You are on page 1of 12

Polymer

Chemistry
View Article Online
PAPER View Journal | View Issue

Methacrylate-ended polypeptides and


Published on 26 September 2017. Downloaded by Iowa State University on 1/24/2019 7:45:54 AM.

Cite this: Polym. Chem., 2017, 8,


polypeptoids for antimicrobial and antifouling
6386 coatings†
Qiang Gao,a Peng Li, *b Hongyang Zhao,c Yashao Chen,d Liu Jiangd and
Peter X. Ma*a,e,f,g,h

Methacrylate-ended polypeptides/polypeptoids were successfully synthesized via ring-opening


polymerization (ROP) of N-carboxyanhydrides (NCA). These oligomers were further initiated under ultra-
violet (UV) irradiation by a polydopamine ( pDA) layer which is attachable to the surface of virtually all
materials to generate a polymer brush coating. This brush-like polymer coating comprising cationic anti-
microbial polypeptides (MePpep) and antifouling polysarcosine (MePsar) exhibited effective antimicrobial
activity against four pathogens (Staphylococcus aureus, Escherichia coli, Pseudomonas aeruginosa, and
Candida albicans), as well as antifouling activity in the resistance to protein and platelet adhesion, and
thus prevented biofilm formation for up to 7 days. An in vitro cytotoxicity study showed that this coating is
Received 3rd September 2017, biocompatible with mouse fibroblast (L929) cells. More importantly, this coating exhibited significant
Accepted 25th September 2017
anti-infectivity in vivo. This dual-functional polymer brush coating can be immobilized on the surface of
DOI: 10.1039/c7py01495c multiple categories of materials through the mussel-inspired pDA coating, and thus should be widely
rsc.li/polymers applicable for combating infection in many classes of bio-medical materials.

1. Introduction human beings.1 According to the estimation in a review com-


missioned by the UK Government, 10 million people could die
The worldwide overuse of antibiotics has become a recent each year from drug-resistant bacteria by 2050.2 Also, the
cause for concern due to its link to the emergence of resistant development of new antibiotics is far surpassed by the rate of
microbial strains. “Superbugs” that show resistance to most development of bacterial resistance. To update the arsenal
current antibiotics are a prospective unimaginable threat to all against bacteria, scientists have taken different approaches.
Instead of traditional antibiotics, natural antimicrobial pep-
tides (AMPs, also known as host defense peptides), which are
a less likely to trigger resistance, have attracted more attention
Center of Biomedical and Engineering and Regenerative Medicine, Frontier Institute
of Science and Technology, Xi’an Jiaotong University, Xi’an 710054, China. recently.3 However, AMPs exist in very low concentrations natu-
E-mail: mapx@umich.edu; Fax: +(86)29-83395131; Tel: +(86)29-83395361 rally, and the production of these species is expensive. Hence,
b
Key Laboratory of Flexible Electronics (KLOFE) and Institute of Advanced Materials a number of chemical synthetic AMPs and their mimics have
(IAM), Jiangsu National Synergetic Innovation Center for Advanced Materials been developed.4–13 Recently, the ring-opening polymerization
(SICAM), Nanjing Tech University, Nanjing 211816, China.
(ROP) of α-/N-substituted amino acid N-carboxyanhydride
E-mail: iampli@njtech.edu.cn; Fax: +(86) 25-83587982; Tel: +(86) 25-83587982
c
Center of Applied Chemical Research, Frontier Institute of Science and Technology, (NCA) monomers for the synthesis of polypeptides and poly-
Xi’an Jiaotong University, Xi’an 710054, China peptoids has gained prominence.14,15 Although this approach
d
Key Laboratory of Applied Surface and Colloid Chemistry, School of Chemical and cannot generate precise amino acid sequences, it allows the
Chemical Engineering, Shaanxi Normal University, Xi’an 710119, China facile and low-cost synthesis of random or block co-polypep-
e
Department of Biomedical Engineering, University of Michigan, Ann Arbor,
tides/polypeptoids with potential applications, such as in anti-
Michigan 48109, USA
f
Department of Biologic and Materials Sciences, University of Michigan, Ann Arbor, microbials, tissue engineering, drug delivery, coatings, etc.16–22
Michigan 48109, USA Microbial colonization and growth on the surface of
g
Macromolecular Science and Engineering Center, University of Michigan, Ann Arbor, implantable materials and medical devices (e.g., catheters,
Michigan 48109, USA indwelling needles, hip replacements, etc.) is a serious
h
Department of Materials Science and Engineering, University of Michigan,
problem that imperils human health, especially when caused
Ann Arbor, Michigan 48109, USA
† Electronic supplementary information (ESI) available. See DOI: 10.1039/ by drug-resistant “superbugs”.23 Surface-associated bacterial
c7py01495c communities embedded in the extracellular polymeric matrix,

6386 | Polym. Chem., 2017, 8, 6386–6397 This journal is © The Royal Society of Chemistry 2017
View Article Online

Polymer Chemistry Paper

i.e., so-called biofilms, further increase the resistance towards studies indicated that PEG can be easily oxidatively degraded,
conventional drugs.24 Recent efforts have been dedicated to and its degradation products may have unexpected toxicity.49,50
preparing AMP coatings in order to inhibit bacterial attach- Polysarcosine (Psar), a hydrophilic homo-polypeptoid, has
ment and thereby prevent subsequent biofilm formation.25,26 exhibited excellent antifouling activity comparable with PEG.51
AMPs have been immobilized on surfaces by surface graft Furthermore, Psar is originating from the natural and non-
polymerization,27,28 click chemistry,29 layer-by-layer assem- toxic amino acid, which is more safe and reliable for bio-
bly,30 etc. However, most of the current methods involve multi- medical usage. Thus Psar is an excellent alternative to PEG. In
step treatments or post-synthesis.31,32 The post-synthesis of addition, the synthesis of block copolymers is also tedious.
AMPs usually requires rigorous chemistry conditions and may Hence, our main target herein is to design facile synthesized
Published on 26 September 2017. Downloaded by Iowa State University on 1/24/2019 7:45:54 AM.

also change the biological properties of the AMPs. functional biomolecules for the development of anti-infective
Furthermore, these surface modifications generally do not coatings that are attachable to multiple categories of surfaces
fulfill the requirements of different categories of substances. via a simple method under ambient conditions without using
Recently, research groups have expended more effort to harsh chemicals or conditions.
develop a simple and universal method of modifying the In order to avoid the post-synthesis of biomolecules, metha-
surface of biomaterials. Mussel-inspired materials for bio- crylate-ended peptides, as well as peptoids, were synthesized
medical applications (e.g., coatings, hydrogels and drug car- via NCA chemistry (Scheme 1A). These methacrylate-ended
riers) are being extensively studied.33–35 Zhou et al. found that polymers (MePs) were ready-made for further polymerization
a mussel-inspired pDA coating which was attachable to vir- to form a functional surface coating via facile mussel-inspired
tually all inorganic and organic surfaces, could generate free surface modification (Scheme 1B). 2-Aminoethyl methacrylate
radicals under irradiation and initiate the surface grafting hydrochloride (AEMA) was used to initiate the methacrylate-
reaction of vinyl monomers.36 ended antimicrobial polypeptides (MePpep) and antifouling
Although immobilized antimicrobial molecules such as polypeptoids (MePsar). Cationic antimicrobial MePpep, which
AMPs and their synthetic mimics were capable of killing the was synthesized by random copolymerization of lysine (Lys)
bacteria on contact, these coatings were often stuck with and phenylalanine (Phe) residues, had an excellent minimal
fouling issues.37 Non-specific adhesion of protein, dead cells inhibitory concentration (MIC) (25 μg mL−1, 4.9 µM) toward
and debris on the surface blocks the coating from contact with Gram-positive bacteria (S. aureus), Gram-negative bacteria
bacteria, resulting in the loss of its efficacy. Thus the use of (E. coli and P. aeruginosa), and fungus (C. albicans) (ESI
just antimicrobial molecules was not enough for an effective and Table S1†). MePsar with excellent protein/cell repellent pro-
long-lasting anti-infective coating.38 Coatings which retain not perties and biocompatibility was synthesized by the polymeriz-
only antimicrobial activity but also antifouling or cell-releasing ation of sarcosine-NCA monomers.52 The methacrylate groups
features have been reported lately.39–45 In our recent work, at the end of these biomolecules undergo initiation by free
block copolymers containing bactericidal and cell/protein- radicals generated on the pDA film under UV, and then
repellent segments were synthesized.46–48 Polyethylene glycol polymerized into surface polymer brushes. Two brush coat-
(PEG) was used as the antifouling segment. However, some ings, poly(Ppep) and poly(Ppep/Psar), were successfully pre-

Scheme 1 Synthesis of methacrylate-ended polymers (MePs) and pDA-assisted grafting of MePs. (A) Synthesis of methacrylate-ended polypeptide
(MePpep) and polysarcosine (MePsar). (B) Schematic illustration of thin-coating deposition of pDA (Step 1) and surface grafting of MeP polymer
brushes through UV irradiation (Step 2).

This journal is © The Royal Society of Chemistry 2017 Polym. Chem., 2017, 8, 6386–6397 | 6387
View Article Online

Paper Polymer Chemistry

pared, as confirmed by various characterization techniques. drous DMF and subsequently added to the reaction mixture.
Comparatively, the poly(Ppep/Psar) coating not only exhibited The reaction solution was stirred at 60 °C for 72 h. After com-
exceptional antimicrobial activity with an inhibition rate of pletion of the reaction, the methacrylate-ended polypeptide
above 94.6% for all four clinically significant pathogens, but also precursor (Lys(CBz)12.5-ran-Phe12.5) was precipitated in 300 mL
greatly reduced non-specific protein adsorption and platelet of cold ether and centrifuged (3200g, 5 min), and then dried
adhesion in vitro. Additionally, this polymer brush coating was under reduced pressure (white solid, 3.11 g, 92%). 1H-NMR
non-toxic to mammalian cells in vitro, and retained its excel- (400 MHz, d6-DMSO): δH 1.22–1.77 (m, –CH2CH2CH2–), 1.86 (s,
lent biological properties in a rat infection model, successfully –CH3–), 2.94–3.22 (m, –CH2–), 4.21–4.48 (m, –CH–), 4.99
preventing implant-related infection by S. aureus. Besides its (C6H5–CH2–), 5.66 (s, vCHH), 6.05 (s, vCHH), 7.21–7.33 (m,
Published on 26 September 2017. Downloaded by Iowa State University on 1/24/2019 7:45:54 AM.

use as an anti-infective coating, we anticipate that this category –C6H5–) ppm (ESI Fig. S2†).
of methacrylate-ended polypeptide/polypeptoid oligomers will Deprotection of the benzyl-protected polypeptide precursor
also be applicable in other areas, such as tissue engineering, (MePpep(z), 200 mg) was performed by the addition of HBr
drug delivery, etc. solution (300 µL, 33% in acetic acid), followed by continuous
agitation at 0 °C for 1 h. The deprotected polypeptide (Lys
(NH3+)12.5-ran-phe12.5) was precipitated by the addition of cold
2. Experimental section ether (10 mL) and centrifuged (3200g, 5 min), and then col-
lected solid precipitate was washed with ether thrice (10 mL).
2.1. Materials
Finally, the polypeptide was dialyzed (MWCO 3500) against de-
H-Lys(z)-OH (99%), L-phenylalanine (99%), dopamine hydro- ionized water for 2 days and subsequently freeze-dried (fluffy
chloride (DA, 98%), and 2-aminoethyl methacrylate hydro- white solid, 175 mg, 87.5%). 1H-NMR (400 MHz, d6-DMSO):
chloride (AEMA, 90%) were obtained from Sigma-Aldrich δH 1.35–1.65 (m, 70H, –CH2CH2CH2–), 1.86 (s, 3H, –CH3–),
(USA). Anhydrous N,N-dimethylformamide (DMF, >99.8%), tri- 2.95–3.16 (m, 48H, –CH2–), 4.21–4.56 (m, 24H, –CH–), 5.70 (s,
phosgene (98%), hydrogen bromide (33 wt% solution in acetic 1H, vCHH), 6.06 (s, 1H, vCHH), 7.27 (m, 60H, –C6H5–) ppm.
acid), anhydrous sodium sulfate (99%), bovine serum albumin Mw = 5080 g mol−1 (ESI Fig. S2†).
(BSA, Mn = 66 000 g mol−1, 98%), and lysozyme (Mn = 14 400
g mol−1, 98%) were purchased from J&K Scientific Ltd, China. 2.4. Synthesis of methacrylate-ended polysarcosine (MePsar)
Anhydrous tetrahydrofuran (THF), ethyl ether, and n-hexane Sar-NCA (1.56 g, 50 mmol) was dissolved in 16 mL of
were obtained by using a solvent-drying system (PureSolv anhydrous DMF. AEMA (166 mg, 1 mmol) was dissolved in
MD-5, Innovative Technology). Staphylococcus aureus (ATCC another 5 mL of anhydrous DMF and subsequently added to
29253), Escherichia coli (ATCC 25922), Pseudomonas aeruginosa the reaction mixture. The reaction solution was stirred at
(ATCC 27853), and Candida albicans (ATCC 10231) were 60 °C for 72 h. Directly after completion of the reaction,
obtained from the American Type Culture Collection (ATCC). the polymer was precipitated in 200 mL of cold ether,
The AlamarBlue® cell viability assay reagent, LIVE/DEAD® centrifuged (3200g, 5 min), and then isolated polysarcosine
BacLight™ bacterial viability kit, and Micro BCA™ protein was dried under reduced pressure, dissolved in water, and
assay kit were purchased from Thermo Fisher Scientific Inc. subsequently freeze-dried (light yellow solid, 1.42 g, 91%).
Phosphate buffered saline (PBS) was obtained from Sigma- 1
H-NMR (400 MHz, d6-DMSO): δH 1.86 (s, 3H, –CH3), 2.75–3.25
Aldrich and sterilized in an autoclave at 121 °C before use. (m, 145H, N–CH3), 4.14–4.44 (m, 90H, –CH2–), 5.67 (s, 1H,
Deionized water used in the experiments was purified by using vCHH), 6.07 (s, 1H, vCHH) ppm. Mw = 3365 g mol−1
a Millipore water purification system with a minimum resis- (ESI Fig. S3†).
tivity of 18.2 MΩ cm. Poly(dimethylsiloxane) (PDMS) was pre-
pared by mixing a silicone elastomer and a curing agent in a 2.5. Measurement of minimal inhibitory concentration (MIC)
weight ratio of 10 : 1 (Sylgard 184, Dow Corning, MI, USA) and The antimicrobial activity of the synthesized MePs against
curing in 90 mm Petri dishes for polymerization at 80 °C S. aureus (Gram positive), E. coli and P. aeruginosa (Gram
for 12 h. negative), and C. albicans (fungi) was investigated by using a
broth micro-dilution MIC assay. Briefly, 100 μL of MeP solu-
2.2. Synthesis of NCA monomers
tion in Mueller-Hinton broth (MHB, for bacteria) or Yeast-Malt
Lys(z)-NCA, Phe-NCA, and Sar-NCA were synthesized as pre- broth (YMB, for fungi) at different concentrations (3.1, 6.3,
viously reported in the literature and confirmed by 1H-NMR 12.5, 25, 50, 100, 200, and 400 µg mL−1) was filled into the
(ESI Fig. S1†).48,53 All synthesized NCA monomers were herme- wells of a 96-well culture plate by a 2-fold dilution method.
tically stored at −80 °C before use. Another 100 µL of 106 CFU mL−1 bacterial suspension was
then carefully added to each well. The plate was incubated at
2.3. Synthesis of methacrylate-ended polypeptides (MePpep) 37 °C for 24 h (28 °C for C. albicans), and the turbidity was
Lys(z)-NCA (2.08 g, 12.5 mmol) and Phe-NCA (1.30 g, read by a microplate reader (SpectraMax Paradigm, Molecular
12.5 mmol) were dissolved in 35 mL of anhydrous DMF under Devices) at 600 nm. The MIC was determined to be the poly-
a nitrogen atmosphere in a glovebox. The initiator, AEMA peptide concentration at which no bacterial growth could be
(166 mg, 1 mmol), was dissolved in another 5 mL of anhy- observed.

6388 | Polym. Chem., 2017, 8, 6386–6397 This journal is © The Royal Society of Chemistry 2017
View Article Online

Polymer Chemistry Paper

2.6. Photo-polymerization of polymer brush coatings of MePs dried under a stream of nitrogen. All samples were pre-coated
Pristine PDMS films were cut into 1 × 1 cm squares and sub- with gold and observed using FE-SEM.
sequently ultrasonically cleaned in hexane for 1 h, then rinsed 2.9. Protein adsorption assay
thoroughly with a copious amount of acetone, and finally
vacuum dried overnight. pDA coating was carried out by The coating was directly prepared on the bottom surface of the
immersing the substrate in a buffer solution (10 mM Tris-HCl, wells in a 12-well plate according to the method described in
section 2.6; the protein adsorption properties were determined
pH 8.5) containing 2 mg mL−1 of DA, followed by shaking (200
by the Micro BCA assay. Briefly, the uncoated and polymer
rpm) at room temperature for 2 h with exposure to air. All
brush-coated surfaces were equilibrated in PBS ( pH 7.4) for
Published on 26 September 2017. Downloaded by Iowa State University on 1/24/2019 7:45:54 AM.

samples were then rinsed with deionized water and dried


12 h; 2 mL of BSA and lysozyme protein solution (5 mg mL−1
under a stream of nitrogen. The pDA-coated substrates were
in PBS) was added to the wells and incubated at 37 °C for 24 h.
immersed in 2 mL of MePpep solution (25 wt%, 4.92 µM) or
After this protein adsorption period, each well was washed
MePpep/MePsar mixed solution (12.5/8.3, wt%, 2.46/2.46 µM)
with PBS and distilled water twice, respectively. A 1 mL aliquot
respectively, then placed in a quartz reaction vessel and
of 2.0 wt% aqueous sodium dodecyl sulfate (SDS) solution was
degassed with nitrogen for 30 min. The reaction vessel was
then added to each well, followed by shaking for 2 h and sub-
subjected to UV irradiation (365 nm, 10 mW cm−2) under
sequent sonication for 1 h in an ice bath to detach the surface-
nitrogen protection for 1 h in a photochemical reactor (XPA-5,
absorbed protein. A 100 μL sample of the protein solution was
Xujiang Machine-electronic Plant, China). The coated sub-
collected from each well and transferred to a new 96-well plate,
strates were then rinsed with deionized water and dried under
supplemented with 100 μL of BCA reagent and incubated at
a stream of nitrogen.
60 °C for 1 h. The protein concentration was measured based
2.7. In vitro antimicrobial test on the absorbance at 560 nm by using a microplate reader
(SpectraMax Paradigm, Molecular Devices). Here, the efficiency
The in vitro bactericidal potency of the coatings was deter-
of the protein resistance was determined by applying the
mined by following a contact protocol. Firstly, the bacteria equation: protein resistance (%) = [(O.D.TCPS − O.D.coatings)/
(S. aureus, E. coli, and P. aeruginosa) were pre-cultured in MHB O.D.TCPS] × 100.
at 37 °C (YMB for C. albicans at 28 °C) and subcultured until
their optical density at 600 nm reached 0.5–0.7. The bacteria 2.10. Platelet adhesion study
were then harvested by centrifugation (1400g, 5 min), washed Platelet-rich plasma (PRP) was obtained by centrifuging
with PBS thrice, and resuspended in PBS at a final concen- citrated whole blood (424g, 15 min).55 PRP (50 μL) was intro-
tration of ∼108 CFU mL−1. A 10 μL-aliquot of the inoculum duced onto the uncoated, poly(Ppep)-, and poly(Ppep/Psar)-
suspension was pipetted and spread onto the uncoated coated (1 × 1 cm2) samples. After incubation at 37 °C for 2 h,
(control), poly(Ppep)- and poly(Ppep/Psar)-coated PDMS sur- all samples were washed thrice with PBS. The adherent plate-
faces (1 × 1 cm2). After a contact time of 1 h, all samples were lets were then fixed with 2.5% glutaraldehyde at 4 °C for 2 h,
placed in 1 mL of PBS and vigorously washed and sonicated dehydrated through a series of graded ethanol solutions (10%,
for 10 min; the surviving bacteria were plated with 10-fold 30%, 50%, 70%, 90%, and 100%, each for 15 min), and dried
serial dilutions and counted. The antimicrobial rate was calcu- under a stream of nitrogen. All samples were pre-coated with
lated as: kill% = [(CFUControl − CFUCoatings)/(CFUControl )] × 100. gold and observed using FE-SEM.
All experiments were carried out in triplicate for each
formulation. LIVE/DEAD staining was also performed to 2.11. In vitro biocompatibility of poly(Ppep/Psar) coating
investigate the viability of the bacteria upon contact with these All samples were cut into disks with a diameter of 5 mm and
coatings. sterilized with 70% ethanol prior to use. L929 cells were
seeded on a tissue culture polystyrene (TCPS) plate at a density
2.8. Field-emission scanning electron microscopy (FE-SEM)
of 6000 cells per cm2 and cultured for 4 h to allow attachment
of biofilms
of the cells. The samples were then gently placed onto the
The bacterial attachment and biofilm formation by S. aureus well, allowing the surface coating to make contact with the
and E. coli on the surfaces were observed by FE-SEM.54 Fresh cells. The cells were cultured for 4 days, and the medium was
bacteria at the logarithmic stage of growth were collected and changed every two days. The cell viability was determined by
resuspended (to OD600 = 0.01) in MHB. The uncoated, poly AlamarBlue and LIVE/DEAD assays at day 1 and 4.56
(Ppep)-, and poly(Ppep/Psar)-coated PDMS films (1 × 1 cm2)
were individually immersed in 2 mL of the bacterial suspen- 2.12. In vivo evaluation of infections
sion and incubated at 37 °C for 7 days to allow biofilm growth. All procedures were carried out according to the Regulation on
The broth was changed every day. At day 1 and 7, the samples the Administration of Laboratory Animals of China (2017 revi-
were washed thrice with sterile PBS to remove the unattached sion) and approved by the Animal Ethical Committee of the
bacteria, then fixed with 2.5% glutaraldehyde for 2 h, de- Xi’an Jiaotong University. Female Sprague-Dawley rats
hydrated through a series of graded ethanol solutions (10%, (8 weeks, ∼200 g) were purchased from the Laboratory Animal
30%, 50%, 70%, 90%, and 100% each for 15 min), and then Unit of the College of Medicine, Xi’an Jiaotong University and

This journal is © The Royal Society of Chemistry 2017 Polym. Chem., 2017, 8, 6386–6397 | 6389
View Article Online

Paper Polymer Chemistry

allowed to acclimatize for 7 days in the lab. A rat subcutaneous protected MePpep appeared at 5.66 and 6.05 ppm, and the
S. aureus infection model was generated using a pre-seeding signal of the Cbz groups appeared at 4.99 ppm and dis-
protocol. The uncoated PDMS (control) and poly(Ppep/Psar)- appeared after the deprotection process (ESI Fig. S2†). For
coated PDMS films (0.2 × 0.6 cm) were firstly inoculated with MePsar, two obvious and assignable signals were apparent at
10 µL of S. aureus suspension (108 CFU mL−1 in PBS) and then 2.75–3.15 ppm (m, Ha) and 3.85–4.45 ppm (m, Hb) and the
subcutaneously implanted into the back of the rats. Post- degree of polymerization was determined to be about 45 (ESI
operatively, all rats were monitored daily. All of the implants Fig. S3†). The FTIR results further confirmed the successful
were retrieved after 5 days, then individually placed in 1 mL of polymerization. The two characteristic stretching vibrations of
sterile PBS and sonicated for 5 min to detach the adhered bac- the NCA monomers at 1850 cm−1 and 1790 cm−1 disappeared
Published on 26 September 2017. Downloaded by Iowa State University on 1/24/2019 7:45:54 AM.

terial cells. The bacteria in the resulting suspensions were after initiation with AEMA (Fig. S4†).
counted after serial dilutions and plating. For histological In order to achieve potent antimicrobial activity, the Lys/
observations, the tissues around the implants were sectioned Phe ratio was set as 1 : 1 and the molar feed ratio of the
to a thickness of 1–2 mm with a sterilized pair of scissors and monomer/initiator ([MLys + MPhe]/[IAEMA]) was 25, according to
immediately fixed in 4% neutral formalin at 4 °C for 4 days, the optimized composition described in previous studies.16,48
and then paraffin embedded. The paraffin embedded tissues The NMR results confirmed that the degree of polymerization
were sectioned and stained with hematoxylin–eosin. of MePpep was about 24 and the content of hydrophilic
groups (Lys) was about 50% (Fig. S2C†). The antimicrobial
2.13. Characterization activity of the synthesized MePs was determined with four
1 pathogenic microbes. Cationic MePpep had a dramatically
H-NMR spectra were recorded on a Bruker Avance 400 MHz
spectrometer. Fourier transform infrared (FTIR) spectra were lower MIC of 25 μg mL−1 (4.9 µM) for the four clinically signifi-
collected on a Nexus 670 FTIR ESP spectrometer (Nicolet cant pathogens: Gram-positive (S. aureus), Gram-negative
Instrument Corp., Waltham, USA) at a resolution of 4 cm−1 (E. coli and P. aeruginosa), and fungi (C. albicans). MePsar,
using 32 scans. FE-SEM analysis was performed on a Hitachi which lacks the cationic and hydrophobic groups, did not
S-3000N instrument. The surface morphology and roughness exhibit significant inhibition of the four pathogens at concen-
of the samples were analyzed by atomic force microscopy trations of up to 1000 μg mL−1 (297 μM).
(AFM). AFM measurements were performed with a MultiMode
3.2. Polymer brush coatings of MePs
STM microscope controlled by the NanoScope III form the
Digital Instruments system, operating in tapping mode at a fre- Methacrylate-ended molecules are capable of polymerizing in
quency of 1 Hz. The surface chemical composition was ana- the presence of free radicals. The mussel-inspired pDA film is
lyzed by X-ray photoelectron spectroscopy (XPS) using an Axis attachable to virtually all organic, inorganic and metallic sur-
UltraDLD spectrometer (Kratos Analytical Ltd, UK) with a mono- faces, and is capable of generating free radicals (e.g., CO•
chromatic Al-Kα X-ray source (1486.71 eV photons). The static and C•) under UV irradiation to initiate the graft polymeriz-
contact angle of the surface was measured at room tempera- ation of vinyl monomers.36 Inspired by the universality of the
ture by the sessile drop method using a 2 μL water droplet in a photo-initiated polymerization using the pDA coating, we
Ramé-Hart telescopic goniometer. decided to use a similar strategy to graft the MePs onto the
surface of pDA pre-coated materials to generate antimicrobial
2.14. Statistical analysis and antifouling coatings. To produce a polymer brush coating,
The results are expressed as the mean ± standard deviation. pDA-coated PDMS was immersed in an aqueous solution com-
All experimental data were analyzed using the Student’s t test; prising a mixture of MePpep and MePsar and irradiated with
p < 0.05 was considered to be statistically significant. UV light in a nitrogen environment. As shown in Fig. 1A, the

3. Results and discussion


3.1. Synthesis, characterization, and biological activities of
MePs
Advancements in the ROP-NCA method have offered a facile
route for the synthesis of various polypeptides or polypeptoids.
Ammonium salts are relatively stable and easy to handle and
purify and can initiate the ROP of NCA monomers.57 Here,
methacrylate-ended antimicrobial polypeptides (MePpep) and
antifouling polypeptoids (MePsar) were successfully syn-
thesized using AEMA as the initiator and characterized by
1
H-NMR and FTIR analyses (ESI Fig. S1–S4†). 1H-NMR analysis
confirmed the polymerization and provided further infor- Fig. 1 (A) ATR-FTIR spectra of the uncoated PDMS (i), pDA (ii) and poly
mation. The 1H-NMR signals of the methacrylate group of the (Ppep/Psar) (iii) coated PDMS surfaces. (B) Contact angle measurement.

6390 | Polym. Chem., 2017, 8, 6386–6397 This journal is © The Royal Society of Chemistry 2017
View Article Online

Polymer Chemistry Paper

attenuated total reflection (ATR)-FTIR spectra of the uncoated The presence of the polymer brush coating was also deter-
PDMS, pDA-coated PDMS, and poly(Ppep/Psar)-coated mined by AFM and XPS characterization. Fig. 2A-i and ii show
PDMS surfaces presented three groups of well-defined that the uncoated PDMS surface was flat, with a roughness
peaks, indicating the successful formation of the coating. (root mean square, RMS) of ∼0.6 nm. The nitrogen peak was
The peaks around 3400 cm−1 were ascribed to the aromatic almost undetectable under XPS (Fig. 2A-iii). Fig. 2B-i and ii
–NHx and –OH stretching vibrations, and the intense, broad show a homogeneous pDA coating with an increased rough-
peaks at 3000–2750 cm−1 in the spectrum of the poly(Ppep/ ness of 58.4 nm. Two new peaks were observed at 399.2 and
Psar) coating were assigned to the CH2 stretching vibrations 401.6 eV (Fig. 2B-iii), which are attributed to the major second-
of Group I. The peaks at 1720 cm−1, 1640 cm−1, and ary amine and minor amine groups in the pDA coating.60 As
Published on 26 September 2017. Downloaded by Iowa State University on 1/24/2019 7:45:54 AM.

1530 cm−1 are the representative peaks of the CvO, amide shown in Fig. 2C-i and ii, the poly(Ppep) coating exhibited a
I, and amide II bonds of the polypeptides (Group II). 48,58 roughness of 35.4 nm, which was a bit lower than the pDA
The peaks at 1260 cm−1 (–Si–CH3), 1000 cm−1 (Si–O–Si), and coating. However, the intensity of the peak (401.8 eV) which
780 cm−1 (–Si(CH3)2) are attributed to the groups of the represented amine groups increased significantly owing to the
uncoated PDMS. The intensity of Group III decreased sig- abundant lysine residues existing in the poly(Ppep) coating
nificantly due to the presence of the poly(Ppep/Psar) (Fig. 2C-iii).61 The poly(Ppep/Psar) coating showed a roughness
coating covering the PDMS substrate.59 Grafting with the of 44.1 nm (Fig. 2D-i and ii). The incorporation of Psar
MePs also altered the wettability of the material surface reduced the concentration of the lysine residues on the
(Fig. 1B). MePpep comprises 50% hydrophilic lysine resi- surface, thus leading to a decreased intensity of the amine
dues; the contact angle of uncoated PDMS was reduced peak at 402.1 eV for the poly(Ppep/Psar) coating (Fig. 2D-iii).
from 102 ± 1° to 50 ± 1° after coating with the MePpep To further verify that the surface-attached pDA film can initiate
brush. The introduction of MePsar, which is composed of polymerization of the MePs, we fabricated a patterned poly
100% hydrophilic sarcosine residues, further decreased the (Ppep/Psar) coating by using the photomask technique; the
contact angle to 44 ± 4°. formed coating and its thickness could be easily observed with

Fig. 2 AFM images (i), topography (nm) ( profile corresponding to the white line in panel i) (ii), and high resolution N 1s XPS spectra (iii) of the
(A) uncoated, (B) pDA-coated, (C) poly(Ppep)-coated, and (D) poly(Ppep/Psar)-coated PDMS surfaces. (E) Schematic illustration of patterned
poly(Ppep/Psar) brush coating (i); AFM images (ii), topography (nm) ( profile corresponding to the white line in panel ii) (iii) of the patterned
poly(Ppep/Psar) brush coating.

This journal is © The Royal Society of Chemistry 2017 Polym. Chem., 2017, 8, 6386–6397 | 6391
View Article Online

Paper Polymer Chemistry

AFM (Fig. 2E). Fig. 2E-ii and iii show successful coating with One important aspect to consider is that although efficient
the patterned poly(Ppep/Psar) brushes; the average thickness contact-killing coatings have been developed, they often fail to
was 310 nm. These results indicated that MeP brushes were maintain their bactericidal activity as they can be easily
successfully grafted onto the surface of the pDA-coated masked by an absorbed conditioning film of organic com-
materials under UV irradiation. pounds such as proteins or remnants of dead cells. Thus, the
adherence of dead cells to the surface after the contact-killing
3.3. Antimicrobial activities of the coatings assay was observed using the LIVE/DEAD assay (Fig. 3B). The
After the successful preparation of the MeP polymer brush uncoated, poly(Ppep)-, and poly(Ppep/Psar)-coated PDMS sur-
faces were respectively brought into contact with E. coli for 1 h,
Published on 26 September 2017. Downloaded by Iowa State University on 1/24/2019 7:45:54 AM.

coatings, their contact-active bactericidal ability was evaluated


with four pathogenic microbes (S. aureus, E. coli, P. aeruginosa, then rinsed with PBS to remove the un-adhered bacteria or
and C. albicans). The antimicrobial activity of the two brush components, and stained with a mixture of SYTO-9 and propi-
coatings, poly(Ppep) and poly(Ppep/Psar), grafted on PDMS dium iodide (PI). SYTO-9 enters all bacterial cells and stains
was assessed by confrontation with four pathogens through a them green, whereas PI is only able to permeate the damaged
surface-contact method; the uncoated PDMS was used as a membrane and reduce the fluorescence of SYTO-9, staining
control.62 As shown in Fig. 3A, the poly(Ppep) coating, totally the cells red. As shown in Fig. 3B-i, an abundance of live E. coli
composed of cationic polypeptides, exhibited higher than cells (stained green) adhered to the uncoated surface. In con-
99.6% reduction against all four pathogens. The poly(Ppep/ trast, no live cells were present on the surface of the poly
Psar) coating also exhibited robust bactericidal ability against (Ppep) coating due to its strong bactericidal activity, whereas
the four pathogens, and exhibited 97.6%, 99.7%, 94.6%, and numerous dead cells (stained red) adhered to the surface
95.6% reduction of S. aureus, E. coli, P. aeruginosa, and (Fig. 3B-ii). This result indicated that the bacterial cell mem-
C. albicans, respectively. The bactericidal activity for E. coli was brane was damaged by our poly(Ppep) coating, thus allowed
the same for both coatings, whereas the poly(Ppep/Psar) the penetration of PI. Previously, Zhou et al. reported that
coating exhibited slightly lower reduction for the other 3 bac- similar synthesized polypeptides exhibit their antimicrobial
teria and fungus. activity by disrupting the bacterial membrane in solution.16
The cell membrane of bacteria contains abundant negative
charged lipids.63 Once the bacteria came into contact with
poly(Ppep) brushes, the positively charged lysine residues and
hydrophobic phenylalanine segments could interact with the
lipid bilayers strongly, thus disrupting the cell membrane and
leading them to death. The strong interaction between the
poly(Ppep) coating and the bacteria also results in the adher-
ence of the dead cells. The accumulation of dead cells on the
surface may block contact of the coating with other pathogens,
thus leading to the failure of the coating in long-term appli-
cations.37 To prevent the fouling of the surface, MePsar was
copolymerized with MePpep to form a dual-functional anti-
microbial and antifouling coating. Polysarcosine is a very
hydrophilic peptoid with excellent cell/protein-resistant pro-
perties.51 As shown in Fig. 3B-iii, no live or dead cells were
found on the surface of the poly(Ppep/Psar) coating. Therefore,
the antimicrobial cationic polypeptides killed the live cells and
antifouling polysarcosine prevented adhesion of the dead
cells.

3.4. Biofilm-resistant activity of the coatings


After adhesion on the surface of the biomaterials, bacteria can
form a biofilm. Bacteria in this form are embedded in a poly-
meric matrix, which makes them much harder to kill than
planktonic individuals. Often, surgical removal of the biofilm-
infected implants may even fail to solve this problem because
Fig. 3 Contact bactericidal activity of MeP coatings. (A) Antimicrobial the residual bacteria in the body cause recurring infections.64
activity of the poly(Ppep) and poly(Ppep/Psar) coatings against various Thus, the biofilm-resistance is a key factor for antimicrobial
pathogens (each data point represents the mean ± standard deviation
coatings to prevent biomaterial-associated infections. The
for three separately prepared samples). (B) LIVE/DEAD bacterial viability
assay of E. coli on the uncoated (i), poly(Ppep) (ii), and poly(Ppep/Psar)
uncoated and poly(MeP)-coated PDMS were subjected to
(iii) brush coatings after 1 h incubation at 37 °C and washing (scale bar = biofilm growth for 7 days and then analyzed by FE-SEM and
20 µm). LIVE/DEAD staining. As shown in Fig. 4A, numerous S. aureus

6392 | Polym. Chem., 2017, 8, 6386–6397 This journal is © The Royal Society of Chemistry 2017
View Article Online

Polymer Chemistry Paper


Published on 26 September 2017. Downloaded by Iowa State University on 1/24/2019 7:45:54 AM.

Fig. 4 Formation of the biofilm by S. aureus (left) and E. coli (right) on the uncoated, poly(Ppep) and poly(Ppep/Psar) coatings, observed with
(A) FE-SEM (scale bar = 10 μm), and (B) LIVE/DEAD bacterial viability assay (scale bar = 20 µm).

or E. coli cells were adhered to the surface of the pristine PDMS colored complex (Fig. 5A). The protein adsorption was quantitat-
(control) at day 1 and 7. The poly(Ppep) coating, which has ively determined by the absorbance at 560 nm (Fig. 5B). Taking
potent bactericidal activity, showed a reduction in the adhered the pristine PDMS as a control, the protein absorption was
bacteria compared with the uncoated PDMS. However, the poly greatly reduced for both the poly(Ppep) and poly(Ppep/Psar) coat-
(Ppep) coating was unable to prevent biofilm formation by ings. The adsorption of protein on the poly(Ppep) coating was
S. aureus and E. coli. More bacteria adhered to the surface of reduced by 54.7% (BSA) and 64.8% (lysozyme) respectively com-
poly(Ppep)-coated PDMS from day 1 to day 7. This may be due pared with that on the pristine PDMS. For the poly(Ppep/Psar)
to fouling of the surface by dead cells and other biomolecules coating, which has more hydrophilic antifouling Psar chains, the
such as proteins. Comparatively, the poly(Ppep/Psar) coating adsorption of the protein was further reduced by 83.4% for BSA
was capable of preventing biofilm formation by S. aureus and and 87.8% for lysozyme (Fig. 5C).
E. coli up to day 7; no bacteria adhered on the surface. This The platelet adhesion test was performed by placing the
result may be due to the existence of numerous poly(Psar) surface in contact with platelet-rich plasma (PRP) for 2 h, fol-
chains in the surface polymer brushes, which prevents fouling lowed by FE-SEM imaging. As shown in Fig. 5D-i, the platelet
of the surface. LIVE/DEAD bacterial staining was performed to adhesion was extensive, most of which were highly active and
further study the biofilm-resistance of the coatings. As shown in spread out with many typical pseudopods littering the surface
Fig. 4B, a bacterial biofilm was formed on the surface of of pristine PDMS. A few platelets adhered to the surface of
uncoated PDMS after 7 days of incubation. The bacteria adher- poly(Ppep), but with a small degree of deformation and a
ing to the poly(Ppep) coating were also alive at day 7, which may small amount of pseudopods (Fig. 5D-ii). No platelets adhered
be attributed to the remnants of adhered dead cells that to the surface of poly(Ppep/Psar) (Fig. 5D-iii). Therefore, the
blocked the coating, thus causing it to lose its antimicrobial presence of the poly(Psar) component greatly improved the
activity. No live or dead bacteria adhered to the poly(Ppep/Psar) antifouling activity of the polymer brush coating to enhance
coating, which indicated that this coating has excellent biofilm- the resistance to protein and platelet adhesion. In summary,
resistance. Therefore, although the poly(Ppep/Psar) coating has the poly(Ppep/Psar) brush coating itself exhibits antimicrobial
slightly poorer contact bactericidal potency than the poly(Ppep) and anti-biofilm ability as well as excellent resistance to
coating, the biofilm-resistance of the poly(Ppep/Psar) coating is protein/platelet adhesion. These advanced features of the poly
better than that of the poly(Ppep) coating. (Ppep/Psar) coating are expected to support long-lasting and
effective anti-infective effects in the body.
3.5. Antifouling activities
For implantable biomaterials, the fouling of their surfaces with 3.6. In vitro biocompatibility of surface coating of MePs
proteins and platelets may promote biofilm formation and may Prior to animal testing, the in vitro biocompatibility of the poly
also result in thrombosis and inflammation.65 Apart from the (Ppep/Psar) coating was evaluated by placing the coating in
antimicrobial activity, the antifouling activity is also a very impor- contact with L929 cells for up to 4 days. A tissue culture poly-
tant factor for the successful coating of implantable materials. styrene dish (TCPS) was used as a control. AlamarBlue and LIVE/
Thus, the protein-resistant ability of the two polymer brush coat- DEAD cell viability assays showed that the L929 cells proliferated
ings was evaluated by the Micro BCA protein assay (Fig. 5A–C). well in contact with the poly(Ppep/Psar) coating (Fig. 6). The
BSA and lysozyme were used as model proteins for adsorption AlamarBlue assay measures the mitochondrial activity and cell
on the surfaces. After adsorption for 24 h, the surface-attached viability and proliferation. Fig. 6A shows that the absorbance of
protein was chelated with the BCA reagent and formed a violet- AlamarBlue increased with the culture time, indicating that the

This journal is © The Royal Society of Chemistry 2017 Polym. Chem., 2017, 8, 6386–6397 | 6393
View Article Online

Paper Polymer Chemistry


Published on 26 September 2017. Downloaded by Iowa State University on 1/24/2019 7:45:54 AM.

Fig. 5 Antifouling ability of polymer brush coatings. (A) BCA reagent (i, blank control) chelated with protein (BSA) from the surface of uncoated
PDMS (ii), poly(Ppep) (iii) and poly(Ppep/Psar) (iv) coatings formed a violet-colored complex. (B) The protein (BSA and lysozyme) adsorption quanti-
tatively determined from the absorbance at 560 nm. (C) Reduction in the protein adsorption of poly(Ppep) and poly(Ppep/Psar) coatings versus that
of the pristine PDMS surface (error bars represent mean ± standard deviation of mean for n = 5, **p < 0.01). (D) Platelet adhesion on the uncoated
PDMS (i), poly(Ppep) (ii), and poly(Ppep/Psar) (iii) coatings observed with FE-SEM (scale bar = 20 µm (lower), 5 µm (upper)).

an animal subcutaneous infection model. S. aureus, one of the


key pathogenic microbes in infections associated with bio-
medical implants, was used as model bacteria to induce infec-
tion using a pre-seeding protocol.66 Incisions (1 cm) were care-
fully cut on two sides of the back of each of eight rats, and
uncoated and poly(Ppep/Psar)-coated implants (size: 0.2 ×
0.6 cm) inoculated with 10 µL (∼106 CFU) of fresh S. aureus
bacteria were placed into the incisions (Fig. 7A-i). After 5 days,
the conditions of the wounds were observed. Fig. 7A-ii shows a
serious infection in the incisions with the uncoated implants,
Fig. 6 In vitro biocompatibility study of L929 cells after 1 and 4 days of
and there were obvious pyogenic fluids. However, the incisions
culture with poly(Ppep/Psar) coating. Cell viability determined with (A) with the poly(Ppep/Psar)-coated implants looked different
AlamarBlue assay (error bars represent mean ± standard deviation of from that of the control groups, where no obvious sign of
mean for n = 5); (B) LIVE/DEAD assay (scale bar = 100 µm). infection was observed (Fig. 7A-iii). All implants were retrieved
and washed with PBS under ultrasonication to determine the
number of viable bacteria on the implants. As shown in
cells proliferated in the presence of the coating. In Fig. 6B, the Fig. 7A-iv and v, there were numerous S. aureus growing on the
cells were mostly stained green, indicating that they remained uncoated implants, but few bacteria were retrieved from the
alive in contact with the coating, suggesting that the poly(Ppep/ poly(Ppep/Psar)-coated implants. The bacterial count for the
Psar) coating is in vitro biocompatible. eight uncoated control implants ranged from 2.6 × 105 to 6.2 × 107
CFU per implant. In contrast, two out of the eight implants
3.7. In vivo anti-infective activity of surface coating of MePs with the poly(Ppep/Psar) coating had no visible bacteria, and
The anti-infective activity of this poly(Ppep/Psar) brush-coated the other six implants had fewer than 320 CFU per implant.
PDMS (a commonly used catheter material) was evaluated in The bacterial count for the poly(Ppep/Psar)-coated implants

6394 | Polym. Chem., 2017, 8, 6386–6397 This journal is © The Royal Society of Chemistry 2017
View Article Online

Polymer Chemistry Paper


Published on 26 September 2017. Downloaded by Iowa State University on 1/24/2019 7:45:54 AM.

Fig. 7 In vivo anti-infective activity of the poly(Ppep/Psar) coating. (A) Schematic of the rat subcutaneous infection model: photographs of surgical
wound with implants seeded with S. aureus at day 1 (i) and day 5 (ii and iii), bacterial cultures from uncoated (iv) and poly(Ppep/Psar)-coated
(v) implants. (B) Numbers of viable S. aureus cells recovered from the uncoated and poly(Ppep/Psar)-coated implants (error bars represent mean ±
standard deviation of mean for n = 8, **p < 0.01). (C) Histological study of the infected tissues around the uncoated (i and ii) and poly(Ppep/Psar)-
coated (iii and iv) implants.

was significantly lower than that of the control group ( p < malian cells in vitro, and retained its excellent biological pro-
0.01), with a log reduction of 5.2. Additionally, histological perties in a rat infection model by successfully preventing
analysis was used to evaluate the anti-infective activity of the implant-related infections caused by S. aureus. This polymer
poly(Ppep/Psar) coating. Inflammatory cells (blue color, as brush coating can be easily UV-grafted onto the surface of
indicated with the arrows) in the tissue around the implants medical materials and exhibits antimicrobial and antifouling
represent the severity of the inflammation and infection. As capabilities that can potentially combat biomaterial-associated
shown in Fig. 7C, more inflammatory cells were observed in infections.
the tissue around the uncoated implants (i an ii) than in the
case of the poly(Ppep/Psar)-coated implants (iii and iv). Thus,
the poly(Ppep/Psar) coating greatly reduced the bacterial count Conflicts of interest
and the inflammation in this rat subcutaneous infection
model. There are no conflicts to declare.

Acknowledgements
4. Conclusions
This work was funded and supported by the Natural Science
A dual-functional poly(Ppep/Psar) brush coating was developed Foundation of China (No. 51403173), the Natural Science Basic
as a viable strategy to deal with biomaterial-associated infec- Research Plan in Shaanxi Province of China (No. 2015JQ5139),
tions. The rationally synthesized MePs not only possess excel- and the Fundamental Research Funds for the Central
lent antimicrobial or antifouling groups, but also have metha- Universities of China. The authors thank Prof. Peng Yang
crylate groups capable of facile grafting onto the pDA-coated (Shaanxi Normal University, China) for their kind assistance
surface under UV irradiation. Two brush coatings, poly(Ppep) with AFM experiments.
and poly(Ppep/Psar), were successfully prepared as confirmed
by ATR-FTIR, water contact angle, AFM, and XPS characteriz-
ation. Comparatively, the poly(Ppep/Psar) coating showed References
excellent antimicrobial activity with an inhibition rate above
94.6% for all four clinically significant Gram-positive/-negative 1 E. D. Brown and G. D. Wright, Nature, 2016, 529, 336–343.
bacteria and fungi, and also prevented biofilm formation for 2 J. O’Neill, Tackling Drug-Resistant Infections Globally: final
up to 7 days. The poly(Ppep/Psar) coating greatly reduced non- report and recommendations, UK Department of Health,
specific protein adsorption and platelet adhesion in vitro. Review on Antimicrobial Resistance, London, 2016.
Additionally, this polymer brush coating is non-toxic to mam- 3 M. Zasloff, Nature, 2002, 415, 389–395.

This journal is © The Royal Society of Chemistry 2017 Polym. Chem., 2017, 8, 6386–6397 | 6395
View Article Online

Paper Polymer Chemistry

4 Z. Y. Ong, N. Wiradharma and Y. Y. Yang, Adv. Drug 26 Q. Yu, Z. Q. Wu and H. Chen, Acta Biomater., 2015, 16, 1–
Delivery Rev., 2014, 78, 28–45. 13.
5 P. Li, X. Li, R. Saravanan, C. M. Li and S. S. J. Leong, RSC 27 K. Yu, J. C. Y. Lo, M. Yan, X. Yang, D. E. Brooks,
Adv., 2012, 2, 4031–4044. R. E. W. Hancock, D. Lange and J. N. Kizhakkedathu,
6 X. B. Qi, C. C. Zhou, P. Li, W. X. Xu, Y. Cao, H. Ling, Biomaterials, 2017, 116, 69–81.
W. N. Chen, C. M. Li, R. Xu, M. Lamrani, Y. G. Mu, 28 C. C. Zhou, P. Li, X. B. Qi, A. R. M. Sharif, Y. F. Poon,
S. S. J. Leong, M. W. Chang and M. B. Chan-Park, Biochem. Y. Cao, M. W. Chang, S. S. J. Leong and M. B. Chan-Park,
Biophys. Res. Commun., 2010, 398, 594–600. Biomaterials, 2011, 32, 2704–2712.
7 R. H. Liu, X. Y. Chen, Z. Hayouka, S. Chakraborty, 29 R. T. C. Cleophas, M. Riool, H. C. Quarles van Ufford,
Published on 26 September 2017. Downloaded by Iowa State University on 1/24/2019 7:45:54 AM.

S. P. Falk, B. Weisblum, K. S. Masters and S. H. Gellman, S. A. J. Zaat, J. A. W. Kruijtzer and R. M. J. Liskamp, ACS
J. Am. Chem. Soc., 2013, 135, 5270–5273. Macro Lett., 2014, 3, 477–480.
8 R. H. Liu, X. Y. Chen, S. Chakraborty, J. J. Lemke, 30 G. Xu, D. Pranantyo, B. Zhang, L. Xu, K. G. Neoh and
Z. Hayouka, C. Chow, R. A. Welch, B. Weisblum, E. T. Kang, RSC Adv., 2016, 6, 14809–14818.
K. S. Masters and S. H. Gellman, J. Am. Chem. Soc., 2014, 31 G. Gao, D. Lange, K. Hilpert, J. Kindrachuk, Y. Zou,
136, 4410–4418. J. T. J. Cheng, M. Kazemzadeh-Narbat, K. Yu, R. Wang,
9 R. H. Liu, X. Y. Chen, S. P. Falk, B. P. Mowery, S. K. Straus, D. E. Brooks, B. H. Chew, R. E. W. Hancock
A. J. Karlsson, B. Weisblum, S. P. Palecek, K. S. Masters and and J. N. Kizhakkedathu, Biomaterials, 2011, 32, 3899–
S. H. Gellman, J. Am. Chem. Soc., 2014, 136, 4333–4342. 3909.
10 R. H. Liu, J. M. Suarez, B. Weisblum, S. H. Gellman and 32 C. Yang, X. Ding, R. J. Ono, H. Lee, L. Y. Hsu, Y. W. Tong,
S. M. McBride, J. Am. Chem. Soc., 2014, 136, 14498–14504. J. Hedrick and Y. Y. Yang, Adv. Mater., 2014, 26, 7346–7351.
11 R. H. Liu, X. Y. Chen, S. P. Falk, K. S. Masters, B. Weisblum 33 G. Pan, S. Sun, W. Zhang, R. Zhao, W. Cui, F. He, L. Huang,
and S. H. Gellman, J. Am. Chem. Soc., 2015, 137, 2183– S. H. Lee, K. J. Shea, Q. Shi and H. Yang, J. Am. Chem. Soc.,
2186. 2016, 138, 15078–15086.
12 A. Muñoz-Bonilla and M. Fernández-García, Prog. Polym. 34 H. Cheng, K. Yue, M. Kazemzadeh-Narbat, Y. Liu,
Sci., 2012, 37, 281–339. A. Khalilpour, B. Li, Y. S. Zhang, N. Annabi and
13 A. Muñoz-Bonilla and M. Fernández-García, Eur. Polym. J., A. Khademhosseini, ACS Appl. Mater. Interfaces, 2017, 9,
2015, 65, 46–62. 11428–11439.
14 T. J. Deming, Chem. Rev., 2016, 116, 786–808. 35 H. M. Yuan, B. R. Yu, L. H. Fan, M. Wang, Y. W. Zhu,
15 N. Gangloff, J. Ulbricht, T. Lorson, H. Schlaad and X. K. Ding and F. J. Xu, Polym. Chem., 2016, 7, 5709–5718.
R. Luxenhofer, Chem. Rev., 2016, 116, 1753–1802. 36 W. B. Sheng, B. Li, X. L. Wang, B. Dai, B. Yu, X. Jia and
16 C. C. Zhou, X. B. Qi, P. Li, W. N. Chen, L. Mouad, F. Zhou, Chem. Sci., 2015, 6, 2068–2073.
M. W. Chang, S. S. J. Leong and M. B. Chan-Park, 37 W. Hartleb, J. S. Saar, P. Zou and K. Lienkamp, Macromol.
Biomacromolecules, 2010, 11, 60–67. Chem. Phys., 2016, 217, 225–231.
17 M. H. Xiong, Z. Y. Han, Z. Y. Song, J. Yu, H. Z. Ying, 38 E. K. Riga, M. Vöhringer, V. T. Widyaya and K. Lienkamp,
L. C. Yin and J. J. Cheng, Angew. Chem., Int. Ed., 2017, 56, Macromol. Rapid Commun., 2017, DOI: 10.1002/marc.
10826–10829. 201700216.
18 H. Lu, J. Wang, Z. Y. Song, L. C. Yin, Y. F. Zhang, 39 T. Wei, Q. Yu, W. J. Zhan and H. Chen, Adv. Healthcare
H. Y. Tang, C. L. Tu, Y. Lin and J. J. Cheng, Chem. Mater., 2016, 5, 449–456.
Commun., 2014, 50, 139–155. 40 T. Wei, W. J. Zhan, L. M. Cao, C. M. Hu, Y. C. Qu, Q. Yu
19 D. Pranantyo, L. Q. Xu, Z. Hou, E. T. Kang and M. B. Chan- and H. Chen, ACS Appl. Mater. Interfaces, 2016, 8, 30048–
Park, Polym. Chem., 2017, 8, 3364–3373. 30057.
20 P. Li, C. C. Zhou, S. Rayatpisheh, K. Ye, Y. F. Poon, 41 T. Wei, W. J. Zhan, Q. Yu and H. Chen, ACS Appl. Mater.
P. T. Hammond, H. W. Duan and M. B. Chan-Park, Adv. Interfaces, 2017, 9, 25767–25774.
Mater., 2012, 24, 4130–4137. 42 S. S. Yuan, Y. G. Li, S. F. Luan, H. C. Shi, S. J. Yan and
21 C. Tian, J. Ling and Y. Q. Shen, Chin. J. Polym. Sci., 2015, J. H. Yin, J. Mater. Chem. B, 2016, 4, 1081–1089.
33, 1186–1195. 43 S. J. Yan, H. C. Shi, L. J. Song, X. H. Wang, L. Liu,
22 Y. Shen, Z. B. Li and H. A. Klok, Chin. J. Polym. Sci., 2015, S. F. Luan, Y. M. Yang and J. H. Yin, ACS Appl. Mater.
33, 931–946. Interfaces, 2016, 8, 24471–24481.
23 H. C. Flemming, J. Wingender, U. Szewzyk, P. Steinberg, 44 B. L. Wang, Q. W. Xu, Z. Ye, H. H. Liu, Q. K. Lin, K. H. Nan,
S. A. Rice and S. Kjelleberg, Nat. Rev. Microbiol., 2016, 14, Y. Z. Li, Y. Wang, L. Qi and H. Chen, ACS Appl. Mater.
563–575. Interfaces, 2016, 8, 27207–27217.
24 D. Davies, Nat. Rev. Drug Discovery, 2003, 2, 114–122. 45 B. L. Wang, Z. Ye, Q. W. Xu, H. H. Liu, Q. K. Lin, H. Chen
25 M. Riool, A. de Breij, L. de Boer, P. H. S. Kwakman, and K. Nan, Biomater. Sci., 2016, 4, 1731–1741.
R. A. Cordfunke, O. Cohen, N. Malanovic, N. Emanuel, 46 Z. L. Zhi, Y. J. Su, Y. W. Xi, L. Tian, M. Xu, Q. Wang,
K. Lohner, J. W. Drijfhout, P. H. Nibbering and S. A. J. Zaat, S. Padidan, P. Li and W. Huang, ACS Appl. Mater. Interfaces,
Adv. Funct. Mater., 2017, 27, 1606623. 2017, 9, 10383–10397.

6396 | Polym. Chem., 2017, 8, 6386–6397 This journal is © The Royal Society of Chemistry 2017
View Article Online

Polymer Chemistry Paper

47 Y. J. Su, Z. L. Zhi, Q. Gao, M. H. Xie, M. Yu, B. Lei, P. Li and 59 B. Dhananjay and K. M. Chantal, Microelectron. Eng., 2006,
P. X. Ma, Adv. Healthcare Mater., 2017, 6, 1601173. 83, 1277–1279.
48 Q. Gao, M. Yu, Y. J. Su, M. H. Xie, X. Zhao, P. Li and 60 F. Bernsmann, A. Ponche, C. Ringwald, J. Hemmerle,
P. X. Ma, Acta Biomater., 2017, 51, 112–124. J. Raya, B. Bechinger, J. C. Voegel, P. Schaaf and V. Ball,
49 R. H. Wang, T. Hughes, S. Beck, S. Vakil, S. Li, P. Pantano J. Phys. Chem. C, 2009, 113, 8234–8242.
and R. K. Draper, Nanotoxicology, 2013, 7, 1272–1281. 61 R. T. C. Cleophas, J. Sjollema, H. J. Busscher,
50 V. Gaberc-Porekar, I. Zore, B. Podobnik and V. Menart, J. A. W. Kruijtzer and R. M. J. Liskam, Biomacromolecules,
Curr. Opin. Drug Discovery Dev., 2008, 11, 242–250. 2014, 15, 3390–3395.
51 K. H. A. Lau, C. L. Ren, T. S. Sileika, S. H. Park, I. Szleifer 62 P. Li, Y. F. Poon, W. F. Li, H. Y. Zhu, S. H. Yeap, Y. Cao,
Published on 26 September 2017. Downloaded by Iowa State University on 1/24/2019 7:45:54 AM.

and P. B. Messersmith, Langmuir, 2012, 28, 16099–16107. X. B. Qi, C. C. Zhou, M. Lamrani, R. W. Beuerman,
52 S. T. Xuan, S. Gupta, X. Li, M. Bleuel, G. J. Schneider and E. T. Kang, Y. G. Mu, C. M. Li, M. W. Chang, S. S. J. Leong
D. H. Zhang, Biomacromolecules, 2017, 18, 951–964. and M. B. Chan-Park, Nat. Mater., 2011, 10, 149–156.
53 A. Birke, D. Huesmann, A. Kelsch, M. Weilbacher, J. Xie, 63 G. J. Gabriel, A. Som, A. E. Madkour, T. Eren and G. N. Tew,
M. Bros, T. Bopp, C. Becker, K. Landfester and M. Barz, Mater. Sci. Eng., R, 2007, 57, 28–64.
Biomacromolecules, 2014, 15, 548–557. 64 G. G. Anderson, J. J. Palermo, J. D. Schilling, R. Roth,
54 Z. X. Voo, M. Khan, K. Narayanan, D. Seah, J. L. Hedrick J. Heuser and S. J. Hultgren, Science, 2003, 301, 105–107.
and Y. Y. Yang, Macromolecules, 2015, 48, 1055–1064. 65 D. C. Leslie, A. Waterhouse, J. B. Berthet, T. M. Valentin,
55 J. Gu, Y. J. Su, P. Liu, P. Li and P. Yang, ACS Appl. Mater. A. L. Watters, A. Jain, P. Kim, B. D. Hatton, A. Nedder,
Interfaces, 2017, 9, 198–210. K. Donovan, E. H. Super, C. Howell, C. P. Johnson, T. L. Vu,
56 Y. J. Su, L. Tian, M. Yu, Q. Gao, D. Wang, Y. W. Xi, P. Yang, D. E. Bolgen, S. Rifai, A. R. Hansen, M. Aizenberg,
B. Lei, P. X. Ma and P. Li, Polym. Chem., 2017, 8, 3788–3800. M. Super, J. Aizenberg and D. E. Ingber, Nat. Biotechnol.,
57 C. D. Vacogne and H. Schlaad, Chem. Commun., 2015, 51, 2014, 32, 1134–1140.
15645–15648. 66 M. Riool, L. de Boer, V. Jaspers, C. M. van der Loos,
58 S. H. Wibowo, A. Sulistio, E. H. H. Wong, A. Blencowe and W. J. B. van Wamel, G. Wu, P. H. S. Kwakman and
G. G. Qiao, Adv. Funct. Mater., 2015, 25, 3147–3156. S. A. J. Zaat, Acta Biomater., 2014, 10, 5202–5212.

This journal is © The Royal Society of Chemistry 2017 Polym. Chem., 2017, 8, 6386–6397 | 6397

You might also like