Electro-responsive and dielectric characteristics of graphene sheets decorated with TiO2 nanorods

Jin-Yong Hong , Eunwoo Lee and Jyongsik Jang *
World Class University (WCU) program of Chemical Convergence for Energy & Environment (C2E2), School of Chemical and Biological Engineering, College of Engineering, Seoul National University (SNU), Seoul, Korea. E-mail: jsjang@plaza.snu.ac.kr; Fax: +82 2 888 1604; Tel: +82 2 880 7069

Received 12th September 2012 , Accepted 23rd October 2012

First published on 24th October 2012


Abstract

Titanium dioxide (TiO2) nanorod-decorated graphene sheets have been synthesized by a simple non-hydrolytic sol–gel approach and demonstrated to be a highly effective dispersing material for electrorheological fluids. Electron microscopy and X-ray diffraction analysis indicated that the TiO2 nanorods with a high crystallinity are well-dispersed and successfully anchored on the graphene sheet surface through the formation of covalent bonds between Ti and C atoms. Furthermore, electro-responsive properties of graphene sheets decorated with TiO2 nanorods are investigated on the basis of the dielectric loss model. The dielectric property analysis based on the dielectric loss model clarifies that an introduction of TiO2 nanorods has been coupled with the larger achievable polarizability and short relaxation time of interfacial polarization. Consequently, the TiO2 nanorod-decorated graphene sheets exhibit consistently higher ER efficiency than that of GO sheets and show unprecedented electro-responsive performance in extremely low concentration ER materials (<0.2 wt%).


1 Introduction

Graphene, a two-dimensional single layer sheet of sp2-hybridized carbon, has attracted tremendous worldwide attention in recent years because of its unique properties, including a tunable bandgap, high carrier mobility, ambipolar field effect and quantum Hall effect.1 In particular, there is now a growing interest in the use of graphene as an electrorheological (ER) material due to its several attractive features, such as its peculiar nanostructure, high surface area, and good mechanical strength.2

ER fluids, typically composed of polarizable solid particles in an insulating medium, undergo dramatic structural changes under an external electric field. Under the electric field strength, the dispersed particles are polarized and attracted to each other to form fibril structures along the external electric field direction, and these electrically induced structures vanish when the electric field strength is removed. ER fluids have received considerable attention due to their several attractive features, such as simple mechanics, low power consumption, rapid response time, and reversibility.

Recently, graphene has stimulated a great deal of interest as a good candidate for the reliable ER property. However, although there have been several recent reports describing the possible use of graphene as an ER material, many of these cases have resulted in insufficient shear strength and low colloidal stability, which could limit their practical or industrial applications.

For this reason, the need to modify the polarizability of graphene with a high relative dielectric constant has been highlighted. A convenient way to enhance the polarizability is to introduce a hybrid nanostructure that interconnects the graphene sheet with high dielectric materials. Recently, titanium dioxide (TiO2) has been extensively studied because of its high dielectric constant, abundance, low cost, and structural stability against external stimulation.3 These features make TiO2 particularly attractive for use in ER materials.

Herein, we have primarily focused on the fabrication of TiO2 nanorod-decorated graphene sheets (TNGSs) by means of a simple non-hydrolytic sol–gel reaction and applied as a dispersing material for ER fluids. Graphene sheets with a large specific surface area could also potentially serve as a supporting material to anchor TiO2 nanorods. The TiO2 nanorods were introduced onto the graphene surface so as to decrease its current density and induce strong interfacial polarization. Furthermore, dielectric properties of the TNGSs and GO sheets were investigated to give deep insights into the primary role of TiO2 nanorods that determine the electro-responsive characteristics.

2 Materials and methods

2.1 Materials

Graphite (flakes, <20 μm, synthetic), titanium(IV) chloride, oleylamine (70%) and toluene (99.5%) were purchased from Sigma-Aldrich. Other reagents including KMnO4, HCl, H2SO4, H2O2 and ethanol were also obtained from Aldrich Chemical Co. All regents were used as received. For electrorheological (ER) fluid application, silicone oil (Aldrich, poly(methylphenyl-siloxane), viscosity = 100 cSt) was used as a dispersing medium.

2.2 Fabrication of graphene oxide (GO) sheets

GO was synthesized from graphite using a modified Hummers method. Typically, graphite (1.0 g) was added to 70 mL of H2SO4 in an ice bath, which was followed by the addition of KMnO4 (3.0 g) and NaNO3 (0.5 g). After stirring for 4 h, 70 mL of distilled water was slowly added and maintained at that temperature for 30 min. Subsequently, H2O2 solution was added to the solution until the color turned a brilliant brown indicating a fully oxidizing state. The as-prepared graphite oxide slurry was exfoliated to generate GO nanosheets by sonication using an ultrasonic generator (42 kHz, 100 W, Branson 3510, Branson Cleaning Equipment Co., Shelton, CT, USA) for 3 h. Finally, the mixture was separated by centrifugation, washed repeatedly with 5% HCl and distilled water, and dried in a vacuum oven at 40 °C for 24 h.

2.3 Synthesis of TiO2 nanorod-decorated graphene sheets

TNGSs were fabricated via a non-hydrolytic sol–gel reaction. A variable amount of GO was dispersed in 30.5 mL of oleylamine solution. For a complete exfoliation of GO, the mixture was ultrasonicated for 3 h. Next, the exfoliated GO solution was transferred into a three-neck round-bottom flask and heated to 290 °C with nitrogen purging. A TiO2 precursor (TiCl4) was then injected into the solution under vigorous stirring. The weight ratio of GO/TiCl4 was 5 wt%. The color of the solution changed from dark purple to yellow during the reaction, indicating that TiO2 nanorods were formed by the non-hydrolytic sol–gel reaction. After 15 min, the reaction was quenched by the addition of 6 mL of toluene and then the reaction mixture was allowed to cool to room temperature. The resultant product was separated from the surfactants by centrifugation using an excess of acetone. The products were ultimately retrieved and allowed to dry in a vacuum oven for 12 h. Finally, the products were calcinated for 2 h at 500 °C under a N2 atmosphere to achieve high crystallinity.

2.4 Characterization

The TEM images were taken with a JEOL EM-2000 EX II microscope. To observe TEM images, the TNGSs diluted in toluene were deposited on a carbon film coated copper grid. TNGSs were characterized by high-power X-ray diffraction (XRD, M18XHF-SRA (Mac Science Co.)) with a Cu Kα radiation source (λ = 1.5406 Å) at 40 kV and 300 mA (12 kW), and a field emission scanning electron microscope (FE-SEM, JEOL-6700) equipped with an energy dispersive X-ray spectrometer (EDS). X-ray photoelectron spectra (XPS) were obtained by a Sigma probe (Thermo).

2.5 Investigation of electrorheological properties

The ER properties were examined via a rheometer (AR 2000 Advanced Rheometer, TA Instruments) with a concentric cylinder conical geometry of 15 mm cup radius with a gap distance of 1.00 mm, a high-voltage generator (Trek 677B), and a temperature controller. To start a run, an ER fluid is placed between the cup and rotor, and DC voltage is applied. An electric field was applied for 3 min to obtain an equilibrium columnar structure before applying shear. All measurements were made at room temperature, and the shear rate was fixed to 0.1 s−1.

3 Results and discussions

3.1 Fabrication of TiO2 nanorod-decorated graphene sheets

The formation of TNGSs was confirmed by transmission electron microscopy (TEM) analysis. Fig. 1a and b showed large quantities of well-dispersed TiO2 nanorods, with a dimension of 25 nm (length) × 4 nm (diameter) (L/D ratio ≈ 6), successfully decorated on the graphene surface. Interestingly, as-synthesized TiO2 nanorods are highly anisotropic with the aspect ratio of ∼6. Under our experimental condition, the TiO2 nano-rods have crystallographically anisotropic structures where surface energy can play a crucial factor. The TiO2 nanorod has a {001} surface with higher energy. Because the growth rate is exponentially proportional to the surface energy under the kinetic growth process, the energy difference between the higher energy surface and other lower energy surfaces can promote preferential growth along the 〈001〉 directions of TiO2.4
(a) TEM image of overall TNGSs morphology. (b) Higher magnification of the boxed area in (a). (c) EDS mapping of carbon (C), titanium (Ti), and merged image (Ti + C) of TNGSs (scale bar: 1 μm). EDS analysis on the selected area of the TNGSs surface. (d) XRD patterns of TNGSs (red line indicates the crystalline phase of anatase TiO2) (inset: deconvoluted XPS peak for TNGSs at the Ti2p core level).
Fig. 1 (a) TEM image of overall TNGSs morphology. (b) Higher magnification of the boxed area in (a). (c) EDS mapping of carbon (C), titanium (Ti), and merged image (Ti + C) of TNGSs (scale bar: 1 μm). EDS analysis on the selected area of the TNGSs surface. (d) XRD patterns of TNGSs (red line indicates the crystalline phase of anatase TiO2) (inset: deconvoluted XPS peak for TNGSs at the Ti2p core level).

In addition, decoration of the surface of graphene with TiO2 nanorods was analyzed with elemental mapping by energy dispersive spectrometry (EDS) (Fig. 1c). The TNGSs were enriched with titanium (Ti) and carbon (C), indicating a uniform distribution of the TiO2 nanorods throughout the whole graphene sheet. The EDS analysis also indicated the presence of C (5.57%), O (49.81%), and Ti (44.62%), which proved that the TiO2 nanorods were successfully introduced on the carbonaceous graphene sheets. Interestingly, it is possible that the TiO2 nanorods located at the surface of graphene could result in a relatively lower intensity of carbon content compared to the higher intensity of titanium content.

The TNGSs were further characterized by XRD and X-ray photoelectron spectroscopy (XPS). The TNGSs exhibited XRD peaks corresponding to the (101), (004), (200), (105), (211), (204), (116), (204) and (215) planes for the anatase structure of TiO2 (Fig. 1d).5 The most predominant peak centered at 2θ = 26.5° was representative of the (101) anatase phase reflections with an interlayer spacing of 3.5 Å, namely, the diffraction pattern of TNGSs revealed the crystallinity and anatase phase of the TiO2 nanorods. XPS analysis was utilized to investigate the hybridization of graphene and TiO2 nanorods (Fig. 1d inset). In the case of the Ti2p XPS spectra in Fig. 1d, two bands were located at binding energies of 464.5 and 458.9 eV, which were assigned to Ti2p1/2 and Ti2p3/2, respectively. The deconvoluted Ti2p spectrum confirmed two low-intensity peaks centered at 465.8 and 460.2 eV, which were assigned to the Ti–C bond.6 Judging from these data, it can be concluded that TiO2 nanorods with a high crystallinity were chemically anchored onto the graphene sheet surface through the non-hydrolytic sol–gel reaction.

3.2 Electrorheological activities of TiO2 nanorod-decorated graphene sheets

Fig. 2a plots the flow curves of the shear stress versus shear rate for the ER fluid (both TNGSs and GO sheets 1 wt% in silicone oil). The shear stresses were measured as a function of shear rate (ŕ) under an applied electric field strength (3 kV mm−1). Under an applied electric field, both TNGSs and GO sheet-based ER fluids show a shear stress plateau region for a broad range of shear rates, indicating typical Bingham plastic behavior. In this plateau region, the electrostatic interactions within ER materials and hydrodynamic force caused by shear flow are competing with each other.7 However, beyond the critical shear rate (ŕcrit), the shear stress converged to the zero electric field strength value, giving a typical curve of Newtonian fluid behavior. This phenomenon can be attributed to the enhanced hydrodynamic force under high shear.8
(a) Shear stress as a function of shear rate for 1 wt% of various graphene-based ER fluids under an applied electric field strength (3 kV mm−1). (b) Dynamic yield stress as a function of weight fraction for various graphene-based ER fluids under 3 kV mm−1 of electric field [inset: dynamic yield stress of various graphene-based ER fluids as a function of electric field strength (1 wt% in silicone oil)].
Fig. 2 (a) Shear stress as a function of shear rate for 1 wt% of various graphene-based ER fluids under an applied electric field strength (3 kV mm−1). (b) Dynamic yield stress as a function of weight fraction for various graphene-based ER fluids under 3 kV mm−1 of electric field [inset: dynamic yield stress of various graphene-based ER fluids as a function of electric field strength (1 wt% in silicone oil)].

To gain an insight into the ER activity, the influence of the weight fraction (ω) of the ER material on the dynamic yield stress was evaluated by changing the weight fractions of ER fluids in the ω = 0–1 wt% range. Interestingly, the TNGSs represent typical Bingham fluid behavior in an extremely low concentration of ER materials (<0.2 wt%), which is an unprecedented result. The dynamic yield stress value increased with increasing weight fraction of ER materials at a fixed electric field strength. The dynamic yield stress value of TNGSs is about 40.6 Pa at 1 wt%, which is 3.9 times that of the same weight fraction of pristine GO sheets.

In addition, the relationship between the yield stress (τy) and the electric field strength is also investigated (Fig. 2b inset). If the ω value is constant, τy shows the dependence on the electric field strength. Similar to most ER fluids, the yield stress values of ER fluids are directly proportional to the increase in the applied electric field strength. In our experimental condition, both TNGSs and GO sheet-based ER fluid showed that τy is proportional to E02 at low E0 (<1 kV mm−1) and typically approaches E03/2 at high E0 (>1 kV mm−1). Even if the two different types of samples (TNGSs and GO sheets) had the same electro-responsive behavior, they displayed significantly different ER activity in terms of ER efficiency (defined as (τEτ0)/τ0 × 100 or Δτ/τ0 × 100, where τE is the shear stress with an applied electric field strength and τ0 is the shear stress without an electric field strength). When the shear rate reached 102 s−1, the TNGSs provided much higher ER efficiency than that of GO sheets and the corresponding ER efficiencies were about 554 and 113%, respectively.

To evaluate the real-time responses of the two different types of graphene-based ER fluids, an applied electric field (1 kV mm−1) was alternately turned on and off, and their shear stress values were monitored in real time (Fig. 3a). When the electric field was applied, the shear stress values of both GO sheets and TNGSs increased instantaneously. In contrast, the shear stress values dropped rapidly back to their original level when the electric field was removed. This result indicates that the responses of the graphene-based ER fluids, upon the application of an electric field, are reversible and reproducible. However, the two different types of graphene-based ER fluids showed different response times (tres) and recovery times (trec) (defined as the time required for the response or recovery from 0% to 90% of its final value). Under the electric field strength, the TNGSs-based ER fluid tends to react more quickly than GO sheets. It is noted that the TNGSs ensure very short response times closely related to their intrinsic characteristics of the decorated TiO2 nanorods.


(a) Effect of switching the applied electric field on the shear stress of various graphene-based ER fluids (1 wt% in silicone oil). (b) Microscope images of chain formation in a silicone oil suspension of TNGSs (1 wt% in silicone oil) without an electric field and with an applied electric field of 1 kV mm−1. The gap distance between the two electrodes was 1.0 mm.
Fig. 3 (a) Effect of switching the applied electric field on the shear stress of various graphene-based ER fluids (1 wt% in silicone oil). (b) Microscope images of chain formation in a silicone oil suspension of TNGSs (1 wt% in silicone oil) without an electric field and with an applied electric field of 1 kV mm−1. The gap distance between the two electrodes was 1.0 mm.

In addition, a microstructural transition of TNGSs-based ER fluid was observed using an optical microscope (OM) under an applied electric field as shown in Fig. 3b. Randomly dispersed TNGSs began to move rapidly toward the electrodes (within 100 milliseconds) and then formed a fibrillated structure along the applied electric field direction. This aligned fibrous structure, which is dominated by sufficient electrostatic interaction between TNGSs, provides rapid structure reformation under a shear force as well as a better resistance to the shear flow.

3.3 Dielectric properties of TiO2 nanorod-decorated graphene sheets

A possible explanation for this difference in ER efficiency and response (recovery) time may reside with the dielectric properties of ER materials, which are closely related to their polarizability under an applied electric field. It has been reported that the ER property depends on the dielectric properties of the ER materials.9 In order to preliminarily understand the enhancement of ER efficiency, the dielectric properties of the two different types of samples are investigated. Fig. 4 presents the dielectric spectra (ε′ and ε′′) as a function of electric field frequency and Cole–Cole plots for the TNGSs and GO sheet-based ER fluids and their dielectric parameters are listed in Table 1. The Cole–Cole curve, which is a dielectric loss model for analyzing the dielectric characteristics of materials, has been introduced to describe the relationship between the ER activity and dielectric properties.10 Fitting lines in Fig. 4 are given by the following equation:
 
ε* = ε′ + ′′ = ε + (ε0ε)/[1 + (iελ)1 − α], (0 ≤ α < 1)(1)
here, ε′ is the dielectric constant, ε′′ is the dielectric loss factor, ε0 is the static permittivity (ƒ → 0), and ε is the fictitious permittivity (ƒ → ∞). It is widely accepted that good ER materials should have a large achievable polarizability (Δε = ε0ε) and short relaxation time of interfacial polarization (λ). The Δε values of GO sheets and TNGSs were 0.49 and 1.89, respectively. The Δε is related to the electrostatic interaction between ER materials. The higher Δε value of TNGSs indicates that it has an enhanced electrostatic interaction and can provide an improved ER property.

(a) Permittivity (ε′) and loss factor (ε′′) as a function of frequency for various graphene-based ER fluids (1 wt%). Open and closed symbols indicate permittivity and loss factor, respectively. (b) Cole–Cole plots for various graphene-based ER fluids. The fitting lines are obtained from eqn (1) with parameters given in Table 1.
Fig. 4 (a) Permittivity (ε′) and loss factor (ε′′) as a function of frequency for various graphene-based ER fluids (1 wt%). Open and closed symbols indicate permittivity and loss factor, respectively. (b) Cole–Cole plots for various graphene-based ER fluids. The fitting lines are obtained from eqn (1) with parameters given in Table 1.
Table 1 Electrorheological values and dielectric parameters for various graphene-based ER fluids
Samples Electrorheological valuea Dielectric parameterb
τ y τ E at 102 s−1 τ 0 ER efficiency ε 0 ε Δε ƒ max λ
a Electrorheological properties were acquired with a particle weight fraction of 1 wt% under 3 kV mm−1 electric field strength. τE and τ0 values were obtained at a shear rate of 102 s−1. b These values were measured by a Solartron SI 1260 Impedence/gain-phase analyzer with a Solartron 1296 dielectric interface. c The local frequency of the peak on the dielectric loss factor ε′′ and the fmax values were obtained by nonlinear regression analysis using OriginPro. d The relaxation time, denoted by λ = 1/2πfmax (fmax is the frequency of the loss peak).
TNGSs 40.62 Pa 65.72 Pa 10.05 Pa 553.93% 4.49 2.60 1.89 6.92 Hz 0.02 s
GO sheets 10.32 Pa 21.34 Pa 10.03 Pa 112.76% 3.01 2.52 0.49 0.89 Hz 0.18 s


Furthermore, the λ value is associated with the proper interfacial polarization response denoted by the relaxation time. The λ values, λ = 1/2πƒmax, were 0.18 and 0.02 s for GO sheets and TNGSs based ER fluids (Table 1). In general, the relaxation time is connected with the proper interfacial polarization response.11 The polarization rate of TNGSs might be faster than that of GO sheets under an electric field strength in terms of the dielectric loss model. Taking these results into account, it is concluded that the large achievable polarizability and the short relaxation time for interfacial polarization give a combined or synergistic contribution to superior ER activity of TNGSs.

4 Conclusion

In conclusion, titanium dioxide (TiO2) nanorod-decorated graphene sheets have been synthesized by a simple non-hydrolytic sol–gel reaction and demonstrated to be a highly effective dispersing material for ER fluids. In particular, the TNGSs exhibited consistently higher ER efficiency than that of GO sheets and showed unprecedented electro-responsive performance in an extremely low concentration of ER materials (<0.2 wt%). Furthermore, the dielectric analysis based on the dielectric loss model clarified that the decoration of the graphene surface with TiO2 nanorods has been coupled with the larger achievable polarizability and short relaxation time of interfacial polarization. Consequently, the synergistic contribution of both a large specific surface area of graphene sheets and high dielectric constant of TiO2 nanorods has a strong influence on the ER activity, which provides an outstanding enhancement in the ER efficiency compared with the other research results of graphene-based ER materials.

Acknowledgements

This research was supported by the WCU (World Class University) program through the National Research Foundation of Korea funded by the Ministry of Education, Science and Technology (R31-10013).

Notes and references

  1. (a) J.-Y. Hong and J. Jang, J. Mater. Chem., 2012, 22, 8179 RSC; (b) J.-Y. Hong, K.-Y. Shin, O. S. Kwon, H. Kang and J. Jang, Chem. Commun., 2011, 47, 7182 RSC; (c) D. Long, J.-Y. Hong, W. Li, J. Miyawaki, L. Ling, I. Mochida, S.-H. Yoon and J. Jang, ACS Nano, 2011, 5, 6254 CrossRef CAS; (d) K.-Y. Shin, J.-Y. Hong, S. Lee and J. Jang, J. Mater. Chem., 2012, 22, 7871 RSC; (e) K.-Y. Shin, J.-Y. Hong and J. Jang, Chem. Commun., 2011, 47, 8527 RSC; (f) K.-Y. Shin, J.-Y. Hong and J. Jang, Adv. Mater., 2011, 23, 2113 CrossRef CAS.
  2. (a) J.-Y. Hong and J. Jang, Soft Matter, 2012, 8, 3348 RSC; (b) J. Yin, W. Wang, R. Chang and X. Zhao, Soft Matter, 2012, 8, 294 RSC; (c) W. L. Zhang, Y. D. Liu, H. J. Choi and S. G. Kim, ACS Appl. Mater. Interfaces, 2012, 4, 2267 CrossRef CAS; (d) W. L. Zhang and H. J. Choi, Chem. Commun., 2011, 47, 12286 RSC; (e) H. Hu, X. Wang, J. Wang, F. Liu, M. Zhang and C. Xu, Appl. Surf. Sci., 2011, 257, 2637 CrossRef CAS; (f) W. L. Zhang, Y. D. Liu and H. J. Choi, J. Mater. Chem., 2011, 21, 6916 RSC; (g) W. L. Zhang, B. J. Park and H. J. Choi, Chem. Commun., 2010, 46, 5596 RSC.
  3. (a) J. Yin, X. Zhao, L. Xiang, X. Xia and Z. Zhang, Soft Matter, 2009, 5, 4687 RSC; (b) C. H. Hong, H. J. Choi and J. H. Kim, J. Mater. Sci., 2008, 43, 5702 CrossRef CAS; (c) L. Xiang and X. Xhao, J. Colloid Interface Sci., 2006, 296, 131 CrossRef CAS.
  4. (a) B. Koo, J. Park, Y. Kim, S.-H. Choi, Y.-E. Sung and T. Hyeon, J. Phys. Chem. B, 2006, 110, 24318 CrossRef CAS; (b) J.-H. Seo, Y.-W. Jun, S. J. Ko and J. Cheon, J. Phys. Chem. B, 2005, 109, 5389 CrossRef CAS.
  5. (a) E. Lee, J.-Y. Hong, H. Kang and J. Jang, J. Hazard. Mater., 2012, 219, 219 CrossRef; (b) S. H. Hwang, J. Song, Y. J. Jung, O. Y. Kweon, H. Song and J. Jang, Chem. Commun., 2011, 47, 9164 RSC; M. Choi, C. Kim, S. O. Jeon, K. S. Yook, J. Y. Lee and J. Jang, Chem. Commun., 2011, 47, 7092 Search PubMed.
  6. (a) O. Akhavan and E. Ghaderi, J. Phys. Chem., 2009, 47, 20214 Search PubMed; (b) Y. Huang, W. Ho, S. Lee, L. Zhang, G. Li and J. C. Yu, Langmuir, 2008, 24, 3510 CrossRef CAS.
  7. (a) J.-Y. Hong and J. Jang, Soft Matter, 2010, 6, 4669 RSC; (b) J.-Y. Hong, E. Kwon and J. Jang, Soft Matter, 2009, 5, 951 RSC; (c) J.-Y. Hong, M. Choi, C. Kim and J. Jang, J. Colloid Interface Sci., 2010, 347, 177 CrossRef CAS.
  8. (a) S. Park, M. Cho, S. Lim, H. J. Choi and M. Jhon, Macromol. Chem. Phys., 2005, 26, 1563 CrossRef CAS; (b) B. Wang and X. Zhao, Langmuir, 2005, 21, 6553 CrossRef CAS.
  9. (a) J. Yin, X. Zhao, L. Xiang, X. Xia and Z. Zhang, Soft Matter, 2009, 5, 4687 RSC; (b) P. Tan, J. Huang, D. Liu, W. Tian and L. Zhou, Soft Matter, 2010, 6, 4800 RSC.
  10. (a) W. L. Zhang and H. J. Choi, Langmuir, 2012, 28, 7055 CrossRef CAS; (b) W. Wen, X. Huang and P. Sheng, Soft Matter, 2008, 4, 200 RSC; (c) Y. G. Ko and U. S. Choi, Soft Matter, 2012, 8, 253 RSC.
  11. (a) S. Park, M. Cho, S. Lim, H. J. Choi and M. Jhon, Macromol. Chem. Phys., 2005, 26, 1563 CrossRef CAS; (b) B. Wang and X. Zhao, Langmuir, 2005, 21, 6553 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2013