You are on page 1of 8

PROCEEDINGS OF SPIE

SPIEDigitalLibrary.org/conference-proceedings-of-spie

Self-consistent field theory of directed


self-assembly in laterally confined
lamellae-forming diblock copolymers

Nabil Laachi, Hassei Takahashi, Kris Delaney, Su-Mi Hur,


David Shykind, et al.

Nabil Laachi, Hassei Takahashi, Kris T. Delaney, Su-Mi Hur, David Shykind,
Corey J. Weinheimer, Glenn H. Fredrickson, "Self-consistent field theory of
directed self-assembly in laterally confined lamellae-forming diblock
copolymers," Proc. SPIE 8323, Alternative Lithographic Technologies IV,
83230K (21 March 2012); doi: 10.1117/12.916577

Event: SPIE Advanced Lithography, 2012, San Jose, California, United States

Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 09 Jul 2023 Terms of Use: https://www.spiedigitallibrary.org/terms-of-use


Self-Consistent Field Theory of Directed Self-Assembly in Laterally
Confined Lamellae-Forming Diblock Copolymers
Nabil Laachia, Hassei Takahashia, b, Kris T. Delaneya, Su-Mi Hura, b, David Shykindc, Corey
Weinheimerc and Glenn H. Fredricksona, b*
a
Materials Research Laboratory, University of California, Santa Barbara, CA, USA.
b
Department of Chemical Engineering, University of California, Santa Barbara, CA, USA.
c
Intel Corporation, Hillsboro, OR, USA.

ABSTRACT
We use Self-Consistent Field Theory (SCFT) to study the directed self-assembly of laterally confined diblock
copolymers. In this study, we focus on systems where the self-assembled lamellae are oriented parallel to the selective
sidewalls of the channel. While well-ordered, perfect lamellae are observed both experimentally and numerically,
undesirable defective structures also emerge. We therefore investigate the energetics of two isolated defects, dislocations
and disclinations, for various chain lengths and channel dimensions and establish conditions that favor the formation of
defects. We also determine the energy barrier and the transition path between the defective and perfect state using the
string method.

Keywords: Self-Consistent Field Theory (SCFT), graphoepitaxy, directed self-assembly, defects, energy barriers, kinetic
pathways, string method.

1. INTRODUCTION
With optical lithography rapidly approaching its scaling limits, progress in nanofabrication is increasingly relying on the
development of alternative patterning routes. Interest lies in particular in nano-patterning tools that yield features below
the ~ 40 nm (half-pitch) limit of conventional line-space lithography. The self-assembly of different materials is arguably
one of the most promising high resolution patterning techniques that can replace current state-of-the-art optical
lithography-based technologies. Due to their 10 nm scale of microdomain ordering and the variety of shapes they
produce, block copolymers are excellent candidates as templates for lithography. However, without careful control or
guidance of the assembly, thin films of block copolymers will organize into randomly oriented grains exhibiting poor
long-range order.1-3Thus the prevalence of highly undesirable defective structures remains a major challenge that affects
the performance and viability of the technique for future applications.1

To overcome such a challenge, directed self-assembly (DSA), and graphoepitaxy (or lateral confinement) in particular,
offers an inexpensive route with the capability of dramatically reducing defect populations. While many successful
examples of using graphoepitaxy to produce well-ordered features of different shapes have been reported, a number of
experimental challenges still need to be addressed.4 The most important of which are high defect concentrations that,
although much lower than those in thin films, remain orders of magnitude higher than ITRS targets of < 0.01 10 nm
defect/cm2.1 The need to understand the formation of defects has thus motivated several groups to study defectivity in a
number of graphoepitaxial systems.5-7 In layered systems, such as lamellae-forming block copolymers, it was found both
theoretically8 and experimentally9 that isolated defects occur at a density,
−2 − ED / kT
nd ≈ ac e ,
where the key quantity ED is the formation energy of the defect and ac is the size of the defect; in the case of block
copolymers assembling in ~10 nm microdomains, 10 nm defects will therefore involve only a few chains per unit height
of the channel. The metastable defects we consider here are thermodynamic in nature, and hence their concentrations
follow the above scaling. Non-equilibrium defects that are kinetically frozen during the coarsening process will not be

*
To whom correspondence should be addressed : ghf@mrl.ucsb.edu

Alternative Lithographic Technologies IV, edited by William M. Tong, Douglas J. Resnick, Proc. of SPIE
Vol. 8323, 83230K · © 2012 SPIE · CCC code: 0277-786X/12/$18 · doi: 10.1117/12.916577

Proc. of SPIE Vol. 8323 83230K-1


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 09 Jul 2023
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Figure 1: a) Density profile of a well-ordered lamellae-forming diblock copolymer between two
confining walls. b) Plot of the excess free energy, Fex, where nc is the number of chains in the
trench and w is the width of the channel. c) Examples of density profiles of dislocations (top)
and disclinations (bottom) we consider in this study.

considered here. We will focus instead on kinetic pathways, or mechanisms that lead to transitions from metastable
defects to defect-free structures. Two key objectives thus motivate our present study. Our first goal is to determine
factors controlling equilibrium defect concentrations through an investigation of the energetics of two isolated defects
under confinement by means of SCFT. Our second objective is to study the kinetics of the transition between defect-
laden and defect-free states. The mechanism and the related barrier height of the transition pathway are of great
relevance and will be investigated using the string method.

2. ENERGETICS OF ISOLATED DISLOCATIONS AND DISCLINATIONS


SCFT has been successfully employed as a predictive tool for studying various phase-separated morphologies10,
including, laterally confined block copolymers11,12. In the present context, the chains, composed of N monomers, are
perfectly symmetric and interactions between A- and B-monomers are mediated through a Flory parameter χ. In the case
of PS-PMMA for example, the value of χ is slowly varying with temperature and a reference value of χN = 25 roughly
corresponds to a polymer with molecular weight, Mw ~ 36K-36K g/mol and a radius of gyration Rg ~ 6 nm.13 The melt is
confined inside a narrow trench where the impenetrable sidewalls are implemented through a mask that enforces a
vanishingly small polymer density at the wall region. The walls also preferentially interact with one of the blocks so as to
induce and favor parallel alignment of the lamellae in the longitudinal direction of the channel. These interactions are
modeled via a Flory-like parameter χw = (χwA – χwB)/2. In the mean-field approximation, the free energy of the system is
approximated by the value of the microscopic Hamiltonian at the saddle point configurations. Details of the method can
be found elsewhere.10-12

We should also mention that all the calculations presented here rely on two key assumptions: i) The melts are perfectly
monodisperse and, ii) the bottom substrate of the channel is completely neutral to both polymer blocks. The latter
assumption ensures that the self-assembled morphologies are homogeneous in the direction normal to the substrate and
can therefore be fully described using two-dimensional (2D) simulations. The free energies we present were nevertheless
computed by extrapolating our 2D results to a homogeneous 3D channel with a depth of about one lamellar period. In
doing so, the number of chains in the channel was estimated using a monomer density of 1 g/cm3, very close to
experimental values of PS and PMMA. Additional refinements through the investigation of the effects of polydispersity
and bottom substrate interactions are the subject of a separate publication.14

Proc. of SPIE Vol. 8323 83230K-2


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 09 Jul 2023
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
130 130

χN =21
120
χN =25
Excess free energy (in units of kT)

110

Excess free energy (in units of kT)


110 χN =30

100
90

90

80 70
n=3
70 n=4
n=5 50
60 n=6
n=7
50
30
12 17 22 27 32 12 13 14 15 16 17 18
Channel width (in units of Rg) Channel width (in units of Rg)

Figure 2: a) Excess free energy for dislocations (bottom) and disclinations (top) for various channel
widths and different number of periods, n. b) Excess free energy of dislocations in channels with 3
lamellar periods for various χN parameters.

2.1. DOMAINS OF STABILITY OF DEFECT-FREE STRUCTURES


Before we discuss defects in laterally confined block copolymers, let us focus first on perfect lamellae-forming systems.
This step is necessary for two reasons. First, we need to determine the optimal circumstances that favor the emergence
and the stability of such structures. Second, any study of defect energetics must be carried relative to defect-free
structures since the concentration of defects, as mentioned above, is related to differences in free energies between
defective and defect-free states. A representative density profile of such a perfectly aligned system is shown in Fig. 1a.
The total free energy of the perfectly ordered system can be decomposed as Fp = Fb + Fex = Fb + Fel + Fw, where Fb is
the free energy of the polymer in bulk and the total excess free energy Fex can be further decomposed into an excess free
energy, Fel, due to confinement (of elastic origin) and an excess free energy, Fw, due to wall interactions (surface
tension). Note that, by definition, Fw depends only on local interactions near the walls and therefore, variations of the
total free energy with respect to channel dimensions mainly result from the elastic component of the free energy. Fig. 1b
shows a plot of the excess free energy as a function of channel width. In this plot and in all what follows, a single period
refers to a repeat unit that consists of a pair of one A- and one B-lamella.

As shown in the figure, perfectly ordered structures with a different number of periods will arise for various channel
widths. The domains of stability, whereby an assembly of n periods is more stable than one with n+1 or n-1 periods, are
also evident. Inside a given interval with n periods, the excess free energy is minimized at an optimal, or
commensurability width, wn, where the assembly is most stable. At wn, residual elastic strains due to confinement are
nearly suppressed and the excess free energy is mainly dominated by the width-independent surface tension. The
periodicity wc, of the block copolymer under confinement can be obtained from the sequence of optimal channel widths
(or from cross-sections of density profiles). In the case of χN = 25, we found that wn = (n-1) wc + 2b, where wc ~ 4.15Rg,
about 3% smaller than in bulk (wcbulk ~ 4.29Rg), and the parameter b ~2Rg is the size of a brush of the A-component near
the walls that arises from the imposed mask. As the width of the channel increases or decreases away from the optimal
width for n periods, tensile or compressive strains build up, and at some critical value, their relaxation requires a
rearrangement of the lamellae resulting in the inclusion or the suppression of a period. As can be seen from Fig. 1b, the
stability domain of n periods is centered around the optimal width wn and extends half a period for increasing and
decreasing values away from wn.

Proc. of SPIE Vol. 8323 83230K-3


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 09 Jul 2023
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
25 12

11

0 10
Relative free energy (in units of kT)

0 0.2 0.4 0.6 0.8 1

Barrier height (in units of kT)


9

-25 8

-50 6

-75 4

-100 2
12 13 14 15 16 17 18 19 20 21 22
Reaction coordinate
Channel width (in units of Rg)

Figure 3: a) Example of a pathway for the transition between the defective (dislocations) and the
defect-free state obtained using the string method. The kinetic barrier to the transition occurs at an
intermediate state where the reaction coordinate is about 0.18. b) Plot of the barrier height as a
function of the channel width for dislocations in channels with 3 (left) and 4 (right) lamellar periods.

2.2. DEFECT FREE ENERGIES: COMMENSURABILIY AND POLYMER SIZE EFFECTS


In addition to well-ordered systems, SCFT simulations also produce a rich variety of defective metastable structures. The
characterization of such defects poses however several challenges, both conceptually and numerically. Because of the
high compliance of melts of block copolymers, defective morphologies are numerous and often span large domains,
contrasting with discrete, well localized defects in crystals. On numerical grounds, defects are also hard to stabilize
inside the channel as they are continuously evolving both in nature and in size and often come with a slow convergence
rate compared to defect-free structures. We therefore chose to focus our study on two isolated defects, namely
dislocations and disclinations, as they abound experimentally and they were more robust to changes in channel and
polymer parameters. While dislocations and disclinations come in different sizes and magnitudes, we limit our
discussion to the elementary defects represented in Fig. 1c.
The results of our simulations, summarized in Fig. 2, indicate that several tens of kT in gain in free energy are necessary
to the formation of defects from the pristine state. The gain is even larger for disclinations than dislocations because they
are accompanied by a larger distortion of the lamellae and stronger strain fields. Furthermore, the plots in Fig. 2a also
indicate that for a given number of periods, the excess free energy due to the presence of the defect is highest at a width
slightly larger than the commensurability width at which perfect lamellae exist. Deviations away from this optimal width
lead to a few kT decrease in the excess free energy and hence, an increase in the concentration of defects. We also note
that, as the number of periods increases, the maximum excess free energy also increases until it saturates to a period-
independent, bulk value for very large periods. This is primarily due to the rigid walls that hinder the propagation of the
strain field generated by the defect at small channel widths. As the trench gets wider, defects located near the center of
the channel will generate a distortion field that now fully propagates, unaffected by the presence of the walls.

The effect of χN on the propensity of defects in confining channels was also considered. In the case of PS-PMMA, the
variations of χN predominantly result from variations of the molecular weight of the block copolymer. Fig. 2b shows
excess free energies due to isolated dislocations in different channels for various χN parameters. Since higher χN values
produce a larger domain spacing, we observe a shift of the peaks to wider channels. We also found that for a given
period, increasing χN values results in significantly higher excess free energies..This gain in free energy mainly results
from the increased A-B interfacial area in the defect core relative to a perfect lamella.

Proc. of SPIE Vol. 8323 83230K-4


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 09 Jul 2023
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
In addition to width variations, we also looked at the effect of changing the strength of the interactions at the walls.
While increasing the strength of wall interactions does affect the absolute free energy of a defective or a defect-free
configuration, the excess free energy is not very sensitive to the magnitude of the selectivity. The main effect of the
presence of the walls is to nucleate a preferential alignment of the lamellae longitudinally, which will then propagate to
the center of the channel. Because the magnitude of the interactions only manifests as a local effect near the walls, as
long as the core of the defect is far away from the walls, there is not a substantial difference between this contribution
and that of a defect-free structure. In other words, what seems to matter in reducing defect concentrations is the
promotion of a preferential wetting at the wall, not the strength of this interaction.
It becomes thus apparent from the above thermodynamic study that the formation of isolated defects from the pristine
state is governed by an extremely small Boltzmann weight. Kinetic aspects will therefore be essential in our
understanding of the ubiquity of defects. The mechanisms leading to defect annihilation and melting to a perfect state are
equally important and will be discussed in the next section.

3. KINETIC PATHWAYS FOR DEFECT MELTING


The string method was first introduced by E et al.15 as a computational algorithm to determine the transition pathways
for complex systems when the mechanism for the transition is not known in advance. The two states, corresponding to
two local minima of the free energy landscape, are connected by a string of intermediate states in configuration space.
The minimum energy path (MEP) then corresponds to the most energetically probable trajectory for the kinetic transition
pathway between the two states. The details of the MEP calculations are beyond the scope of this document and can be
found elsewhere.15,16 Briefly, in the context of a field-based theory, the continuous string is discretized into a
parameterized set of field configurations using a “reaction coordinate”, α, such that α=0 is the defective metastable
configuration and α=1 is the defect-free structure. The string is then initialized, typically using a linear interpolation
between the two local minima, and relaxed to the MEP in a two-step process. First, each field configuration is updated
using a potential force explicitly obtained from the overall Hamiltonian of the system in the mean field approximation.
This generates a new set of states with new coordinates along the string. To avoid clustering of the updated
configurations near the two local minima of the free energy, a second step of reparameterization of the string by
interpolation is performed to evenly redistribute the configurations along the path. The converged solution that results
from successive iterations of the above two steps will lead to the MEP. The trajectories thus obtained not only inform
about the two local minima of the free energy landscape but also about the height of the barrier that needs to be crossed
for a transition between the two minima.

The string method outlined above was implemented to study transitions from defective assemblies to well-ordered
assemblies of block copolymers in confining channels. An example of such calculation is shown in Fig. 3a. In the case of
transitions from dislocations to perfect lamellar systems, all our string calculations indicate a single melting mechanism,
similar to that depicted in Fig. 4. Starting from the defective state, the two arms of the dislocation must first break into
separate microdomains. The rearrangement of the microdomains into well connected lamellae then occurs when the
polymeric chains in the defect core distort, resulting in an unfavorable stretching energy and an excess interfacial area.
The contribution of this transition state leads to an increase in the free energy, corresponding to the energy barrier shown
in the MEP plot in Fig. 3a. Since the mechanism described here involves essentially a diffusion of the chains parallel to
the A-B interface and therefore requires little mixing of the blocks, the kinetic barrier is relatively small and only a few
kT are sufficient for the transition between the two local minima.

Beyond the description of the melting of dislocations, we also investigated the commensurability effects on the transition
barrier. Variations of the kinetic barriers as a function of channel width are plotted in Fig. 3b for dislocations with n=3
and n=4 lamellae. Similar to defect free energies in Fig. 2, kinetic barriers are non-monotonic functions of channel
widths with a minimum value where defect annihilation is the fastest. However, for a given stability domain, the barrier
is now minimal at a width slightly smaller than the commensurability width. For channels slightly narrower than a
commensurability width, chains at the defect core are more compressed than their counterparts in the commensurate
case. As a result, the stretching of polymers that occurs at the transition configuration requires a smaller strain energy
and hence, a lower barrier when the channels are slightly narrower.

Let us focus now on the effect of the segregation strength on transition barriers, with string calculations performed at
commensurability widths for χN = 21, 25 and 30. We found that the height of the barrier strongly increases with χN, with
values of 0.22kT, 3.25kT and 7.86kT, respectively. This suggests an increased difficulty in producing well-ordered

Proc. of SPIE Vol. 8323 83230K-5


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 09 Jul 2023
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Figure 4: Snapshots of the melting mechanism of an isolated dislocation. The two arms of the
dislocation first break and then rearrange to form perfect lamellae.

lamellae with increased molecular weight, consistent with comparable experimental studies.17 Note that other kinetic
factors, such as entanglement effects, can also dramatically affect transition barriers and freeze the long polymers in
long-lived defective configurations.

While calculations using the string method revealed a single mechanism of the melting process of dislocations, the
uniqueness of MEP trajectories and the corresponding transition pathways, is not in general inherent to the string
method. Rather, early applications of the string method highlight examples where multiple MEP solutions can be
obtained depending on the initialization of the string.16,18 Contrary to dislocations, our string calculations show that
disclinations annihilate to perfect states via two distinct pathways. The first mechanism is symmetric and occurs when
both terminated lamellae simultaneously meet the disclination bend. The second mechanism however, is asymmetric and
produces an intermediate metastable state and a second barrier that needs to be crossed before the complete melting of
the disclination. Interestingly, the intermediate state appears when one of the two arms connects with the bend before the
other, thus, essentially forming the elementary dislocation. The barriers from the two mechanisms are also different. On
one hand, we found that the barrier for the symmetric mechanism is higher than that of a pure dislocation for all the
parameter sets we studied; a one-step melting of disclinations is much slower than that of dislocations. On the other
hand, the first barrier of the asymmetric transition is much lower than that of a dislocation; the melting of the
intermediate dislocation is therefore the limiting factor in the two-step melting process of disclinations.

Beyond the 2D melting mechanisms described above, more complex pathways can emerge from full 3D calculations
when inhomogeneities in defects are present. Such calculations are extremely costly in the context of the string method
and our current 2D results can be viewed as an upper bound of the predictions of kinetic barriers.

4. CONCLUSIONS
We used SCFT simulations to investigate the self-assembly of lamellae-forming block copolymers confined in narrow
trenches. Domains of stability of well-ordered structures were identified. The free energy associated with the formation
of two isolated defects was computed under various conditions of commensurability and segregation strengths. In the
case of transitions from defective to well-ordered states, we found that only a few kT of energy is necessary to overcome
the kinetic barrier and suppress the defect, sharply contrasting with the large gain in free energy (tens of kT) that is
necessary for the formation of the defect from the pristine state. Our results should ultimately provide useful guidelines
for the design of optimized graphoepitaxial processes where defect concentrations are minimized.

ACKNOWLEDGMENTS
This work was funded by Intel Corporation and the Focus Center Research Program (FCRP)-Center on Functional
Engineered Nano Architectonics (FENA). The calculations presented in this work were largely conducted using

Proc. of SPIE Vol. 8323 83230K-6


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 09 Jul 2023
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
computational facilities at Intel. Partial calculations of defect energetics were conducted using the computational
resources of the California NanoSystems Institute (CNSI) and Materials Research Laboratory (MRL) at the University of
California-Santa Barbara. The MRL Central Facilities are supported by the MRSEC Program of the NSF under Award
No. DMR 1121053; a member of the NSF-funded Materials Research Facilities Network (www.mrfn.org). We finally
wish to thank Robert Bristol, Intel Components Research, for experimental and theoretical conversations about DSA and
lithography.

REFERENCES
[1] Herr, D. J. C., “Directed block copolymer self-assembly for nanoelectronics fabrication,” J. Mater. Res., 26, 122
(2003).
[2] Park, C., Yoon, J., and Tomas, E. L., “Enabling nanotechnology with self assembled block copolymer patterns,”
Polymer, 44, 6725 (2003).
[3] Segalman, R. A., “Patterning with block copolymer thin films,” Mater. Sci. Eng. R, 48, 191 (2005).
[4] Bang, J., Jeong, U., Ryu, D. Y., Russell, T. P. and Hawker, C. J., “Block copolymer nanolithography: translation of
moleculaer level control to nanoscale patterns,” Adv. Mater., 21, 4769 (2009).
[5] Tiron, R., Chevalier, X., Couderc, C., Pradelles, J., Bustos, J., Pain, L., Navarro, C., Magnet, S., Fleury, G., and
Hadziioannou, G., “Optimization of block copolymer self-assembly through graphoepitaxy: A defectivity study,”
J. Vac. Sci. Technol. B, 29, 06F206 (2011)
[6] Bosse, A. W., Sides, S. W., Katsov, K., Garcia-Cervera, C. J. and Fredrickson, G. H., “Defects and their removal
in block copolymer thin film simulations,” J. Polym. Sci. part B: Pol. Phys., 44, 2495 (2006).
[7] Cheng, J. Y., Mayes, A. M. and Ross, C. A., “Nanostructure engineering by templated self-assembly of block
copolymers,” Nat. Mater., 3, 823 (2004).
[8] Toner, J. and Nelson, D. R., “Smectic, cholesteric, and Rayleigh-Benard order in two dimensions,” Phys. Rev. B,
23, 316 (1981).
[9] Hammond, M. R., Cochran, E., Fredrickson, G. H., and Kramer, E. J., ”Temperature dependence of order,
disorder, and defects in laterally confined diblock copolymer cylinder monolayers,” Macromolecules, 38, 6575
(2005).
[10] Fredrickson, G. H., [The equilibrium theory of inhomogeneous polymers], Oxford University Press, USA, (2006).
[11] Bosse, A. W., Garcia-Cervera, C. J. and Fredrickson, G. H., “Microdomain ordering in laterally confined block
copolymer thin films,” Macromolecules, 40, 9570 (2007).
[12] Hur, S. M., Garcia-Cervera, C. J., Kramer, E. J. and Fredrickson, G. H. “Scft simulations of thin film blends of
block copolymer and homopolymer laterally confined in a square well,” Macromolecules, 42, 5861 (2009).
[13] Russell, T. P., Hjelm Jr, R. P. and Seeger, P. A., “Temperature dependence of the interaction parameter of
polstyrene and poly (methyl methacrylate),” Macromolecules, 23, 890 (1990).
[14] Hassei, T., Laachi, N., Hur, S., Shykind, D., Weinheimer, C. and Fredrickson, G. H., “Directed self-assembly of
laterally confined lamellae-forming diblock copolymers: Polydispersity and substrate interaction effects,” Proc.
SPIE, paper 8323-58 (2012).
[15] Weinan, E., Ren, W. and Vanden-Eijnden, E., “String method for the study of rare events,” Phys. Rev. B, 66,
052301 (2002).
[16] Weinan, E., Ren, W. and Vanden-Eijnden, E., “Simplified and improved string method for computing the
minimum energy paths in barrier-crossing events,” J. Chem. Phys., 126, 164103 (2007).
[17] Park, S. M., Stoykovich, M. P., Ruiz, R., Zhang, Y., Black, C. T. and Nealey, P. F., “Directed Assembly of
Lamellae-Forming Block Copolymers by Using Chemically and Topographically Patterned Substrates,“ Adv.
Mater., 19 , 607 (2007).
[18] Cheng, X., Lin, L., Zhang, P. and Shi, A. C., “Nucleation of ordered phases in block copolymers,” Phys. Rev.
Lett., 104, 148301 (2010)

Proc. of SPIE Vol. 8323 83230K-7


Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 09 Jul 2023
Terms of Use: https://www.spiedigitallibrary.org/terms-of-use

You might also like