Compatibilization of Immiscible Polymer Blends Using in Situ Formed

Nov 30, 2015 - The in situ grafting of PLLA chains onto P((S-co-GMA)-g-MMA) ... opens new opportunities to exploit interfacial emulsification using JN...
1 downloads 0 Views 7MB Size
Letter pubs.acs.org/macroletters

Compatibilization of Immiscible Polymer Blends Using in Situ Formed Janus Nanomicelles by Reactive Blending Hengti Wang, Wenyong Dong, and Yongjin Li* College of Materials, Chemistry and Chemical Engineering, Hangzhou Normal University, Hangzhou 310036, P. R. China S Supporting Information *

ABSTRACT: Block or graft copolymers located at polymer−polymer interfaces have been considered as ideal compatibilizers for immiscible polymer blends. Herein, we report a novel compatibilization mechanism using Janus nanomicelles (JNMs) formed in situ at the polymer−polymer interface in immiscible polyvinylidene fluoride (PVDF)/polylactic acid (PLLA) blends. A small amount of a reactive graft copolymer, poly(styrene-co-glycidyl methacrylate)-graf t-poly(methyl methacrylate) (P((S-co-GMA)-g-MMA)), is incorporated into the PLLA/PVDF blends by simple melt mixing. The in situ grafting of PLLA chains onto P((S-co-GMA)-g-MMA) during melt mixing leads to the formation of numerous JNMs with a shell structure consisting of PLLA and PMMA hemispheres. These JNMs are located at the PLLA/PVDF interface, where they behave as effective compatibilizers for the immiscible PLLA/PVDF blends. This interfacial micelle compatibilization (IMC) mechanism opens new opportunities to exploit interfacial emulsification using JNMs and should be of great significance in the compatibilization of polymer alloys.

B

for immiscible blends.27−29 A type of colloidal particle, JNPs combine their inherent character (Pickering effect)30,31 with amphipathy and thus locate exclusively at polymer−polymer interfaces. Müller et al.32−34 achieved efficient compatibilization of immiscible polymer blends using premade JNPs. They found that JNPs with cross-linked PB as the core and PMMA/PS chains as the Janus shell are effective compatibilizers for PMMA/PS and SAN/PPE blends. The JNPs immigrated against high shear to anchor the polymer−polymer interface, which was attributed to their high interfacial activity. Nevertheless, both the complicated prefabrication of JNPs and their large content (usually ≥8 wt %) used in the strategy make polymer compatibilization extremely expensive. Meanwhile, the harsh compounding conditions involving high shear, ultrasound, or high pressure needed to anchor the JNPs to the interface impede the possible industrial application of this approach. Motivated by the formation of micelles in reactive polymer blending and the ability of JNPs to stabilize the polymer− polymer interface in blends, we expected that Janus nanomicelles (JNMs) formed by reactive blending should behave as effective compatibilizers for immiscible polymer blends. Scheme 1 shows that the premade reactive graft or comb polymers can readily react with a component polymer during melt mixing to form double graft/comb polymers. The main chain and two side chains are not miscible with each other, so reactive blending may form Janus micelles located exclusively at

lending polymers is a practical, efficient technique to produce novel materials with advantageous combinations of desired properties.1 Appropriate compatibility of the component pair could provide materials with uncontemplated properties and diverse applications.2−5 Paradoxically, the overwhelming majority of commercial polymer pairs is thermodynamically immiscible, so they exhibit poor mechanical properties because of the weak interactions between phases.6−9 Therefore, improvement of the compatibility of polymer pairs10 has attracted considerable attention in industry and academia. Reactive blending is one of the most accessible strategies to improve the compatibility of polymer blends.11−13 This technique is based on the in situ formation of block/graft copolymers at their interface,14−16 which substantially lowers polymer−polymer interface tension, suppresses particle coalescence, and ultimately improves interfacial adhesion.17 However, enhanced by shear force (pull-out or -in) during melt mixing, even graft copolymers with symmetric structure are readily converted into micelles in a single-like phase during reactive blending.18,19 As a result, only a proportion of copolymers at the polymer−polymer interface are involved in compatibilization. Obviously, the formation of micelles decreases the compatibilization efficiency of the reactive compatibilizers, which is a serious waste with regard to copolymer cost. Therefore, many studies have attempted to minimize the formation of micelles in reactive blending.20−23 Nevertheless, some questions arise: Are the micelles formed in conventional reactive blending of potential benefit? Do the micelles enhance polymer miscibility if the micelles are located at the interface? Janus nanoparticles (JNPs)24−26 with heterogeneous surface composition have recently been used as an emulsifying agent © 2015 American Chemical Society

Received: October 27, 2015 Accepted: November 23, 2015 Published: November 30, 2015 1398

DOI: 10.1021/acsmacrolett.5b00763 ACS Macro Lett. 2015, 4, 1398−1403

Letter

ACS Macro Letters Scheme 1. Proposed Formation of in Situ Formed JNMs for Compatibilizers in Immiscible Polymer Blends

poly(methyl methacrylate) (P((S-co-GMA)-g-MMA)) copolymer (Mn = 1.9 × 104, PDI = 4.3, Mn,PMMA = 6300, f PMMA = 1/ chain, GMA % = 20 wt %) via the “grafting-through” strategy (see Supporting Information (Tables S1−S2, Figures S1−S6)). The graft copolymer is incorporated in a PVDF/PLLA 50/50 blend by melt mixing at 190 °C using a batch mixer (Haake Polylab QC) with a rotation speed of 50 r/min. The PVDF with a Mn of 1.05 × 105 and a Mw/Mn of 2.0 and PLLA with a Mn of 8.93 × 104 and a Mw/Mn of 1.77 used in this text are provided by Kureha and Nature Works, respectively. The long PMMA side chains (Mw 14 500) are miscible with PVDF because of their CF2 and carbonyl groups.10,35 The COOH groups of PLLA react with epoxide groups of the graft copolymer during melt mixing, covalently grafting the long PLLA chains onto the PS backbone (see Figure S7). Note that the χ interaction parameters between PLLA, PS, and PMMA are high,36−39 indicating strong immiscibility between the PS backbone, PMMA side chains, and grafted PLLA chains. Figure 1 shows the phase morphologies of PLLA/PVDF blends with and without the reactive graft copolymers. The

the interface between the two phases. The Janus micelles should be able to compatibilize immiscible polymer blends. Such a one-step strategy may open a new possibility for polymer compatibilization. In this communication, we investigate the formation of JNMs by in situ reactive blending and the effects of JNMs on the structure and properties of immiscible PVDF/PLLA blends. We first synthesized a poly(styrene-co-glycidyl methacrylate)-graf t-

Figure 1. JNM compatibilized PVDF/PLLA blends via an in situ reaction of melt blending. (a)−(c) SEM images of (a) PVDF/PLLA (50/50) and (b),(c) PVDF/PLLA/P((S-co-GMA)-g-MMA) (50/50/3); (d)−(f) TEM images of (d) PVDF/PLLA (50/50) and (e),(f) PVDF/PLLA/P((S-coGMA)-g-MMA) (50/50/3). The blend samples are ultramicrotomed with a thickness of 80−90 nm and then stained by RuO4 for 4 h in order to selective stain the PVDF phase as well as the graft copolymer (the white phase is PLLA, the black phase is graft copolymer, and the gray phase is PVDF). (g) Scanning transmission electron microscopy (STEM) images illustrating the selective dispersion of nanomicelles (appearing white, stained by RuO4). (h),(i) Elemental mapping images (EMI) on PVDF/PLLA/P((S-co-GMA)-g-MMA) (50/50/3) blends showing the signals of fluorine and oxygen element, respectively. 1399

DOI: 10.1021/acsmacrolett.5b00763 ACS Macro Lett. 2015, 4, 1398−1403

Letter

ACS Macro Letters

Figure 2. TEM images (a) and SEM images (b) of the PVDF/PLLA blends with 3 phr P((S-co-GMA)-g-MMA) that melt mixed for the indicated time (along the rows) at different magnifications (along the column). The TEM images in row 4 show the micelles in the PLLA phase in different reaction times.

mixing time was 10 min for all samples in Figure 1. Both SEM image (Figure 1a) and TEM image (Figure 1d) indicate that PLLA and PVDF are totally immiscible, as evidenced by the large domain size and very weak interfacial bonding for the blends without a reactive graft copolymer. PVDF has higher density and viscosity than PLLA. Therefore, PVDF forms domains in the PLLA matrix even at 50/50 weight ratio. Incorporation of the 3 wt % graft copolymer changes the phase structure. A cocontinuous structure with a smaller phase size is formed (Figure 1b). The interface between the phases is also drastically improved with clear adhesion layers (Figure 1c). Interestingly, numerous nanomicelles (shown in black) are located at the interface as well as in the PLLA matrix (Figure 1e). The number density of micelles at the phase interface is much higher than that in PLLA, and the micelles form a dense layer at the interface between the PLLA and PVDF phases. Such a “cocontinuous encapsulated nanomicelles” blend

formed by simple reactive melt blending is unprecedented. The nanomicelles within PLLA and those at the interface possess different structures (Figure 1f). Compared with the nanomicelles in PLLA, the interfacial ones are larger with darker cores, which may be related to different self-assembly behavior (discussed in more detail later). To identify the chemical composition of nanomicelles, we performed elemental mapping image (EMI) analysis; the results are shown in Figure 1g−i. F signals are only observed in the PVDF domains, indicating that the micelles do not contain PVDF chains. This is further confirmed by the O EMI of the nanomicelles. Obviously, the micelles originate from the microphase separation of PLLA-grafted P((S-co-GMA)-gMMA) (double-grafted polymers). Thus, we speculate that the micelles located at the interface bridge the two phases and act as a compatibilizer. 1400

DOI: 10.1021/acsmacrolett.5b00763 ACS Macro Lett. 2015, 4, 1398−1403

Letter

ACS Macro Letters

Scheme 2. Schematic View of the in Situ Formation of JNMs and the Morphology Development of Binary Blends from “Droplet Stacked Domains” to “Cocontinuous Encapsulated Nanomicelles” via Reactive Blending

pulled into the PLLA phase to form the micelles observed in the PLLA phase; this process can also be considered as interfacial corrosion. For the sample prepared by melt mixing for 15 min, few micelles are observed at the PLLA−PVDF interface because almost all of the micelles have been pulled into the PLLA phase by almost full reaction. The results show that PVDF/PLLA blends could be emulsified through gradual formation of interfacial JNMs, as illustrated in Scheme 2. P((S-co-GMA)-g-MMA) contains many reactive epoxide groups (20 wt % GMA content) and long PMMA side chains. The epoxide groups readily react with COOH end groups of PLLA during melt mixing to form double-graft copolymers with PLLA and PMMA as side chains and PS as the backbone. Before PLLA is grafted onto P((S-coGMA)-g-MMA), P((S-co-GMA)-g-MMA) is located at the PLLA−PVDF interface as stacked domains (Scheme 2, State 1). Therefore, the reaction between PLLA and the graft copolymer can easily occur at the PLLA−PVDF interface. Through shearing force and the grafting reaction, the stacked domains break up to form small JNMs with both PMMA side chains and grafted PLLA chains (Scheme 2, State 2). The JNMs have a hemisphere structure with PMMA and PLLA chains on their two sides. The PMMA chains are miscible with PVDF because of the well-known specific interaction between CF2 and carboxyl groups, while the grafted PLLA chains of the micelles are readily entangled with themselves in the PLLA phase. Therefore, the JNMs are effective as compatibilizers of PLLA and PVDF. A distinct transition from island to lamellar morphology with lower circularity was observed as the reaction proceeded (Figure 1a,b). Presumably, the decrease in Laplace pressure caused by lowering of interfacial tension originating from the encapsulation of JNMs allows cocontinuous domains to form more readily as mixing time lengthens, thus explicating the morphology evolution detected. The micelles at the PLLA−PVDF interface are pulled into the PLLA phase under shear as more PLLA is grafted onto the micelles. Therefore, more micelles are observed in the PLLA phase as mixing time increases (Scheme 2, State 3). In addition, the structure of the micelles in the PLLA phase is different from that of the micelles at the PLLA−PVDF interface. The interfacial micelles have a Janus structure with PMMA and

To clarify the formation mechanism of nanomicelles and their ability to compatibilize PLLA/PVDF blends, we carried out TEM and SEM measurements for the compatibilized PVDF/PLLA blends prepared with different mixing times, as shown in Figure 2. We have focused on the interface between the PVDF and PLLA phases. Figure 2(a1−5) reveals large domains ranging from several hundred nanometers to two micrometers at the interface during the initial mixing stage before the reaction between glycidyl methacrylate (GMA) of P((S-co-GMA)-g-MMA) and PLLA occurs. P((S-co-GMA)-gMMA) immediately forms large droplets under shear, and the droplets are exclusively located at the interface between the two major phases before the reaction. This is a typical behavior for ternary polymer blends, where one component forms domains at the interface of the other two polymers because of its suitable spreading coefficient.40,41 As mixing time increases from 1 to 10 min, a morphological transition is observed as the reaction proceeds. The size of the domains at the interface decreases gradually. In addition, numerous micelles with a size of about 20 nm are observed in the PLLA phase, and their number increases with mixing time. After 10 min, a shell composed of numerous micelles formed between the PVDF and PLLA phases. At the same time, a large amount of small micelles in PLLA are also observed. The reaction between COOH end groups of PLLA and epoxy groups of copolymer was confirmed by Fourier transform infrared and torque data (Figures S7 and S8). As expected, double-grafted copolymers are generated quickly, which lower the interfacial tension between the copolymers and PLLA phase. As a result, the large droplets break up to form smaller ones. PS segments in the copolymers are rejected by both PLLA and PVDF phases. To minimize the interfacial energy, the copolymers are forced to locate at the PLLA−PVDF interface, where they act as a micelle-based emulsifier. Thus, the formation of “micelle capsules” occurs gradually. These micelles, which we term JNMs, may possess two distinct surface chemistries with one side rich in PMMA and the other in PLLA side chains to allow interfacial adhesion. The JNMs at the PLLA−PVDF interface contribute to the compatibilization of the immiscible PVDF/PLLA blends. As the PLLA chains gradually graft onto the micelles at the PLLA−PVDF interface, the double-grafted copolymers are 1401

DOI: 10.1021/acsmacrolett.5b00763 ACS Macro Lett. 2015, 4, 1398−1403

Letter

ACS Macro Letters PLLA as the two hemispheres and a PS backbone as the core. However, it is rational that the micelles in the PLLA phase should have a PLLA shell to mix with the bulk PLLA phase, a PMMA intermediate shell, and PS core due to higher χ of PLLA/PS than PLLA/PMMA.42,43 This speculation is consistent with the high-magnification TEM image in Figure 1f, which shows distinctive structural differences between nanomicelles in the PLLA phase and those at the PLLA−PVDF interface. Obviously, only the JNMs at the PLLA−PVDF interface are effective as compatibilizers of PLLA/PVDF blends. This may be the reason why the blend prepared by melt mixing for 15 min has lower interfacial bonding between PVDF and PLLA than that prepared by melt mixing for 10 min, which is evidenced by the poorer mechanical properties of the former sample compared with those of the latter (discussed in the next section). Notably, compared with the core−shell nanomicelles in the PLLA phase that display a discrete distribution, the JNMs at the PLLA−PVDF interface are positioned much closer together. Muller et al.25−27 compared the desorption energies of premade nanoparticles with Janus and homogeneous surface structures from the interface of an immiscible polymer blend. They concluded that the JNPs were 20 times less likely to immigrate from the interface than those with homogeneous surface structure. In our case, as PLLA was gradually grafted onto the micelles, the micelles immigrated from the interface into the PLLA phase. Such immigration is attributed to both dynamic shear and the increased interaction between the micelles and PLLA. Figure 3a and 3b depict the loss tangent and storage modulus (E′) curves as a function of temperature, respectively, for neat PVDF, neat PLLA, and compatibilized PVDF/PLLA blends prepared by melt mixing for different times. The relevant thermal parameters are given in Table S3. In the tan delta curves (Figure 3a), the change of glass transition temperature Tg (i.e., ΔTg) between PLLA and PVDF phases decreases markedly compared with that of the blend without compatibilizer. ΔTg of the PVDF/PLLA/graft (50/50/3) sample prepared by melt mixing for 10 min is much lower than that of other reactive blends, indicating improved compatibility of PVDF with PLLA. Furthermore, PLLA chains may entangle with PVDF chains at the interface, which might possibly explain the slightly shifted Tg of PLLA. Figure 3b reveals that E′ of neat PLLA and the blends decreases abruptly at 60 °C because of the glass transition, followed by an increase of E′ originating from the cold crystallization of PLLA.44−46 The sample with graft copolymer prepared by melt mixing for 10 min has the smallest E′ decrease in the range of 70−100 °C of the samples. A higher E′ means that the sample can be self-supported even at temperatures above the Tg of PLLA, which is attributed to the cocontinuous morphology of the blend. Both PLLA and PVDF are continuous phases because of the compatibilization effect of the JNMs. Therefore, the blend samples still exhibit high E′ because of the continuous PVDF phase when PLLA becomes soft above its Tg. The mechanical properties of PLLA, PVDF, and the blends are presented in Figure 4. Markedly enhanced ductility is observed for the blend containing the reactive graft copolymer prepared by melt mixing for 10 min. The elongation at break of this blend is 320%, compared with 5% for the blend without graft copolymer and 8% for the sample formed by melt mixing for 0.5 min. The substantially improved mechanical properties of the blend containing reactive graft copolymer indicate that

Figure 3. (a) Loss tangent (tan delta) and (b) storage modulus as a function of temperature, obtained by dynamic thermomechanical analysis (DMA) at 5 Hz.

Figure 4. Mechanical characterization of JNMs compatibilized with PVDF/PLLA blends and related materials, including neat PVDF, neat PLLA, and the free PVDF/PLLA (50/50) blend. Strain−stress curves of blends with various reaction time. The inset shows partial enlargement of the stress−strain curves.

the interfacial JNMs effectively compatibilize PVDF/PLLA blends. For the sample fabricated by melt mixing for 15 min, the elongation at break is about 140%, which is lower than that of the sample obtained following 10 min of melt mixing. This is 1402

DOI: 10.1021/acsmacrolett.5b00763 ACS Macro Lett. 2015, 4, 1398−1403

Letter

ACS Macro Letters

(12) Jahani, Y.; Ghetmiri, M.; Vaseghi, M. R. RSC Adv. 2015, 5, 21620. (13) Song, Y. H.; Xu, C. F.; Zheng, Q. Soft Matter 2014, 10, 2685. (14) Jeon, H. K.; Macosko, C. W.; Moon, B.; Hoye, T. R.; Yin, Z. H. Macromolecules 2004, 37, 2563. (15) Retsos, H.; Anastasiadis, S. H.; Pispas, S.; Mays, J. W.; Hadjichristidis, N. Macromolecules 2004, 37, 524. (16) Shi, H. C.; Shi, D. A.; Wang, X. Y.; Yin, L. G.; Yin, J. H.; Mai, Y. W. Polymer 2010, 51, 4958. (17) Eklind, H.; Schantz, S.; Maurer, F. H. J.; Jannasch, P.; Wesslén, B. Macromolecules 1996, 29, 984. (18) Adedeji, A.; Lyu, S.; Macoscko, C. W. Macromolecules 2001, 34, 8663. (19) Zhang, J. B.; Ji, S. X.; Song, J.; Lodge, T. P.; Macosko, C. W. Macromolecules 2010, 43, 7617. (20) Macosko, C. W.; Guégan, P.; Khandpur, A. K.; et al. Macromolecules 1996, 29, 5590. (21) Charoensirisomboon, P.; Inoue, T.; Weber, M. Polymer 2000, 41, 6907. (22) Pan, L. H.; Chiba, T.; Inoue, T. Polymer 2001, 42, 8825. (23) Pan, L. H.; Inoue, T.; Hayami, H.; Nishikawa, S. Polymer 2002, 43, 337. (24) Yan, L. T.; Popp, N.; Ghosh, S. K.; Böker, A. ACS Nano 2010, 4, 913. (25) Walther, A.; Müller, A. H. E. Chem. Rev. 2013, 113, 5194. (26) Lattuada, M.; Hatton, T. A. Nano Today 2011, 6, 286. (27) Walther, A.; Hoffmann, M.; Müller, A. H. E. Angew. Chem. 2008, 120, 723. (28) Xu, K. L.; Guo, R. H.; Dong, B. J.; Yan, L. T. Soft Matter 2012, 8, 9581. (29) Huang, M. X.; Li, Z. Q.; Guo, H. X. Soft Matter 2012, 8, 6834. (30) Binks, B. P. Curr. Opin. Colloid Interface Sci. 2002, 7, 21. (31) Pickering, S. U. J. Chem. Soc., Trans. 1907, 91, 2001. (32) Walther, A.; Matussek, K.; Müller, A. H. E. ACS Nano 2008, 2, 1167. (33) Bahrami, R.; Löbling, T. I.; Gröschel, A. H.; Schmalz, H.; Müller, A. H. E.; Altstädt, V. ACS Nano 2014, 8, 10048. (34) Bryson, K. C.; Löbling, T. I.; Müller, A. H. E.; Russell, T. P.; Hayward, R. C. Macromolecules 2015, 48, 4220. (35) Nishi, T.; Wang, T. T. Macromolecules 1975, 8, 909. (36) Shirahase, T.; Komatsu, Y.; Tominaga, Y.; Asai, S.; Sumita, M. Polymer 2006, 47, 4839. (37) Eguiburu, J. L.; Iruin, J. J.; Fernandez-Berridi, M. J.; Sanroman, J. Polymer 1998, 39, 6891. (38) Russell, T. P.; Hjelm, R. P.; Seeger, P. A. Macromolecules 1990, 23, 890. (39) Yuan, Z.; Favis, B. D. Biomaterials 2004, 25, 2161. (40) Virgilio, N.; Desjardins, P.; Espérance, G. L.; Favis, B. D. Macromolecules 2009, 42, 7518. (41) Horiuchi, S.; Matchariyakul, N.; Yase, K.; Kitano, T. Macromolecules 1997, 30, 3664. (42) She, M. S.; Lo, T. Y.; Ho, R. M. Macromolecules 2014, 47, 175. (43) Samuel, C.; Barrau, S.; Lefebvre, J.-M.; Raquez, J.-M.; Dubois, P. Macromolecules 2014, 47, 6791. (44) Bai, H. W.; Xiu, H.; Gao, J.; Deng, H.; Zhang, Q.; Yang, M. B.; Fu, Q. ACS Appl. Mater. Interfaces 2012, 4, 897. (45) Tsuji, H.; Sawada, M.; Bouapao, L. ACS Appl. Mater. Interfaces 2009, 1, 1719. (46) Zhou, D. D.; Shao, J.; Li, G.; Sun, J. R.; Bian, X. C.; Chen, X. S. Polymer 2015, 62, 70.

because almost all nanomicelles have been pulled into the PLLA phase, so fewer JNMs are present to compatibilize the PLLA and PVDF phases. In conclusion, we present a novel compatibilization strategy by interfacial micelle compatibilized immiscible polymer blends using in situ formed nanomicelles with Janus structure. The JNMs formed from the stacked droplets via an in situ reaction during melt mixing. The nanomicelles located at the polymer− polymer interface acted as an effective compatibilizer for an immiscible polymer blend. In consequence, a high-performance PVDF/PLLA (50/50) blend with superior mechanical strength and thermodynamic properties was obtained by controlling reaction time. This novel approach reveals new opportunities to exploit interfacial emulsification using nanomicelles, which is of great significance for engineering plastic alloys. Although we did not achieve JNMs located exclusively at the interface of immiscible polymer blends as compatibilizers, this should be possible through careful design of the molecular structure of the reactive graft copolymer (length of PMMA side chains, density of reactive epoxide groups, and molecular weight of PS backbone) and processing conditions.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsmacrolett.5b00763. Synthetic details of P((S-co-GMA)-g-MMA) and additional PVDF/PLLA/graft copolymer characterizations (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Fax: +86 571 28867899. Tel.: +86 571 28867026. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was financially supported by the National Natural Science Foundation of China (51173036, 21374027) and Program for New Century Excellent Talents in University (NCET-13-0762).



REFERENCES

(1) Utracki, L. A. Commercial Polymer Blends; Chapman and Hall: London, 1998. (2) Li, Y. J.; Shimizu, H. ACS Appl. Mater. Interfaces 2009, 1, 1650. (3) Singh, A. K.; Prakash, R.; Pandey, D. RSC Adv. 2013, 3, 15411. (4) Song, P. G.; Shen, Y.; Du, B. X.; Peng, M.; Shen, L.; Fang, Z. P. ACS Appl. Mater. Interfaces 2009, 1, 452. (5) Cao, Y. W.; Zhang, J.; Feng, J. C.; Wu, P. ACS Nano 2011, 5, 5920. (6) Xu, Y. W.; Loi, J.; Delgado, P.; Topolkaraev, V.; Mceneany, R. J.; Macosko, C. W.; Hillmyer, M. Ind. Eng. Chem. Res. 2015, 54, 6108. (7) Freluche, M.; Iliopoulos, I.; Flat, J. J.; Ruzette, A. V.; Leibler, L. Polymer 2005, 46, 6554. (8) Leibler, L. Prog. Polym. Sci. 2005, 30, 898. (9) Dong, W. Y.; Wang, H. T.; He, M. F.; Ren, F. L.; Wu, T.; Zheng, Q. R.; Li, Y. J. Ind. Eng. Chem. Res. 2015, 54, 2081. (10) Ruzette, A. V.; Leibler, L. Nat. Mater. 2005, 4, 19. (11) Ojijo, V.; Ray, S. S.; Sadiku, R. ACS Appl. Mater. Interfaces 2013, 5, 4266. 1403

DOI: 10.1021/acsmacrolett.5b00763 ACS Macro Lett. 2015, 4, 1398−1403