Scalable Fabrication of Polydopamine Nanotubes Based on Curcumin

Mar 30, 2016 - Herein we for the first time demonstrated a scalable and simple fabrication of polydopamine nanotubes by using curcumin crystal as temp...
1 downloads 0 Views 4MB Size
Letter pubs.acs.org/journal/abseba

Scalable Fabrication of Polydopamine Nanotubes Based on Curcumin Crystals Junhui Xue,‡ Weichao Zheng,‡ Le Wang, and Zhaoxia Jin* Department of Chemistry, Renmin University of China, No. 59, Zhongguancun Street, Haidian District, Beijing 100872, P. R. China S Supporting Information *

ABSTRACT: Herein we for the first time demonstrated a scalable and simple fabrication of polydopamine nanotubes by using curcumin crystal as templates. BET surface area of obtained polydopamine nanotubes is 51.9 m2/g. Polydopamine nanotubes show great potential as biocompatible porous nanomaterials.

KEYWORDS: polydopamine, nanotubes, curcumin, template-fabrication

P

fabrication cost. (2) There are only a few reports of onedimensional polydopamine nanostructures fabrication in literatures, mainly based on electrospinning technique.4,42 One-dimensional nanotubes with large aspect ratio possess advantages in internalization by cells, leading to a good performance as drug delivery system44,45 and biosensors inside cells. Polydopamine 1D nanostructures also have advanced the study of nerve tissue engineering. A low-cost fabrication of 1D polydopamine nanostructures will benefit their emerging applications. Herein we for the first time demonstrate a unique, easily scalable fabrication method to produce polydopamine nanotube. Their length is varied from several to several tens of micrometers. Obtained hollow polydopamine nanotubes show significantly improved Brunauer−Emmett−Teller (BET) surface area 51.9 m2/g, and mixed porous structure including mesoporous (∼20 nm) and microporous (∼2 nm) parts. In a typical experiment, 100 mg of curcumin and 500 mg of dopamine were first dispersed in 100 mL of ethanol/acetone mixed solution (v/v 1:1), and then 400 mL water was added. Because curcumin is insoluble in water, the solvent exchange induces the precipitation of long curcumin crystals. Pure curcumin crystals generated via solvent exchange were presented in Figure S1. When the precipitation was completed, 3 mL Tris-HCl buffer solution (1.5 M, pH 8.8) was added to keep its concentration at 9 mM. The solution was kept for 24 h under tender stirring and its color gradually becomes dark. After desired reaction time, the sediment (PDA@curcumin)

olydopamine and its copolymers have attracted growing attention in the past decade because they combine biocompatibility and unique optoelectronic properties together, showing great potential as bioelectronic devices, biosensors, bioimaging agents, and drug delivery in biomedical area, tissue engineering, and photothermal therapy.1−19 Universal adhesion of polydopamine further expands its application as multifunctional surface-modified platforms, which demonstrate versatile applications not only in biomedical areas but also in environmental science, such as catalyst carriers for decomposing organic pollutants20−23 or efficient adsorbents for water purification.24 In addition, polydopamine-derived carbon nanomaterials also show their potential applications in energy storage, solar water splitting, and various engineering fields.25 For these applications, hollow capsules with enlarged surface area will greatly improve their performance, for instance, the sensitivity of biosensors, loading capacity for target drugs, catalysts or pollutants. Besides, hollow capsules can be used as unique nanoreactors.22,26−28 However, fabrication methods of polydopamine capsules are limited, which are mainly through sacrificial templates such as silica, CaCO3 or polymer microparticles.22,27,29−39 Using emulsion as soft templates is an alternative fabrication path of polydopamine hollow capsules.39−42 By using tetrahydrofuran-tris buffer mixture as soft template, polydopamine hollow capsules have also been produced.40 Electronspun polymer micro- or nanofibers are often used as templates to further coating polydopamine, thus generating one-dimensional polydopamine nanostructure.41−43 However, main problems hampering the practical application of polydopamine capsules are as follows: (1) The removing of hard template requires harsh experimental conditions, such as alkali or HF that limits their scalable production, and increases © XXXX American Chemical Society

Received: February 18, 2016 Accepted: March 30, 2016

A

DOI: 10.1021/acsbiomaterials.6b00102 ACS Biomater. Sci. Eng. XXXX, XXX, XXX−XXX

Letter

ACS Biomaterials Science & Engineering

polydopamine nanotubes because many self-polymerized polydopamine nanoparticles are mixed with nanotubes (Figure S3). The in situ formation of curcumin crystal is critical in this fabrication, because the existence of ethanol and acetone in solution effectively hampers the formation of self-polymerized polydopamine nanoparticles. In addition, the use of mixed solvent (ethanol/acetone) also improves the quality of obtained polydopamine nanotubes, compared with using single solvent (ethanol), in which produced PDA nanotubes have more defects (Figure S4). However, curcumin crystals generated from acetone/water are not uniform, the PDA product using acetone as solvent showed mixed morphologies (Figure S5). The yields of polydopamine nanotubes from mixed solvent (ethanol/acetone) or single solvent (ethanol) are comparable as shown in Figure S6. The morphological details varied in different solvent systems can be utilized to modulate the surface areas and pore-sizes of polydopamine nanotubes. Scheme 1 demonstrated the fabrication process of PDA nanotubes based on curcumin crystals template. The most

was separated from its solution, then cleaned by fresh water three times and collected by sedimentation or centrifugation. Then PDA@curcumin nanostructures were soaked in ethanol to dissolve curcumin crystal template, pure PDA nanotubes are obtained. Figure 1 presents SEM and TEM images of obtained PDA nanotubes with long (several tens of micrometers) or short

Scheme 1. Illustration of the Fabrication Process of PDA Nanotubes

Figure 1. (a−c) SEM and TEM images of long PDA nanotubes generated from PDA@curcumin composites. Their length-scale is over several-tens micrometers. (d−f) SEM and TEM images of slightly short PDA nanotubes. Their length is several micrometers.

(several micrometers) length. We found that severe stirring in removing curcumin template will reduce the tube length, from several-tens micrometers (Figure 1a−c) to several micrometers (Figure 1d−f). On the basis of magnified SEM and TEM images, we observed that tube-wall thickness is around 40 nm, the diameter of PDA nanotubes is 200−400 nm depending on crystal size (Figure 1c, f, Figure S2). Compared with PDA capsules reported in literatures, the mechanical strength of these PDA nanotubes is high enough to avoid structural collapse, no matter how long these nanotubes are. This may be the result of thick tubewall (∼40 nm) of PDA nanotubes. It is worth mentioning that this thickness of PDA nanotubes can be achieved in a single polymerization-cycle. In literatures, the thickness of PDA coating on silica particles is about 19 nm,29 and thick shell (∼75 nm) requires three polymerization cycles.39 Although the formation of polydopamine coating on curcumin crystals happens in several hours, the generation of polydopamine nanotubes with perfect structures requires more than 12 h. The formation kinetics is shown in Figure S2, in which polydopamine nanotubes obtained from different reaction times (from 2.5 to 24 h) were compared. Generally, we chose 24 h for reaction time to produce PDA nanotubes with perfect structures, and the yield of PDA nanotubes is over 25%. Used curcumin can be recycled through recrystallization from ethanol, and over 80% curcumin can be recovered. If curcumin crystals are directly used as template to produce polydopamine nanotubes, it will be hard to fabricate clean

important advantage of this fabrication is its being easily handled and scalable. In particular, this study demonstrates a development for interfacial polymerization of dopamine. Previous studies utilized oil/water, air/water, and liquid/solid interface to attract oxidative self-polymerization of dopamine. The in situ nucleation and growth of crystals creates a fresh liquid/solid interface which also induces adhesion and polymerization of dopamine. This study demonstrated that in situ formed interface effectively attracts polydopamine coating. We further conducted chemical and structural characterization of obtained PDA nanotubes. Figure 2 presented XPS characterization (Figure 2a−c), Raman spectrum (Figure 2d), BET surface area and pore size distribution of PDA nanotubes (Figure 2e, f). The N/C ratio in PDA nanotube is 0.12 based on XPS survey scan. C 1s high-resolution spectrum was fitted into three components: CHx or CC (284.6 eV, 51.2%), C− N (285.8 eV, 40.5%,) and CO (287.7 eV, 8.2%). The N 1s peak was fitted with three components: primary (R-NH2, 401.8 eV, 4.7%), secondary (R1-NH-R2, 399.9 eV, 89.9%) and tertiary/aromatic (CN-R, 398.6 eV, 5.4%) amine functionalities. The content of secondary amine is very high that indicates its formation follows cyclized self-polymerization path via 5, 6-dihydroxyindole.46 Because of the heterogeneous polymeric nature and fluorescent background, Raman characB

DOI: 10.1021/acsbiomaterials.6b00102 ACS Biomater. Sci. Eng. XXXX, XXX, XXX−XXX

Letter

ACS Biomaterials Science & Engineering

Previous studies presented that DOPA or polydopamine coating on nanoparticles has advantages of increasing their intracellular delivery and elongating retention in tumor.44,45 Caruso et al. have tested the cytotoxicity of PDA capsules generated on the basis of silica particle’s template.29 They have predicted that PDA capsules will be convenient and inexpensive candidates for therapeutic delivery. In addition, polydopamine coating has been proved to be a good UV-shielding, which is particularly important for UV-irradiation- instable drugs, such as avermectin58 and doxorubicin. Here we use doxorubicin hydrochloride (DOX) as model drugs to study the loading and releasing of DOX@PDA nanotubes. Based on the surface charge of PDA nanotubes (Figure S9), we found that in a neutral environment, negative PDA nanotubes have strong electrostatic interaction with positive DOX. In aqueous solution the loading capacity of DOX in PDA nanotubes was 24.2%. The electrostatic interaction will be weakened in acidic environment, in which the surface charge of PDA nanotubes turns to positive. We have tested the releasing behaviors of DOX@PDA nanotubes in different buffered saline at three pH values: 4, 5, and 7.4 (Figure 3). At neutral condition (pH 7.4),

Figure 2. (a−c) XPS of PDA nanotubes. (a) Full scan, (b) C 1s highresolution spectrum, (c) N 1s high-resolution spectrum. (d) Raman spectrum of polydopamine nanotubes. (e) BET surface area of PDA nanotubes. (f) Pore size distribution of PDA nanotubes.

terization of polydopamine and melenins is difficult and there are few reports in the literature.47−50 In this polydopamine nanotube sample, broad and strong peaks at 1300−1600 cm−1 were recorded (Figure 2d). The peak centered at 1579 cm−1 can be assigned to ν(CC) aromatic coupled with pyrrole ring stretching vibration or indole ring vibration. Another peak at 1355 cm−1 is due to the aromatic ν(C−N) stretching mixed with indole ring stretching.47 They partly represent the structure feature of monomers composing polydopamine,50 and they are in agreement with the reported values in literatures.47,48 FTIR and XRD of PDA nanotubes were also characterized (Figures S7 and S8). These characterizations confirmed that in polydopamine nanotubes, the cyclized indole (FTIR spectrum 1592, 1514, and 1343 cm−1) works as structural units,47 and there is π−π interaction connecting their oligomer’s components.51−54 Zeta potential of PDA nanotubes is listed in Figure S9. The surface charge of PDA nanotubes is varied with pH value, from +42.72 mV (pH 2) to −35.86 mV (pH 8.5). Their isoelectric point is at pH 4.41. BET surface area is an important parameter for porous nanostructures. PDA nanotube (short) shows surface area 51.9 m2/g. This value is much higher than that of natural and synthetic melanin nanoparticles reported in literatures (9.2−19.9 m2/ g).55 It also exceeds the CO2-supercritically dried Sepia eumelanin (37.5 m2/g)56 and melanin samples after special treatments (29−32 m2/g).57 On the other hand, it is worth noting that the pore-size-distribution of PDA nanotubes has a mesoporous part at ∼20 nm and a microporous part at ∼2 nm. The mesoporous part is due to the inner lumen, and the microporous part may be due to microporous structure on tubewalls. This is the first report of existence of microporous part in polydopamine nanostructure, which may further enrich their applications as porous materials.

Figure 3. Release kinetics of DOX@PDA nanotubes at different pH values.

the release amount of DOX is only about 2% within 30 h. However, the cumulative releases of DOX significantly increased at pH 5 and pH 4 environments, rising to 10 and 20%, respectively. Bettinger et al. have observed that melaninbased implants are fully degraded after 8 weeks.12 If these PDA nanotubes can be degraded in several months, they will be indeed an ideal candidate for drug delivery with UV-protecting, pH- and degradation-controlled release. Besides, the successful fabrication of polydopamine nanotubes provides us an opportunity for studying their applications in tissue engineering. Porous structures of polydopamine nanotubes are suitable as scaffolds to culture cells; the surface functional groups of polydopamine nanotubes make the postmodification of scaffolds easier; and the large aspect ratio of polydopamine nanotubes endows them advantages of being easily aligned. In summary, we have presented a simple and scalable fabrication of polydopamine nanotubes with tunable lengths. These PDA nanotubes showed largest surface area (51.9 m2/g) compared with that of reported polydopamine nanostructures. PDA nanotubes present both the microporous and mesoporous features. Doxorubicin was used as model drug to be loaded on PDA nanotubes and its releasing behaviors were studied in vitro. A pH-related release of DOX@PDA was demonstrated. The strong affinity and adsorption of polydopamine nanotubes to versatile materials endow it great potential in improving the load of drugs and enhancing sensitivity of biosensors. In addition, microporous and mesoporous property of polydopC

DOI: 10.1021/acsbiomaterials.6b00102 ACS Biomater. Sci. Eng. XXXX, XXX, XXX−XXX

Letter

ACS Biomaterials Science & Engineering

for a Hybrid Biomimetic interface. J. Mater. Chem. C 2014, 2, 573− 582. (11) Lynge, M. E.; van der Westen, R.; Postma, A.; Städler, B. PolydopamineA Nature-inspired Polymer Coating for Biomedical Science. Nanoscale 2011, 3, 4916−4928. (12) Bettinger, C. J.; Bruggeman, J. P.; Misra, A.; Borenstein, J. T.; Langer, R. Biocompatibility of Biodegradable Semiconducting Melanin Films for Nerve Tissue Engineering. Biomaterials 2009, 30, 3050− 3057. (13) Zheng, R.; Wang, S.; Tian, Y.; Jiang, X.; Fu, D.; Shen, S.; Yang, W. Polydopamine-Coated Magnetic Composite Particles with an Enhanced Photothermal Effect. ACS Appl. Mater. Interfaces 2015, 7, 15876−15884. (14) Nurunnabi, M.; Khatun, Z.; Nafiujjaman, M.; Lee, D. G.; Lee, Y. K. Surface Coating of Graphene Quantum Dots Using Mussel-inspired Polydopamine for Biomedical Optical Imaging. ACS Appl. Mater. Interfaces 2013, 5, 8246−8253. (15) Lin, L. S.; Cong, Z. X.; Cao, J. B.; Ke, K. M.; Peng, Q. L.; Gao, J.; Yang, H. H.; Liu, G.; Chen, X. Multifunctional Fe3O4@ Polydopamine Core−shell Nanocomposites for Intracellular mRNA Detection and Imaging-guided Photothermal Therapy. ACS Nano 2014, 8, 3876−3883. (16) Lv, R.; Li, Q.; Botello-Méndez, A. R.; Hayashi, T.; Wang, B.; Berkdemir, A.; Hao, Q.; Elías, A. L.; Cruz-Silva, R.; Gutiérrez, H. R.; Kim, Y. A.; Muramatsu, H.; Zhu, J.; Endo, M.; Terrones, H.; Charlier, J. C.; Pan, M.; Terrones, M. Nitrogen-doped Graphene: Beyond Single Substitution and Enhanced Molecular Sensing. Sci. Rep. 2012, 2, 586. (17) Liu, Y.; Ai, K.; Lu, L. Polydopamine and Its Derivative Materials: Synthesis and Promising Applications in Energy, Environmental, and Biomedical Fields. Chem. Rev. 2014, 114, 5057−5115. (18) Park, J.; Brust, T. F.; Lee, H. J.; Lee, S. C.; Watts, V. J.; Yeo, Y. Polydopamine-based Simple and Versatile Surface Modification of Polymeric Nano Drug Carriers. ACS Nano 2014, 8, 3347−3356. (19) Madhurakkat Perikamana, S. K.; Lee, J.; Lee, Y. B.; Shin, Y. M.; Lee, E. J.; Mikos, A. G.; Shin, H. Materials from Mussel-Inspired Chemistry for Cell and Tissue Engineering Applications. Biomacromolecules 2015, 16, 2541−2555. (20) Liu, R.; Guo, Y.; Odusote, G.; Qu, F.; Priestley, R. D. Core− shell Fe3O4 Polydopamine Nanoparticles Serve Multipurpose as Drug Carrier, Catalyst Support and Carbon Adsorbent. ACS Appl. Mater. Interfaces 2013, 5, 9167−9171. (21) Xie, Y.; Yan, B.; Xu, H.; Chen, J.; Liu, Q.; Deng, Y.; Zeng, H. Highly Regenerable Mussel-Inspired Fe3O4@ Polydopamine-Ag Core−Shell Microspheres as Catalyst and Adsorbent for Methylene Blue Removal. ACS Appl. Mater. Interfaces 2014, 6, 8845−8852. (22) Yu, B.; Wang, D. A.; Ye, Q.; Zhou, F.; Liu, W. Robust Polydopamine Nano/Microcapsules and Their Loading and Release Behavior. Chem. Commun. 2009, 6789−6791. (23) Zhou, J.; Duan, B.; Fang, Z.; Song, J.; Wang, C.; Messersmith, P. B.; Duan, H. Interfacial Assembly of Mussel-Inspired Au@ Ag@ Polydopamine Core-Shell Nanoparticles for Recyclable Nanocatalysts. Adv. Mater. 2014, 26, 701−705. (24) Gao, H.; Sun, Y.; Zhou, J.; Xu, R.; Duan, H. Mussel-Inspired Synthesis of Polydopamine-Functionalized Graphene Hydrogel as Reusable Adsorbents for Water Purification. ACS Appl. Mater. Interfaces 2013, 5, 425−432. (25) Kong, J.; Shahabadi, S. I. S.; Lu, X. Integration of Inorganic Nanostructures with Polydopamine-Derived Carbon: Tunable Morphologies and Versatile Applications. Nanoscale 2016, 8, 1770−1788. (26) Gaitzsch, J.; Huang, X.; Voit, B. Engineering Functional Polymer Capsules toward Smart Nanoreactors. Chem. Rev. 2016, 116, 1053. (27) Wang, Z.; Li, C.; Xu, J.; Wang, K.; Lu, X.; Zhang, H.; Qu, S.; Zhen, G.; Ren, F. Bioadhesive Microporous Architectures by SelfAssembling Polydopamine Microcapsules for Biomedical Applications. Chem. Mater. 2015, 27, 848−856. (28) Zhang, C.; Yang, H. C.; Wan, L. S.; Liang, H. Q.; Li, H.; Xu, Z. K. Polydopamine-Coated Porous Substrates as a Platform for Mineralized β-FeOOH Nanorods with Photocatalysis under Sunlight. ACS Appl. Mater. Interfaces 2015, 7, 11567−11574.

amine also benefits the differential exchange of materials inside and outside PDA nanotubes depending on their sizes. The simplicity, mass production, and low cost of this fabrication of PDA nanotubes will stimulate their applications not only in biomedical nanotechnology but also in environmental science and energy field.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsbiomaterials.6b00102. Experimental section and supplementary figures (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Author Contributions ‡

J.H. Xue and W.C. Zheng contributed equally.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors gratefully acknowledge the National Natural Science Foundation of China (Grants 21374132, 51173201) for financial support.



REFERENCES

(1) Wünsche, J.; Cicoira, F.; Graeff, C. F.; Santato, C. Eumelanin Thin Films: Solution-processing, Growth, and Charge Transport Properties. J. Mater. Chem. B 2013, 1, 3836−3842. (2) Mostert, A. B.; Powell, B. J.; Pratt, F. L.; Hanson, G. R.; Sarna, T.; Gentle, I. R.; Meredith, P. Role of Semiconductivity and Ion Transport in the Electrical Conduction of Melanin. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 8943−8947. (3) Ambrico, M.; Ambrico, P. F.; Ligonzo, T.; Cardone, A.; Cicco, S. R.; d’Ischia, M.; Farinola, G. M. From Commercial Tyrosine Polymers to a Tailored Polydopamine Platform: Concepts, Issues and Challenges en Route to Melanin-based Bioelectronics. J. Mater. Chem. C 2015, 3, 6413−6423. (4) Yu, X.; Fan, H.; Wang, L.; Jin, Z. Formation of Polydopamine Nanofibers with the Aid of Folic Acid. Angew. Chem., Int. Ed. 2014, 53, 12600−12604. (5) Ambrico, M.; Ambrico, P. F.; Cardone, A.; Della Vecchia, N. F.; Ligonzo, T.; Cicco, S. R.; Talamo, M. M.; Napolitano, A.; Augelli, V.; Farinola, G. M. Engineering Polydopamine Films with Tailored Behaviour For Next-generation Eumelanin-related Hybrid Devices. J. Mater. Chem. C 2013, 1, 1018−1028. (6) Mostert, A. B.; Powell, B. J.; Gentle, I. R.; Meredith, P. On the Origin of Electrical Conductivity In the Bio-electronic Material Melanin. Appl. Phys. Lett. 2012, 100, 093701. (7) Ambrico, M.; Ambrico, P.; Ligonzo, T.; Cardone, A.; Cicco, S.; Lavizzera, A.; Augelli, V.; Farinola, G. Memory-like Behavior as a Feature of Electrical Signal Transmission in Melanin-like Biopolymers. Appl. Phys. Lett. 2012, 100, 253702. (8) Wünsche, J.; Cardenas, L.; Rosei, F.; Cicoira, F.; Gauvin, R.; Graeff, C. F.; Poulin, S.; Pezzella, A.; Santato, C. In Situ Formation of Dendrites in Eumelanin Thin Films between Gold Electrodes. Adv. Funct. Mater. 2013, 23, 5591−5598. (9) Meredith, P.; Bettinger, C.; Irimia-Vladu, M.; Mostert, A.; Schwenn, P. Electronic and Optoelectronic Materials and Devices Inspired by Nature. Rep. Prog. Phys. 2013, 76, 034501. (10) Ambrico, M.; Ambrico, P. F.; Cardone, A.; Cicco, S. R.; Palumbo, F.; Ligonzo, T.; Di Mundo, R.; Petta, V.; Augelli, V.; Favia, P. Melanin-like Polymer Layered on a Nanotextured Silicon Surface D

DOI: 10.1021/acsbiomaterials.6b00102 ACS Biomater. Sci. Eng. XXXX, XXX, XXX−XXX

Letter

ACS Biomaterials Science & Engineering

Through a Quinone Control Point. Adv. Funct. Mater. 2013, 23, 1331−1340. (47) Centeno, S. A.; Shamir, J. Surface Enhanced Raman Scattering(SERS) and FTIR Characterization of the Sepia Melanin Pigment Used in Works of Art. J. Mol. Struct. 2008, 873, 149−159. (48) Capozzi, V.; Perna, G.; Gallone, A.; Biagi, P.; Carmone, P.; Fratello, A.; Guida, G.; Zanna, P.; Cicero, R. Raman and Optical Spectroscopy of Eumelanin Films. J. Mol. Struct. 2005, 744-777, 717− 721. (49) Powell, B.; Baruah, T.; Bernstein, N.; Brake, K.; McKenzie, R. H.; Meredith, P.; Pederson, M. A First-Principles Density-Functional Calculation of the Electronic and Vibrational Structure of the Key Melanin Monomers. J. Chem. Phys. 2004, 120, 8608−8615. (50) Roldán, M. L.; Centeno, S. A.; Rizzo, A. An Improved Methodology for the Characterization and Identification of Sepia in Works of Art by Normal Raman and SERS, Complemented by FTIR, Py-GC/MS, and XRF. J. Raman Spectrosc. 2014, 45, 1160−1171. (51) Clancy, C. M.; Simon, J. D. Ultrastructural Organization of Eumelanin from Sepia Officinalis Measured by Atomic Force Microscopy. Biochemistry 2001, 40, 13353−13360. (52) Capozzi, V.; Perna, G.; Carmone, P.; Gallone, A.; Lastella, M.; Mezzenga, E.; Quartucci, G.; Ambrico, M.; Augelli, V.; Biagi, P.; et al. Optical and Photoelectronic Properties of Melanin. Thin Solid Films 2006, 511-512, 362−366. (53) Cheng, J.; Moss, S. C.; Eisner, M.; Zschack, P. X-Ray Characterization of MelaninsI. Pigm. Cell Res. 1994, 7, 255−262. (54) Cheng, J.; Moss, S. C.; Eisner, M.; Zschack, P. X-Ray Characterization of MelaninsII. Pigm. Cell Res. 1994, 7, 263−273. (55) Kim, Y. J.; Wu, W.; Chun, S. E.; Whitacre, J. F.; Bettinger, C. J. Biologically Derived Melanin Electrodes in Aqueous Sodium-Ion Energy Storage Devices. Proc. Natl. Acad. Sci. U. S. A. 2013, 110, 20912−20917. (56) Liu, Y.; Simon, J. D. The Effect of Preparation Procedures on the Morphology of Melanin from the Ink Sac of Sepia Officinalis. Pigm. Cell Res. 2003, 16, 72−80. (57) Liu, Y.; Hong, L.; Kempf, V. R.; Wakamatsu, K.; Ito, S.; Simon, J. D. Ion-Exchange and Adsorption of Fe(III) by Sepia Melanin. Pigm. Cell Res. 2004, 17, 262−269. (58) Sheng, W. B.; Li, W.; Zhang, G. X.; Tong, Y. B.; Liu, Z. Y.; Jia, X. Study on the UV-shielding and Controlled-Release Properties of a Polydopamine Coating for Avermectin. New J. Chem. 2015, 39, 2752− 2757.

(29) Postma, A.; Yan, Y.; Wang, Y.; Zelikin, A. N.; Tjipto, E.; Caruso, F. Self-polymerization of Dopamine as a Versatile and Robust Technique to Prepare Polymer Capsules. Chem. Mater. 2009, 21, 3042−3044. (30) Cui, J.; Yan, Y.; Such, G. K.; Liang, K.; Ochs, C. J.; Postma, A.; Caruso, F. Immobilization and Intracellular Delivery of an Anticancer Drug Using Mussel-Inspired Polydopamine Capsules. Biomacromolecules 2012, 13, 2225−2228. (31) Hosta-Rigau, L.; York-Duran, M. J.; Zhang, Y.; Goldie, K. N.; Städler, B. Confined Multiple Enzymatic(Cascade) Reactions within Poly(dopamine)-based Capsosomes. ACS Appl. Mater. Interfaces 2014, 6, 12771−12779. (32) Zhang, Y.; Teo, B. M.; Goldie, K. N.; Städler, B. Poly(Nisopropylacrylamide)/Poly(dopamine) Capsules. Langmuir 2014, 30, 5592−5598. (33) Chen, X.; Yan, Y.; Müllner, M.; van Koeverden, M. P.; Noi, K. F.; Zhu, W.; Caruso, F. Engineering Fluorescent Poly(dopamine) Capsules. Langmuir 2014, 30, 2921−2925. (34) Ochs, C. J.; Hong, T.; Such, G. K.; Cui, J.; Postma, A.; Caruso, F. Dopamine-Mediated Continuous Assembly of Biodegradable Capsules. Chem. Mater. 2011, 23, 3141−3143. (35) Sheng, W.; Li, W.; Li, B.; Li, C.; Xu, Y.; Guo, X.; Zhou, F.; Jia, X. Mussel-Inspired Photografting on Colloidal Spheres: A Generalized Self-Template Route to Stimuli-Responsive Hollow Spheres for Controlled Pesticide Release. Macromol. Rapid Commun. 2015, 36, 1640−1645. (36) Kang, S.; Baginska, M.; White, S. R.; Sottos, N. R. Core−Shell Polymeric Microcapsules with Superior Thermal and Solvent Stability. ACS Appl. Mater. Interfaces 2015, 7, 10952−10956. (37) Zhang, L.; Shi, J.; Jiang, Z.; Jiang, Y.; Meng, R.; Zhu, Y.; Liang, Y.; Zheng, Y. Facile Preparation of Robust Microcapsules by Manipulating Metal-Coordination Interaction Between Biomineral Layer and Bioadhesive Layer. ACS Appl. Mater. Interfaces 2011, 3, 597−605. (38) Shi, J.; Yang, C.; Zhang, S.; Wang, X.; Jiang, Z.; Zhang, W.; Song, X.; Ai, Q.; Tian, C. Polydopamine Microcapsules with Different Wall Structures Prepared by a Template-Mediated Method for Enzyme Immobilization. ACS Appl. Mater. Interfaces 2013, 5, 9991− 9997. (39) Cui, J.; Wang, Y.; Postma, A.; Hao, J.; Hosta-Rigau, L.; Caruso, F. Monodisperse Polymer Capsules: Tailoring Size, Shell Thickness, and Hydrophobic Cargo Loading via Emulsion Templating. Adv. Funct. Mater. 2010, 20, 1625−1631. (40) Ni, Y. Z.; Jiang, W. F.; Tong, G. S.; Chen, J. X.; Wang, J.; Li, H. M.; Yu, C. Y.; Huang, X. H.; Zhou, Y. F. Preparation of Polydopamine Nanocapsules in a Miscible Tetrahydrofuran−Buffer Mixture. Org. Biomol. Chem. 2015, 13, 686−690. (41) Kong, J.; Zhao, C.; Wei, Y.; Phua, S. L.; Dong, Y.; Lu, X. Nanocups-on-Microtubes: A Unique Host Towards High-Performance Lithium Ion Batteries. J. Mater. Chem. A 2014, 2, 15191−15199. (42) Kong, J.; Yee, W. A.; Wei, Y.; Yang, L.; Ang, J. M.; Phua, S. L.; Wong, S. Y.; Zhou, R.; Dong, Y.; Li, X. Silicon Nanoparticles Encapsulated in Hollow Graphitized Carbon Nanofibers for Lithium Ion Battery Anodes. Nanoscale 2013, 5, 2967−2973. (43) Kong, J.; Yao, X.; Wei, Y.; Zhao, C.; Ang, J. M.; Lu, X. Polydopamine-Derived Porous Nanofibers as Host of ZnFe2O4 Nanoneedles: Towards High-Performance Anodes for Lithium-Ion Batteries. RSC Adv. 2015, 5, 13315−13323. (44) Geng, Y.; Dalhaimer, P.; Cai, S.; Tsai, R.; Tewari, M.; Minko, T.; Discher, D. E. Shape Effects of Filaments versus Spherical Particles in Flow and Drug Delivery. Nat. Nanotechnol. 2007, 2, 249−255. (45) Ernsting, M. J.; Murakami, M.; Roy, A.; Li, S. D. Factors Controlling the Pharmacokinetics, Biodistribution and Intratumoral Penetration of Nanoparticles. J. Controlled Release 2013, 172, 782− 794. (46) Della Vecchia, N. F.; Avolio, R.; Alfè, M.; Errico, M. E.; Napolitano, A.; d’Ischia, M. Building-Block Diversity in Polydopamine Underpins a Multifunctional Eumelanin-Type Platform Tunable E

DOI: 10.1021/acsbiomaterials.6b00102 ACS Biomater. Sci. Eng. XXXX, XXX, XXX−XXX