You are on page 1of 32

Subscriber access provided by READING UNIV

Article
Capsaicin-Inspired Thiol-Ene Terpolymer
Networks Designed for Antibiofouling Coatings
Haiye Wang, Joshua Jasensky, Nathan Ulrich, Junjie Chen, Hao Huang, Zhan Chen, and Chunju He
Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.7b03098 • Publication Date (Web): 03 Nov 2017
Downloaded from http://pubs.acs.org on November 6, 2017

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 free 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 accessible to all
readers and 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.

Langmuir 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 Langmuir

1
2
3
4 Capsaicin-Inspired Thiol−Ene Terpolymer Networks Designed
5
6 for Antibiofouling Coatings
7
8
9 Haiye Wang,1,2 Joshua Jasensky,2 Nathan W. Ulrich, 2 Junjie Cheng, 2 Hao Huang, 2 Zhan Chen, 2 Chunju He1
10
11 1
College of Materials Science and Engineering, Donghua University, Shanghai, 201620, P. R. China
12
13 2
14 Department of Chemistry, University of Michigan, Ann Arbor, Michigan 48109, United States
15
16 ABSTRACT: Novel photocurable ternary polymer networks were prepared by incorporating
17
18
19 N-(4-hydroxy-3-methoxybenzyl)-acrylamide (HMBA) into a cross-linked thiol−ene network
20
21 based on poly(ethylene glycol) diacrylate (PEGDA) and (mercaptopropyl) methylsiloxane
22
23
24 homopolymers (MSHP). The ternary network materials displayed bactericidal activity against
25
26 Escherichia coli, Staphylococcus aureus, and reduced the attachment of marine organism
27
28
29 Phaeodactylum tricornutum. Extensive soaking of the polymer networks in aqueous solution
30
31 indicated that no active antibacterial component leached out from the materials, and thus the
32
33
34 ternary thiol-ene coating killed the bacteria by surface contact. The surface structures of the
35
36 polymer networks with varied content ratios were studied by sum frequency generation (SFG)
37
38
39 vibrational spectroscopy. The results demonstrated that the PDMS Si-CH3 groups and
40
41
mimic-capsaicine groups are predominantly present at the polymer-air interface of the
42
43
44 coatings. Surface reorganization was apparent after polymers were placed in contact with
45
46
D2O: the hydrophobic PDMS Si-CH3 groups left the surface and returned to the bulk of the
47
48
49 polymer networks, while the hydrophilic PEG chains cover the polymer surfaces in D2O. The
50
51
capasaicine methoxy groups are able to segregate to the surface in an aqueous environment,
52
53
54 depending upon the ratio of HMBA/PEGDA. SFG measurements in situ showed that the
55
56
57
antibacterial HMBA chains, rather than the “non-fouling” PEG, played a dominant role in
58
59 1
60
ACS Paragon Plus Environment
Langmuir Page 2 of 31

1
2
3
4 mediating the antibiofouling performance in this particular polymer system.
5
6 KEYWORDS: thiol-ene network; capsaicin-mimic, antibiofouling; surface reorganization,
7
8
9 SFG vibrational spectroscopy
10
11 Introduction
12
13 Over the past decades, the development of antifouling coatings for marine ships has been of
14
15 great interest due to adhesion of marine organisms onto ship surfaces causing increased
16
17
18
hydrodynamic drag, greater fuel consumption, as well as an increase in ship maintenance and
19
20 environment compliance costs.1-5 Biocide-containing antifouling coatings have been
21
22 extensively used in order to resist the settlement of marine organisms.6-8 The most widely
23
24 used antifouling compounds are mainly based on heavy metals such as copper, lead, tin, and
25
26 arsenic.9-11 Even though they have been shown to be effective antifouling materials, their
27
28
29 toxicity has limited their use in recent years.12-14 The restriction on the use of such toxic
30
31 compounds has promoted the research on natural and eco-friendly synthetic biocides.6, 15-16
32
33 Since adhesion of marine organisms to a surface is known to be highly related to the
34
35 surface chemical composition or structure,17-18 great effort has been made to design surfaces
36
37
38
with non-fouling performance (with or without biocides), with a focus on varying the surface
39
40 chemical composition.19-22 In particular, a marine biofouling process could start from the
41
42 adsorption of protein-containing glues secreted by marine organisms to a surface. Thus,
43
44 protein-resistant materials have been employed to prevent adhesion of marine organisms.
45
46 Poly(ethylene glycol) (PEG) and zwitterionic polymers are the most popular nonfouling
47
48
49 materials used for protein resistance.23-28 The strong surface hydration of PEG is known to
50
51 play an important role in its mechanism of protein resistance. Such a strong hydration was
52
53 formed at the PEG surface through extensive hydrogen bonding between water molecules and
54
55 the ether oxygen atoms.29-30 The tightly bound hydration layer acts as a physical barrier,
56
57
58
where the removal of water by biomolecules poses as a significant energy cost. This leads to
59 2
60
ACS Paragon Plus Environment
Page 3 of 31 Langmuir

1
2
3 the excellent non-fouling property of the material.31
4
5 However, incorporation of the “non-fouling” PEG chains to the surface can only
6
7
effectively resist protein adsorption, which may not be able to reduce bacterial
8
9
10 colonization.32-33 Indeed, bacterial colonization on a surface is another crucial factor to induce
11
12 the marine biofouling process.1 The fouling process is highly dynamic and the adhered
13
14 bacteria incubate rapidly to form a biofilm, which contribute to the microfouling
15
16 community.34 Thus, effective strategies to prevent biofouling may comprise (1) preparation of
17
18
19 “non-fouling” surfaces that can resist initial protein adsorption, (2) fabrication of antibacterial
20
21 surfaces that can kill bacteria when they are deposited. We believe that an effective way to
22
23 design an antifouling coating is to combine both protein-resistant and antibacterial properties
24
25 into the surface.35 In order to ensure a long-term antibacterial effectiveness for the coating,
26
27
the leaching of biocides from the coating should be avoided as much as possible. To do so it
28
29
30 is necessary to introduce the polymer with antimicrobial activity by binding the antibacterial
31
32 agent covalently to the polymer.33, 36-37
33
34 Capsaicin, the main active component of chili peppers, has strong deterrence toward
35
36 microorganisms. Previous research has shown that its derivatives and analogs could possess
37
38
39 much more attractive properties than the current toxic antifoulants,32 and have consequently
40
41 been used as an additive for anti-fouling coating on ships and water craft in ocean
42
43 transportation.38-40 As one of the capsaicin-mimic materials, a functional monomer
44
45 N-(4-hydroxy-3-methoxybenzyl)-acrylamide (HMBA) was derived from acrylamide by
46
47
reaction with guaiacol through the Friedel-Crafts reaction. Also, the carbon–carbon double
48
49
50 bond can be used for polymerization with other monomers to produce polymer network
51
52 antifouling coatings.41-43
53
54 This study aimed to develop a new concept of using hot and spicy surfaces for
55
56 antibiofouling. In the long run, the materials need to be optimized with better performance. In
57
58
59 3
60
ACS Paragon Plus Environment
Langmuir Page 4 of 31

1
2
3 this work, a simple and versatile method to prepare bio-inspired ternary polymer networks
4
5 was developed. This method was based on the photo-curing of polysiloxane macromonomer
6
7
(PSMM), capsaicin-mimic HMBA and poly(ethylene glycol) diacrylate (PEGDA). The above
8
9
10 polymer network compositions were chosen due to their photo-curing behavior and
11
12 extensively characterized physicochemical properties. The protein resistance and antibacterial
13
14 property of the prepared polymer network films were investigated and found to be dependent
15
16 on the ratios of the different components in the films. Also, a study on the antibiofouling
17
18
19 activity of the prepared polymer network films against the diatom Phaeodactylum
20
21 tricornutum, taken as a commonly accepted model organism, was performed. In addition, the
22
23 surface structures of the polymer network films in different chemical environments were
24
25 studied by sum frequency generation (SFG) vibrational spectroscopy in situ; the effect of the
26
27
polymer bulk composition on the surface structure of the polymer network was elucidated.
28
29
30 The surface structural results can be well correlated to the antibiofouling performance of
31
32 various polymer samples, providing an in-depth understanding of structural – function
33
34 relationships of the prepared materials as well as the antibiofouling mechanisms of such
35
36 materials.
37
38
39 Experimental Section
40
41
42
Materials. All chemical reagents and solvents were obtained at the highest purity available
43
44 from Aldrich Chemical Co. and used without further purification unless otherwise specified.
45
46
47
2,2-dimethoxy-2-phenylacetophenone (DMPA), dimethyl formamide (DMF), ethanol,
48
49 poly(ethylene glycol) diacrylate, N-hydroxymethylacrylamide, and guaiacol
50
51
52 (2-methoxyphenol) were purchased from Aladdin. (Mercaptopropyl) methylsiloxane
53
54 homopolymer (MSHP) (Mn=4000-7000, average numbers of thiols were 40) was obtained
55
56
57 from Gelest and used as received. Bovine serum albumin (BSA) and lysozyme were
58
59 4
60
ACS Paragon Plus Environment
Page 5 of 31 Langmuir

1
2
3
4 purchased from Sinopharm Chemical Reagent Co. Phaeodactylum tricornutum was supplied
5
6 by the Ocean University of China.
7
8
9 Synthesis of the HMBA by Friedel-Craft Reaction. The synthesis procedure of HMBA was
10
11 similar to that presented in the previous report with some modifications.33 10 g of
12
13
14 N-hydroxymethylacrylamide and 12.5 g of guaiacol were dissolved in 50 ml of absolute
15
16 ethanol. The mixture was cooled with an ice-bath, and 5 mL of H2SO4 was added dropwise
17
18
19 under stirring over a period of 1 h. The reaction mixture was kept stirring for 7 days at room
20
21 temperature. Then the product was washed with water, and concentrated under vacuum to
22
23
24 give a yellow liquid, which was poured into diethyl ether and white precipitate of HMBA was
25
26 formed. The crude product was recrystallized from ethanol. The results of the 1H NMR and
27
28
29 FT-IR measurements of the synthesized HMBA are presented in the supplementary material.
30
31 Synthesis of Ternary Thiol-Ene Polymer Networks. A series of ternary thiol-ene polymer
32
33
34 networks containing different molar ratios of PEGDA and HMBA were synthesized from
35
36 resins maintaining a 1:1 thiol/total alkene functional group stoichiometry (The bulk
37
38
39 concentration of MSHP in the coating was 37 wt%). HMBA and poly(ethylene glycol)
40
41
diacrylate were successively added to the (Mercaptopropyl) methylsiloxane homopolymer to
42
43
44 prepare the three-component mixtures. For all cases, 1 wt % photoinitiator DMPA was first
45
46
dissolved in MSHP, and subsequently HMBA and PEGDA were added with the proper
47
48
49 amounts (see Table 1). Once completely dissolved, the mixture was transferred onto a glass
50
51
slide and then cured by exposure to UV radiation under N2 purge.
52
53
54
55
56
57
58
59 5
60
ACS Paragon Plus Environment
Langmuir Page 6 of 31

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16 Scheme 1. Schematic showing the photo-cured ternary thiol−ene/capsaicin polymer
17
18
19 networks.
20
21 Table 1 Reaction conditions and the obtained ternary thiol-ene polymer networks. Each
22
23
24 sample contains a 1:1 molar ratio of siloxane vs. the sum of HBMA and PEGDA, but with a
25
26 different HBMA/PEGDA molar ratio.
27
28
Molar ratio of Coating composition
29
Light intensity
30 Sample HMBA HMBA (wt %) Curing time color
31 and distance
32 /PEGDA
33
34 MHP (1:9) 1:9 5 80%, 12 cm 16 min colorless
35 MHP (3:7) 3:7 15 80%, 12 cm 14 min colorless
36 MHP (5:5) 5:5 26.9 80%, 12 cm 12 min colorless
37
38 MHP (7:3) 7:3 39.3 80%, 12 cm 10 min colorless
39 MHP (9:1) 9:1 53 80%, 12 cm 8 min yellow
40
41 Characterization
42
43
Attenuated Total Reflectance FTIR (ATR-FTIR). A Nicolet 7600 FTIR spectrometer was
44
45
46 used to collect ATR-FTIR spectra in this research. The spectra were collected in the
47
48
wavenumber range between 400 and 5000 cm-1 with a resolution of 4 cm-1.
49
50
51 Contact Angle Measurements. A contact angle goniometer (Dataphysics OCA40, Germany)
52
53
54
was used to measure static water contact angles on sample surfaces using the sessile drop
55
56 method. The measurements were performed at room temperature. For each measurement, a
57
58
59 6
60
ACS Paragon Plus Environment
Page 7 of 31 Langmuir

1
2
3
4 syringe was used to drop 3 µl of deionized water onto the sample surface. The contact angle
5
6 data presented in this research for each sample were the average of five contact angles
7
8
9 measured on different regions of the sample.
10
11 SFG Vibrational Spectroscopy. As a second order nonlinear optical spectroscopic method,
12
13
14 SFG has a submonolayer surface specificity, which comes from its selection rule.44-54
15
16 Previous papers have extensively reported the details of the SFG instrumentation, date
17
18
19 collection geometry, and SFG experimental procedures,55-57 which will not be repeated here.
20
21 The polymer films used in this SFG study were prepared by spin coating polymer solutions
22
23
24 onto fused silica substrates. A spin coater from Specialty Coating Systems (Indianapolis, IN)
25
26 was used to coat thin films of MHP-3/7, MHP-5/5, or MHP-7/3 with their 2 wt % solutions in
27
28
29 DMF. The samples were then cured by exposure to UV radiation under N2 purge. The
30
31 thickness of the resulting films was ~55 nm. In SFG experiments, the two input beams, the
32
33
34 visible (30 µJ pulse energy) and the frequency tunable infrared (IR, 100 µJ pulse energy)
35
36 beams, overlapped spatially and temporally on the sample surface in air or in water after
37
38
39 penetrating through a right angle prism. The incident angles of the beams were 60° (visible)
40
41
and 54° (IR) with respect to the surface normal, respectively. A monochromator along with a
42
43
44 photomultiplier tube was used to collect the reflected SFG signal from the sample surface. In
45
46
this research, SFG ssp (s-polarized sum frequency output, s-polarized visible input, and
47
48
49 p-polarized IR input) signals were collected. Such signals were normalized by using the input
50
51
visible and IR beam intensities.
52
53
54 Optical Properties. A Shimadzu UV-1800 UV/Vis spectrophotometer was used for the optical
55
56
57
absorption measurements of 400-700 nm. The quantitative transmittance reported below was
58
59 7
60
ACS Paragon Plus Environment
Langmuir Page 8 of 31

1
2
3
4 measured at 550 nm.
5
6 Fluorescence Microscopy. BSA with FITC-label was prepared according to a published
7
8
9 procedure.58 Each polymer coating sample with a size of 1×1 cm2 was incubated with 4 ml
10
11 FITC labeled BSA solution in a vial. After samples were left to shake for 12 at 4 °C in the
12
13
14 dark, the polymer sample was thoroughly washed using PBS so that any loosely adsorbed
15
16 proteins on the surface could be removed. An Olympus BX-51 inverted microscope equipped
17
18
19 with a 20× objective, a mercury arc lamp, and an Olympus DP52 digital camera was used for
20
21 the fluorescence imaging experiments. The same experimental conditions such as collection
22
23
24 time, CCD gain, and optics set (U-MWB2 filter cube with excitation of 450-480 nm and
25
26 emission of 510-520 nm) were used in all the experiments. Therefore it is reasonable to
27
28
29 compare images from frame-to-frame.
30
31 Protein Adsorption Tests. Protein adsorption was determined using an ELISA method
32
33
34 according to a previous procedure59 (see Supporting Information).
35
36
Biofouling Assays
37
38
39 Phaeodactylum tricornutum Settlement and Adhesion Assays. Phaeodactylum tricornutum
40
41
was used to test the antifouling behaviors of various materials. Supplied by the Ocean
42
43
44 University of China, Phaeodactylum tricornutum has been grown in f/2 medium for several
45
46
days at 15°C. In this study, the diatoms were cultivated in a climate chamber at 20°C with a
47
48
49 light intensity of 2000 lux. Then diatom suspension (5 ml, 104 cells/ml) was added to the
50
51
surface of each polymer sample (coated on glass slide) in an individual dish. After 7 days, all
52
53
54 the samples were gently dipped into a beaker with 3% salinity, 0.22 µm FSW to wash off any
55
56
57
unattached cells. This washing step was repeated three times. Then a hemocytometer was
58
59 8
60
ACS Paragon Plus Environment
Page 9 of 31 Langmuir

1
2
3
4 used to determine the density of cells on each polymer sample surface and digital images of
5
6 diatom growth were recorded.
7
8
9 Antimicrobial Test. Antimicrobial tests were performed according to standard antimicrobial
10
11 test protocols described in detail elsewhere.35 E. coli and S. aureus were chosen as
12
13
14 representative Gram-negative and Gram-positive bacteria, respectively.
15
16 Antimicrobial Activity of Polymer Leached from Ternary Polymer Networks.
17
18
19 Polymer-coated prisms were incubated with 5 mL of PBS (pH 7.4, without calcium and
20
21 magnesium) in a sterile 15 mL Falcon tube at 37°C with orbital rotation at 200 rpm for 1 hour.
22
23
24 An untreated prism was used as a control. The PBS solution (4.5 mL) was removed from
25
26 each Falcon tube and added to another tube containing E. coli fresh culture (0.5mL, OD =
27
28
29 ~0.05), then incubated at 37 °C for 2 hour. The density of cells was determined using a cell
30
31 density meter.
32
33
34 Some other test methods of physical and chemical properties, such as hardness, adhesion
35
36 force, chemical resistance and swelling, are presented in supporting information.
37
38 Results and Discussion
39
40
Photo-Curing of Ternary Thiol−Ene/Capsaicin Polymer Networks. In order to investigate
41
42
43 the compositional effect on the polymer surface structure, as well as the physicochemical and
44
45 biological properties of the resulting ternary networks, thiol-ene polymer networks containing
46
47 different ratios of HMBA and PEGDA were synthesized. HMBA and di-functional
48
49 poly(ethylene glycol) diacrylate (PEGDA) were used in the polymer network sample along
50
51
52
with a linear siloxane polymer (mercaptopropyl) methylsiloxane homopolymers (MSHP) to
53
54 study the structure-property relationships of the thiol-ene networks. Reaction conditions on
55
56 the ternary thiol-ene polymer networks and different samples prepared are shown in Table 1.
57
58
59 9
60
ACS Paragon Plus Environment
Langmuir Page 10 of 31

1
2
3 The coating MHP(9:1) containing more than 53 wt% HMBA appeared rough and light yellow.
4
5 Although all coatings were clear and smooth, they were difficult to cure and could not
6
7
solidify completely when the HMBA content was less than 5 %. Perhaps this is due to the fact
8
9
10 that the PEGDA content is too high, which prevents the complete curing. For these reasons,
11
12 only optically transparent and flexible coatings having HMBA between 5-53 wt% were
13
14 studied in detail.
15
16 Stability of the Ternary Thiol-Ene Polymer Networks.
17
18
19 Although there exist many classes of marine antifouling coatings ranging from resins to
20
21 hydrogels, the ideal characteristics of highly active and long lasting materials include low
22
23 water-uptake and high mechanical stability.60 After immersion in water for 48 hrs, thiol-ene
24
25 cross-linked networks with 15-39.3 wt% HMBA were found to maintain visual integrity on
26
27
the substrates, exhibiting excellent stability in water, as shown in Fig.1(a). We also studied
28
29
30 the stability for a week, and no change was observed. All the coatings were still transparent
31
32 (88.4%~89.3% transmittance measured at 550 nm), only a slight decrease was observed upon
33
34 the increased addition of HMBA (Fig. 1(b)).
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51 Fig. 1 (a) Stability of the ternary thiol-ene polymer network coatings against water immersion.
52
53 (b) Transmission spectra of the control and corresponding coatings. (The samples are
54
55
56 transparent and colorless. The papers underneath the samples are green).
57
58 Physicochemical Characters of the Ternary Thiol-Ene Polymer Networks. Cross-hatch
59 10
60
ACS Paragon Plus Environment
Page 11 of 31 Langmuir

1
2
3 adhesion, hardness, seawater resistance, and chemical resistance were determined to evaluate
4
5 the physicochemical performance of these network samples. As shown in Table 2, all three
6
7
coatings yielded similar results for the seawater and 20% HCl resistance, where the coatings
8
9
10 with different compositions kept their intact morphology. Adhesion tests were used to
11
12 measure the adhesive properties of the ternary polymer networks to glass. With an increased
13
14 HMBA content, the adhesive property of the MHP (7:3) coating improved to 0B, where
15
16 absolutely no film was removed from the surface of the substrate. This originates from the
17
18
19 special structure of HMBA, which was very similar to a widely studied strong bioadhesive
20
21 3,4-dihydroxyphenylalanine (DOPA). In hardness tests, the rigidity of the ternary polymer
22
23 networks also increased, which was ascribed to the stiffness of the benzene ring from the
24
25 HMBA. Despite the slight deficiency in alkali and acetone resistance, other measured
26
27
properties all demonstrated outstanding physicochemical performance for these samples.
28
29
30 Table 2 Results of thiol-ene polymer network coatings from the measurements on cross-hatch
31
32 adhesion, hardness, seawater resistance, chemical resistance, water-uptake, and transmittance
33
34 at 550 nm.
35
36
20%
37 Seawater 20% HCl Aceton Water Transmit
38 cross-hatch Hard NaOH
Sample resistance resistance resistan uptake tance
39 adhesion ness resistance
40 (24h) (24h) ce (%) (%)
(24h)
41
MHP(3:7) 3B HB E R E R 1.12 89.35
42
43 MHP(5:5) 1B HB E C E R 0.76 88.74
44 MHP(7:3) 0B 1H E F E C 0.28 88.49
45 Remark: E, F, C, R, P, corresponding to the intact, bubble, wrinkle, fracture, peeling
46
47 phenomenon, respectively.
48
49 Static Water Contact Angle. It has been shown that wettability is a critical parameter for an
50
51 amphiphilic system to resist protein adsorption.61 The amphiphilic balance of the ternary
52
53 thiol-ene polymer networks can be controlled by altering the ratio of the HMBA and PEGDA
54
55 monomers, which should lead to the variation of surface wettability. As shown in Fig. 2, the
56
57
static water contact angles depended on the ratio of the HMBA and PEGDA monomers in the
58
59 11
60
ACS Paragon Plus Environment
Langmuir Page 12 of 31

1
2
3 ternary network (e.g. the contact angle increased from 85.2° for MHP(3:7) to 93.6° for
4
5 MHP(5:5) and further to 106.4° for MHP(7:3)). This indicated that the surface became more
6
7
hydrophobic with an increasing content of HMBA, leading to an enhanced bacteria killing
8
9
10 capability by surface contact.37,38 In contrast, the PEGDA segment played a significant role in
11
12 increasing the surface hydrophilicity, which was favorable to reduce the nonspecific adhesion
13
14 of proteins.
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32 Fig. 2 Static water contact angle of thiol-ene polymer network coatings.
33
34
35
36
37 Nonspecific Protein Adsorption Resistance
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 12
60
ACS Paragon Plus Environment
Page 13 of 31 Langmuir

1
2
3 Fig. 3 (a) Fluorescence microscopic images showing BSA and lysozyme adsorption results
4
5 on the thiol-ene polymer networks. (b) The quantitative results of protein adsorption on
6
7
polymer networks.
8
9
10 The adsorption of protein on a material surface was used as one of the most important criteria
11
12 to evaluate the antifouling performance of the material, e.g., for using as a marine coating.62
13
14 In this study, both BSA and lysozyme adsorptions onto polymer networks were tested (Fig. 3).
15
16 The Polydimethylsiloxane elastomer (PDMSe: Silastic T2) coating (as a control surface)
17
18
19 exhibited a high protein adsorption on the surface. This is reasonable because it is
20
21 hydrophobic which likes to adsorb more protein molecules. The protein adsorption amounts
22
23 on amphiphilic polymer networks were measured to be less, because their surface PEG
24
25 functionality could resist BSA and lysozyme adsorption, especially when compared to the
26
27
control PDMSe sample. It was also observed that with an increase in PEG content within the
28
29
30 coating, the protein adsorption amount was reduced. For the amphiphilic coating with a high
31
32 PEG content, barely any BSA or lysozyme protein was found to be adsorbed on the surface.
33
34 Interestingly, on each coating surface, the adsorption amounts for BSA and lysozyme were
35
36 slightly different. This likely was due to that fact that BSA and lysozyme had different net
37
38
39 charges in a PBS solution (BSA was negative while lysozyme was positive). The electrostatic
40
41 interaction effect between the protein and a coating surface would also contribute to protein
42
43 adsorption. That is why all the coatings always have a slightly higher lysozyeme adsorption.
44
45 Nevertheless, the improved protein adsorption resisting property will have benefits to
46
47
enhance the antibiofouling performance of a material.
48
49
50
51
52
53
54
55
56
57
58
59 13
60
ACS Paragon Plus Environment
Langmuir Page 14 of 31

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
Fig. 4 Results of bactericidal tests against E. coli and S. aureus.
19
20
21
22
23 Table 3 Antibacterial efficiency for the ternary thiol-ene polymer networks.
24
E. coli S. aureus
25
26 Sample No. of bacteria average Eb No. of bacteria
average Eb (%)a
27 colonies (%)a colonies
28 Blank 97 0 124 0
29
30 MHP-3/7 49 49.4 40 67.7
31 MHP-7/3 6 93.9 5 96.0
32
33 Antibacterial Activity of MHP Coating. Escherichia coli and Staphylococcus aureus were
34
35
36
chosen as representative gram-negative and gram-positive bacteria to evaluate the
37
38 antimicrobial activities of MHP coatings. With the pristine glass serving as a control, the
39
40
41
adherent bacteria colonies within 24 h incubation were shown in Fig. 4. There were
42
43 considerable reductions in E. coli and S. aureus colonies after 24 h incubation with the MHP
44
45
46 coatings, indicating that the presence of HMBA in the coating significantly inhibited bacteria
47
48 growth. With an increase of the HMBA content in MHP coating, the antibacterial effect
49
50
51 against both S. aureus and E. coli became stronger. The antibacterial efficiencies (Eb) of the
52
53 different coatings are presented in Table 3. For the MHP coatings, the average antibacterial
54
55
56 efficiency of the MHP-7/3 sample surface for E. coli and S. aureus reached 93.9% and 96.0%,
57
58 respectively, which had a greater killing effect than the MHP-3/7 coating. As the
59 14
60
ACS Paragon Plus Environment
Page 15 of 31 Langmuir

1
2
3
4 Gram-negative E. coli possesses an additional lipopolysaccharide-containing outermost
5
6 membrane as compared to Gram-positive S. aureus, the polymer coatings had a slightly lower
7
8
9 killing efficiency against E. coli as compared to S. aureus.63 The high bacteria killing
10
11 efficiencies further confirmed that the MHP cross-linked coating should be especially useful
12
13
14 for marine antibiofouling.
15
16 To determine whether any bacteria were killed by the antimicrobial contents leached out of
17
18
19 the coatings in solution, each of the prisms with coated polymer films was placed in PBS
20
21 buffer without bacteria and then the buffer solution was added to an E. coli suspension. If the
22
23
24 antimicrobial polymer components were leached out from a coating film into the buffer
25
26 solution, then the biocidal activity of the buffer could be detected.64 As shown in Table 4, the
27
28
29 PBS buffer solution incubated with the coatings of MHP-3/7, MHP-5/5, MHP-7/3 exhibited
30
31 no antimicrobial activity. This suggests that no antimicrobial polymer contents leached to the
32
33
34 solution. Therefore the biocidal activity of MHP coatings must come from the contact
35
36 between the bacteria and the coating surface.
37
38
39 Table 4 Antibacterial effects of possible leachables from the ternary thiol-ene polymer
40
41
networks.
42
43 Sample solution OD of the E. coli. cells
44
Pretreatment (a) 0.05±0.01
45
46 Control (b) 0.44±0.02
47 MHP-3/7 0.43±0.03
48 MHP-5/5 0.42±0.02
49
50 MHP-7/3 0.43±0.03
51
52 a) The PBS solution with E.coli.before incubation. b) The PBS solution was incubated with a
53
54 blank (control) and polymer coated prisms after two hours.
55
56
57 Phaeodactylum tricornutum Adhesion Assays. Two ternary thiol-ene polymer networks with
58
59 15
60
ACS Paragon Plus Environment
Langmuir Page 16 of 31

1
2
3
4 different contents of HMBA as well as two controls (glass and PDMSe) were chosen for
5
6 testing their antifouling performance against Phaeodactylum tricornutum. The purpose of this
7
8
9 research was to understand whether the “non-fouling” PEG component or the antibacterial
10
11 HMBA chains played a dominating role in the determination of the antibiofouling
12
13
14 performance of a polymer network material. As shown in Fig. 5A, the numbers of
15
16 Phaeodactylum tricornutum diatoms attached on the surfaces of MHP-3/7 and MHP-7/3
17
18
19 coatings are 195 × 103 and 85 × 103/cm2 respectively, which are much lower than those on the
20
21 hydrophilic glass control (290 × 103 /cm2), and the hydrophobic commercially available
22
23
24 PDMS coating (370 × 103 /cm2) surfaces. The results indicated that both the polymer network
25
26 amphiphilic coatings exhibited significantly better resistance to Phaeodactylum tricornutum
27
28
29 settlement. More importantly, the chemical composition of the polymer films markedly
30
31 affected their antifouling properties. The sample containing a higher HBMA exhibited a
32
33
34 better antifouling effect, indicating that here in this type of polymer network polymers the
35
36 anti-bacterial component is primarily responsible for the improved anti-biofouling
37
38
39 performance.
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 16
60
ACS Paragon Plus Environment
Page 17 of 31 Langmuir

1
2
3
4 Fig. 5 (A) The average numbers of Phaeodactylum tricornutum cells attached on various
5
6 sample surfaces. (B) Example optical microscopic images of the settlement of
7
8
9 Phaeodactylum tricornutum on the surfaces after 7 days. The bars are 10 µm.
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24 Fig. 6. SFG measurement of the polymer networks on a right angle silica prism in contact
25
26
with diluted water.
27
28
29 Molecular Surface Structures of Thiol-Ene Polymer Networks in Air and diluted Water. To
30
31 understand the varied antifouling performance of the polymer networks with different
32
33 components, molecular surface structures of these materials were probed using sum
34
35 frequency generation (SFG) vibrational spectroscopy systematically in air and in an aqueous
36
37
38 environment (Fig. 6). Fig. 7 (a) shows SFG ssp spectra detected in the frequency region
39
40 between 2800 and 3050 cm−1 observed from the surfaces of various coating samples in air.
41
42 The SFG spectra collected from the surfaces of the three samples are similar, dominated by
43
44 peaks centered at 2835, 2880, 2910, 2940, and 2985 cm-1, which can be assigned to the C-H
45
46
symmetric stretching modes of the methoxy group (OCH3), the regular methyl (C-CH3) group,
47
48
49 and the Si−CH3 group, the Fermi resonance modes of the regular methyl and/or the Si-CH3
50
51 groups, and the asymmetric of the methoxy group, respectively. Therefore the peaks at 2835
52
53 cm-1 and 2985 cm-1 are believed to originate from the HMBA component, and the 2880, 2910,
54
55 and 2940 cm-1 peaks are from the PDMS. This shows that both HMBA and PDMS
56
57
58 components cover the polymer network surfaces in air. The presence of a strong 2910 cm-1
59 17
60
ACS Paragon Plus Environment
Langmuir Page 18 of 31

1
2
3 peak indicates that the PDMS chains are dominating the polymer-air interface. The relative
4
5 intensities of the methoxy signals increased as the HMBA/PEG ratio in the sample increased
6
7
from 3:7 to 7:3. This suggests that the surface coverage of HMBA increased as a function of
8
9
10 bulk content, as expected. It is worth mentioning that the above peaks are overlapped with
11
12 each other (Fig. 7), thus we cannot exclude the possible presence of the signals around
13
14 2850-2860 cm-1 range, where the signals from the PEGDA methylene groups should be
15
16 located. Therefore, we could not exclude the possibility that PEGDA also was present on the
17
18
19 sample surface in air.
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36 Fig. 7. SFG ssp spectra of the polymer surfaces in air.
37
38
39 Since the antifouling polymers need to be used in an aqueous environment, we then studied
40
41 the surface restructuring behavior of the polymer networks in water. Previously we have
42
43
44 demonstrated that SFG is a powerful tool to elucidate the surface restructuring behavior of
45
46 polymer materials in water in situ in real time. 57, 65-70
To exclude spectral interference from
47
48
49 the water molecules, D2O was used. The SFG spectrum collected from the 7/3 polymer/water
50
51 interface was markedly different from that collected from the surface in air (Fig. 8a), showing
52
53
54 that the polymer surface structure changed in water. In the SFG spectrum collected from the
55
56
polymer/D2O interface (Fig. 8b), the strong peak at 2910 cm-1 disappeared, showing that the
57
58
59 18
60
ACS Paragon Plus Environment
Page 19 of 31 Langmuir

1
2
3
4 hydrophobic PDMS backbone component returned to the bulk in water. On the other hand,
5
6 the spectrum from the MHP-7/3 surface in water exhibited a distinct peak around 2850 cm−1,
7
8
9 indicative of the good ordering of the PEG component at the polymer/water interface,
10
11 showing that PEG component was clearly present on the surface in water. This is reasonable
12
13
14 because PEG is very hydrophilic; its presence on the surface reduced the free energy of the
15
16 entire system. The detection of the 2880 cm-1 signal indicates that the end groups of the
17
18
19 PDMS still can exist on the polymer surface in water.
20
21 This surface restructuring behavior is reversible upon drying of the surface. The SFG signal
22
23
24 recovered to the initial feature after the sample was removed from water and exposed to air
25
26 again. Moreover, the polymer surface structure can recover to the initial state after repeating
27
28
29 the process of placing the sample in water and then drying multiple times. We believe that
30
31 this result also supports the results obtained from the previous leaching experiment, showing
32
33
34 that no materials leached or removed from the sample in water, otherwise the surface
35
36 structure should gradually altered as a function of time during many repeat washings.
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53 Fig. 8. SFG spectra collected from the MHP-7/3 coating surface in (a) air, (b) in contact with
54
55 D2O, (C) in air, after removal from the D2O.
56
57
58 To understand the surface restructuring behaviors of all the polymer networks in water, SFG
59 19
60
ACS Paragon Plus Environment
Langmuir Page 20 of 31

1
2
3
4 spectra were also collected from the other two polymer networks with the 3/7 and 5/5 ratios
5
6 of HMBA and PEG in D2O. All the three SFG spectra collected from the polymer/D2O
7
8
9 interface are shown in Fig. 9. As presented above, the peak at 2850 cm-1 is assigned to the
10
11 CH2,s of the PEGDA chains. The peak observed at 2880 cm-1 is ascribed to the CH3,s of the
12
13
14 terminal methyl group of the PDMS side chains. The spectral features of the MHP coating
15
16 surfaces in water are similar, dominated by the PEG and PDMS end group. However, the
17
18
19 methoxy C-H symmetric stretching signal at 2835 cm-1 varied. At the MHP-3/7 interface, no
20
21 SFG signal at ~2835 cm-1 was detected in water, indicating that the lack of methoxy group at
22
23
24 the polymer/D2O interface. However, on the MHP-5/5 and 7/3 surfaces in D2O, the peak
25
26 centered at ~2835 cm−1 was clearly detected, indicating that the HMBA component migrated
27
28
29 to the interface, because of the increased overall content in the sample. The different surface
30
31 structures of the 3/7 and 7/3 samples showed that the different bulk contents led to the varied
32
33
34 surface molecular structures in water: More HMBA contents in the bulk resulted in a higher
35
36 surface coverage in water. We believe that the better antifouling performance against
37
38
39 Phaeodactylum tricornutum by the 7/3 sample is due to the high surface coverage of HMBA,
40
41
which presents a higher bacteria killing capability.
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58 Fig. 9. SFG spectra in the C−H stretching frequency region collected from the interfaces
59 20
60
ACS Paragon Plus Environment
Page 21 of 31 Langmuir

1
2
3
4 between D2O and (a) MHP-3/7, (b) MHP-5/5, and (c) MHP-7/3.
5
6 Conclusions
7
8
9 A simple method was developed to prepare novel bio-inspired thiol-ene polymer networks for
10
11 antibiofouling application. More specifically, ternary amphiphilic polymer materials were
12
13
14 synthesized via photo-polymerization of a polysiloxane macromonomer MSHP carrying
15
16 mercaptopropyl side groups with poly(ethylene glycol) diacrylate PEGDA and
17
18
19 capsaicin-mimic N-(4-hydroxy-3-methoxybenzyl)-acrylamide HMBA monomer, without
20
21 using the conventional heavy-metal catalyzed crosslinking reactions. The measured results on
22
23
24 cross-hatch adhesion, hardness, and stability indicated that such materials exhibit desired
25
26 physicochemical properties. It was found that the varied compositions of the two components
27
28
29 in the bulk led to different surface structures and surface properties. These materials were
30
31 found to be able to resist protein adsorption, likely due to the PEG component. The
32
33
34 incorporation of the capsaicin-mimic HMBA enables the polymer network materials to
35
36 demonstrate good antibacterial activity against E. coli and S. aureus. The molecular structural
37
38
39 information obtained from the SFG data provides direct experimental evidence to interpret
40
41
the mechanism of the antibiofouling performance of the thiol-ene polymer network materials.
42
43
44 The better antibacterial activity of MHP-7/3 was due to the surface presence of more HMBA
45
46
content. The diatom adhesion assay results indicated that the antibacterial component HMBA
47
48
49 played a dominating role in the improvement of antifouling performance of a polymer
50
51
network material against P. tricornutum. This research demonstrates that substantial presence
52
53
54 of the antibacterial component in a polymer material on the surface is needed to ensure
55
56
57
excellent antibiofouling performance of a polymer material.
58
59 21
60
ACS Paragon Plus Environment
Langmuir Page 22 of 31

1
2
3
4 ASSOCIATED CONTENT
5
6 Supporting Information
7
8
9 The Supporting Information is available free of charge on the ACS Publications website.
10
11 AUTHOR INFORMATION
12
13
14 Corresponding Authors
15
16 *E-mail zhanc@umich.edu
17
18
19 *E-mail chunjuhe@dhu.edu.cn
20
21 ORCID
22
23
24 Zhan Chen: 734-615-4189
25
26 Chunju He: +86-21-67792842
27
28
29 ACKNOWLEDGMENTS
30
31 This work was financially supported by Fundamental Research Funds for the Central
32
33
34 Universities (CUSF-DH-D-2016037) and US Office of Naval Research (N00014-16-1-3115).
35
36
REFERENCES
37
38
39 (1) Callow, J. A.; Callow, M. E., Trends in the Development of Environmentally Friendly
40
41 Fouling-Resistant Marine Coatings. Nature Communications 2011, 2.
42
43
44 (2) Kim, S.; Gim, T.; Kang, S. M., Versatile, Tannic Acid-Mediated Surface Pegylation for Marine
45
46
Antifouling Applications. ACS Appl Mater Interfaces 2015, 7, 6412-6.
47
48
49 (3) Shivapooja, P.; Wang, Q.; Szott, L. M.; Orihuela, B.; Rittschof, D.; Zhao, X.; Lopez, G. P., Dynamic
50
51
Surface Deformation of Silicone Elastomers for Management of Marine Biofouling: Laboratory and Field
52
53
54 Studies Using Pneumatic Actuation. Biofouling 2015, 31, 265-74.
55
56
(4) Pollack, K. A.; Imbesi, P. M.; Raymond, J. E.; Wooley, K. L., Hyperbranched
57
58
59 22
60
ACS Paragon Plus Environment
Page 23 of 31 Langmuir

1
2
3
4 Fluoropolymer-Polydimethylsiloxane-Poly(Ethylene Glycol) Cross-Linked Terpolymer Networks
5
6 Designed for Marine and Biomedical Applications: Heterogeneous Nontoxic Antibiofouling Surfaces. ACS
7
8
9 Appl Mater Interfaces 2014, 6, 19265-74.
10
11 (5) Wang, H. Y.; Zhang, C. F.; Wang, J. X.; Feng, X. F.; He, C. J., Dual-Mode Antifouling Ability of
12
13
14 Thiol-Ene Amphiphilic Conetworks: Minimally Adhesive Coatings Via the Surface Zwitterionization. Acs
15
16 Sustainable Chemistry & Engineering 2016, 4, 3803-3811.
17
18
19 (6) Lejars, M.; Margaillan, A.; Bressy, C., Fouling Release Coatings: A Nontoxic Alternative to Biocidal
20
21 Antifouling Coatings. Chemical Reviews 2012, 112, 4347-4390.
22
23
24 (7) Dafforn, K. A.; Lewis, J. A.; Johnston, E. L., Antifouling Strategies: History and Regulation, Ecological
25
26 Impacts and Mitigation. Marine Pollution Bulletin 2011, 62, 453-465.
27
28
29 (8) Magin, C. M.; Cooper, S. P.; Brennan, A. B., Non-Toxic Antifouling Strategies. Materials Today 2010,
30
31 13, 36-44.
32
33
34 (9) Fusetani, N., Antifouling Marine Natural Products. Natural Product Reports 2011, 28, 400-410.
35
36 (10) Krishnan, S.; Weinman, C. J.; Ober, C. K., Advances in Polymers for Anti-Biofouling Surfaces.
37
38
39 Journal of Materials Chemistry 2008, 18, 3405-3413.
40
41 (11) Qian, P. Y.; Xu, Y.; Fusetani, N., Natural Products as Antifouling Compounds: Recent Progress and
42
43
44 Future Perspectives. Biofouling 2010, 26, 223-234.
45
46
(12) Thomas, K. V.; Brooks, S., The Environmental Fate and Effects of Antifouling Paint Biocides.
47
48
49 Biofouling 2010, 26, 73-88.
50
51
(13) Sonak, S.; Pangam, P.; Giriyan, A.; Hawaldar, K., Implications of the Ban on Organotins for
52
53
54 Protection of Global Coastal and Marine Ecology. Journal of Environmental Management 2009, 90,
55
56
S96-S108.
57
58
59 23
60
ACS Paragon Plus Environment
Langmuir Page 24 of 31

1
2
3
4 (14) Yebra, D. M.; Kiil, S.; Dam-Johansen, K., Antifouling Technology - Past, Present and Future Steps
5
6 Towards Efficient and Environmentally Friendly Antifouling Coatings. Progress in Organic Coatings
7
8
9 2004, 50, 75-104.
10
11 (15) Salta, M.; Wharton, J. A.; Stoodley, P.; Dennington, S. P.; Goodes, L. R.; Werwinski, S.; Mart, U.;
12
13
14 Wood, R. J. K.; Stokes, K. R., Designing Biomimetic Antifouling Surfaces. Philosophical Transactions of
15
16 the Royal Society a-Mathematical Physical and Engineering Sciences 2010, 368, 4729-4754.
17
18
19 (16) Wang, C. C.; Feng, R. R.; Yang, F. L., Enhancing the Hydrophilic and Antifouling Properties of
20
21 Polypropylene Nonwoven Fabric Membranes by the Grafting of Poly(N-Vinyl-2-Pyrrolidone) Via the Atrp
22
23
24 Method. Journal of Colloid and Interface Science 2011, 357, 273-279.
25
26 (17) Bowen, J.; Pettitt, M. E.; Kendall, K.; Leggett, G. J.; Preece, J. A.; Callow, M. E.; Callow, J. A., The
27
28
29 Influence of Surface Lubricity on the Adhesion of Navicula Perminuta and Ulva Linza to Alkanethiol
30
31 Self-Assembled Monolayers. Journal of the Royal Society Interface 2007, 4, 473-477.
32
33
34 (18) Holland, R.; Dugdale, T. M.; Wetherbee, R.; Brennan, A. B.; Finlay, J. A.; Callow, J. A.; Callow, M.
35
36 E., Adhesion and Motility of Fouling Diatoms on a Silicone Elastomer. Biofouling 2004, 20, 323-329.
37
38
39 (19) Wan, F.; Pei, X. W.; Yu, B.; Ye, Q.; Zhou, F.; Xue, Q. J., Grafting Polymer Brushes on Biomimetic
40
41 Structural Surfaces for Anti-Algae Fouling and Foul Release. Acs Applied Materials & Interfaces 2012, 4,
42
43
44 4557-4565.
45
46
(20) Cho, Y. J.; Sundaram, H. S.; Weinman, C. J.; Paik, M. Y.; Dimitriou, M. D.; Finlay, J. A.; Callow, M.
47
48
49 E.; Callow, J. A.; Kramer, E. J.; Ober, C. K., Triblock Copolymers with Grafted Fluorine-Free,
50
51
Amphiphilic, Non-Ionic Side Chains for Antifouling and Fouling-Release Applications. Macromolecules
52
53
54 2011, 44, 4783-4792.
55
56
(21) Wan, F.; Ye, Q.; Yu, B.; Pei, X. W.; Zhou, F., Multiscale Hairy Surfaces for Nearly Perfect Marine
57
58
59 24
60
ACS Paragon Plus Environment
Page 25 of 31 Langmuir

1
2
3
4 Antibiofouling. Journal of Materials Chemistry B 2013, 1, 3599-3606.
5
6 (22) Zhu, X.; Janczewski, D.; Lee, S. S.; Teo, S. L.; Vancso, G. J., Cross-Linked Polyelectrolyte
7
8
9 Multilayers for Marine Antifouling Applications. ACS Appl Mater Interfaces 2013, 5, 5961-8.
10
11 (23) Banerjee, I.; Pangule, R. C.; Kane, R. S., Antifouling Coatings: Recent Developments in the Design of
12
13
14 Surfaces That Prevent Fouling by Proteins, Bacteria, and Marine Organisms. Advanced Materials 2011, 23,
15
16 690-718.
17
18
19 (24) Zhang, L.; Cao, Z. Q.; Bai, T.; Carr, L.; Ella-Menye, J. R.; Irvin, C.; Ratner, B. D.; Jiang, S. Y.,
20
21 Zwitterionic Hydrogels Implanted in Mice Resist the Foreign-Body Reaction. Nature Biotechnology 2013,
22
23
24 31, 553-+.
25
26 (25) Zhu, Y. H.; Xu, X. W.; Brault, N. D.; Keefe, A. J.; Han, X.; Deng, Y.; Xu, J. Q.; Yu, Q. M.; Jiang, S.
27
28
29 Y., Cellulose Paper Sensors Modified with Zwitterionic Poly(Carboxybetaine) for Sensing and Detection in
30
31 Complex Media. Analytical Chemistry 2014, 86, 2871-2875.
32
33
34 (26) Zhu, Y. H.; Sundaram, H. S.; Liu, S. J.; Zhang, L.; Xu, X. W.; Yu, Q. M.; Xu, J. Q.; Jiang, S. Y., A
35
36 Robust Graft-to Strategy to Form Multifunctional and Stealth Zwitterionic Polymer-Coated Mesoporous
37
38
39 Silica Nanoparticles. Biomacromolecules 2014, 15, 1845-1851.
40
41 (27) Cao, Z. Q.; Zhang, L.; Jiang, S. Y., Superhydrophilic Zwitterionic Polymers Stabilize Liposomes.
42
43
44 Langmuir 2012, 28, 11625-11632.
45
46
(28) Zhao, Y. F.; Zhu, L. P.; Jiang, J. H.; Yi, Z.; Zhu, B. K.; Xu, Y. Y., Enhancing the Antifouling and
47
48
49 Antimicrobial Properties of Poly(Ether Sulfone) Membranes by Surface Quaternization from a Reactive
50
51
Poly(Ether Sulfone) Based Copolymer Additive. Industrial & Engineering Chemistry Research 2014, 53,
52
53
54 13952-13962.
55
56
(29) Galvin, C. J.; Dimitriou, M. D.; Satija, S. K.; Genzer, J., Swelling of Polyelectrolyte and
57
58
59 25
60
ACS Paragon Plus Environment
Langmuir Page 26 of 31

1
2
3
4 Polyzwitterion Brushes by Humid Vapors. Journal of the American Chemical Society 2014, 136,
5
6 12737-12745.
7
8
9 (30) Leung, B. O.; Yang, Z.; Wu, S. S. H.; Chou, K. C., Role of Interfacial Water on Protein Adsorption at
10
11 Cross-Linked Polyethylene Oxide Interfaces. Langmuir 2012, 28, 5724-5728.
12
13
14 (31) Chen, S. F.; Li, L. Y.; Zhao, C.; Zheng, J., Surface Hydration: Principles and Applications toward
15
16 Low-Fouling/Nonfouling Biomaterials. Polymer 2010, 51, 5283-5293.
17
18
19 (32) Xu, J.; Feng, X.; Hou, J.; Wang, X.; Shan, B. T.; Yu, L. M.; Gao, C. J., Preparation and
20
21 Characterization of a Novel Polysulfone Uf Membrane Using a Copolymer with Capsaicin-Mimic
22
23
24 Moieties for Improved Anti-Fouling Properties. Journal of Membrane Science 2013, 446, 171-180.
25
26 (33) Liu, H. F.; Lepoittevin, B.; Roddier, C.; Guerineau, V.; Bech, L.; Herry, J. M.; Bellon-Fontaine, M. N.;
27
28
29 Roger, P., Facile Synthesis and Promising Antibacterial Properties of a New Guaiacol-Based Polymer.
30
31 Polymer 2011, 52, 1908-1916.
32
33
34 (34) Galhenage, T. P.; Hoffman, D.; Silbert, S. D.; Stafslien, S. J.; Daniels, J.; Miljkovic, T.; Finlay, J. A.;
35
36 Franco, S. C.; Clare, A. S.; Nedved, B. T.; Hadfield, M. G.; Wendt, D. E.; Waltz, G.; Brewer, L.; Teo, S. L.
37
38
39 M.; Lim, C. S.; Webster, D. C., Fouling-Release Performance of Silicone Oil-Modified
40
41 Siloxane-Polyurethane Coatings. Acs Applied Materials & Interfaces 2016, 8, 29025-29036.
42
43
44 (35) Chen, X.; Zhang, G. F.; Zhang, Q. H.; Zhan, X. L.; Chen, F. Q., Preparation and Performance of
45
46
Amphiphilic Polyurethane Copolymers with Capsaicin-Mimic and Peg Moieties for Protein Resistance and
47
48
49 Antibacteria. Industrial & Engineering Chemistry Research 2015, 54, 3813-3820.
50
51
(36) Adelmann, R.; Mennicken, M.; Popescu, D.; Heine, E.; Keul, H.; Moeller, M., Functional
52
53
54 Polymethacrylates as Bacteriostatic Polymers. European Polymer Journal 2009, 45, 3093-3107.
55
56
(37) Kenawy, E. R.; Worley, S. D.; Broughton, R., The Chemistry and Applications of Antimicrobial
57
58
59 26
60
ACS Paragon Plus Environment
Page 27 of 31 Langmuir

1
2
3
4 Polymers: A State-of-the-Art Review. Biomacromolecules 2007, 8, 1359-1384.
5
6 (38) Wang, J. B.; Shi, T.; Yang, X. L.; Han, W. Y.; Zhou, Y. R., Environmental Risk Assessment on
7
8
9 Capsaicin Used as Active Substance for Antifouling System on Ships. Chemosphere 2014, 104, 85-90.
10
11 (39) Shen, X.; Zhao, Y. P.; Feng, X.; Bi, S. X.; Ding, W. B.; Chen, L., Improved Antifouling Properties of
12
13
14 Pvdf Membranes Modified with Oppositely Charged Copolymer. Biofouling 2013, 29, 331-343.
15
16 (40) Gao, X. L.; Wang, H. Z.; Wang, J.; Huang, X.; Gao, C. J., Surface-Modified Psf Uf Membrane by
17
18
19 Uv-Assisted Graft Polymerization of Capsaicin Derivative Moiety for Fouling and Bacterial Resistance.
20
21 Journal of Membrane Science 2013, 445, 146-155.
22
23
24 (41) Dizman, B.; Elasri, M. O.; Mathias, L. J., Synthesis, Characterization, and Antibacterial Activities of
25
26 Novel Methacrylate Polymers Containing Norfloxacin. Biomacromolecules 2005, 6, 514-520.
27
28
29 (42) Moreau, O.; Portella, C.; Massicot, F.; Herry, J. M.; Riquet, A. M., Adhesion on Polyethylene Glycol
30
31 and Quaternary Ammonium Salt-Grafted Silicon Surfaces: Influence of Physicochemical Properties.
32
33
34 Surface and Coatings Technology 2007, 201, 5994-6004.
35
36 (43) Wang, H.; Qin, A.; Li, X.; Zhao, X.; Liu, D.; He, C., Biocompatible Amphiphilic Conetwork Based on
37
38
39 Crosslinked Star Copolymers: A Potential Drug Carrier. Journal of Polymer Science Part A: Polymer
40
41 Chemistry 2015, 53, 2537-2545.
42
43
44 (44) Shen, Y. R., Surface Properties Probed by Second-Harmonic and Sum-Frequency Generation. Nature
45
46
1989, 337, 519-525.
47
48
49 (45) Richmond, G. L., Molecular Bonding and Interactions at Aqueous Surfaces as Probed by Vibrational
50
51
Sum Frequency Spectroscopy. Chemical Reviews 2002, 102, 2693-2724.
52
53
54 (46) Chen, Z.; Shen, Y. R.; Somorjai, G. A., Studies of Polymer Surfaces by Sum Frequency Generation
55
56
Vibrational Spectroscopy. Annual Review of Physical Chemistry 2002, 53, 437-465.
57
58
59 27
60
ACS Paragon Plus Environment
Langmuir Page 28 of 31

1
2
3
4 (47) Okuno, M.; Ishibashi, T.-a., Heterodyne-Detected Achiral and Chiral Vibrational Sum Frequency
5
6 Generation of Proteins at Air/Water Interface. The Journal of Physical Chemistry C 2015, 119, 9947-9954.
7
8
9 (48) Hu, D.; Chou, K. C., Re-Evaluating the Surface Tension Analysis of Polyelectrolyte-Surfactant
10
11 Mixtures Using Phase-Sensitive Sum Frequency Generation Spectroscopy. Journal of the American
12
13
14 Chemical Society 2014, 136, 15114-15117.
15
16 (49) Anim-Danso, E.; Zhang, Y.; Alizadeh, A.; Dhinojwala, A., Freezing of Water Next to Solid Surfaces
17
18
19 Probed by Infrared–Visible Sum Frequency Generation Spectroscopy. Journal of the American Chemical
20
21 Society 2013, 135, 2734-2740.
22
23
24 (50) Ye, H.; Abu-Akeel, A.; Huang, J.; Katz, H. E.; Gracias, D. H., Probing Organic Field Effect
25
26 Transistors in Situ During Operation Using Sfg. Journal of the American Chemical Society 2006, 128,
27
28
29 6528-6529.
30
31 (51) Dhar, P.; Khlyabich, P. P.; Burkhart, B.; Roberts, S. T.; Malyk, S.; Thompson, B. C.; Benderskii, A. V.,
32
33
34 Annealing-Induced Changes in the Molecular Orientation of Poly-3-Hexylthiophene at Buried Interfaces.
35
36 The Journal of Physical Chemistry C 2013, 117, 15213-15220.
37
38
39 (52) Walter, S. R.; Youn, J.; Emery, J. D.; Kewalramani, S.; Hennek, J. W.; Bedzyk, M. J.; Facchetti, A.;
40
41 Marks, T. J.; Geiger, F. M., In-Situ Probe of Gate Dielectric-Semiconductor Interfacial Order in Organic
42
43
44 Transistors: Origin and Control of Large Performance Sensitivities. Journal of the American Chemical
45
46
Society 2012, 134, 11726-11733.
47
48
49 (53) Hirata, T.; Matsuno, H.; Kawaguchi, D.; Hirai, T.; Yamada, N. L.; Tanaka, M.; Tanaka, K., Effect of
50
51
Local Chain Dynamics on a Bioinert Interface. Langmuir 2015, 31, 3661-3667.
52
53
54 (54) Hirata, T.; Matsuno, H.; Kawaguchi, D.; Yamada, N. L.; Tanaka, M.; Tanaka, K., Effect of Interfacial
55
56
Structure on Bioinert Properties of Poly(2-Methoxyethyl Acrylate)/Poly(Methyl Methacrylate) Blend
57
58
59 28
60
ACS Paragon Plus Environment
Page 29 of 31 Langmuir

1
2
3
4 Films in Water. Physical Chemistry Chemical Physics 2015, 17, 17399-17405.
5
6 (55) Loch, C. L.; Ahn, D.; Chen, Z., Sum Frequency Generation Vibrational Spectroscopic Studies on a
7
8
9 Silane Adhesion-Promoting Mixture at a Polymer Interface. Journal of Physical Chemistry B 2006, 110,
10
11 914-918.
12
13
14 (56) Chen, C. Y.; Wang, J.; Loch, C. L.; Ahn, D.; Chen, Z., Demonstrating the Feasibility of Monitoring the
15
16 Molecular-Level Structures of Moving Polymer/Silane Interfaces During Silane Diffusion Using Sfg.
17
18
19 Journal of the American Chemical Society 2004, 126, 1174-1179.
20
21 (57) Wang, J.; Woodcock, S. E.; Buck, S. M.; Chen, C. Y.; Chen, Z., Different Surface-Restructuring
22
23
24 Behaviors of Poly(Methacrylate)S Detected by Sfg in Water. Journal of the American Chemical Society
25
26 2001, 123, 9470-9471.
27
28
29 (58) Cho, Y.; Cho, D.; Park, J. H.; Frey, M. W.; Ober, C. K.; Joo, Y. L., Preparation and Characterization of
30
31 Amphiphilic Triblock Terpolymer-Based Nanofibers as Antifouling Biomaterials. Biomacromolecules
32
33
34 2012, 13, 1606-14.
35
36 (59) Li, S.-S.; Xie, Y.; Xiang, T.; Ma, L.; He, C.; Sun, S.-d.; Zhao, C.-S., Heparin-Mimicking
37
38
39 Polyethersulfone Membranes – Hemocompatibility, Cytocompatibility, Antifouling and Antibacterial
40
41 Properties. Journal of Membrane Science 2016, 498, 135-146.
42
43
44 (60) Nurioglu, A. G.; Esteves, A. C. C.; de With, G., Non-Toxic, Non-Biocide-Release Antifouling
45
46
Coatings Based on Molecular Structure Design for Marine Applications. J. Mater. Chem. B 2015, 3,
47
48
49 6547-6570.
50
51
(61) Colak, S.; Tew, G. N., Amphiphilic Polybetaines: The Effect of Side-Chain Hydrophobicity on Protein
52
53
54 Adsorption. Biomacromolecules 2012, 13, 1233-1239.
55
56
(62) Pollack, K. A.; Imbesi, P. M.; Raymond, J. E.; Wooley, K. L., Hyperbranched
57
58
59 29
60
ACS Paragon Plus Environment
Langmuir Page 30 of 31

1
2
3
4 Fluoropolymer-Polydimethylsiloxane-Poly(Ethylene Glycol) Cross-Linked Terpolymer Networks
5
6 Designed for Marine and Biomedical Applications: Heterogeneous Nontoxic Antibiofouling Surfaces. Acs
7
8
9 Applied Materials & Interfaces 2014, 6, 19265-19274.
10
11 (63) Voo, Z. X.; Khan, M.; Xu, Q.; Narayanan, K.; Ng, B. W. J.; Bte Ahmad, R.; Hedrick, J. L.; Yang, Y.
12
13
14 Y., Antimicrobial Coatings against Biofilm Formation: The Unexpected Balance between Antifouling and
15
16 Bactericidal Behavior. Polym. Chem. 2016, 7, 656-668.
17
18
19 (64) Han, H.; Wu, J.; Avery, C. W.; Mizutani, M.; Jiang, X.; Kamigaito, M.; Chen, Z.; Xi, C.; Kuroda, K.,
20
21 Immobilization of Amphiphilic Polycations by Catechol Functionality for Antimicrobial Coatings.
22
23
24 Langmuir 2011, 27, 4010-9.
25
26 (65) Wang, J.; Paszti, Z.; Even, M. A.; Chen, Z., Measuring Polymer Surface Ordering Differences in Air
27
28
29 and Water by Sum Frequency Generation Vibrational Spectroscopy. Journal of the American Chemical
30
31 Society 2002, 124, 7016-7023.
32
33
34 (66) Hankett, J. M.; Lu, X. L.; Liu, Y. W.; Seeley, E.; Chen, Z., Interfacial Molecular Restructuring of
35
36 Plasticized Polymers in Water. Physical Chemistry Chemical Physics 2014, 16, 20097-20106.
37
38
39 (67) Hankett, J. M.; Liu, Y. W.; Zhang, X. X.; Zhang, C.; Chen, Z., Molecular Level Studies of Polymer
40
41 Behaviors at the Water Interface Using Sum Frequency Generation Vibrational Spectroscopy. Journal of
42
43
44 Polymer Science Part B-Polymer Physics 2013, 51, 311-328.
45
46
(68) Lu, X. L.; Zhang, C.; Ulrich, N.; Xiao, M. Y.; Ma, Y. H.; Chen, Z., Studying Polymer Surfaces and
47
48
49 Interfaces with Sum Frequency Generation Vibrational Spectroscopy. Analytical Chemistry 2017, 89,
50
51
466-489.
52
53
54 (69) Leng, C.; Sun, S. W.; Zhang, K. X.; Jiang, S. Y.; Chen, Z., Molecular Level Studies on Interfacial
55
56
Hydration of Zwitterionic and Other Antifouling Polymers in Situ. Acta Biomaterialia 2016, 40, 6-15.
57
58
59 30
60
ACS Paragon Plus Environment
Page 31 of 31 Langmuir

1
2
3
4 (70) Leng, C.; Buss, H. G.; Segalman, R. A.; Chen, Z., Surface Structure and Hydration of
5
6 Sequence-Specific Amphiphilic Polypeptoids for Antifouling/Fouling Release Applications. Langmuir
7
8
9 2015, 31, 9306-9311.
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27 TOC graphy
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 31
60
ACS Paragon Plus Environment

You might also like