Antimicrobial Polymers: Molecular Design as Synthetic Mimics of Host

Jul 8, 2013 - Amphiphilic synthetic polymers have been utilized as a new molecular platform to mimic the structural features and antimicrobial functio...
0 downloads 8 Views 1MB Size
Chapter 19

Antimicrobial Polymers: Molecular Design as Synthetic Mimics of Host-Defense Peptides Downloaded by PURDUE UNIVERSITY on July 20, 2013 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch019

Edmund F. Palermo,1 Satyavani Vemparala,2 and Kenichi Kuroda*,1,3 1Macromolecular

Science and Engineering Center, University of Michigan, Ann Arbor, Michigan 48109, United States 2The Institute of Mathematical Sciences, CIT Campus, Taramani, Chennai 600113, India 3Department of Biologic and Materials Sciences, School of Dentistry, University of Michigan, Ann Arbor, Michigan 48109, United States *E-mail: [email protected].

The development of new antimicrobials effective against drug-resistant bacterial infections is a significant scientific challenge, requiring new molecular targets and antimicrobial strategies. Amphiphilic synthetic polymers have been utilized as a new molecular platform to mimic the structural features and antimicrobial functions of naturally occurring host-defense antimicrobial peptides. We have developed cationic amphiphilic random methacrylate copolymers which displayed a broad spectrum of inhibitory activity while maintaining low toxicity and low susceptibility to resistance in bacteria. The polymer-based antimicrobial design will provide a new versatile strategy for creation of a diversity of antimicrobial agents to fight against resistant bacteria.

Introduction Emerging Issue of Antibiotic Resistance The emergence of drug-resistance in pathogenic bacteria is significant threat to public health (1–6). These resistances can arise from antibiotic misuse, giving pressure for selection of resistant subpopulation in bacteria. These resistant strains spread to communities and public places including schools, causing invasive infections. In recent years, the number of new antibiotics in © 2013 American Chemical Society In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by PURDUE UNIVERSITY on July 20, 2013 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch019

the development pipeline dramatically declined, reducing available treatment options (7). Therefore, there is an urgent need to create new antibiotics which can overcome the existing resistance mechanisms of bacteria without contributing to the development of multi-drug resistance. However, the development of new antimicrobials effective against drug-resistant bacteria is a scientific challenge because the new antibiotic design requires new molecular targets and antimicrobial strategies rather than modifying existing drugs or their formulations.

Figure 1. α-helical cationic antimicrobial peptide LL-37. Cationic residues (blue) and hydrophobic reduces (yellow) are segregated into the opposite sides of helix. The backbone structure was colored gray. These images were created using the structure of human LL-37 from coordinates in PDB ID 2K6O (Wang, G. Structures of human host defense cathelicidin LL-37 and its smallest antimicrobial peptide KR-12 in lipid micelles, J. Biol. Chem. 2008, 283, 32637−32643) with Pymol Open-Source 0.99rc6.

Antimicrobial Peptides as New Antibiotics One approach has been centered in implementation of host-defense antimicrobial peptides, which are components of the innate immune system. Host-defense antimicrobial peptides act as “Nature’s antibiotics” which protect the host from bacterial infections by controlling growth of bacteria or killing them without harming the host (mammalian) cells (8–14). These peptides target bacterial cell membranes and disrupt membrane structures, causing leakage of cellular components, breakdown of membrane potential, and cell death. These peptides are relatively small (20-50 amino acids) and have generally cationic and hydrophobic residues, which provide cationic, amphiphilic properties. One of the classes is α -helical peptides including magainins and cathelicidin LL-37 (Figure 1). These peptides are in the form of random coil in solution, but they adopt helical conformation when bound to bacterial cell membranes. The unique feature of these peptides is the amphiphilic structure of helix, in which cationic and hydrophobic residues are segregated into the different sides of helix. The peptides selectively attack bacteria over human cells because of preferential electrostatic binding of cationic peptides to bacterial cell surfaces with high net negative charges as compared to human cells. The amphiphilic helical peptides accumulate on bacterial membrane surfaces and insert into the membrane, resulting in disintegration of membrane structures or formation of 320 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by PURDUE UNIVERSITY on July 20, 2013 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch019

membrane pores. Several molecular mechanisms have been proposed to describe the membrane disruption (Figure 2). These modes of actions contrast to those of traditional antibiotics such as penicillin, which are generally enzyme inhibitors or DNA replication inhibitors with specific molecular targets.

Figure 2. Proposed models for membrane disruption by α-helical antimicrobial peptides. Cationic residues of peptides are colored blue while hydrophobic residues are yellow.

Synthetic Mimics of Antimicrobial Peptides Since the discovery of host-defense peptides, many efforts have been made to develop natural antimicrobial peptides as new antibiotic drugs because of their reduced susceptibility to bacterial resistance. However, implementation of antimicrobial peptides and derivatives meets significant obstacles associated with poor pharmacokinetics in vivo, low stability due to susceptibility to proteolysis, and high manufacturing cost, preventing large-scale production (15). One approach to address these issues is to create synthetic mimics of antimicrobial peptides. The unique feature of antimicrobial peptides is that the molecular mechanism of antimicrobial action does not rely on the specific interactions with receptors or enzymes in bacteria. This was initially evidenced by the fact that D-enantiomer magainin showed the same level of activity compared to natural magainin (16). On the other hand, a class of antimicrobial peptides has cationic and hydrophobic side chains in common and adopt secondary conformations such as helix upon binding to lipid membranes, displaying segregated amphiphilic structures (Figure 1). It is, however, interesting that no common consensus sequences have been found yet (10). These findings suggest that the amphiphilic conformations with segregated cationic and hydrophobic residues, rather than the exact amino acid sequences or chemical structures, are likely key determinants for the antimicrobial activity. This inspired researchers to create synthetic molecules which can display segregated amphiphilic structures based on synthetic backbones 321 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by PURDUE UNIVERSITY on July 20, 2013 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch019

as new antimicrobial agents. The examples include antimicrobial β-peptides (17) and peptoids, which can adopt helical conformations with segregated amphiphilic structures (Figure 3) (18, 19). In general, these peptidomimetics are resistant to proteolysis and expected to have prolonged lifetime and activity in physiological conditions. This design principle was also applied to arylamide derivatives (Figure 3) (20–22). Although these arylamides are not designed to form helical structures, the rigid backbone presents the segregated amphiphilic conformation. These studies suggest that the amphiphilic structures are key determinants for antimicrobial activity, not requiring natural peptide sequences and backbones.

Figure 3. Synthetic mimics of antimicrobial peptides. (A) β-peptide β-17 (23), (B) β-peptide (24), (C) peptoid (19), and (D) arylamide (20).

Antimicrobial Polymers Although these approaches using peptide-like oligomers showed promising results, it has been of great interest to utilize molecular frameworks which can facilitate the development and use of antimicrobials for potential applications. To that end, the molecular design was further extended to include synthetic polymer backbones such as polymethacrylate (Figure 4) (25–30). These polymers are designed to have low molecular weights (MWs) (a few kDa) and primary ammonium and hydrophobic groups in side chains to mimic antimicrobial peptides. This contrasts to traditional antimicrobial or disinfecting polycations with quaternary ammonium groups modified with hydrophobic moieties such as long alkyls and relatively high MWs (30–32). The antimicrobial peptide-mimetic polymers showed potent activity against a panel of bacteria and minimal hemolytic activity against human erythrocytes. These polymers are not designed to form defined secondary conformations such as the α-helix, and the polymer conformation is rather random coils. The polymers have also random sequences of cationic and hydrophobic groups along the polymer backbone as well as distribution in the molecular size. Despite these heterogeneities in terms of conformation and chemical structures, these polymers are potent antimicrobial, suggesting that random amphiphilic structures can also mimic the antimicrobial activity of natural peptides. 322 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by PURDUE UNIVERSITY on July 20, 2013 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch019

Figure 4. Examples of antimicrobial polymers based antimicrobial peptide-mimetic design: (A) polymethacrylate (33), (B) polynorbornene (34), and (C) nylon (35, 36).

Amphiphilic Random Copolymers with Antimicrobial Activity Our research has focused on evaluating the potential for amphiphilic polymers to serve as antimicrobial agents (26, 28). We sought to develop a rational design of synthetic polymers based on a methacrylate platform. Previously, we demonstrated that the methacrylate random copolymers displayed antimicrobial activity against a broad spectrum of bacteria including methicillin-resistant Staphylococcus aureus (MRSA) without adverse toxicity to human cells (37, 38). Bacteria cultured with the copolymers did not develop resistance after 21 passages while the inhibitory concentration of conventional antibiotics increased up to 500 times (37). The broad spectrum of inhibitory activity while maintaining low toxicity and low susceptibility to resistance in bacteria are the hallmarks of naturally occurring host defense peptides (8, 10, 14).

Results and Discussion Snorkeling Effect Our previous research indicated that the primary ammonium groups of antimicrobial methacrylate copolymers bind effectively to lipids because of their complexation with phosphate lipid headgroups (39). In addition, sum frequency generation (SFG) vibrational spectra of polymers in a lipid bilayer showed that the hydrophobic groups of polymers are oriented parallel to the surface normal of lipid bilayer, suggesting that the hydrophobic groups of polymers are inserted into the hydrophobic core of lipid bilayers (40). Accordingly, we hypothesized that the amphiphilic structure of random copolymers ought to be capable of a “snorkeling effect” exerted by elongated spacer arms in the cationic side chains. In transmembrane helical peptides, the snorkel effect refers to stabilization of transmembrane conformation of peptides upon elongation of the cationic linker groups, which reach to the lipid water interface while the hydrophobic peptide is located in the hydrophobic membrane core (41, 42). As an analogy to this concept, we supposed that elongating the cationic side chain spacer groups of copolymers 323 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by PURDUE UNIVERSITY on July 20, 2013 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch019

might enhance insertion of the polymer backbone into the hydrophobic membrane core, facilitating membrane disruption ultimately. To test this hypothesis, we synthesized a library of cationic, amphiphilic copolymers with different length of cationic spacer groups (Figure 5) (38).

Figure 5. Tuning antimicrobial activity of amphiphilic methacrylate copolymers by snorkeling effect. (A) Chemical structures of polymers with different lengths of cationic spacer arms. The copolymers have ethyl groups as hydrophobic side chains (0 to 50 mole %). (B) Minimum inhibitory concentration (MIC) of polymers for E. coli. Data are from Ref. (38). For each spacer group (E2: ethylene, E4: butylene, E6: hexylene), we first prepared a series of copolymers with varying percentages of ethyl side chains as hydrophobic co-monomer ethyl methacrylate (EMA) to examine the snorkeling effect on their antimicrobial activity (38). The elongated spacer groups endowed the copolymers with much more potent inhibitory ability against Escherichia coli and S. aureus. The copolymers with ethylene spacers (E2 polymers) and butylene (E4 polymers) showed potent antimicrobial activity against E. coli (MIC < 20 µg/mL) when the percentage of ethyl methacrylate (EMA) comonomer is grater than 45 mol. % or 30 mol.% respectively. On the other hand, the copolymers with hexylene spacer groups (E6 polymers) exhibited high antimicrobial activity (MIC ~ 10 µg/mL) even with 0% EMA, and the activity is not sensitive to the EMA contents. In addition to the antimicrobial activity, we also evaluated the lytic activity of polymers to human red blood cells (hemolysis). In general, the hemolytic activity was increased as the spacer length was increased. The E6 polymers were highly hemolytic with HC50 (polymer concentration for 50% hemolysis) ~ 3 µg/mL. Among the polymer series in this study, the E4 copolymers showed the highest selectivity index (HC50/MIC) which implies that the butylene spacer groups are the most effective against bacteria without causing harm to human cells. Computational Modeling Inspired by the antimicrobial activity controlled by the new designed based on the cationic spacer groups, we further examined the details of membrane binding of polymers via molecular dynamic simulations (38). 324 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by PURDUE UNIVERSITY on July 20, 2013 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch019

For this study, we used a pre-equilibrated bilayer patch of zwitterionic POPE (1-palmitoyl-2-oleoyl phosphatidylethanolamine) and anionic POPG (1-palmitoyl-2-oleoyl phosphatidylglycerol) in a 7:3 mixture, which is a common formulation used to mimic bacterial cell membranes (43). Three model polymers with spacer arm lengths (E2model, E4model, and E6model) were constructed (Figure 6). These polymers have the degree of polymerization (DP) of 10 and fraction of EMA comonomer, fHB of 0.3. The mole fraction fHB = 0.3 was selected because we are interested in the polymers with potent antimicrobial activity and low hemolytic activity (38).

Figure 6. Monomer sequence and structures of model copolymers. The copolymers have different length of cationic spacer arms: ethylene (m = 2) for E2 model, butylene (m = 4) for E4model, and hexane (m = 6) for E6model. The copolymer structures have an isotactic sequence, and the ratio of cationic monomers and EMA co-monomers is 7:3. These groups are unevenly distributed, modeling the random sequence of polymers (38).

Figure 7. The average conformation of the copolymer E4model in water (A) and at the water-lipid interface (B). The cationic and EMA ethyl side chains are colored red and blue, respectively. The polymer backbones are colored green. All three systems were simulated under isothermal-isobaric ensemble conditions (305 K temperature and 1 atm pressure). See Reference (38) for experimental details. Adapted with permission from Ref. (38). Copyright 2012 American Chemical Society. 325 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by PURDUE UNIVERSITY on July 20, 2013 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch019

Each polymer was first placed in aqueous phase above POPE-POPG (7:3) bilayers, and the three systems were simulated for 100 ns each. Counter-ions were added to each system for overall charge neutrality. In general, the three polymers in water adopted compact structures, and the side groups are randomly oriented with respect to the polymer backbone. This is likely because the exposure of hydrophobic polymer backbone and ethyl side chains of EMA to water needs to be minimized, resulting in the compact conformation in water. The results also indicated that no predominant conformations were observed for polymers in water. On the other hand, simulations of polymers interacting with lipid-water interface revealed that the polymer backbones are predominantly oriented parallel to the membrane surface, with the hydrophobic groups projected into the membrane core, and the ammonium groups projected to the water-membrane interface. These results were represented by the two snapshots of conformations of polymer E4model in water and at the lipid-water interface as shown in Figure 7. E4model polymer adopts an extended conformation with well-separated hydrophobic and charged side groups near the lipid interface, whereas the conformation in water is crescent-shaped and no predominant amphiphilic structure can be observed. Yethiraj and coworkers also previously demonstrated that random copolymer models of β-peptides form the segregated amphiphilic structure upon binding to lipid membrane (44). Simulations of the three polymers with bacteria-type POPE-POPG bilayers suggest that the depth of insertion of the polymers into the lipid bilayer depends on the cationic length of the spacer arm groups (Figure 8); the center of mass of the copolymer chains along the membrane normal moves towards the hydrophobic membrane core as the spacer arms are elongated. The polymer backbone of E4model and E6model polymers are stretched in the membranes. The pendant ammonium groups are located near the water-lipid interface while the ethyl groups of EMA comonomers are rather oriented to the hydrophobic core of membranes. Apparently, the cationic and hydrophobic ethyl side chains of the copolymer are segregated in the opposite side of the polymer backbones.

Figure 8. Amphiphilic copolymers with different spacer arm lengths (E2model, E4 model, and E6model), interacting with bacteria-type lipid bilayer at the end of 100 ns simulations. Water is not shown for clarity. The ammonium spacer arm groups and hydrophobic ethyl groups of EMA comonomers are colored orange and green, respectively. Reproduced with permission from Ref. (38). Copyright 2012 American Chemical Society. 326 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by PURDUE UNIVERSITY on July 20, 2013 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch019

Figure 9. (A) End-to-end distances of the polymers in the last 20ns of simulation (B) distance between the centers of mass of the ethyl and amine groups in the polymers Insets: average conformation of the copolymers at the end of 100 ns of simulation in the (A)XY plane (perpendicular to membrane normal) (B) parallel to the membrane normal. The cationic and EMA ethyl side chains are colored red and blue, respectively. The polymer backbones are colored green. Adapted with permission from Ref. (38). Copyright 2012 American Chemical Society.

The development of global amphiphilic structure by methacrylate copolymers was also quantified by examining the parameters of polymer conformations. The end-to-end distance of polymer chains is a measure to describe how much polymers chains are stretched (Figure 9). The end-to-end distance of E4model and E6model polymers in the last 20ns of simulation appears to be constant to 327 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by PURDUE UNIVERSITY on July 20, 2013 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch019

be ~20 Å (Figure 9A). The theoretical length for a fully extended conformation of polymer chains with DP = 10 is 23-24 Å based on the chemical structure. This result indicates that E4model and E6model polymers are in highly stretched conformations. Once these polymer chains are inserted in the hydrophobic core of membranes, the polymer chains extend to maximize the interaction between the hydrophobic polymer backbone and acyl chains of lipids for stabilization. The average distances between the center of mass of the ethyl and amine groups in the polymers are ~7 and ~9 Å for E4model and E6model. The theoretical maximum distances for complete segregation of these side chains would be ~16 and ~18 Å, respectively (see Figure 9B). This results support the distinctive segregation of cationic spacer arms and hydrophobic ethyl groups of EMA comonomer to opposite faces (Figure 9). In analogy to the snorkeling effect observed with transmembrane peptide helices, the increased spatial separation between the cationic ammonium and hydrophobic ester groups would facilitate deeper membrane insertion. These random copolymers are not designed to adopt defined secondary structures such as α-helix. These results may indicate that the formation of such amphiphilic structures in lipid membranes may determine their antimicrobial activity rather than the random sequence and molecular weight distribution, and random coil conformation although the exact molecular details of antimicrobial mechanism is not clear at this point. Gellman and coworkers previously demonstrated that cationic amphiphilic random nylon copolymers showed potent activity by tuning the balance of cationic and hydrophobic groups (35, 36). Their studies also reached the similar conclusion that such amphiphilic conformation of polymers and peptides in lipid membranes is a general phenomenon for antimicrobial activity.

Summary and Outlook Cationic amphiphilic random copolymers showed potent antimicrobial activity and minimal toxicity to human cells, demonstrating their potential as new antimicrobials. The copolymers are also effective against drug resistant bacteria and do not contribute to the development of new resistance mechanisms in bacteria. These polymers likely act on bacterial cell membranes and disrupt the membrane integrity, which mimic the function of natural antimicrobial peptides. The antimicrobial activity can be tuned by altering the cationic spacer arm length, which control the depth of polymer insertion into the lipid membrane, representing the snorkeling effect. The random copolymers have flexible backbones and distribution in the molecular length or molecular weight, which are distinctively different from the defined structures and sequences of antimicrobial peptides. However, many types of polymers have already been utilized as platforms of antimicrobial polymers, and they showed potent activity against a panel of bacteria. The polymer-based strategy to create new antimicrobials will be versatile and can provide a diversity of antimicrobial agents to fight against resistant bacteria. 328 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Acknowledgments We would like to thank Dr. Gregory Caputo at Rowan University and Dr. Haruko Takahashi at University of Michigan for help to prepare figures.

Downloaded by PURDUE UNIVERSITY on July 20, 2013 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch019

Conflict of interest K. Kuroda is a co-inventor on a patent application filed by the University of Pennsylvania covering “Antimicrobial Copolymers and Uses Thereof”. The patent application has been licensed to PolyMedix Inc. (Radnor, PA). PolyMedix did not play a role in the design and conduct of this study, in the collection, analysis, or interpretation of the data, or in the preparation, review, or approval of the article.

References 1. 2. 3. 4.

5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18.

19.

Fernandes, P. Nat. Biotechnol. 2006, 24, 1497–1503. Fischbach, M. A.; Walsh, C. T. Science 2009, 325, 1089–1093. Levy, S. B. The Antibiotic Paradox. How Miracle Drugs Are Destroying the Miracle; Plenum Press: New York, 1992. Levy, S. B. Antibiotic resistance: An Ecological Imbalance. In Antibiotic Resistance: Origins, Evolution, Selection and Spread, 207 ed.; Chadwick, D. J., Goode, J., Eds. Wiley: Chichester, 1997; pp 1−14. Levy, S. B.; Marshall, B. Nat. Med. 2004, 10, S122–S129. Nathan, C. Nature 2004, 431, 899–902. Rodriguez de Castro, F.; Naranjo, O. R.; Marco, J. A.; Violan, J. S. Semin. Respir. Crit. Care Med. 2009, 30, 161–71. Brogden, K. A. Nat. Rev. Microbiol. 2005, 3, 238–250. Fjell, C. D.; Hiss, J. A.; Hancock, R. E. W.; Schneider, G. Nat. Rev. Drug Discovery 2012, 11, 37–51. Hancock, R. E. W.; Sahl, H. G. Nat. Biotechnol. 2006, 24, 1551–1557. Matsuzaki, K. Biochim. Biophys. Acta, Biomembr. 1999, 1462, 1–10. Nguyen, L. T.; Haney, E. F.; Vogel, H. J. Trends Biotechnol. 2011, 29, 464–472. Shai, Y. Biopolymers 2002, 66, 236–248. Zasloff, M. Nature 2002, 415, 389–395. Marr, A. K.; Gooderham, W. J.; Hancock, R. E. W. Curr. Opin. Pharm. 2006, 6, 468–472. Wade, D.; Boman, A.; Wahlin, B.; Drain, C. M.; Andreu, D.; Boman, H. G.; Merrifield, R. B. Proc. Natl. Acad. Sci. U.S.A. 1990, 87, 4761–4765. Cheng, R. P.; Gellman, S. H.; DeGrado, W. F. Chem. Rev. 2001, 101, 3219–3232. Chongsiriwatana, N. P.; Patch, J. A.; Czyzewski, A. M.; Dohm, M. T.; Ivankin, A.; Gidalevitz, D.; Zuckermann, R. N.; Barron, A. E. Proc. Natl. Acad. Sci. U.S.A. 2008, 105, 2794–2799. Patch, J. A.; Barron, A. E. J. Am. Chem. Soc. 2003, 125, 12092–12093.

329 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.

Downloaded by PURDUE UNIVERSITY on July 20, 2013 | http://pubs.acs.org Publication Date (Web): July 8, 2013 | doi: 10.1021/bk-2013-1135.ch019

20. Choi, S.; Isaacs, A.; Clements, D.; Liu, D. H.; Kim, H.; Scott, R. W.; Winkler, J. D.; DeGrado, W. F. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 6968–6973. 21. Goodman, C. M.; Choi, S.; Shandler, S.; DeGrado, W. F. Nat. Chem. Biol. 2007, 3, 252–262. 22. Liu, D. H.; Choi, S.; Chen, B.; Doerksen, R. J.; Clements, D. J.; Winkler, J. D.; Klein, M. L.; DeGrado, W. F. Angew. Chem., Int. Ed. 2004, 43, 1158–1162. 23. Porter, E. A.; Wang, X.; Lee, H.-S.; Weisblum, B.; Gellman, S. H. Nature 2000, 404, 565–565. 24. Liu, D.; DeGrado, W. F. J. Am. Chem. Soc. 2001, 123, 7553–7559. 25. Gabriel, G. J.; Som, A.; Madkour, A. E.; Eren, T.; Tew, G. N. Mater. Sci. Eng., R 2007, 57, 28–64. 26. Kuroda, K.; Caputo, G. A. WIREs Nanomed. Nanobiotechnol. 2012. 27. Li, P.; Li, X.; Saravanan, R.; Li, C. M.; Leong, S. S. J. RSC Adv. 2012, 2, 4031–4044. 28. Palermo, E. F.; Kuroda, K. Appl. Microbiol. Biotechnol. 2010, 87, 1605–1615. 29. Timofeeva, L.; Kleshcheva, N. Appl. Microbiol. Biotechnol. 2011, 89, 475–492. 30. Munoz-Bonilla, A.; Fernandez-Garcia, M. Prog. Polym. Sci. 2012, 37, 281–339. 31. Kenawy, E. R.; Worley, S. D.; Broughton, R. Biomacromolecules 2007, 8, 1359–1384. 32. Tashiro, T. Macromol. Mater. Eng. 2001, 286, 63–87. 33. Kuroda, K.; DeGrado, W. F. J. Am. Chem. Soc. 2005, 127, 4128–4129. 34. Lienkamp, K.; Madkour, A. E.; Musante, A.; Nelson, C. F.; Nusslein, K.; Tew, G. N. J. Am. Chem. Soc. 2008, 130, 9836–9843. 35. Mowery, B. P.; Lee, S. E.; Kissounko, D. A.; Epand, R. F.; Epand, R. M.; Weisblum, B.; Stahl, S. S.; Gellman, S. H. J. Am. Chem. Soc. 2007, 129, 15474–15476. 36. Mowery, B. P.; Lindner, A. H.; Weisblum, B.; Stahl, S. S.; Gellman, S. H. J. Am. Chem. Soc. 2009, 131, 9735–9745. 37. Sovadinova, I.; Palermo, E. F.; Urban, M.; Mpiga, P.; Caputo, G. A.; Kuroda, K. Polymers 2011, 3, 1512–1532. 38. Palermo, E. F.; Vemparala, S.; Kuroda, K. Biomacromolecules 2012, 13, 1632–1641. 39. Palermo, E. F.; Lee, D. K.; Ramamoorthy, A.; Kuroda, K. J. Phys. Chem. B 2011, 115, 366–375. 40. Avery, C. W.; Palermo, E. F.; McLaughin, A.; Kuroda, K.; Chen, Z. Anal. Chem. 2011, 83, 1342–1349. 41. Mishra, V. K.; Palgunachari, M. N.; Segrest, J. P.; Anantharamaiah, G. M. J. Biol. Chem. 1994, 269, 7185–7191. 42. Monne, M.; Nilsson, I.; Johansson, M.; Elmhed, N.; von Heijne, G. J. Mol. Biol. 1998, 284, 1177–1183. 43. Wilkinson, S. G. Microbial Lipids; Academic Press: London, 1988; Vol. 1. 44. Mondal, J.; Zhu, X.; Cui, Q. A.; Yethiraj, A. J. Phys. Chem. B 2010, 114, 13585–13592.

330 In Tailored Polymer Architectures for Pharmaceutical and Biomedical Applications; Scholz, C., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 2013.