of PFS48-b-PI264/PFS50 stained with Karstedts catalyst. The solid black line indicates where EDX
line scan was acquired, and the dashed black line indicates EDX peak position. Panels: A, carbon; B, silicon; C, iron; D, platinum; E, titanium.

measure the lengths of the bers and the areas of the central
lamellae for multiple unstained micelles of each type of
aggregate. Lines were traced by hand along the bers and
around each lamella (further details are provided in Supporting
Information and Figure S1). While we were careful to be
consistent, there is some arbitrariness to the outlining of
lamellar areas. However, we are looking for dierences in
features that far exceed any error in these measurements.
From these results, histograms of the ber lengths and
lamellar areas were constructed with bin sizes chosen to
accommodate dierences in the standard deviation of each
measurement (for more details, see Supporting Information).
Histograms of number normalized distributions of the lamellar
areas of the PFS48-b-PI264/PFS50 blends are presented in Figure
S9. The distributions for the 10:90, 25:75, and 50:50 blends are
all relatively narrow. However, for the 75:25 blend, the shape of
the area histogram is much broader, tailing toward larger
lamella areas. From this analysis, we obtained and plot the
average area of the lamellae (in m2) versus the PFS50 content
in the blends (Figure 6A). The graph shows that the lamellar
areas increase linearly with the PFS50 content.
Corresponding histograms of the area distributions for
PFS48-b-PI264/PFS20 blends are presented in Figure S10. For
lamellar areas of a blend of 10:90 PFS48-b-PI264/PFS20, the
histogram was centered at 0.07 m2, more or less the same as
the histogram for the 25:75 blend. Interestingly, for the PFS48b-PI264/PFS20 blends, the average lamellar area does not show
any dramatic trend with increasing PFS20 content, even when
large amounts of homopolymer were present. The mean area as
a function of homopolymer content is also plotted in Figure
6A. In this graph we can observed that the value of the average
area (0.072 m2) is also similar to the area value obtained for
the smallest PFS50 homopolymer content for the PFS48-b-PI264/
PFS50 blend (for the 10:90 mol ratio, the area was 0.077 m2).

nm in the unstained sample to ca. 25 nm after the sample was


stained. In the TEM image of the bers themselves, there is
very little contrast from the PI corona. The width one observes
is that of the PFS core. After treatment with Karstedts catalyst,
the presence of Pt provides contrast, and one sees the
contribution of the PI corona to the ber width. In contrast,
after treatment with Karstedts catalyst, there was no signicant
change in the width of the lamellae. This result indicates that
the relative contribution of the PI corona to the width of the
lamellae is small. For PFS48-b-PI264/PFS20 blends, we obtained
similar results (Figure S8).
Energy-dispersive X-ray spectroscopy (EDX) was used to
determine the location of dierent elements in self-assembled
structures formed from blends of PFS48-b-PI264/PFS. Figure 5
presents a TEM image of a 75:25 blend of PFS48-b-PI264/PFS50.
The line scan was performed on overlapping lamellae to
increase the strength of the EDX signal. The black solid line
traces the line scan, and the black dashed line refers to the
position in the structures where the peaks were observed in the
EDX spectra. The intensities for ve elements (carbon, silicon,
iron, platinum, and titanium) were collected. Both silicon and
iron are present only in the PFS polymer, and the platinum
comes from the Karstedts catalyst that selectively stains the PI
block. Titanium was used as negative control. The results
indicate that carbon, silicon, iron, and platinum were all present
in large quantities where the lamella crossed. As expected,
titanium was not present in the sample. The main conclusion
from this experiment is that Pt was present and prominent in
the lamellar region of the structures; i.e., the block copolymer is
also present in the lamellae and not conned to the bers.
Quantifying the Dimensions of Self-Assembled Blend
Structures. In order to gain further insights into the nature of
the coself-assembly of the PFS homopolymer and PFS48-b-PI264
block copolymer, we used the software program ImageJ to
712

DOI: 10.1021/ma502279b
Macromolecules 2015, 48, 707716

Article

Macromolecules

of PFS block copolymers in which the corona chains are equal


in length or shorter than the PFS block. For example, PFS76-bPI76 forms long narrow rectangular platelet micelles in decane
(containing a small amount of xylene).15 PFS114-b-PDMS81
(PDMS = polydimethylsiloxane) forms wider rectangular
platelet micelles with a much smaller aspect ratio.21 In both
instances, the ends of the long axis serve as initiating sites for
the growth of ber-like PFS-b-PI micelles. The nal structures
resemble scarves, with tassels of the ber-forming PFS-b-PI at
the ends of the rectangular PFS-b-PI or PFS-b-PDMS cores.
The latter system provides another piece of important
information about how the ber-forming PFS block copolymer
interacts with the edges of the rectangular core. When the
scarves consisting of the PFS114-b-PDMS81 core and PFS53-bPI637 tassels were treated with Karstedts catalyst to cross-link
the PI component, the PFS-b-PDMS core could be selectively
dissolved away to reveal a thin perimeter of cross-linked PFS-bPI. From this result, we learn that some PFS-b-PI attaches to
the long (lateral) edge of the crystalline planar PFS core, but no
further growth occurs in this direction.
Blends. The structures obtained from the blends of PFS50
and PFS20 with PFS48-b-PI264 as shown in the TEM images in
Figures 2 and 3 are very dierent from those of the crystals
formed by the homopolymers themselves. From this
perspective, one imagines that the lamellar cores of these
spider-like structures are not composed exclusively of
homopolymer and that there may be a gradient of composition
as the cores taper into elongated bers. Another indication of
PI incorporation into the lamellae was the low incidence of
lamella stacking observed in the blends. Stacking of lamellar
crystals was prominent feature of the TEM images obtained
from the pure homopolymers. This association between the
PFS crystals would minimize interfacial interactions between
the homopolymer and the solvent. However, in the blends, PI
corona chains protruding from the top and bottom surfaces of
the lamellae would provide steric stabilization against stacking
in decane solution. The PI corona also forms an exclusion layer,
prohibiting the growth of PFS crystals on the planar surfaces of
the growing crystal. This feature helps to explain why the blend
micelles are so thin.
Further information comes from the staining experiments
with Karstedts catalyst, which stains PI with Pt. The lack of
contrast dierence between the lamellae and the bers after
staining provides a strong suggestion for the incorporation of
PFS48-b-PI264 in the lamellae. This suggestion is conrmed in
EDX traces across the lamellar cores of the blend micelles. As
shown in Figure 5, there is a pronounced presence of Pt in the
center of the lamellae of micelles formed by a 75:25 blend of
PFS50 and PFS48-b-PI264, demonstrating the presence of PI
chains. We can conclude that the lamellar cores of these block
copolymer blend micelles are formed by the cocrystallization of
block copolymer and homopolymer. The feature of these
blends that remain to be explained is how the shapes of the
micelles vary with changes in composition of the blends.
A Pictorial Model. Lamellar Micelles. In the Alexander and
de Gennes model,23,24 chains tethered to a surface can be
described in terms of a brush of blobs (each of energy kT) that
extend from the surface when the separation between anchor
points d is smaller than twice the radius of gyration (Rg) of the
chains. All the blobs have the same size because the surface is
considered to be innite, and edge eects are ignored. On the
basis of this model, Vilgis and Halperin described the
aggregates formed by coil-crystalline diblock copolymers in

Figure 6. (A) Plot of the lamellar area and (B) of the ber length of
the micelles vs PFS content in the blends of PFS with PFS48-b-PI264.
The error bars represent the standard deviation in the histograms of
the (A) area and (B) length distributions.

Histograms of number normalized distributions of the ber


lengths of blends of PFS48-b-PI264/PFS50 and PFS48-b-PI264/
PFS20 are presented in Figures S11 and S12, respectively. The
average ber length decreases with increasing PFS50 content in
the blends. This dependence is plotted in Figure 6B. For these
blends, the ber length varies inversely with the measured
lamella area (Figure S13). For PFS48-b-PI264/PFS20 blends, the
mean ber length shows no signicant change with PFS20
content in the blends. The dierence between the largest mean
value of length (25:75, 3270 nm) and the smallest length value
(10:90, 3180 nm) is less than 100 nm.

DISCUSSION
Homopolymer and Block Copolymer Self-Assembly.
When crystallized from decane, PFS50 itself forms thin
rectangular crystals (Figure 1A,B) that tend to stack. Under
the same conditions, PFS20 forms clusters of ber-like crystals
(Figure 1C). The fact that the structures are elongated and
remain thin suggests that the primary crystal growth direction is
from the ends of the long axis of the lamellae with a slower
growth in width, with negligible growth perpendicular to the
lamellar faces. The thin lamellae tend to stack or aggregate to
minimize unfavorable interactions with the solvent.
In addition, we observed that PFS48-b-PI264 itself forms long
ber-like micelles of uniform width (Figure 1D),13 consistent
with previous observations that the crystal growth direction for
the block copolymer micelle is exclusively from the ends of the
bers. When a solution of PFS48-b-PI264 in THF is added to the
suspension of PFS50 crystals, ber growth is initiated at the ends
of the long axis of the rectangular crystals. The ends of the
crystal corresponding to the preferential growth direction for
PFS50 homopolymer crystals are also the face where growth of
PFS48-b-PI264 micelles is nucleated. While there may be some
attachment of the block copolymer to the two long edges of the
lamellar crystals, there is no measurable change in width for
these structures.
In many ways, the results obtained for PFS48-b-PI264 in THF
added to the suspension of PFS50 crystals mirror the behavior
713

DOI: 10.1021/ma502279b
Macromolecules 2015, 48, 707716

Article

Macromolecules

Figure 7. Schematic of crystalline-coil self-assembled structures in corona-selective solvents. Two views are presented: a 3-dimensional projection
(above) and an edge on projection (below). The red lines represent the semicrystalline PFS component. The blue lines represent the corona chains.
The blue circles represent the blob sizes of solvent swollen corona chains. The dashed blue line depicts the edge of the corona layer. The black
arrows represent the preferred crystal growth direction. The dashed black arrow depicts the secondary lateral growth direction. (A) Cross section of
a micelle formed by a block copolymer with a short corona forming block. (B) Cross section of a micelle formed by a block copolymer with a long
corona forming block. (C) Self-assembled cocrystallized blend of a block copolymer with a long corona forming block (as in part B) with
homopolymer present in the semicrystalline core.

solution.25 They treated the semicrystalline core of cylindrical


micelles as rectangular in cross section, consisting of folded
polymer chains, and analyzed the shape of the corona in terms
of a blob model. In an attempt to capture the essential features
of this model, we present pictorial models for the structure of
lamellar PFS micelles in Figure 7A, ber-like PFS micelles in
Figure 7B, and blends of ber-forming PFS micelles with PFS
homopolymer in Figure 7C.
In Figure 7A we show the case for a block copolymer micelle
in which the lamellar dimension is large, i.e., edge eects are
small. For PFS block copolymers, this case corresponds to
samples in which the length of the corona block is shorter than
the core-forming PFS block. For example, PFS114-b-PDMS81
forms wide rectangular platelet micelles.21 (For PFS block
copolymers in which the length of the corona block is similar in
length to the core-forming PFS block, e.g. PFS76-b-PI76, one
obtains lamellar micelles, but these have a broad, elongated
shape.15) We depict the corona chains as blobs of uniform size
on the top and bottom faces of the lamella. For asymmetric
lamellae, the long axis of the semicrystalline core is determined
by the principal directions of crystal growth, with slower growth
along the perpendicular secondary growth direction.
It is likely that rectangular lamellar crystals of PFS
homopolymer grow in a similar manner.26 From this
perspective, the size of individual lamellae is determined
primarily by the number of nuclei formed initially and the
relative rates of growth along the principal and secondary
directions. PFS50 formed rectangular crystals (Figure 1B). The
PFS20 crystals (Figure 1C) had rectangular features but often
tapered toward into single or multiple bers at one or both
ends. The branching of the PFS20 crystals into tapering bers
may originate from defects at the crystal growth front that aect
the deposition of PFS blocks at the crystal growth fronts,27,28

but we have no clear explanation for the dierences in crystal


growth for PFS50 and PFS20.29
For lamellar-forming PFS block copolymers with a long
corona block, edge eects start to become important. The
aspect of Figure 7B,C pertinent to this case is the depiction of
corona blobs at the edge of the platelet, expanding to ll the
additional volume available. At the edge of the platelet, the
blobs expand to ll accessible volume at increasing distance
from the small spherical core.30
Elongated PFS-b-PI Block Copolymer Micelles. In Figure 7B
we draw a representation of the ber-like micelles formed by
PFS48-b-PI264. For the specic case of PFS48-b-PI264 micelles, we
found the core width to be 8 nm. We can imagine that the
polymer can crystallize as a chain of connected cubes. In our
system, each cube will represent a ferrocenyldimethylsilane
unit.31 The edge of each cube is 0.65 nm.32 The packing density
of the chains has to satisfy the measured mass per unit length of
these micelles, equal to approximately 3 polymer molecules per
nanometer of micelle length.33 This number of chains per
nanometer corresponds to the case where two chains are
displayed side by side. The surface occupied by these two
chains is ca. 40 nm2 (0.65 nm 0.65 nm 48 repeat units 2
chains side by side). If the core cross-section is rectangular, we
can deduce that the thickness of the core is ca. 5 nm. The
corona chains surrounding the core impart an overall cylindrical
cross section to the micelle. While we do not know the dspacing (separation between anchor points of the corona
chains), it is likely to be signicantly smaller than the estimated
unperturbed radius of gyration (4 nm) of PI with a (1,2 and
3,4)-rich architecture and a degree of polymerization of 264.34
Thus, we assume that the corona chains in the PFS48-b-PI264
micelles are signicantly stretched.
Returning to the Vilgis and Halperin model of the
rectangular core, we envision that the long chains expand
714

DOI: 10.1021/ma502279b
Macromolecules 2015, 48, 707716

Article

Macromolecules
into the volume at the sides of the ber to minimize interchain
repulsion. This decrease in corona density can be depicted in
terms of increasing blob size. Edge eects become important,
and the polymer in these blobs acts as a barrier to lateral growth
on the sides of the crystal. We assume that they have a much
smaller inuence on the principal direction of ber growth.
This eect produces long thin bers of uniform width (Figure
7B).
HomopolymerBlock Copolymer Blends. For blends of
PFS48-b-PI264 with PFS homopolymer, the key question is how
the homopolymer is incorporated into the cocrystallized
structure. The degree of polymerization of PFS50 matches
almost exactly that of the PFS chains of the block copolymer.
PFS50 should be less soluble in decane than PFS48-b-PI264
because of the solubilizing eect of the PI block. When the
blend solutions are cooled, crystallization is likely initiated by
homogeneous nucleation of the homopolymer, which acts as
nucleation point for the block copolymer cocrystallization. This
process leads to formation of the lamellar core and has the
eect of increasing the d-spacing between anchor points of the
corona chains and reducing corona chain repulsion. In addition,
this eect also reduces the tendency for corona chains to
interfere with crystal growth in the lateral direction as we
indicate in the drawing in Figure 7C. Since the d-spacing
increases with increasing PFS50 homopolymer content in the
cocrystal, this model rationalizes the increase in lamellar width
and lamellar area (Figure 6A) of the core portion of the blends
of block copolymer with PFS50 homopolymer.
We also believe that PFS50 is not easily incorporated into thin
bers, and it tends to be depleted more rapidly from the
supersaturated blend solution. Thus, it becomes concentrated
into the lamellar core. From this perspective, cocrystallization
of PFS50 with the block copolymer can aect the length of the
tassels at the ends of the lamellae in two ways. First, as the
homopolymer content in the blends is increased, the lamellae
become wider. The broader ends provide more sites for seeded
growth of the ber-like structures, formed primarily of block
copolymer, i.e., the formation of more bers that are on average
shorter in length. Second, we imagine that the ber-like tassels
grow o the ends of the core only after most of the
homopolymer in solution is depleted. As the amount of
homopolymer in the blend is increased, a larger fraction of the
PFS-b-PI block copolymer is incorporated into the lamellar
core. This would leave less block copolymer available to grow
bers o the ends of the core. Both factors explain why when
the lamella area increased, the mean length of the bers
decreased (Figure S13).
For blends of PFS48-b-PI264 with PFS20, neither the mean
width or area of the lamellae nor the mean length of bers
(Figure 6) varied with the amount of PFS20 in the blend. PFS20
is much shorter than the PFS block of the block copolymer.
This will aect how it packs in the cocrystal. While PFS20 itself
forms elongated rectangular crystals (Figure 1C), these crystals
often have frayed ends that taper into elongated bers (Figure
S2C). Thus, we imagine that it is much easier for PFS20 than for
PFS50 to be incorporated deeper into the ber-like structures
formed from cocrystallization of the homopolymer and block
copolymer.

homopolymer. In both sets of examples, colloidally stable


structures were formed. The micelles had an insect-like shape,
consisting of elongated lamellae with long protruding bers. We
established that for both type of blends not only the bers but
also the core of the blends micelles contained block copolymer,
implying the cocrystallization of both components. For the
PFS48-b-PI264/PFS20 blends, the areas of the lamellar cores and
the lengths of the bers did not change as the homopolymer
content was varied. This observation suggests that the both
components became intimately mixed in both the core and
ber portions of the structure.
For the case of the PFS48-b-PI264/PFS50 homopolymer
block copolymer blends, the mean width and area of the
lamellar core increased with the homopolymer content. We
interpret this result to suggest that the longer PFS 50
homopolymer remained largely conned to the lamellar core,
where it cocrystallized with block copolymer molecules. Here
the degree of polymerization of PFS50 matched almost exactly
with that of the PFS chains of the block copolymer. When the
amount of PFS50 in the blend was increased, the space between
the anchor points of the PFS48-b-PI264 molecules in the lamellar
core also increased due to the presence of the PFS 50
homopolymer. This increment aects the size of the micelles,
reducing the corona chain repulsion and allowing the lamellar
width and lamellar area to grow.
With this study we have demonstrated that the cocrystallization plays a crucial role in the modication of the structure of
micelles formed. The shape of PFS-b-PI block copolymer
micelles can be changed by blending with a homopolymer. This
nature of this change depends upon the amount and the relative
length of the homopolymer compared to that of the
crystallizable block in the block copolymer.

ASSOCIATED CONTENT

S Supporting Information
*

TEM images of PFS20 and PFS50 homopolymer crystals, of


micelles obtained by the addition of PFS48-b-PI264 to a
suspension of PFS50 homopolymer crystals in decane, and of
the dierent PFS48-b-PI264/PFS20 and PFS48-b-PI264/PFS50
blends; a table containing the average lamellar areas measured
in multiple TEM images of blends of PFS48-b-PI264/PFS50 and
PFS48-b-PI264/PFS20; histograms of number normalized distributions of the lamellar areas of blends of PFS48-b-PI264/PFS50
and PFS48-b-PI264/PFS20; histograms of number normalized
distributions of ber lengths of blends of PFS48-b-PI264/PFS50
and PFS48-b-PI264/PFS20; plots of lamella area vs ber length for
blends of PFS with PFS48-b-PI264. This material is available free
of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION

Corresponding Authors

*E-mail Ian.Manners@bristol.ac.uk (I.M.).


*E-mail mwinnik@chem.utoronto.ca (M.A.K.).
Author Contributions

G.C. and M.J.G.-A. contributed equally to this work.


Notes

The authors declare no competing nancial interest.

CONCLUSIONS
In this study we examined the self-assembled structures formed
in decane solution by blends of crystalline-coil PFS48-b-PI264
block copolymer plus either PFS20 or PFS50 semicrystalline

ACKNOWLEDGMENTS
The Toronto authors thank NSERC Canada for their nancial
support. I.M. acknowledges the EU for an Advanced
715

DOI: 10.1021/ma502279b
Macromolecules 2015, 48, 707716

Article

Macromolecules
Investigator Grant. The authors also thank Amy Petretic and
Elsa Lu for fruitful discussions.

REFERENCES

(1) Koizumi, S.; Hasegawa, H.; Hashimoto, T. Macromolecules 1994,


27, 6532.
(2) Hashimoto, T.; Tanaka, H.; Hasegawa, H. Macromolecules 1990,
23, 4378.
(3) Tanaka, H.; Hasegawa, H.; Hashimoto, T. Macromolecules 1991,
24, 240.
(4) Zhang, L.; Eisenberg, A. J. Am. Chem. Soc. 1996, 118, 3168.
(5) Cai, C.; Lin, J.; Chen, T.; Wang, X.-S.; Lin, S. Chem. Commun.
2009, 2709.
(6) Ouarti, N.; Viville, P.; Lazzaroni, R.; Minatti, E.; Schappacher, M.;
Deffieux, A.; Borsali, R. Langmuir 2005, 21, 1180.
(7) Kamps, A. C.; Fryd, M.; Park, S.-J. ACS Nano 2012, 6, 2844.
(8) Cai, C.; Li, Y.; Lin, J.; Wang, L.; Lin, S.; Wang, X.-S.; Jiang, T.
Angew. Chem., Int. Ed. 2013, 52, 7732.
(9) Chen, L.; Jiang, T.; Lin, J.; Cai, C. Langmuir 2013, 29, 8417.
(10) Xu, G.-K.; Feng, X.-Q.; Li, Y. J. Phys. Chem. B 2010, 114, 1257.
(11) Schmelz, J.; Schacher, F. H.; Schmalz, H. Soft Matter 2013, 9,
2101.
(12) Zheng, J. X.; Xiong, H.; Chen, W. Y.; Lee, K.; Van Horn, R. M.;
Quirk, R. P.; Lotz, B.; Thomas, E. L.; Shi, A.-C.; Cheng, S. Z. D.
Macromolecules 2006, 39, 641.
(13) Wang, X.; Guerin, G.; Wang, H.; Wang, Y.; Manners, I.; Winnik,
M. A. Science 2007, 317, 644.
(14) Gilroy, J. B.; Gadt, T.; Whittell, G. R.; Chabanne, L.; Mitchels, J.
M.; Richardson, R. M.; Winnik, M. A.; Manners, I. Nat. Chem. 2010, 2,
566.
(15) Gadt, T.; Ieong, N. S.; Cambridge, G.; Winnik, M. A.; Manners,
I. Nat. Mater. 2009, 8, 144.
(16) Mohd Yusoff, S. F.; Gilroy, J. B.; Cambridge, G.; Winnik, M. A.;
Manners, I. J. Am. Chem. Soc. 2011, 133, 11220.
(17) Qiu, H.; Cambridge, G.; Winnik, M. A.; Manners, I. J. Am.
Chem. Soc. 2013, 135, 12180.
(18) Rizis, G.; van de Ven, T. G. M.; Eisenberg, A. Angew. Chem., Int.
Ed. 2014, 53, 9000.
(19) Ni, Y.; Rulkens, R.; Manners, I. J. Am. Chem. Soc. 1996, 118,
4102.
(20) Massey, J. A.; Temple, K.; Cao, L.; Rharbi, Y.; Raez, J.; Winnik,
M. A.; Manners, I. J. Am. Chem. Soc. 2000, 122, 11577.
(21) Rupar, P. A.; Cambridge, G.; Winnik, M. A.; Manners, I. J. Am.
Chem. Soc. 2011, 133, 16947.
(22) Stein, J.; Lewis, L. N.; Gao, Y.; Scott, R. A. J. Am. Chem. Soc.
1999, 121, 3693.
(23) de Gennes, P. G. Macromolecules 1980, 13, 1069.
(24) Alexander, S. J. Phys. (Paris) 1977, 38, 983.
(25) Vilgis, T.; Halperin, A. Macromolecules 1991, 24, 2090.
(26) Hsiao, M.-S.; Yusoff, S. F. M.; Winnik, M. A.; Manners, I.
Macromolecules 2014, 47, 2361.
(27) Ma, Y.; Qi, B.; Ren, Y.; Ungar, G.; Hobbs, J. K.; Hu, W. J. Phys.
Chem. B 2009, 113, 13485.
(28) Ungar, G.; Putra, E. G. R.; de Silva, D. S. M.; Shcherbina, M. A.;
Waddon, A. J. Adv. Polym. Sci. 2005, 180, 45.
(29) Muthukumar, M. Lect. Notes Phys. 2007, 714, 1.
(30) Daoud, M.; Cotton, J. P. J. Phys. (Paris) 1982, 43, 531.
(31) Reiter, G. Chem. Soc. Rev. 2014, 43, 2055.
(32) Qi, F.; Guerin, G.; Cambridge, G.; Xu, W.; Manners, I.; Winnik,
M. A. Macromolecules 2011, 44, 6136.
(33) Guerin, G.; Raez, J.; Manners, I.; Winnik, M. A. Macromolecules
2005, 38, 7819.
(34) Mark, J. E. Physical Properties of Polymers Handbook; AIP Press:
Woodbury, NY, 1996.

716

DOI: 10.1021/ma502279b
Macromolecules 2015, 48, 707716

Anda mungkin juga menyukai