A structure-activity study of antibacterial poly(ester urethane)s with

6 days ago - Chao Peng , Apoorva Vishwakarma , Steven Mankoci , Hazel A. Barton , and Abraham Joy. Biomacromolecules , Just Accepted Manuscript...
0 downloads 0 Views 1MB Size
Subscriber access provided by UNIV OF TEXAS DALLAS

Article

A structure-activity study of antibacterial poly(ester urethane)s with uniform distribution of hydrophobic and cationic groups Chao Peng, Apoorva Vishwakarma, Steven Mankoci, Hazel A. Barton, and Abraham Joy Biomacromolecules, Just Accepted Manuscript • DOI: 10.1021/acs.biomac.9b00029 • Publication Date (Web): 07 Mar 2019 Downloaded from http://pubs.acs.org on March 10, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

A Structure-Activity Study of Antibacterial Poly(Ester Urethane)s with Uniform Distribution of Hydrophobic and Cationic Groups Chao Penga, Apoorva Vishwakarmaa, Steven Mankocia, Hazel A. Barton*b, Abraham Joy*a aDepartment

of Polymer Science and bDepartment of Biology, The University of Akron,

Akron, Ohio 44325, United States

ABSTRACT. Infections associated with antibiotic-resistant bacteria have become a threat to the global public health. Antimicrobial polymers, which are synthetic mimics of antimicrobial peptides, have gained increasing attention as they may have a lower chance of inducing resistance. The cationic-hydrophobic balance and distribution of cationic and hydrophobic moieties of these polymers is known to have a major effect on antimicrobial activity. We studied the properties of a series of facially amphiphilic antimicrobial surfactant-like poly(ester urethane)s with different hydrophobic pendant groups (P1, P2, and P3) and cationic groups distributed uniformly along the polymer chain. These polymers exhibited bactericidal activity against Gram-negative Escherichia coli and Pseudomonas aeruginosa as well as Gram-positive Staphylococcus aureus and Staphylococcus

ACS Paragon Plus Environment

1

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 31

epidermidis. Microscopy and dye release assays demonstrated that these polymers cause membrane disruption, which is dependent on the cationic-hydrophobic ratio in the polymer. Membrane permeability assays revealed that these polymers can permeabilize the outer membrane of E. coli and damage the cytoplasmic membrane of both E. coli and S. aureus. In addition, our results indicate that the three polymers exhibit different extent of membrane disruption against E. coli. P1 caused minor damage to the cytoplasmic membrane integrity, but it was able to dissipate the cytoplasmic membrane potential, leading to cell death. P2 and P3 depolarized the cytoplasmic membrane and also caused significant damage to the cytoplasmic membrane. Overall, we showed a new class of broad-spectrum bactericidal polymers whose membrane disrupting ability against E. coli correlates with the structural differences of the hydrophobic pendant groups.

INTRODUCTION Antimicrobial peptides (AMPs) are short cationic peptides present in various organisms that are an important part of the host innate response, and have received extensive attention due to their broad-spectrum bactericidal activity and reduced likelihood of resistance compared to traditional antibiotics.1-3 AMPs are believed to cause membrane lysis by binding to negatively charged surfaces on bacterial membranes to disrupt the cell membrane integrity.1,

4-6

Although AMPs are efficient antimicrobial agents, the

commercialization of AMPs has been limited by their relatively high manufacturing cost

ACS Paragon Plus Environment

2

Page 3 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

and inherent proteolytic degradation.4, 7, 8 Consequently, the use of synthetic polymeric mimics of AMP has gathered increased relevance. These AMP analogs (antimicrobial polymers) present several advantages over natural AMPs, such as lower cost, long-term activity, metabolic stability, and the flexibility of chemical modification.9, 10 In recent years, a number of advancements in antimicrobial polymers have been reported based on various platforms, such as poly(meth)acrylates, polycarbonates, polyesters, polyamides, and poly(oxanorbornene)s.11-21 Kuroda et al. developed cationic random copolymers of methacrylates and methacrylamides, and studied the influence of different cationic groups as well as the hydrophilic-hydrophobic ratio on both antimicrobial activity and toxicity against eukaryotic cells.11, 12 Yang et al. also developed biodegradable antimicrobial polycarbonates that displayed high bactericidal efficiency against clinically isolated methicillin-resistant Staphylococcus aureus (MRSA) and showed high efficacy in vivo when tested in a mouse systemic infection model.15 There are several antimicrobial polymers reported in literature in which the cationic and hydrophobic moieties are randomly distributed along the polymer chain.11, 12, 14, 18, 22-27 Recent studies have shown that the distribution of the cationic and hydrophobic pendant groups in the polymer chain can play a role in the resultant antimicrobial property of the polymer.28, 29 For example, according to the study by Sambhy et al. polymers with cationic pyridines and hydrophobic tails in the same repeating unit exhibited much better

ACS Paragon Plus Environment

3

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 31

selectivity to bacterial cells compared to polymers with cationic and hydrophobic groups distributed randomly along the polymer chain.29 In this work, we report a densely functionalized antimicrobial poly(ester urethane) system, in which the cationic and hydrophobic pendant groups are adjacent to each other in the same repeating unit, resulting in a cationic surfactant-like structure. This design provides a uniform local hydrophilic-hydrophobic balance and facilitates the electrostatic and hydrophobic interactions between the polymer and the negatively charged lipid molecules in the bacterial membrane. Three antimicrobial poly(ester urethane)s bearing different hydrophobic pendant groups were synthesized based on a previously developed unsaturated poly(ester urethane) platform.30 The antimicrobial activity of these polymers was investigated in model bacteria: Gram-negative Escherichia coli and Gram-positive Staphylococcus aureus. The mechanism of action of these polymers against bacteria was investigated by microscopy studies, dye release assays using a bacterial membranemimicking liposome model, and membrane permeability assays. The toxicity of these polymers against eukaryotic cells was also examined by a red blood cell hemolysis assay. Overall, the results show a strong correlation between the structure of the polymers and the resultant bacterial activity and mammalian cell toxicity. EXPERIMENTAL SECTION

ACS Paragon Plus Environment

4

Page 5 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

Materials. Tin(II) 2-ethylhexanoate (Sn(Oct)2), 2-hydroxy-4’-(2-hydroxyethoxy)-2methylpropiophenone (Irgacure 2959), sodium phosphate dibasic, potassium phosphate monobasic, potassium chloride, calcium chloride, HEPES, D-lactose monohydrate, and Triton X-100 were purchased from Sigma-Aldrich. Fluorescein isothiocyanate (FITC), hexamethylene diisocyanate (HDI) and 2-nitrophenyl-β-D-galactopyranoside (ONPG) were purchased from Acros Organics. Anhydrous methanol, triethylamine (TEA), and dimethylformamide (DMF) were purchased from EMD Millipore. Cysteamine hydrochloride, N,N-diisopropylethylamine (DIPEA), and 6-carboxyfluorescein were purchased from Chem-Impex International. Dimethyl sulfoxide (DMSO) and sodium chloride were purchased from BDH Chemicals. N-phenyl-1-naphthylamine (NPN), 3,3’dipropylthiadicarbocyanine iodide (DiSC3(5)) were purchased from TCI America. 1Palmitoyl-2-oleoyl-sn-glycero-3-phosphoethanolamine (POPE) and 1-palmitoyl-2-oleoylsn-glycero-3-phosphoglycerol (POPG) were purchased from Cayman Chemical. Mueller Hinton broth (MHB) was purchased from Himedia. Melittin, a membrane-pore forming peptide was purchased from Enzo Life Sciences. Defibrinated sheep blood was purchased from Hardy Diagnostics. Escherichia coli K12 (ATCC 10798), Staphylococcus aureus (ATCC 25923), and Staphylococcus epidermidis (ATCC 12228) were obtained from the American Type Culture Collection (ATCC), with Escherichia coli UB1005 from the E. coli Genetic Stock Center (CGSC) at Yale University. Sephadex G-50 was purchased from Pharmacia

ACS Paragon Plus Environment

5

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 31

Fine Chemicals. Dichloromethane was dried by distilling over calcium hydride. Unless otherwise stated, all reagents were used as received. Analytical Methods NMR spectra were recorded on a Varian NMRS 300 MHz instrument. 1H NMR chemical shifts are reported in ppm relative to the solvent’s residual 1H signal. Size exclusion chromatography (SEC) analysis in DMF was performed on an HLC-8320 GPC from TOSOH equipped with an RI detector using polystyrene (PS) as the standard. Absorbance and fluorescence spectroscopy were performed using a BioTek Synergy H1 plate reader. The morphology of the bacteria was characterized by scanning electron microscope (SEM) using JEOL-JSM-7401F with operating voltage as 1 kV. Fluorescence imaging was performed on a U-TV1X-2 fluorescence microscope (Olympus Co., Japan) with standard FITC/TRITC filter set. Synthesis of Cationic Poly(Ester Urethane)s. The unsaturated poly(ester urethane) precursors were synthesized according to a reported procedure.30 As an example, the synthesis of P2 is as follows: Pre2 (0.21 g), cysteamine hydrochloride (0.57 g) and Irgacure 2959 (57 mg) were dissolved in a mixture of anhydrous methanol (4 mL) and anhydrous dichloromethane (2 mL). Once dissolved, the solution was irradiated at 350 nm for 30 minutes. The polymer was purified by dialysis in methanol to give a colorless solid (yield ≈90%).

ACS Paragon Plus Environment

6

Page 7 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

P1. 1H NMR (300 MHz, MeOD), δ (ppm): 0.91-1.00 (m, 3H), 1.35-1.70 (m, 12H), 2.84-3.21 (m, 11H), 4.27-4.35 (m, 4H), 5.02 (m, 1H). P2. 1H NMR (300 MHz, MeOD), δ (ppm): 1.32-1.66 (m, 10H), 2.59-3.18 (m, 13H), 4.24-4.29 (m, 4H), 5.02 (m, 1H), 7.20-7.28 (m, 5H). P3. 1H NMR (300 MHz, MeOD), δ (ppm): 0.88-0.93 (m, 3H), 1.30-1.64 (m, 24H), 2.82-3.11 (m, 11H), 4.31 (m, 4H), 5.02 (m, 1H). Preparation of FITC-Labeled Polymers. As an example, the synthesis of P1-FITC is as follows: P1 (10 mg) was dissolved in anhydrous DMF (0.5 mL). Then, anhydrous triethylamine (5 μL), FITC stock solution (0.5 mL, 1 mg/mL in anhydrous DMF), and DIPEA (5 μL) were added. The reaction was carried out at room temperature for 2 hours. The product was purified by dialysis in water for 24 hours with water changes at 1h, 2h, 4h, 6h, and 16h. Minimum Inhibitory Concentration (MIC) and Time-Kill Assays. Unless otherwise stated, all bacterial cultures were grown in MHB with shaking (200 rpm) at 37°C. Overnight cultures of either E. coli K12, S. aureus, S. epidermidis, or P. aeruginosa were diluted in MHB to an OD600nm 0.001 (measured by Hach DR 2800 Spectrophotometer), by initially diluting to 0.1 and then serially diluting 100 fold (~1×106 CFU/mL). For the MIC assays, this suspension was mixed with the polymer aqueous solution in a 1:1 ratio in a 96-well plate. The optical density (OD600nm) was measured and followed by incubating at

ACS Paragon Plus Environment

7

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 31

37°C for 16-20 h, whereupon, optical density (OD600nm) was measured again using a BioTek Synergy H1 plate reader. The minimum inhibitory concentration (MIC) was determined as the lowest concentration of polymer that prevented bacteria growth after overnight incubation. For the time-kill assays, the working bacterial suspensions were mixed with the polymer aqueous solution in a 1:1 ratio. 100 L aliquots at different time points were serially diluted onto trypticase soy agar (TSA). The results were expressed as colony forming units (CFU)/mL over time. Hemolysis Assays. Defibrinated sheep red blood cells (RBC) were centrifuged at 500 x g for 10 min at 5 °C. The RBC pellet was washed with phosphate-buffered saline (PBS; pH 7.4) until the supernatant no longer contained any visually observable hemoglobin. The RBC suspension was then diluted 1:50 in PBS and mixed with polymer aqueous solution in a 1:1 ratio. Wells containing no polymer were used as 0% hemolysis control, and 1% Triton X-100 was added to the positive control for 100% hemolysis. After incubation at 37°C for 1 h, the plate was centrifuged at 500 x g for 10 min at 5°C and the hemoglobin release as a relative measure of hemolysis was determined at 450 nm on a BioTek Synergy H1 plate reader. Lactate Dehydrogenase (LDH) Assay for NIH-3T3 Cells. Mouse fibroblast NIH-3T3 cells were cultured using DMEM medium supplemented with 10% FBS and 1% penicillin streptomycin. The cells were grown at 37 °C in 5% CO2 until reaching ~90% confluence.

ACS Paragon Plus Environment

8

Page 9 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

The cells were harvested using 0.25% trypsin solution and seeded in a 96-well plate at a density of 1×104 cells/well. The cells were allowed to attach to the wells overnight before treatment with polymers. The polymer treated cells were incubated at 37 °C for 1 hour before determining the cell viability. The cell viability was determined using a Pierce LDH Cytotoxicity Assay Kit according to the instruction of the manufacturer. Outer Membrane Permeability Assay. A 20 mL culture of E. coli was incubated until the culture reached an OD600nm of 0.5. The culture was centrifuged at 2000 x g for 5 min, and the cell pellet was resuspended in 20 mL of 10 mM HEPES/150 mM NaCl (pH 7.4). 100 μL of this suspension was mixed with 50 μL of polymer solution and 50 μL of N-phenyl1-naphthylamine (NPN) solution (40 μM) in a 96-well plate. Wells containing no polymer were used as the negative control, and with 1% Triton X-100 as a positive control. The fluorescence at 350/420 nm (excitation/emission) was measured on a BioTek Synergy H1 plate reader. Cytoplasmic Membrane Permeability Assay. E. coli K12 was used to inoculate 3 mL of MHB supplemented with 2% lactose and incubated until mid-log phase (OD600nm~0.5). The bacterial suspension was then centrifuged at 2000 xg for 5 min, and the cell pellet was resuspended in PBS (pH 7.4) to an OD600nm of 0.05. 100 μL of this suspension was then added to each well of a 96-well plate along with 50 μL of polymer solution and 50 μL of ONPG solution (5 mM). Wells containing no polymer were used as the negative

ACS Paragon Plus Environment

9

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 31

control, and melittin was used as the positive control. The UV absorbance at 420 nm was recorded at 37 °C for 60 min. Cytoplasmic Membrane Depolarization Assay. A culture of E. coli UB1005 or S. aureus was prepared as described for the cytoplasmic membrane permeability assay and resuspended in 10 mM HEPES/150 mM NaCl (1% DMSO) with 4 μM DiSC3(5) and incubated at 37°C for 60 min. The bacterial suspension was further diluted 2 times in 10 mM HEPES/150 mM NaCl (1% DMSO) and incubated at 37°C for an additional 60 min to give the working suspension. This suspension was then mixed with polymer solutions at a 1:1 ratio in a 96-well plate and the fluorescence at 610/660 nm (excitation/emission) was monitored over a period of 60 min at 37 °C. Wells containing no polymer were used as the negative control, and the wells containing melittin (100 μg/mL) were used as a positive control. SEM Sample Preparation. A 3 mL culture of E. coli K12 with an OD600nm of 0.5 was mixed with polymer aqueous solution (32 μg/mL) in a 1:1 ratio, and incubated at 37°C for 30 min. The mixture was then filtered through a 0.22 μm Durapore® Membrane Filter and fixed with 4% formaldehyde in PBS for 2 hours. The bacteria were then washed with DI water (3X) and dehydrated in a gradient ethanol/water mixture (50%, 60%, 70%, 80%, 90%, 95%, and 100%; 15 min for each step), followed by critical point drying and sputter

ACS Paragon Plus Environment

10

Page 11 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

coating.

The samples were observed using a JEOL-JSM-7401F scanning electron

microscope (SEM). Fluorescence Microscopy. The bacteria were resuspended in 10 mM HEPES with 150 mM NaCl, and diluted to an OD600nm value of 0.1. This suspension was then mixed with 100 μg/mL FITC-labelled polymer solution in a 1:1 ratio and incubated at 37 °C for 30 min. Fluorescence imaging was performed on a U-TV1X-2 fluorescence microscope (Olympus Co., Japan) with standard FITC/TRITC filter set. Liposome Preparation and Dye Release Assay. Liposomes were prepared using POPE/POPG (4:1). The lipid film was hydrated with 10 mM HEPES buffer with 50 mM 6-carboxyfluorescein and 150 mM NaCl, freeze-thawed 5 times, and placed on a shaker overnight. This lipid solution was then extruded 15 times through an Avanti MiniExtruder equipped with filters of 100 nm pore size and placed on a shaker for 1 hour. The unincorporated dye was removed from the liposome using a Sephadex G-50 column and lipid concentration was quantified by phosphorus assay.31 As calculated using a NaH2PO4 standard curve (Figure S5), the final lipid concentration was 30 μM. The liposomes were mixed with polymer solutions. The samples were incubated at 37 °C for 1 hour and fluorescence emission at 530 nm (excitation at 495 nm) was monitored over a period of 60 min at 37 °C on a BioTek Synergy H1 plate reader. Wells containing no

ACS Paragon Plus Environment

11

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 31

polymer were used as 0% dye release control, and wells containing 1% Triton X-100 were the 100% dye release positive control. RESULTS AND DISCUSSION Polymer Design and Synthesis

Scheme 1. Synthetic route of the antimicrobial poly(ester urethane)s (P1, P2, and P3) The novel antimicrobial poly(ester urethane)s reported in this work were synthesized based on an alkene functionalized poly(ester urethane) platform previously reported by our group.30 The synthetic route of the antimicrobial poly(ester urethane)s is depicted in Scheme 1. Briefly, an alkene functionalized diol was first synthesized via a one-step Baylis-Hillman reaction of functionalized aldehyde with 2-hydroxyethyl acrylate in a mixture of 1,4-dioxane and water catalyzed by DABCO according to a reported procedure.30, 32 The resulting diol monomer was then polymerized with hexamethylene diisocyanate (HDI) in anhydrous CH2Cl2 catalyzed by tin(II) octoate to generate an

ACS Paragon Plus Environment

12

Page 13 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

unsaturated poly(ester urethane) precursor. As confirmed by SEC analysis, the molecular weight of these precursors was in the range of 5-9 kDa (Pre1~8.0 kDa, Ð~1.7; Pre2~5.1 kDa, Ð~1.9; Pre3~8.8 kDa, Ð~1.7). The unsaturated polymer precursors were then functionalized by reacting with cysteamine hydrochloride via thiol-ene reaction to provide the final cationic poly(ester urethane)s (P1, P2, P3). The chemical structure of the cationic polymers was confirmed by 1H NMR spectroscopy (Figure S1, S2, and S3). As shown by the polymer structure, the cationic and hydrophobic groups are presented in the same repeating unit, and they are adjacent to each other, giving a cationic surfactantlike structure. The rationale of using such polymer design is to provide a uniform local hydrophilic-hydrophobic balance and increase the affinity of the polymers to the negatively charged lipid molecules in the bacterial membrane, leading to enhanced antimicrobial efficacy. Antimicrobial Activity Determined by MIC Measurement and Time-Kill Test The antimicrobial activity of these poly(ester urethane)s was investigated using the Gram-negative E. coli and P. aeruginosa as well as Gram-positive S. aureus and S. epidermidis. These antimicrobial polymers showed low MIC values for E. coli (MIC = 8-16 μg/mL; Table 1), S. aureus (MIC = 16-32 μg/mL; Table 1), P. aeruginosa (MIC = 16-32 μg/mL; Table 1), and S. epidermidis (MIC = 4-16 μg/mL; Table 1) in MHB. The hydrophobicity of the pendant group does not have a significant influence on the MIC values, as the three

ACS Paragon Plus Environment

13

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 31

polymers exhibit only slightly different MICs (1-2 dilution difference). Similar trend was observed when comparing the MIC results using molar-based concentrations. To understand whether these polymers are bacteriostatic or bactericidal, we monitored bacteria viability over 24 hours for E. coli and S. aureus, as proof-of-concept. As shown in Figure 1 (a), all three polymers are bactericidal toward E. coli, as evidenced by the ~5 log reduction in CFU after treatment with 16 μg/mL polymer. This reduction in CFU values for E. coli increased with increasing hydrophobicity in the pendant groups. The polymers were also bactericidal toward S. aureus, albeit at higher concentrations and lower killing rates (Figure 1(b)). At the concentration of 32 μg/mL, P2 and P3 reduced the number of viable S. aureus by ~5 log after 2 h and 6 h, respectively. P1 was not as effective against S. aureus, being bacteriostatic for up to 8 hours, followed by growth. Nonetheless, P1 was bacteriocidal at a concentration of 64 μg/mL. Similar trend in the killing efficiency would be observed when compared using the same molar-based concentration, as these polymers have similar molecular weight. Also, polymers with more hydrophobic pendant groups showed lower molar-based MIC values (P1~4.0 µM, P2~3.1 µM, P3~1.8 µM, against S. aureus). Table 1. Minimum inhibitory concentration (MIC) of the polymers against E. coli, S. aureus, P. aeruginosa, and S. epidermidis and HC10 of the polymers toward red blood cells

polymer

MIC Gram-negative bacteria

Gram-positive bacteria

HC10

ACS Paragon Plus Environment

14

Page 15 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

P1 P2 P3 Ampicilli n Melittin

E. coli µg/m µM L 16 2.0 8 1.6 8 0.9 1

2.9

25

8.8

P. aeruginosa µg/m µM L 32 4.0 16 3.1 16 1.8 128 371.2 50

17.6

S. aureus µg/m µM L 32 4.0 16 3.1 16 1.8 0.2 0.6 25

8.8

S. epidermidis µg/m µM L 16 2.0 4 0.8 16 1.8 4 11.6 0.8

µg/m L

µM

379.4 19.8 3.5

47.4 3.8 0.4

0.3

Figure 1. Time-kill test of (a) E. coli and (b) S. aureus over a period of 24 hours Toxicity of the Antimicrobial Polymers The toxicity of these polymers was evaluated by hemolysis assay, with the hemolytic activity quantified by percentage of hemoglobin released from lysed red blood cells relative to 100% lysis by 1% Triton X-100.33,

34

HC10 is defined as the polymer

concentration that causes 10% hemolysis and it is commonly used as a parameter to reflect the toxicity of the polymers. The percentage of hemolysis induced by the

ACS Paragon Plus Environment

15

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 31

antimicrobial poly(ester urethane)s is shown in Figure 2, with a lower HC10 value indicating higher toxicity. The polymers showed hydrophobicity-dependent toxicity (Figure 2 and Table 1). P2 and P3 with more hydrophobic pendant groups exhibited higher hemolytic activity than P1. Thus, while increased hydrophobicity leads to better antimicrobial efficacy, it also dramatically increased the hemolytic activity of the polymers, making P2 and P3 less selective to bacterial cells over mammalian cells (Figure 2 and Table 1). The same trend was also observed in the results of cell viability assay using NIH-3T3 fibroblast cells, in which polymers with more hydrophobic pendant groups exhibited higher toxicity (Figure S9).

Figure 2. Hemolytic activity of P1, P2, and P3 Fluorescence Microscopy Study of E. coli

ACS Paragon Plus Environment

16

Page 17 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

In this work, we performed mechanistic studies of these antimicrobial polymers against Gram-negative E. coli and Gram-positive S. aureus to understand their mechanism of action. To determine the mechanism of action of our polymers against the bacterial membrane, we carried out microscopy studies in E. coli, which has an outer (periplasmic) membrane (compared with the Gram-positive S. aureus surrounded by a peptidoglycan cell wall). FITC-labeled antimicrobial polymers were added to E. coli cells and observed via fluorescence microscopy. As shown in Figure 3 and S4, the E. coli cells were labelled with the fluorescent polymer, indicating that the polymer was bound to the cell.

Figure 3. Fluorescence microscopy images of E. coli before treatment (left) and after treatment with P1-FITC for 30 min (right) Membrane Permeabilization of E. coli To determine the location of binding by these polymers, an outer membrane permeability assay for E. coli was carried out using a membrane-impermeable fluorescent probe N-

ACS Paragon Plus Environment

17

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 31

phenyl-1-naphthylamine (NPN), whose fluorescence is weak in an aqueous environment but strong in a hydrophobic environment. When the outer membrane is disrupted, NPN dye is able to enter the cell membrane and accumulate in a hydrophobic environment, leading to an increase in fluorescence intensity, which can be used as an indication of permeability. As shown in Figure 4, all three polymers permeabilized the outer membrane in a concentration-dependent manner, and polymers with more hydrophobic pendant groups caused greater permeability of the outer membrane (P3>P2>P1).

Figure 4. E. coli outer membrane permeability of P1, P2, and P3 measured using NPN as a fluorescent probe A cytoplasmic membrane permeability assay was performed using a colorimetric probe ortho-nitrophenol-β-galactoside (ONPG) to detect β-galactosidase enzymatic activity. In E. coli, β-galactosidase is the enzymatic product of lacZ, which catalyzes the hydrolysis of

ACS Paragon Plus Environment

18

Page 19 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

β-galactosides.35 Disruption of the cytoplasmic membrane would lead to the enzymatic reaction between β-galactosidase and ONPG (caused by the permeabilization of ONPG into the cells or the release of β-galactosidase into the medium), generating a yellow ortho-nitrophenol product, which has absorbance at 420 nm. As shown in Figure 5, P1 caused minimal activity as the absorbance at 420 nm was very close to the negative control even at concentrations significantly above its MIC. On the other hand, P2 and P3 showed significant β-galactosidase activity (P3>P2>P1), again in a concentration dependent fashion. These results indicate that P1 may exhibit weaker membrane disrupting ability compared to P2 and P3, indicating lower permeabilization of ONPG or comparatively lower release of β-galactosidase. However, the assay is not able to provide more detailed information on the mechanism (whether the enzymatic activity is a result of the permeabilization of ONPG or the release of enzyme) other than the overall membrane disrupting ability. Interestingly, Figure 5(d) showed that melittin exhibited very different kinetics compared to P2 and P3, in which the absorbance at 420 nm gradually increased over the test period, suggesting that melittin might have a different mechanism of action against E. coli compared to the described poly(ester urethane)s. Melittin can bind to and insert into the bacterial membrane, leading to the formation of toroidal pores. In the case of P2 and P3, the absorbance increased rapidly in the beginning and reached a plateau within 10 minutes. The results suggested that P2 and P3 would cause the formation of larger pores compared to melittin in the beginning, which led to

ACS Paragon Plus Environment

19

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 31

the rapid increase in the absorbance. The presence of plateau could be caused by the blockage of the pores by the antimicrobial polymers or the potential deactivation of βgalactosidase by the antimicrobial polymers.

Figure 5. E. coli cytoplasmic membrane permeability assay of P1 (a), P2 (b), P3 (c), and melittin (d) over a period of 60 min measured using ONPG as a colorimetric probe Further SEM studies were performed to observe the effect of the polymers on the bacterial morphology. As shown in Figure 6, the polymer treated bacteria exhibited rough and wrinkled surfaces (b, c, d), while the untreated bacterial cells exhibited a relatively smooth surface (a), indicating that these polymers disrupted the bacterial membrane.

ACS Paragon Plus Environment

20

Page 21 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

Figure 6. SEM images of E. coli before treatment (a) and after treatment with 16 μg/mL of P1 (b), P2 (c), and P3 (d) for 30 min; scale bar is 1 μm Membrane Depolarization of E. coli Since P1 was an effective antimicrobial polymer, but did not cause sufficient cytoplasmic membrane permeabilization to allow leakage of the β-galactosidase enzyme, we wondered whether the extent of membrane disruption caused by P1 would affect the ability of the cell to maintain the proton-motive force necessary for cell energetics. To test

ACS Paragon Plus Environment

21

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 31

this, we performed a cytoplasmic membrane depolarization assay for E. coli using DiSC3(5) probe, which is a membrane potential-sensitive cyanine dye. DiSC3(5) can readily enter cell membranes and accumulate in healthy polarized cytoplasmic membranes.36 Once accumulated in the cytoplasmic membrane, DiSC3(5) shows weak fluorescence due to self-quenching; if the antimicrobial polymers disrupt the cytoplasmic membrane and dissipates the membrane potential, DiSC3(5) will be released into the medium, leading to a corresponding increase of fluorescence intensity.37 As shown in Figure 7, all three polymers caused membrane depolarization, which confirmed the interactions between the polymers and the cytoplasmic membrane. This result also indicated that P1 was able to dissipate the membrane potential and cause cell death even though it did not cause sufficient damage to the cytoplasmic membrane to show any significant β-galactosidase activity (Figure 5(a)), since a functioning transmembrane potential is essential for metabolic activity and growth.38

Figure 7. E. coli cytoplasmic membrane depolarization by P1 (a), P2 (b), and P3 (c) over a period of 60 min measured using DiSC3(5) as a fluorescent probe

ACS Paragon Plus Environment

22

Page 23 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

Membrane Disrupting Ability Studied by Dye Release Assay To understand the effect of polymer structure on the membrane disrupting ability, a dye release assay was performed in which fluorescence detection of 6-carboxyfluorescein (CF) leakage from bacterial membrane-mimic liposomes was used as an indication of the membrane disrupting ability. The fluorescence intensity of CF is weak when encapsulated in the liposomes due to self-quenching. However, release of the dye from the liposomes due to liposome disruption would lead to an increase in the fluorescence intensity. In this study, a lipid composition that mimics the membrane of E. coli was used, which consists of 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoethanolamine (POPE) and 1palmitoyl-2-oleoyl-sn-glycero-3-phosphoglycerol (POPG) in a ratio of 4:1. The dye encapsulated liposomes were incubated with the antimicrobial polymers and the fluorescence intensity at 530 nm was monitored (495 nm excitation). As shown in Figure 8, all three polymers disrupted the liposomes, leading to CF release. The dye release was concentration dependent (b, c, d) and polymers with more hydrophobic pendant group caused more dye release (P3>P2>P1) at the same concentration (a), which correlated well with the trend of MIC values and killing rates of the three polymers.

ACS Paragon Plus Environment

23

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 31

Figure 8. (a) Summary of dye release from liposomes that were treated with the antimicrobial polymers at different concentration; dye release kinetics of the liposomes after treatment with P1 (b), P2 (c), and P3 (d) over a period of 60 min. Membrane Permeability of S. aureus In the absence of an outer (periplasmic) membrane in S. aureus, we wanted to determine whether our polymers could still affect the cytoplasmic membrane protected by the thick, peptidoglycan cell wall of this organism. Cytoplasmic membrane disruption of S. aureus was observed for all three polymers using the DiSC3(5) fluorescent probe (Figure 9), increasing with the hydrophobicity of the pendant groups (and correlated with

ACS Paragon Plus Environment

24

Page 25 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

antimicrobial activity), suggesting that these polymers could still access the cytoplasmic membrane, even in the presence of a thick cell wall.

Figure 9. Cytoplasmic membrane depolarization of S. aureus by P1 (a), P2 (b), and P3 (c) over a period of 30 min measured using DiSC3(5) as a fluorescent probe Conclusion A series of antimicrobial surfactant-like poly(ester urethane)s were designed and developed

via

step-growth

polymerization

followed

by

post-polymerization

modification using thiol-ene reaction. These novel antimicrobial polymers exhibited antimicrobial activity against E. coli, S. aureus, P. aeruginosa, and S. epidermidis. Microscopic analysis demonstrated that these polymers bind to and disrupt bacterial membranes, with their membrane disrupting ability depending on the hydrophilichydrophobic balance. The polymers with more hydrophobic pendant groups showed stronger tendency to disrupt the bacterial membrane, which correlated well with their antimicrobial activity. Membrane permeability assays showed that all three antimicrobial polymers were able to permeabilize the outer (periplasmic) and inner (cytoplasmic)

ACS Paragon Plus Environment

25

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 31

membrane of E. coli, and that the cell wall of S. aureus did not prevent these polymers from disrupting the cytoplasmic membrane. P2 and P3 that showed significant damage to the cytoplasmic membrane of E. coli, while P1 caused minor damage to the integrity of cytoplasmic membrane, but even such an extent of disruption was able to dissipate the cytoplasmic membrane potential and cause cell death. The structure-activity relationships and mechanistic insights from these studies can help guide the design and development of new antimicrobial polymers with improved antimicrobial activity and cytocompatiblity. ASSOCIATED CONTENT Supporting Information Molecular weight information, NMR spectra, and fluorescence microscopy images. This material is available free of charge via the Internet at http://pubs.acs.org. AUTHOR INFORMATION Corresponding Authors * (A.J.) E-mail: [email protected] * (H.A.B.) E-mail: [email protected] Notes The authors declare no competing financial interest.

ACS Paragon Plus Environment

26

Page 27 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

ACKNOWLEDGMENT The work described herein was funded in part by an NSF CAREER grant (NSF DMR no. 1352485). REFERENCES 1. Hancock, R. E. W.; Sahl, H.-G., Antimicrobial and host-defense peptides as new anti-infective therapeutic strategies. Nature Biotechnology 2006, 24, 1551. 2. Zhang, L.-j.; Gallo, R. L., Antimicrobial peptides. Current Biology 2016, 26, (1), R14R19. 3. Mahlapuu, M.; Håkansson, J.; Ringstad, L.; Björn, C., Antimicrobial Peptides: An Emerging Category of Therapeutic Agents. Frontiers in Cellular and Infection Microbiology 2016, 6, (194). 4. Marr, A. K.; Gooderham, W. J.; Hancock, R. E. W., Antibacterial peptides for therapeutic use: obstacles and realistic outlook. Current Opinion in Pharmacology 2006, 6, (5), 468-472. 5. Thennarasu, S.; Tan, A.; Penumatchu, R.; Shelburne, C. E.; Heyl, D. L.; Ramamoorthy, A., Antimicrobial and Membrane Disrupting Activities of a Peptide Derived from the Human Cathelicidin Antimicrobial Peptide LL37. Biophysical Journal 2010, 98, (2), 248-257. 6. Yin, L. M.; Edwards, M. A.; Li, J.; Yip, C. M.; Deber, C. M., Roles of Hydrophobicity and Charge Distribution of Cationic Antimicrobial Peptides in Peptide-Membrane Interactions. Journal of Biological Chemistry 2012, 287, (10), 7738-7745. 7. Kang, S.-J.; Park, S. J.; Mishig-Ochir, T.; Lee, B.-J., Antimicrobial peptides: therapeutic potentials. Expert Review of Anti-infective Therapy 2014, 12, (12), 1477-1486. 8. Zasloff, M., Antimicrobial peptides of multicellular organisms. Nature 2002, 415, 389. 9. Kenichi, K.; A., C. G., Antimicrobial polymers as synthetic mimics of host ‐ defense peptides. Wiley Interdisciplinary Reviews: Nanomedicine and Nanobiotechnology 2013, 5, (1), 49-66. 10. Santos, M. R. E.; Fonseca, A. C.; Mendonça, P. V.; Branco, R.; Serra, A. C.; Morais, P. V.; Coelho, J. F. J., Recent Developments in Antimicrobial Polymers: A Review. Materials 2016, 9, (7), 599. 11. Palermo, E. F.; Kuroda, K., Chemical Structure of Cationic Groups in Amphiphilic Polymethacrylates Modulates the Antimicrobial and Hemolytic Activities. Biomacromolecules 2009, 10, (6), 1416-1428.

ACS Paragon Plus Environment

27

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 31

12. Palermo, E. F.; Sovadinova, I.; Kuroda, K., Structural Determinants of Antimicrobial Activity and Biocompatibility in Membrane-Disrupting Methacrylamide Random Copolymers. Biomacromolecules 2009, 10, (11), 3098-3107. 13. Palermo, E. F.; Lee, D.-K.; Ramamoorthy, A.; Kuroda, K., Role of Cationic Group Structure in Membrane Binding and Disruption by Amphiphilic Copolymers. The Journal of Physical Chemistry B 2011, 115, (2), 366-375. 14. Yang, C.; Krishnamurthy, S.; Liu, J.; Liu, S.; Lu, X.; Coady, D. J.; Cheng, W.; Libero, G. D.; Singhal, A.; Hedrick, J. L.; Yang, Y. Y., Broad ‐ Spectrum Antimicrobial Star Polycarbonates Functionalized with Mannose for Targeting Bacteria Residing inside Immune Cells. Advanced Healthcare Materials 2016, 5, (11), 1272-1281. 15. Cheng, J.; Chin, W.; Dong, H.; Xu, L.; Zhong, G.; Huang, Y.; Li, L.; Xu, K.; Wu, M.; Hedrick, J. L.; Yang, Y. Y.; Fan, W., Biodegradable Antimicrobial Polycarbonates with In Vivo Efficacy against Multidrug ‐ Resistant MRSA Systemic Infection. Advanced Healthcare Materials 2015, 4, (14), 2128-2136. 16. Ong, Z. Y.; Yang, C.; Cheng, W.; Voo, Z. X.; Chin, W.; Hedrick, J. L.; Yang, Y. Y., Biodegradable cationic poly(carbonates): Effect of varying side chain hydrophobicity on key aspects of gene transfection. Acta Biomaterialia 2017, 54, 201-211. 17. Chakraborty, S.; Liu, R.; Lemke, J. J.; Hayouka, Z.; Welch, R. A.; Weisblum, B.; Masters, K. S.; Gellman, S. H., Effects of Cyclic vs Acyclic Hydrophobic Subunits on the Chemical Structure and Biological Properties of Nylon-3 Copolymers. ACS Macro Letters 2013, 2, (8), 753-756. 18. Lam, S. J.; O'Brien-Simpson, N. M.; Pantarat, N.; Sulistio, A.; Wong, E. H. H.; Chen, Y.-Y.; Lenzo, J. C.; Holden, J. A.; Blencowe, A.; Reynolds, E. C.; Qiao, G. G., Combating multidrug-resistant Gram-negative bacteria with structurally nanoengineered antimicrobial peptide polymers. Nature Microbiology 2016, 1, 16162. 19. Lienkamp, K.; Madkour, A. E.; Musante, A.; Nelson, C. F.; Nüsslein, K.; Tew, G. N., Antimicrobial Polymers Prepared by ROMP with Unprecedented Selectivity: A Molecular Construction Kit Approach. Journal of the American Chemical Society 2008, 130, (30), 9836-9843. 20. Lienkamp, K.; Madkour, A. E.; Kumar, K. N.; Nüsslein, K.; Tew, G. N., Antimicrobial Polymers Prepared by Ring ‐ Opening Metathesis Polymerization: Manipulating Antimicrobial Properties by Organic Counterion and Charge Density Variation. Chemistry – A European Journal 2009, 15, (43), 11715-11722. 21. Mankoci, S.; Kaiser, R. L.; Sahai, N.; Barton, H. A.; Joy, A., Bactericidal Peptidomimetic Polyurethanes with Remarkable Selectivity against Escherichia coli. ACS Biomaterials Science & Engineering 2017, 3, (10), 2588-2597. 22. Jiang, Y.; Yang, X.; Zhu, R.; Hu, K.; Lan, W.-W.; Wu, F.; Yang, L., Acid-Activated Antimicrobial Random Copolymers: A Mechanism-Guided Design of Antimicrobial Peptide Mimics. Macromolecules 2013, 46, (10), 3959-3964.

ACS Paragon Plus Environment

28

Page 29 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

23. Taresco, V.; Crisante, F.; Francolini, I.; Martinelli, A.; D’Ilario, L.; Ricci-Vitiani, L.; Buccarelli, M.; Pietrelli, L.; Piozzi, A., Antimicrobial and antioxidant amphiphilic random copolymers to address medical device-centered infections. Acta Biomaterialia 2015, 22, 131-140. 24. Hong, S.; Takahashi, H.; Nadres, E. T.; Mortazavian, H.; Caputo, G. A.; Younger, J. G.; Kuroda, K., A Cationic Amphiphilic Random Copolymer with pH-Responsive Activity against Methicillin-Resistant Staphylococcus aureus. PLOS ONE 2017, 12, (1), e0169262. 25. Allison, B. C.; Applegate, B. M.; Youngblood, J. P., Hemocompatibility of Hydrophilic Antimicrobial Copolymers of Alkylated 4-Vinylpyridine. Biomacromolecules 2007, 8, (10), 2995-2999. 26. Nimmagadda, A.; Liu, X.; Teng, P.; Su, M.; Li, Y.; Qiao, Q.; Khadka, N. K.; Sun, X.; Pan, J.; Xu, H.; Li, Q.; Cai, J., Polycarbonates with Potent and Selective Antimicrobial Activity toward Gram-Positive Bacteria. Biomacromolecules 2017, 18, (1), 87-95. 27. Lam, S. J.; Wong, E. H. H.; Boyer, C.; Qiao, G. G., Antimicrobial polymeric nanoparticles. Progress in Polymer Science 2018, 76, 40-64. 28. Gabriel, G. J.; Maegerlein, J. A.; Nelson, C. F.; Dabkowski, J. M.; Eren, T.; Nüsslein, K.; Tew, G. N., Comparison of Facially Amphiphilic versus Segregated Monomers in the Design of Antibacterial Copolymers. Chemistry (Weinheim an der Bergstrasse, Germany) 2009, 15, (2), 10.1002/chem.200801233. 29. Sambhy, V.; Peterson, B. R.; Sen, A., Antibacterial and Hemolytic Activities of Pyridinium Polymers as a Function of the Spatial Relationship between the Positive Charge and the Pendant Alkyl Tail. Angewandte Chemie International Edition 2008, 47, (7), 1250-1254. 30. Peng, C.; Joy, A., Baylis–Hillman Reaction as a Versatile Platform for the Synthesis of Diverse Functionalized Polymers by Chain and Step Polymerization. Macromolecules 2014, 47, (4), 1258-1268. 31. George, R.; Sidney, F.; Akira, Y., Two dimensional thin layer chromatographic separation of polar lipids and determination of phospholipids by phosphorus analysis of spots. Lipids 1970, 5, (5), 494-496. 32. Peng, C.; Joy, A., Self‐emulsion polymerization of baylis–hillman‐derived α ‐ hydroxymethyl ‐ substituted acrylates. Journal of Polymer Science Part A: Polymer Chemistry 2015, 53, (15), 1743-1747. 33. Jiang, N.; Tan, N. S.; Ho, B.; Ding, J. L., Measurement of the red blood cell lysis by bacterial hemolysin. Protocol Exchange 2007. doi:10.1038/nprot.2007.483 34. Preté, P. S. C.; Malheiros, S. V. P.; Meirelles, N. C.; de Paula, E., Quantitative assessment of human erythrocyte membrane solubilization by Triton X-100. Biophysical Chemistry 2002, 97, (1), 1-5.

ACS Paragon Plus Environment

29

Biomacromolecules 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 31

35. Juers, D. H.; Matthews, B. W.; Huber, R. E., LacZ β-galactosidase: Structure and function of an enzyme of historical and molecular biological importance. Protein Science : A Publication of the Protein Society 2012, 21, (12), 1792-1807. 36. Wu, M.; Maier, E.; Benz, R.; Hancock, R. E. W., Mechanism of Interaction of Different Classes of Cationic Antimicrobial Peptides with Planar Bilayers and with the Cytoplasmic Membrane of Escherichia coli. Biochemistry 1999, 38, (22), 7235-7242. 37. Sims, P. J.; Waggoner, A. S.; Wang, C.-H.; Hoffman, J. F., Mechanism by which cyanine dyes measure membrane potential in red blood cells and phosphatidylcholine vesicles. Biochemistry 1974, 13, (16), 3315-3330. 38. Nebe-von-Caron, G.; Stephens, P. J.; Hewitt, C. J.; Powell, J. R.; Badley, R. A., Analysis of bacterial function by multi-colour fluorescence flow cytometry and single cell sorting. Journal of Microbiological Methods 2000, 42, (1), 97-114.

ACS Paragon Plus Environment

30

Page 31 of 31 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

Antimicrobial Poly(ester urethane)s 148x103mm (150 x 150 DPI)

ACS Paragon Plus Environment