11
Feature Article Surface hydration: Principles and applications toward low-fouling/nonfouling biomaterials Shenfu Chen a , Lingyan Li b , Chao Zhao c , Jie Zheng c, * a State Key Laboratory of Chemical Engineering, Department of Chemical and Biological Engineering, Zhejiang University, Hangzhou 310027, PR China b Department of Biomedical Engineering, Cleveland Clinic, Cleveland, OH 44195, USA c Department of Chemical and Biomolecular Engineering, The University of Akron, Akron, OH 44325, USA article info Article history: Received 17 April 2010 Received in revised form 6 August 2010 Accepted 10 August 2010 Available online 18 August 2010 Keywords: Zwitterionic Protein adsorption Biocompatability Nonfouling abstract Surface resistance to nonspecic protein adsorption, cell/bacterial adhesion, and biolm formation is critical for the development and performance of biomedical and analytical devices. Signicant needs and efforts have been made in the development of biocompatible and bioactive materials for antifouling surfaces, but much of the work retains an empirical avor due to the complexity of experiments and the lack of robust theoretical models. In this review, two major classes of nonfouling materials (i.e. hydro- philic and zwitterionic materials) and associated basic nonfouling mechanisms and practical examples are discussed. Highly hydrated chemical groups with optimized physical properties of the surface, along with appropriate surface coating methods, are the keys to developing effective and stable nonfouling materials for long-term biomedical applications. The zwitterionic polymers are promising nonfouling biomaterials due to the simplicity of synthesis, ease of applicability, abundance of raw materials, and availability of functional groups. Ó 2010 Elsevier Ltd. All rights reserved. 1. Introduction There is a signicant need for the development of novel anti- fouling materials to resist nonspecic protein adsorption and cell adhesion for medical implants, drug delivery carriers, biosensors, and marine coatings [1]. It is generally believed that reducing biofouling could signicantly attenuate subsequent adverse inammatory responses including leukocyte activation, tissue brosis, thrombosis coagulation, and infection. On the other hand, even 10 ng/cm 2 brinogen adsorption can induce a full-scale blood platelet adhesion, resulting in implantable device failure and adverse outcomes to the patients. Poly (ethylene glycol) (PEG)- based materials [2] have been well demonstrated for their super- low-fouling ability to resist nonspecic protein adsorption and cell adhesion. But, the PEG-based materials are readily subject to oxidation in most biochemically relevant solutions. Signicant effort has been invested in the search for alternative antifouling materials with stability higher than PEG and this has resulted in substantial progress in developing improved materials. There are two major classes of antifouling materials, namely polyhydrophilic and polyzwitterionic materials as summarized in Table 1 . A number of low-fouling or nonfouling polyhydrophilic materials including PEG, polysaccharides, and polyamides have been found to mostly share some common structural and chemical properties: hydrophilic nature, electrically neutrality, and hydrogen-bond acceptors/donor [3,4]. For the polyzwitterionic materials, they can be further classied into polybetaines carrying a positive and a negative charge on the same monomer unit such as 2-methacryloyloxylethyl phosphorylcholine (MPC), sulfobetaine methacrylate (SBMA), and carboxybetaine methacrylate (CBMA) and polyampholytes carrying 1:1 positive and negative charge on two different monomer units such as mixed charge complex eN þ (CH 3 ) 3 and eSO 3 /eCOO and natural amino acids (Glu, Asp, Lysþ, and Argþ). Since all proteins have positively and negatively charged residues randomly distributed on its surface, they can be easily adhered to either positively or negatively charged surfaces [5]. It is suggested that a nanoscale homogenous mixture of balanced charge groups from polyzwitterionic materials is the key to controlling nonfouling properties. Apart from functional mate- rials, surface packing also plays an important role in achieving nonfouling properties of a surface. Surface coating and modica- tion methods include physical adsorption via hydrophobic inter- actions or spin-coating [6e8], chemical linkage, self-assembled monolayers (SAMs) [9,10], grafted-from-surfacevia surface- initiated atom transfer radical polymerization (ATRP) [11,12], graft-to-surfacevia dipping, cross-linking, or reaction of the * Corresponding author. Tel.: þ1 3309722096. E-mail address: [email protected] (J. Zheng). Contents lists available at ScienceDirect Polymer journal homepage: www.elsevier.com/locate/polymer 0032-3861/$ e see front matter Ó 2010 Elsevier Ltd. All rights reserved. doi:10.1016/j.polymer.2010.08.022 Polymer 51 (2010) 5283e5293

Surface hydration: Principles and applications toward low-fouling

  • Upload
    lethuy

  • View
    226

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Surface hydration: Principles and applications toward low-fouling

lable at ScienceDirect

Polymer 51 (2010) 5283e5293

Contents lists avai

Polymer

journal homepage: www.elsevier .com/locate/polymer

Feature Article

Surface hydration: Principles and applications toward low-fouling/nonfoulingbiomaterials

Shenfu Chen a, Lingyan Li b, Chao Zhao c, Jie Zheng c,*

a State Key Laboratory of Chemical Engineering, Department of Chemical and Biological Engineering, Zhejiang University, Hangzhou 310027, PR ChinabDepartment of Biomedical Engineering, Cleveland Clinic, Cleveland, OH 44195, USAcDepartment of Chemical and Biomolecular Engineering, The University of Akron, Akron, OH 44325, USA

a r t i c l e i n f o

Article history:Received 17 April 2010Received in revised form6 August 2010Accepted 10 August 2010Available online 18 August 2010

Keywords:ZwitterionicProtein adsorptionBiocompatabilityNonfouling

* Corresponding author. Tel.: þ1 3309722096.E-mail address: [email protected] (J. Zheng).

0032-3861/$ e see front matter � 2010 Elsevier Ltd.doi:10.1016/j.polymer.2010.08.022

a b s t r a c t

Surface resistance to nonspecific protein adsorption, cell/bacterial adhesion, and biofilm formation iscritical for the development and performance of biomedical and analytical devices. Significant needs andefforts have been made in the development of biocompatible and bioactive materials for antifoulingsurfaces, but much of the work retains an empirical flavor due to the complexity of experiments and thelack of robust theoretical models. In this review, two major classes of nonfouling materials (i.e. hydro-philic and zwitterionic materials) and associated basic nonfouling mechanisms and practical examplesare discussed. Highly hydrated chemical groups with optimized physical properties of the surface, alongwith appropriate surface coating methods, are the keys to developing effective and stable nonfoulingmaterials for long-term biomedical applications. The zwitterionic polymers are promising nonfoulingbiomaterials due to the simplicity of synthesis, ease of applicability, abundance of raw materials, andavailability of functional groups.

� 2010 Elsevier Ltd. All rights reserved.

1. Introduction

There is a significant need for the development of novel anti-fouling materials to resist nonspecific protein adsorption and celladhesion for medical implants, drug delivery carriers, biosensors,and marine coatings [1]. It is generally believed that reducingbiofouling could significantly attenuate subsequent adverseinflammatory responses including leukocyte activation, tissuefibrosis, thrombosis coagulation, and infection. On the other hand,even 10 ng/cm2

fibrinogen adsorption can induce a full-scale bloodplatelet adhesion, resulting in implantable device failure andadverse outcomes to the patients. Poly (ethylene glycol) (PEG)-based materials [2] have been well demonstrated for their super-low-fouling ability to resist nonspecific protein adsorption and celladhesion. But, the PEG-based materials are readily subject tooxidation in most biochemically relevant solutions. Significanteffort has been invested in the search for alternative antifoulingmaterials with stability higher than PEG and this has resulted insubstantial progress in developing improved materials.

There are two major classes of antifouling materials, namelypolyhydrophilic and polyzwitterionic materials as summarized in

All rights reserved.

Table 1. A number of low-fouling or nonfouling polyhydrophilicmaterials including PEG, polysaccharides, and polyamides havebeen found to mostly share some common structural and chemicalproperties: hydrophilic nature, electrically neutrality, andhydrogen-bond acceptors/donor [3,4]. For the polyzwitterionicmaterials, they can be further classified into polybetaines carryinga positive and a negative charge on the samemonomer unit such as2-methacryloyloxylethyl phosphorylcholine (MPC), sulfobetainemethacrylate (SBMA), and carboxybetaine methacrylate (CBMA)and polyampholytes carrying 1:1 positive and negative charge ontwo different monomer units such as mixed charge complexeNþ(CH3)3 andeSO3

�/eCOO� and natural amino acids (Glu�, Asp�,Lysþ, and Argþ). Since all proteins have positively and negativelycharged residues randomly distributed on its surface, they can beeasily adhered to either positively or negatively charged surfaces[5]. It is suggested that a nanoscale homogenous mixture ofbalanced charge groups from polyzwitterionic materials is the keyto controlling nonfouling properties. Apart from functional mate-rials, surface packing also plays an important role in achievingnonfouling properties of a surface. Surface coating and modifica-tion methods include physical adsorption via hydrophobic inter-actions or spin-coating [6e8], chemical linkage, self-assembledmonolayers (SAMs) [9,10], “grafted-from-surface” via surface-initiated atom transfer radical polymerization (ATRP) [11,12],“graft-to-surface” via dipping, cross-linking, or reaction of the

Page 2: Surface hydration: Principles and applications toward low-fouling

Table 1Overview of hydrophilic and zwitterionic antifouling materials.

Materials Proteinadsorption

Celladhesion

Hydrophilic materialsPEG-based materialsPS-g-PEGMA and PMMA-g-PEGMA [92] Yes NoPEG-poly(phosphonate) terpolymer [93] Yes NoPLL-g-PEG [94,95] Yes YesPEGMA [96,97] Yes NoPPEGxLys [98] Yes NoPOEGMA [99e102] Yes YesPEO-PU-PEO [61,103e105] Yes NoPEO-PPO-PEO [106] Yes NoPEO [33] Yes NoPEG [8,68,107e109] Yes YesPy-g-PEG [63] Yes NomPEG-DOPA [60,110] Yes NomPEG-MAPD [110] Yes NoOEG-SAM [10,87,102] Yes Yes

PMOXA [77,111] Yes NoDendronGlycerol dendron [75] Yes NoHPG [76] Yes No

Tetraglyme [112e115] Yes YesDextran [57,116] Yes NoPolysaccharide [117] Yes NoPoly (HEMA) [118,119] No YesPVA [120] Yes NoPolyamines functionalized with acetyl chloride [121] Yes YesMannitol-SAM [73] Yes YesPeptide-based SAM [74] Yes No

PolybetainePoly(CBAA) [43,62,85,102,122,123] Yes YesPoly(SBMA) [100,102,124,125] Yes YesPoly(CBMA) [83,126] Yes YesPoly(MPC) [119,127,128] Yes YesPC-SAM [87,102] Yes YesOPC-SAM [87,102,129] Yes Yes

PolyampholyteSA/TMA-SAM [19,87,102] Yes YesCA/TMA-SAM [19,87,102] Yes YesPM/TMA-SAM [19,87,102] Yes NoPeptide surfaces derived from natural amino acids [90] Yes NoPoly(TM-SA) [12] Yes NoPoly(METMA-MES) [14] Yes NoPDDA/PSS [130] Yes No

S. Chen et al. / Polymer 51 (2010) 5283e52935284

specific groups of polymers with the substrate [13,14], and plasmastreatments [15].

It is hypothesized that the nonfouling ability of both poly-hydrophilic and polyzwitterionic materials are tightly correlatedwith a hydration layer near the surface [16e18], because a tightlyboundwater layer forms a physical and energetic barrier to preventprotein adsorption on the surface. Water molecules residing onand/or penetrating into nonfouling materials can be classified intotwo types of “surface-bound” waters formed by hydrogen bondingfor hydrophilic materials and by even more strongly ionic solvationfor zwitterionic materials. Expulsion of water molecules from bothsurface and protein is the first and obligatory step to facilitateprotein adsorption by reducing free energy barrier arising fromdehydration entropic effects [19]. This is an intrinsic cause fornonfouling materials. The strength of surface hydration is primarilydetermined by physicochemical property of materials (i.e. molec-ular weight and surface chemistry) and their surface packing (i.e.film thickness, packing density, and chain conformation). Asidefrom surface hydration, chain flexibility also plays an importantrole in protein resistance especially for long-chain polymers. Whenprotein approaches to the surface, the compression of the polymerchains causes steric repulsion to resist protein adsorption due tounfavorable decrease in entropy [20,21]. Although most of water-

soluble polymers can reduce protein adsorption to some extent, thebest nonfouling ability of polymers can only be achieved whensurface hydration and steric repulsion work together. It is specu-lated that if chain flexibility plays a little role, chain hydration is anonly source to resist protein adsorption (Fig. 1c), while if chainflexibility plays a certain role, both chain flexibility and chainhydration are important (Fig. 1a and b).

Several nonfouling material reviews were published recently[22,23]. These articles present a variety of antifouling materialssynthesized by different methods. However, we believe that a moredetailed and comprehensive review that combines design princi-ples and biological applications would be useful for the biomaterialcommunity. This review includes three major parts. The first partdiscusses the importance of surface hydration in protein resistanceusing both experimental and computational evidences. The secondpart focuses on the hydrophilic materials while the third part onthe zwitterionic materials, both following the concept of enhancingsurface hydration. Examples of nonfouling materials and associatedsurface-grafting methods in each part are given to illustrate basicprinciples and practical applications. Finally, recent advances andfuture perspective are discussed.

2. Surface hydration: hydrogen bonding or ionic solvation

2.1. Experimental evidence of hydration layer

Numerous studies have been carried out to characterize thedynamics and structure of interfacial waters and their correlationwith nonfouling properties. Tanaka et al. classified interfacialwaters in poly(2-methoxyethyl acrylate) (PMEA) into threecategories: nonfreezing water (non-crystallize at �100 �C),freezing bound water (crystallize below 0 �C), and free water(crystallize at 0 �C), they found that freezing bound water atabout �40 �C displayed an excellent blood compatibility due tostrong interactions with polymer via hydrogen bondings[24e26]. Similarly, Morisaku et al. measured and compared thecontents of freezable (non-crystallizable at �70 �C) and non-freezable (crystallizable at about 0 �C) waters in poly(MPC) andpoly(methoxy oligo(ethylene glycol)-monomethacrylate) [(Me(EG)nMA)] hydrogels [27]. They found that the poly(MPC)hydrogels had larger non-freezable water than the poly(Me(EG)nMA) hydrogels although the equilibrium water contents weresimilar and they thus proposed that the amount of non-freezablewater around polymer chains may influence the degree ofprotein adsorption resistance. Along the similar line, moreinformation about structure transition of nonfouling materialsbefore and after exposed to water was investigated. Fick et al.showed the swelling behavior of self-assembled monolayers ofalkanethiol-terminated poly(ethylene glycol) by neutron reflec-tometry [28]. Zolk et al. used IRevis SFG to demonstrate theconformational change of OEG-SAMs on gold from an orderedcrystallize state to a disordered amorphous state when the OEG-SAMs were switched from dry air condition to aqueous solution[29]. Wang et al. applied vibrational sum-frequency generation(VSFG) to confirm that interfacial water molecules were able tonot only retain in SAM film even after the sample was removedfrom the fluid, but also recover the disturbed SAM structure afterre-immersing the sample into waters [30]. All theses resultsindicate that water molecules can be associated with thehydrophilic film due to their great capability of hydration.

From a more universal point of view, researchers who pay moreattention on the strong interactions between water molecules andpolymers provided further evidences in such a correlation betweenstrong hydration of polymers and their resistance to proteinadsorption. When hydrophilic polymers contact with bulk water,

Page 3: Surface hydration: Principles and applications toward low-fouling

Fig. 1. Illustration of chain hydration and chain flexibility of (a) hydrophilic polymers, (b) zwitterionic polymers, and (c) SAMs, which attribute to surface resistance to nonspeficprotein adsorption in different ways.

S. Chen et al. / Polymer 51 (2010) 5283e5293 5285

water molecules penetrate into the polymer film to forma hydrogen-bond network in the polymers. The highly hydratedpolymer films exhibit nonfouling property, thus any decrease in thesurface hydration might lead to the decrease of resistance tononspecific protein adsorption. Since hydrogen bonds are relativelyeasily to break and reform, hydrophilic materials such as poly-amide, mannitol, and PEG often experience the transition fromnonfouling to fouling upon the change in surface hydration [31]caused by increase of packing density [10,32], increase of hydro-phobicity when copolymerized with hydrophobic monomers [17],and raise of temperature [33,34].

On the other hand, zwitterionic materials containing bothpositive and negative charged units can bind water moleculeseven more strongly and stably via electrostatically inducedhydration, as compared to those hydrophilic materials to achievesurface hydration via hydrogen bonding. Thus, materials withmore strongly ionic solvation fall into the scope of searchingnonfouling materials. pMPC, a big group of bio-mimetic polymer,has been extensively investigated with a long history in itsbiocompatibility for numerous applications [35e38]. Kitano et al.found that more bound water around the surface of zwitterionicmaterials than other hydrophilic or hydrophobic materials [39],which revealed the differences in the nature and strength ofhydration between ionic solvation and hydrogen bonding.Holmlin et al. [40] demonstrated that the resistance to nonspe-cific protein adsorption of a mixed SAMs formed by a 1:1combination of a thiol terminated in a trimethylammoniumgroup and a thiol terminated in a sulfonate group is comparableto the best known OEG-SAM system. This result indicates that theflexibility in short-chain OEG-SAM systems is not required, onlystrong hydration of SAMs became a necessary prerequisite forprotein resistance. On the other hand, the hydration of zwitter-ionic groups could be greatly compromised to cause the increaseof nonspecific protein adsorption at certain experimentalconditions such as very low ionic solution, very large molecularweight, extreme high or low packing density, or low temperature[10,41e43].

2.2. Molecular simulations of nonfouling materials

In parallel to extensive experimental efforts, molecularsimulations provide atomic-level information to understand thenonfouling mechanism. De Gennes and co-workers [20] reportedthe first theoretical studies of the resistance of poly(ethyleneglycol) (PEG) polymers to protein adsorption. They concludedthat steric repulsion resulting from the compression of PEOchains as protein approaches the surface was mainly responsiblefor prevention of protein adsorption. It was predicted that thelonger chain lengths and higher surface densities lead to betterprotein resistance. Szleifer et al. [44,45] improved the De Gen-nes’s model using the single chain mean field (SCMF) theory.They found that polymers grafted to a hydrophobic surface canreduce protein adsorption simply because the polymers blocked

protein adsorption sites. In the Szleifer model, surface densitywas a key factor to control protein adsorption, but chain lengthhad a little effect on protein adsorption. Thus, the Szleifer modelis able to interpret experimental results for both long-chain PEGand short-chain oligo (ethylene glycol) (OEG) polymers. Both DeGennes and Szleifer models highlight the importance of surfacedensity of grafted polymers to resist protein adsorption, but theirmodels do not provide atomic-level information about non-fouling mechanism. In their work, the protein was modeled asa structureless, spherical particle, while water was treated asa continuous medium in the De Gennes model or as a homoge-neous spherical noninteracting molecule in the Szleifer model.While simplified models allow simulations to be performed inthe time scale of seconds, the detailed conformational change ofa protein and polymer chains was ignored. Later, the Grunzegroup [17,46,47] performed grand canonical Monte Carlo (GCMC)simulations to study the effect of conformation of OEG-termi-nated alkanethiol SAMs on gold and silver substrates on theinteraction of waters with OEG-SAM. They attributed OEG-SAMswith helical conformation on Au (h-SAM) to reduce proteinadsorption, while those SAMs with all trans conformation on Ag(t-SAM) enhance protein adsorption. Their results showed thatmore water molecules penetrated into h-SAM than t-SAM toform hydrogen bonds with OEG chains, leading to the preventionof protein adsorption on the surfaces.

Since molecular details of a surface are of great importance toprotein adsorption, a series of explicit-solvent moleculardynamics (MD) simulations [16,48,49] were performed toexamine the effect of OEG surface density on the lysozymeadsorption by varying the composition of cOEG ¼ 0.2, 0.5, 0.8, and1.0 in a mixed SAMs of S(CH2)4(OCH2CH2)4OH (OEG-SAM) and S(CH2)4OH (OH-SAM). Simulation results show that more watermolecules are tightly bound at the protein/SAMs interface viahydrogen bonding at medium OEG surface densities cOEG ¼ 0.5and 0.8 than those at too high (cOEG ¼ 1.0) or too low (cOEG ¼ 0.2)OEG surface densities. Recent surface plasmon resonance (SPR)studies of protein adsorption on the OEG-SAMs [10] also showthat OEG-SAMs resist protein adsorption within a certain range ofthe surface OEG densities (0.6e0.8), yet adsorb proteins when theOEG surface density is too high or too low (Fig. 2a). By thecomparison of experimental and simulation results, it appearsthat there is a correlation between the nonfouling properties ofthe OEG-SAM surface and the hydration of OEG chains (orappropriate OEG surface density). To directly correlate hydrationforce with nonfouling ability of a surface, restrained MD simula-tions [16,50] are performed to calculate the interaction forcesexerted onto a protein from an OEG-SAMs and interfacial solventas the protein approaches to the surface from a large separationdown to intimate contact. Force vs distance profiles (Fig. 2b) showthat relative strength of repulsive force acting on the protein is inthe decreasing order of OEG-SAMs > OH-SAMs > CH3-SAMs.Moreover, the total repulsive force is decomposed into twocontributions: one from the SAM surface and the other from

Page 4: Surface hydration: Principles and applications toward low-fouling

Fig. 2. (a) The amount of adsorbed fibrinogen as a function of the surface packing density of (ethylene glycol)4 OH-SAM [10]. Copyright 2005 American Chemical Society. (b)Interaction force acting on the Lysozyme from three SAM surfaces [16]. Copyright 2005 Biophysical Journal.

Fig. 3. Hydration capacity (i.e. number of hydrated waters) for a series of nonfoulingmaterials evaluated by experiments and simulations, where OEG represents Oligo(ethylene glycol) with four ethylene glycol repeat units, MAN (mannitol), SORB(sorbitol), GLY (primary amine glycine), 1MG (secondary amine sarcosine), DMG(tertiary amine dimethylglycine), and TMG (quartenary amine trimethylglycine). Dashline indicates a threshold intrinsic hydration capacity for nonfouling surfaces. Copy-right 2008 American Chemical Society.

S. Chen et al. / Polymer 51 (2010) 5283e52935286

interfacial water. Results show that the repulsive force mainlycomes from interfacial waters (over 95% contribution to totalforce), not from the SAM surface. On the other hand, the dynamic,structural, and binding properties of interfacial waters are indeedaffected by the physicochemical properties of surfaces, indicatingthat the interplay interactions between waters and surfaces areessential for nonfouling ability.

Along the same line, although simulations demonstrate thehydration force originated from interfacial waters at the hydro-philic OEG-SAMs is dominant driving force to resist proteinadsorption, “surface hydration hypothesis” derived from hydro-philic surfaces is needed to be further examined for zwitterionicsurfaces to establish the generality of nonfouling mechanism.Additional MD simulations are performed to study the interactionsof zwitterionic phosphorylcholine-terminated SAM [S(CH2)

11PO4(CH2)2N(CH3)3, PC-SAM] [49,51] with a model protein oflysozyme. The PC-SAMs display a much stronger repulsive force toprotein raised from interfacial waters than the OEG-SAMs.Although both PC-SAM and OEG-SAM have been demonstrated asnonfouling surfaces by experiments [10,52], comparison of inter-facial waters near the PC-SAM and OEG-SAM surfaces from simu-lations reveals different structural and dynamic properties. First,interfacial waters near zwitterionic PC-SAM surface have longerresidence time (stay longer) and smaller self-diffusion coefficient(move slower) than those near the hydrophilic OEG-SAM surface,indicating that PC-SAM surface binds water molecules morestrongly than the OEG-SAM surface. Secondly, interfacial waters atthe PC-SAMs havemuch larger reorientational dynamics than thoseat the OEG-SAMs, indicating that the PC-SAM binds waters via ionicsolvation while the OEG-SAM via hydrogen bonding. Furthermore,based on the fact that surface hydration is well correlated withnonfouling properties derived from PEG and PC systems, a simplecomputational method is developed to quickly evaluate theintrinsic hydration capacity of model surface coating functionalgroups by computing hydration volume and number of hydratingsolvent molecules [53]. Fig. 3 shows a range of hydration capacityand nonfouling performance of three class materials of ethyleneglycols, sugar alcohols, and glycine analogues. With known asnonfouling materials from experiments, OEG with four ethyleneglycol repeat units, MAN (mannitol), SORB (sorbitol), and TMG(quartenary amine trimethylglycine) have scaled number of asso-ciatedwater molecules greater than 0.15, while GLY (primary amineglycine), 1 MG (secondary amine sarcosine), and DMG (tertiaryamine dimethylglycine), all known to allow nonspecific proteinadsorption, are less than 0.15. Consistent experimental measure-ments and simulation calculations suggest a critical threshold value

of hydration capacity (w0.15) for initially screening nonfoulingmaterials.

3. Hydrogen bonding-rich materials

Many hydrophilic materials have been demonstrated theirnonfouling ability to reduce not only nonspecific protein adsorp-tion, but also cell and bacteria adhesion. Hydrophilic PEG-basedpolymers [2,54e56] and polymers incorporating oligosaccharidemoieties [57,58] are inherently anti-biofouling in nature. Hydrationof materials via hydrogen bonding is often used to interpret themolecular basis for nonfouling behavior by experimental andcomputational works.

3.1. Short-chain hydrophilic SAMs

Molecular details of the surface are of great importance toprotein adsorption. SAMs are well suitable platforms to studyprotein adsorption at the molecular-level since nanoscale surfaceproperties can be precisely controlled via varying the abundance,type, and spatial (both normal and lateral) distribution of tailgroups of the SAMs. Over the last 10e15 years, amongmany studiesto search for nonfouling materials beyond PEG, Whitesides and hisco-workers have used SPR and SAMs as a platform to systematicallytest and indentify a large number of protein-resistant SAMs with

Page 5: Surface hydration: Principles and applications toward low-fouling

S. Chen et al. / Polymer 51 (2010) 5283e5293 5287

different functional groups [3,4]. Four common features for thosesurfaces resisting protein adsorption are summarized e (i) hydro-philic, (ii) electrically neutral, (iii) hydrogen bond acceptors, and(iv) not hydrogen bond donors. However, this generalization doesnot explain all nonfouling surfaces especially carbohydratesterminated SAMs. For example, a SAM terminated with mannitolgroups only containing H-bond donor were found to prevent theadsorption of several proteins and the attachment of cell and wasindistinguishable from a SAM presenting oligo (ethylene glycol)(e(EG)nOH and e(EG)nOCH3, n ¼ 3e6) groups [50]. This resultfurther indicates that the formation of tightly bound water viahydrogen bonds between interfacial waters and surfaces does notnecessarily require “H-bond acceptors” materials, rather than“H-bond forming” materials.

Both nanoscale surface structural and chemical properties ofSAMs affect protein adsorption, including adsorption kinetics,adsorption isotherm, and protein morphology on surfaces. As canbeen seen in Fig. 4a, COOH-terminated SAMs formed at highertemperature of 50 �C have more compact structures than thoseformed at room temperature of 22 �C. The difference in nanoscalestructure was revealed by low-current STM but was not detectablefrom X-ray photoelectron spectroscopy (XPS) and contact anglemeasurements. Carboxylic-terminated SAMs formed at highertemperatures generally promote protein adsorption because oftheir compact structures and fewer adsorbed water molecules.However, the adsorbed amount of protein on hydrophobicmethyl-terminated SAMs prepared at different solution temperatures(55 �C and 22 �C) is similar because of similar adsorbed watermolecules despite their nanoscale structural difference, high-lighting the importance of surface chemistry. Similarly, surfacedensity and structure can also be controlled by using differentsolvents to assemble SAM chains. XPS reveals that OEG-SAMsassembled from a 95% ethanol and 5% water solution show higherpacking density on gold than those frompure ethanol solution [10].SPR results show that more fibrinogens adsorb on the denselypacked OEG-SAMs prepared by mixed ethanol and water solutionthan those on relatively loosely packed OEG-SAMs prepared bypure ethanol solution (Fig. 4b). The normalized coverage offibrinogen and lysozyme on OEG-SAMs assembled from the 95%ethanol solution is about 8.5% and 6%, respectively, assumed a fullmonolayer coverage of protein on C16SH-SAM surface and use thiscoverage to normalize the amount of protein adsorbed on theOEG-SAMs. Although the adsorbed amount of protein is small (w18 ng/cm2 or w8.5% surface coverage) on the OEG-SAMs prepared fromthe 95% ethanol solution, previous studies show that as little as10 ng/cm2 adsorbedfibrinogen could inducemonocyte adhesion invitro [59]. Early studies [10,17,32] also show that too high or lowOEG surface densities will lead to protein adsorption. AppropriateOEG surface densities are thus required for its nonfouling behavior(Fig. 2).

Moreover, mixed SAMs prepared from a 1:1 mixture of posi-tively charged 11-mercapto-N,N,N-trimethylammonium chloride(TMA) and one of the following negatively charged compounds:SA, 12-mercaptododecanoic acid (CA), 11-mercaptoundecylphos-phonic acid (PA), or methyl 11-mercaptoundecylphosphonate(PM) are further studied to examine their protein-resistantproperties [19,40]. SPR results in Fig. 5 show that all these mixedSAMs (TMA/SA, TMA/CA, and TMA/PM) had been shown to have<0.3 ng/cm�2 protein adsorption from single fibrinogen, lyso-zyme, and BSA solution. Although SAM is a model system to studythe interactions between protein and surface due to ease ofpreparation and modification, non-covalent bonds between thoiland gold may not be stable enough in complex media and anysurface defect in SAM structures can greatly and quickly reducenonfouling property.

3.2. Long-chain hydrophilic polymers

Surface hydration is often used to interpret the nonfoulingbehavior of the short-chain SAMs because the tightly bound watermolecules on the topmost part of SAMs are only source of largerepulsive forces to repel protein adsorption, whereas chain flexi-bility plays little role in protein resistance due to tightly packeddensity and short rigid chains. But, for the long-chain polymer, bothsurface hydration (i.e. water barrier) and chain flexibility (i.e. stericrepulsion) contribute to surface resistance to nonspecific proteinadsorption. Unlike surface hydration, the steric repulsion resultsfrom the compression of long polymer chains to yield repulsiveforces to resist protein adsorption. It is generally accepted that thenonfouling properties of a surface are determined by both func-tional groups and their surface packing. There are two ways tofabricate polymer brushes: physisorption and covalent attachment.Covalent attachment can be accomplished by either “graft-to” or“graft-from” approaches. The “graft-from” approach is a morepromising method in the synthesis of polymer brushes with a highgrafting density andwell-controlled film thickness, while “graft-to”method is more convenient for practical applications [14].

Numerous methods have been developed to immobilize PEG-based materials (i.e. the most commonly used nonfouling materialstoday) on surfaces including physisorption [6,13,60], chemisorption[61,62], covalent grafting of PEG onto surfaces [63,64], and plasmapolymerizationofOEGprecursors [12,15]. Thenonfoulingpropertiesof PEG are primarily due to their high chainmobility, large exclusionvolume, and steric hindrance effect of highly hydrated layer. Luesseand Arnold applied proton and deuterium NMR relaxation timemeasurements to determine the number of water molecules perethylene glycol repeat unit (eCH2CH2eOe) in PEG. They observedthat the maximum water content was one water molecule perethylene glycol unit [65]. The enthalpy of hydration per mole ofoxyethylene group is w7 kJ measured by both microtitration calo-rimetry [66] and NMR [67]. Although PEG displays excellent proteinresistance ability in vitro, in vivo studies of cell/bacteria adhesion onPEG surfaces showed that bacteria can attach to PEG-based surfacesat the low surface density [68] or to the nanopatterned PEG-basedsurfaces [69], highlighting the importance of surface density andtopology for nonfouling properties. The cell/bacteria adhesion onthe PEGylated surfaces might be caused by following possiblereasons: (i) The PEG-basedmaterials are readily subject to oxidationin most biochemically relevant solutions [59,70e72]. The short-term stability of PEG is one of its major limits responsible fordecreased performance in preventing nonspecific protein adsorp-tion and (ii) Defects on the PEG-based surfaces always inducenonspecific protein adsorption and cell attachment. For example,the “graft-to-surface”methods via physical adsorption, dipping, andcross-linking usually result inmore defects and less coverage on thesurface than the “graft-from-surface”method such as SI-ATRP, thusleading to more protein adsorption and cell adhesion. It should benoted that the nonfoulingmechanism of PEG-based surfaces to cell/bacterial adhesion is still not conclusivedue to the complex responseand interactions among PEG, cell/bacterial, and associated envi-ronmental stimuli, especially in complex medium such as in vivo orconditions with full nutrition for biofilm formation.

Recently, substantial efforts have been made toward thedevelopment of alternative bioinert polymers as a substitute forPEG, including polyoxazolines [73,74], polyglycerol (PG) dendrons[75,76], polysaccharides [57], and zwitterionic polymers. Conceptslearned from Nature provide a basis for designing new interfacingmaterials. The blood contacting face of the vascular endothelialcells is nonfouling and non-activating, suggesting biomimeticdesign ideas. Poly-2-methyl-2-oxazoline (PMOXA), a peptide-likepolymer [77], exhibited a significant reduction in full human serum

Page 6: Surface hydration: Principles and applications toward low-fouling

Fig. 4. The effect of surface packing density of SAM chains induced by temperature and solvent on the protein adsorption. (a) Protein adsorption is different on COOH-SAMsprepared at higher or room temperature from SPR experiments. STM images (20 nm � 20 nm) show more uniform SAMs form at higher temperature. (b) Fibrinogen (Fn) has 8.5%monolayer coverage on highly packed OEG-SAMs prepared from mixed 95% ethanol and 5% water solution while loosely packed OEG-SAMs prepared from pure ethanol resist Fnadsorption.

S. Chen et al. / Polymer 51 (2010) 5283e52935288

protein adsorption quantitatively to the same level as for PLL-g-PEG(<2 ng/cm2 as determined by optical waveguide lightmode spec-troscopy). Dendritic polyethers based on glycerol dendrimers andlinear/hyperbranched polyglycidols [78] are highly biocompatiblewith well cytotoxic toleration by mice even when injected in highdoes. Mimicking biological cell membranes containing highlyhydrated glycocalyx provides another approach for conferringantifouling properties to surfaces [57] because carbohydrates e theprincipal component of the glycocalyx e are thought to be mainlyresponsible for its ability to prevent undesirable nonspecificadsorption of protein. A series of oligosaccharide surfactant poly-mers consisting of a flexible polyvinylamine (PVAm) backbonewithhydrophilic dextran or oligomaltose and hydrophobic hexanolyside chains [58,79] were synthesized and coated on the hydro-phobic surface such as graphite, medical-grade polycarbonate, and

OTS via a simple dipping method. On hydrophobic surface, thehydrophobic hexanoly chains assemble through hydrophobicinteractions and are constrained to lie parallel to the substrate,while solvated hydrophilic dextran side chains protruding into theaqueous phase to shield a biomaterial from adhesive bacterialinteractions. The application of thin oligosaccharide surfactantpolymers has been shown to reduce non-adhesive interactionswith fibrinogen and reduce infection and thrombus formation vianon-adhesive interactions with proteins, platelets, cells andbacteria.

4. Zwitterionic materials

Zwitterionic materials are considered as biocompatible becausethey are bio-mimetic ones. Two types of nonfouling

Page 7: Surface hydration: Principles and applications toward low-fouling

Fig. 5. Fibrinogen adsorption on three 1:1 mixed SAMs of positively and negativelycharged components of SA/TMA, CA/TMA, and PM/TMA. All these mixed charged SAMshave demonstrated their highly resistant to Fn adsorption. Copyright 2006 AmericanChemical Society.

Fig. 7. (a) Chemical structures of phosphonate-betaines (PB), sulfonate-betaines (SB),and carboxylate-betaines (CB). (b) Platelet adhesion on OEG-SAM and different poly-mer brush surfaces: polySBMA, polyCBMA, and polyOEGMA. Copyright 2008 ElsevierLtd.

S. Chen et al. / Polymer 51 (2010) 5283e5293 5289

polyzwitterionic materials are polybetaines with positively andnegatively charged moieties on the same monomer unit and pol-yampholytes with positively and negatively charged moieties ondifferent monomer units. A critical factor determining nonfoulingproperty of polyzwitterionic materials is to control both uniformityof charge distribution and charge neutrality of two opposite chargemoieties on the surface (Fig. 6), which is the guide line to designnew nonfouling polyzwitterionic materials. In such a way non-fouling polyzwitterionic materials could maximize the surfacehydration and reduce charge interaction with protein molecules. Inthis section, wewill focus on the most recent advance in the designof various polybetaines and polyampholytes following this designguide line on surfaces for biological applications.

4.1. Polybetaines

Polybetaines can be further classified into three major groupsbased on the negatively charged groups: sulfonate-betaines (SB),carboxylate-betaines (CB), and phosphonate-betaines (PB) (Fig. 7).The first synthetic polybetaine is polycarboxybetaine through thehydrolysis of poly(4-vinylpyridine) reported by Ladenheim andMorawetz in 1957 [80]. A number of polybetaines such as phos-phorylcholine (PC)-based copolymers [35,81,82], polySB, and pol-yCB materials [41,62,64,83e85] have been synthesized andcharacterized for demonstrating their biocompatibility and

Fig. 6. Illustration of the guide line to design new nonfouling polyzwitterionic mate-rials. All these surfaces have both features in uniformity of charge distribution andcharge neutrality of two opposite charge moieties on the surface, which show excellentresistance to proteins.

hemocompatibility to reduce nonspecific protein adsorption andconsequent platelet adsorption in various biological applications.

To provide insights into the molecular-level nonfouling mech-anism of zwitterionic materials, key factors leading to the strongresistance of zwitterionic phosphorylcholine (PC)-terminatedSAMs on gold were examined by both experimental and molecularsimulation methods [49,51,52]. Results indicated that three keyproperties, including the charge balance, the minimized dipole andthe close packing density, lead to fibrinogen adsorption of <0.3 ng/cm2 and BSA adsorption of <0.1 ng/cm2 when 1:1 N/P ratio ispresented to maintain the charge neutrality of PC-SAMs where Natom represents a positive charge and P atom represents a negativecharge in PC [52]. The thickness of PC-SAMs is lower than theconfiguration of PC group pointing to solution, indicating that PCgroups are in an antiparallel orientation through charge and dipoleinteraction. An additional MD simulation revealed that thePC-SAMs could form 2-D crystalline structure with the samepacking configuration as PC head groups of membrane lipids, inagreement with the observation in experiments [49,51,52].

Upon these PC results, the nonfouling property of polySB andpolyCB were explored since those groups possess the similarity incharge distribution and neutrality as PC [11,84]. The nonfoulingproperties of SB and CB have been demonstrated by several well-controlled methods to form polymer thin film via surface-initiatedatom transfer radial polymerization (SI-ATRP) [86], self-assembly[84], free radical polymerization [87], and 3,4-dihydrox-yphenylalanine (DOPA) linkage [14]. Ladd et al. showed the <5 ng/cm2 adsorbed proteins from 100% blood plasma on gold surfacecoated with polyCB by surface-initiated-ATRP [64], Yang et al.showed the <0.3 ng/cm2 adsorbed proteins from 100% bloodplasma when the polyCB film thickness is optimized at w20 nm[43]. Yang et al. also demonstrated that the nano gold particles keepthe unchanged diameters in undiluted human blood serum bydynamic light scattering method [88]. With these three experi-ments, it is believed that the ultralow-fouling of the proteins ona material surface leads to ‘invisible’ nanoparticles in blood, whichcan avoid blood clot process and possibly escape from immuneresponse. These results also indicated that multiple proteins ina complex blood plasma require even more stringent surfaceproperties in nanoscale and macroscale to prevent protein

Page 8: Surface hydration: Principles and applications toward low-fouling

Fig. 8. (a) Illustration of the polyampholytes surface with perfect main structure andpossible surface defects. (b) Anti-fibrinogen adsorption on various hydrogels. Copy-right 2008 Wiley-VCH, Weinheim.

S. Chen et al. / Polymer 51 (2010) 5283e52935290

adsorption. The optimization of surface roughness, pattern, andthickness is critical for avoiding uncertain surface defects and thusminimizing protein adsorption [41].

Up to cell level, the resistance to microbial adhesion on materialsurfaces is also a property being examined since the adhesionof platelet or bacteria could lead to infection when medical devicesare inserted into or implanted in patients. Fig. 7b shows plateletadhesion level of 5.76�103/cm2, 4.71�103/cm2, and 9.57�103/cm2

on three polymer brushes of polySBMA, polyCBMA, and poly-OEGMA, respectively, in comparison to 4.57 � 103/cm2 plateletadhesion on the OEG-SAM. In blood, the presence of adherentplatelets usually shortens the plasma clotting time because plateletscatalyze undesirable thrombin formation. Blood clotting tests havealso shown that all these polymer brushes did not alter plasmasclotting time, especially forpolyCBMAbrush thatexhibits significantanticoagulant activity. PolyCB was further subject to the study oflong-term biofilm formation. The surface coatedwith 30 nmpolyCBcan greatly suppress the adhesion of Pseudomonas aeruginosa overten days, having less than 7% of P. aeruginosa coverage on theuntreated glass surface [89]. Control biofilm experiments ofOEG-SAM that has no protein adsorption and low platelet adhesionfail to resist long-term bacterial formation. Based on the experi-ments illustrated above, it is not necessarily true that the surface toresist protein adsorption is able to prevent platelet adhesion andbiofilm formation, because the latter involves more harsh andcomplex environmental conditions such as bacteria strength andgrowth medium. Meanwhile polyCBMA is a promising polymer forblood-contacting applications, as well as for environmentallyfriendly coatings for anti marine fouling, due to its excellent resis-tance to protein, platelet, and bacteria adhesion [41,64].

4.2. Polyampholytes

Besides three classic zwitterionic polybetaines such as MPC, SBand CB, polyampholytes is another big group of materials with theclosest structure to polybetaine polymers. Polyampholyte polymersare formed by a pair of separate mononers with two opposite

charge moieties respectively. The uniformity of charge distributionand charge neutrality of polymers is mostly achieved by 1:1homogeneous reaction mixture of the two oppositely chargedmonomers before co-polymerization. Although there is a possi-bility to form minor defects with two or more same charge con-nected together in ion-pair comonomer polymer (Fig. 8a), recentworks of ion-pair comonomers polymer [12,14] demonstrates thatsuch minor defects are negligible for protein adsorption. It isbelieved that the statistically distribution of charge groups arerather homogenous in nanometer scale. Possible rearrangement ofcharge groups through soft polymer backbone might also reducethe defects in charge distribution. Thus, it is possible to designa wide range of new nonfouling polyampholytes through mixedcharge monomers.

To demonstrate the importance of uniform charge distributionon the surface in the protein adsorption, different polyampholyteswith net and neutral surface charges were examined for theirresistance to protein (Fig. 8b). Similar to well-known nonfoulingsurfaces of OEG and SBMA, all mixed charge polyampholytes withhomogenous charge balance including 2-carboxyethyl acrylate(CA)-based mixtures with NH2 (NH2/CA), (Dimethylamino)ethylmethacrylate (DM/CA), 2-(Diethylamino)ethyl methacrylate (DE/CA), 2-(Methacryloyloxy)ethyl trimethylammonium chloride (TM/CA) and 3-Sulfopropyl methacrylate potassium salt (SA)-basedmixtures of NH2/SA, DM/SA, DE/SA, TM/SA exhibit extremely lowfibrinogen adsorption at a wide range of pH and ionic strength. Asexpected, deviation from charge neutrality can induce electrostaticinteractions between proteins and surface, leading to proteinadsorption. This fact indicates that charge balanced poly-ampholytes are equivalent to polybetaines with very stable androbust structures at nonphysiological conditions.

SAMs with mixed positively and negatively charge groups areanother format of mixed charge polymers. Even a 1:1 mixture ofcationic and anionic SAM displays excellent resistance to nonspe-cific protein adsorption, Chen et al. closely investigated thesezwitterionic SAMs. They show better resistance to proteinadsorption than the statistical copolymers [19,40] since themobility of thiols on gold surface improved uniformity of chargedistribution. The molecular simulation shows a 2-D crystallinestructure formed in the mixed SAMs of 1:1 mixture of TMA/SA onAu(111). The lattice of mixed TMA/SA SAMs revealed by atomicforce microscopy is about (5.2 � 0.2 �A � 5.2 � 0.2 �A) 60�, which issimilar to the same packing configuration as PC head groupsof membrane lipids. This indicated that the packing structure ofmixed SAMs is determined by strong chargeecharge interactions ofthe terminal groups rather than SeAu and chainechain interac-tions. Simulations also indicate that a balanced charge and mini-mized dipole are essential for enhancing hydration capacity at thetopmost layer via electrostatic interactions [52].

Besides the statistical copolymers from the two chargedcomponents, natural peptides with alternating negatively chargedGlu or Asp and positively charged Lys or Arg are another candidatesas biodegradable nonfouling materials since their final metabolizedproducts are natural amino acids. Based on the design principlelearned from mixed charged polyampholytes polymers and SAMs,results show that peptides with alternating Glu/Lys or Asp/Lys viaself-assembly are highly resistant to the nonspecific adsorption offibrinogen, lysozyme, and albumin (<0.3 ng/cm2) [90], comparableto PEG-based materials. This work demonstrates for the first timethe development of ultralow-fouling materials from naturalpeptides, providing a convenient way to synthesize super stealthnatural peptides by genetic engineering methods. Furthermore,these novel peptides mimic natural proteins and are biodegradableand nontoxicity through metabolism of peptides and may helpunderstand how various peptide sequences play their roles in

Page 9: Surface hydration: Principles and applications toward low-fouling

S. Chen et al. / Polymer 51 (2010) 5283e5293 5291

proteineprotein interactions and protein stability. Overall, stronghydration ability through ionic solvation of polyampholyte playsa key role in their nonfouling properties. A nanometer-scalehomogenous mixture of balanced charge groups provide anexcellent condition for charge groups to be fully ionized throughthe electrical field effects of the opposite ions, which override theweakening effect of weak acid or base in crowded conditions,which leads to the best condition for surface hydration. Thishomogeneity of opposite charge groups could avoid proteinadsorption induced local charge interaction.

5. Conclusions and perspectives

The development of nonfoulingmaterials and associated surfacecoating methods is critically important for many biomedicalapplications. Zwitterionic materials, especially for polyampholytepolymers, are very well positioned to play a role in the develop-ment of next-generation antifouling and antimicrobial materials.The advantages of nonfouling zwitterionic materials include thesimplicity of synthesis, ease of applicability, abundance of rawmaterials, and availability of functional groups. With the under-standing of the nonfouling mechanism and the design principlethat the nonfouling properties are mainly attributed to surfacehydration depending on the surface chemistry and physical packingof materials, future research direction could focus on the devel-opment of nonfouling materials with the general property andspecial function to meet challenge of the long-term stable perfor-mance in more complex condition in vivo [91]. Specially, (1) zwit-terionic functional groups that possess the maximum polarity areused to enhance surface hydration via electrostatic interactions; (2)functional groups, such as carboxyl or amino groups, will be a partof nonfouling materials. For example, dual functional polyCB canachieve both protein resistance and ligand immobilizability[83,85]; (3) apart from surface chemistry, novel synthesizing andsurface coating approaches need to be developed to optimize thephysical properties of the surface. The basic information is criticalto our understanding of the factors involved in nonfouling poly-mers and this information is the basis of efforts to design new andmore effective zwitterionic materials.

Acknowledgement

S.C. thanks for the financial support from the National NaturalScience Foundation of China (Nos. 20874085 and 20974095). J.Z.thanks the financial supports from NSF CAREER Award (CBET0952624), 3M Non-Tenured Faculty Award, and American ChemicalSociety Petroleum Research Fund (48188-G5).

References

[1] Ratner BD, Bryant SJ. Ann Rev Biomed Eng 2004;6(1):41e75.[2] Harris JM. Poly(ethylene glycol) chemistry - biotechnical and biomedical

applications. New York and London: Plenum Press; 1992.[3] Chapman RG, Ostuni E, Takayama S, Holmlin RE, Yan L, Whitesides GM. J Am

Chem Soc 2000;122(34):8303e4.[4] Ostuni E, Chapman RG, Holmlin RE, Takayama S, Whitesides GM. Langmuir

2001;17(18):5605e20.[5] Chen Y, Thayumanavan S. Langmuir 2009;25(24):13795e9.[6] Chang Y, Chu WL, Chen WY, Zheng J, Liu L, Ruaan RC, et al. J Biomed Mater

Res A 2010;93A(1):400e8.[7] Ma C, Hou Y, Liu S, Zhang G. Langmuir 2009;25(16):9467e72.[8] Harbers GM, Emoto K, Greef C, Metzger SW, Woodward HN, Mascali JJ, et al.

Chem Mater 2007;19(18):4405e14.[9] Prime KL, Whitesides GM. Science 1991;252(5009):1164e7.

[10] Li L, Chen S, Zheng J, Ratner BD, Jiang S. J Phys Chem B 2005;109(7):2934e41.[11] Zhang Z, Chen S, Chang Y, Jiang S. J Phys Chem B 2006;110(22):10799e804.[12] Bernards MT, Cheng G, Zhang Z, Chen S, Jiang S. Macromolecules 2008;41

(12):4216e9.

[13] Gao C, Li G, Xue H, Yang W, Zhang F, Jiang S. Biomaterials 2010;31(7):1486e92.

[14] Li GZ, Xue H, Gao CL, Zhang FB, Jiang SY. Macromolecules 2010;43(1):14e6.[15] Johnston EE, Bryers JD, Ratner BD. Langmuir 2004;21(3):870e81.[16] Zheng J, Li L, Tsao HK, Sheng YJ, Chen S, Jiang S. Biophys J 2005;89

(1):158e66.[17] Herrwerth S, Eck W, Reinhardt S, Grunze M. J Am Chem Soc 2003;125

(31):9359e66.[18] Chen S, Li L, Boozer CL, Jiang S. Langmuir 2000;16(24):9287e93.[19] Chen S, Yu F, Yu Q, He Y, Jiang S. Langmuir 2006;22(19):8186e91.[20] Jeon SI, Lee JH, Andrade JD, De Gennes PG. J Colloid Interface Sci 1991;142

(1):149e58.[21] Jeon SI, Andrade JD. J Colloid Interface Sci 1991;142(1):159e66.[22] Gong P, Grainger DW. Methods Mol Biol 2007;381:59e92.[23] Jiang S, Cao Z. Adv Mater 2010;22(9):920e32.[24] Tanaka M, Sackmann E. Nature 2005;437(7059):656e63.[25] Tanaka M, Motomura T, Ishii N, Shimura K, Onishi M, Mochizuki A, et al.

Polym Int 2000;49(12):1709e13.[26] TanakaM,Mochizuki A, Ishii N,Motomura T,HatakeyamaT. Biomacromolecules

2001;3(1):36e41.[27] Morisaku T, Watanabe J, Konno T, Takai M, Ishihara K. Polymer 2008;49

(21):4652e7.[28] Fick J, Steitz R, Leiner V, Tokumitsu S, Himmelhaus M, Grunze M. Langmuir

2004;20(10):3848e53.[29] Zolk M, Eisert F, Pipper J, Herrwerth S, Eck W, Buck M, et al. Langmuir

2000;16(14):5849e52.[30] Wang RY, Himmelhaus M, Fick J, Herrwerth S, Eck W, Grunze M. J Chem Phys

2005;122(16):164702.[31] Gil ES, Hudson SM. Prog Polym Sci 2004;29(12):1173e222.[32] Vanderah DJ, La H, Naff J, Silin V, Rubinson KA. J Am Chem Soc 2004;126

(42):13639e41.[33] Leckband D, Sheth S, Halperin A. J Biomater Sci Polym Ed 1999;10:1125e47.[34] Lutz JF, Akdemir Ö, Hoth A. J Am Chem Soc 2006;128(40):13046e7.[35] Lewis AL. Colloids Surf B Biointerfaces 2000;18(3e4):261e75.[36] Ishihara K, Nomura H, Mihara T, Kurita K, Iwasaki Y, Nakabayashi NJ. Biomed

Mater Res Part A 1998;39(2):323e30.[37] Iwasaki Y, Ijuin M, Mikami A, Nakabayashi N, Ishihara K. J Biomed Mater Res

1999;46(3):360e7.[38] Xu Y, Takai M, Ishihara K. Biomacromolecules 2008;10(2):267e74.[39] Kitano H, Mori T, Takeuchi Y, Tada S, Gemmei-Ide M, Yokoyama Y, et al.

Macromol Biosci 2005;5(4):314e21.[40] Holmlin RE, Chen X, Chapman RG, Takayama S, Whitesides GM. Langmuir

2001;17(9):2841e50.[41] Yang W, Chen S, Cheng G, Vaisocherova�c H, Xue H, Li W, et al. Langmuir

2008;24(17):9211e4.[42] Azzaroni O, Brown AA, Huck WTS. Angew Chem Int Ed 2006;45

(11):1770e4.[43] Yang W, Xue H, Li W, Zhang J, Jiang S. Langmuir 2009;25(19):11911e6.[44] Szleifer I. Biophys J 1997;72(2):595e612.[45] Fang F, Szleifer I. PNAS 2006;103(15):5769e74.[46] Wang RLC, Kreuzer HJ, Grunze M. J Phys Chem B 1997;101(47):9767e73.[47] Pertsin AJ, Grunze M. Langmuir 2000;16(23):8829e41.[48] Zheng J, Li L, Chen S, Jiang S. Langmuir 2004;20(20):8931e8.[49] Zheng J, He Y, Chen S, Li L, Bernards MT, Jiang S. J Chem Phys 2006;125

(17):174714e7.[50] He Y, Chang Y, Hower JC, Zheng J, Chen S, Jiang S. Phys Chem Chem Phys

2008;10(36):5539e44.[51] He Y, Hower J, Chen S, Bernards MT, Chang Y, Jiang S. Langmuir 2008;24

(18):10358e64.[52] Chen S, Zheng J, Li L, Jiang S. J Am Chem Soc 2005;127(41):14473e8.[53] Hower JC, Bernards MT, Chen S, Tsao HK, Sheng YJ, Jiang S. J Phys Chem B

2008;113(1):197e201.[54] Bailey Jr FE, Koleske JY. Poly(ethylene oxide). New York: Academic Press;

1984.[55] Pale-Grosdemange C, Simon ES, Prime KL, Whitesides GM. J Am Chem Soc

1991;113(1):12e20.[56] Prime KL, Whitesides GM. J Am Chem Soc 1993;115(23):10714e21.[57] Holland NB, Qiu YX, Ruegsegger M, Marchant RE. Nature 1998;392

(6678):799e801.[58] Ruegsegger MA, Marchant RE. J Biomed Mater Res 2001;56(2):159e67.[59] Shen M, Pan YV, Wagner MS, Hauch KD, Castner DG, Ratner BD, et al. J

Biomater Sci Polym Ed 2001;12:961.[60] Dalsin JL, Lin LJ, Tosatti S, Voros J, Textor M, Messersmith PB. Langmuir

2005;21(2):640e6.[61] Tan J, Brash JL. J Biomed Mater Res A 2008;85A(4):862e72.[62] Zhang Z, VaisocherovaH, Cheng G, YangW,Xue H, Jiang SY. Biomacromolecules

2008;9(10):2686e92.[63] Otsuka H, Satomi T, Ueno K, Mitamura Y, Tateishi T. Synthesis of poly-

pyridine-graft-PEG copolymer for long-term stability of nonfouling char-acter. In: Somiya S, Doyama M, editors. Transactions of the materialsresearch Society of Japan, vol. 31. Amsterdam: Elsevier Science Bv; 2006. p.627e31. No. 3.

[64] Ladd J, Zhang Z, Chen S, Hower JC, Jiang S. Biomacromolecules 2008;9(5):1357e61.

[65] Lusse S, Arnold K. Macromolecules 1996;29(12):4251e7.

Page 10: Surface hydration: Principles and applications toward low-fouling

S. Chen et al. / Polymer 51 (2010) 5283e52935292

[66] Olofsson G. J Phys Chem 1985;89(8):1473e7.[67] Klose G, Eisenblaetter S, Galle J, Islamov A, Dietrich U. Langmuir 1995;11

(8):2889e92.[68] Kingshott P, Wei J, Bagge-Ravn D, Gadegaard N, Gram L. Langmuir 2003;19

(17):6912e21.[69] Schulte VA, Di�cez M, Mo

ller M, Lensen MC. Biomacromolecules 2009;10(10):2795e801.

[70] Li L, Chen S, Jiang S. J Biomater Sci Polym Ed 2007;18:1415e27.[71] Ostuni E, Chapman RG, Liang MN, Meluleni G, Pier G, Ingber DE, et al.

Langmuir 2001;17(20):6336e43.[72] Herold DA, Keil K, Bruns DE. Biochem Pharmacol 1989;38(1):73e6.[73] Luk YY, Kato M, Mrksich M. Langmuir 2000;16(24):9604e8.[74] Chelmowski R, Ko

ster SD, Kerstan A, Prekelt A, Grunwald C, Winkler T, et al. JAm Chem Soc 2008;130(45):14952e3.

[75] Wyszogrodzka M, Haag R. Biomacromolecules 2009;10(5):1043e54.[76] Yeh PYJ, Kainthan RK, Zou YQ, Chiao M, Kizhakkedathu JN. Langmuir 2008;24

(9):4907e16.[77] Konradi R, Pidhatika B, Muhlebach A, Textort M. Langmuir 2008;24

(3):613e6.[78] Kainthan RK, Janzen J, Levin E, Devine DV, Brooks DE. Biomacromolecules

2006;7(3):703e9.[79] Vacheethasanee K, Marchant RE. J Biomed Mater Res A 2000;50(3):302e12.[80] Ladenheim H, Morawetz H. J Polym Sci 1957;26(113):251e4.[81] Ueda T, Oshida H, Kurita K, Ishihara K, Nakabayashi N. Polym J 1992;24

(11):1259e69.[82] Nakaya T, Li Y-J. Prog Polym Sci 1999;24(1):143e81.[83] Zhang Z, Chen S, Jiang S. Biomacromolecules 2006;7(12):3311e5.[84] Chang Y, Chen S, Zhang Z, Jiang S. Langmuir 2006;22(5):2222e6.[85] Vaisocherova�c H, Yang W, Zhang Z, Cao Z, Cheng G, Piliarik M, et al. Anal

Chem 2008;80(20):7894e901.[86] Ma H, Hyun J, Stiller P, Chilkoti A. Adv Mater 2004;16(4):338e41.[87] Chen SF, Jiang SY. Adv Mater 2008;20(2):335.[88] YangW,ZhangL,WangS,WhiteAD, JiangS.Biomaterials2009;30(29):5617e21.[89] Cheng G, Li G, Xue H, Chen S, Bryers JD, Jiang S. Biomaterials 2009;30

(28):5234e40.[90] Chen SF, Cao ZQ, Jiang SY. Biomaterials 2009;30(29):5892e6.[91] Laughlin RG. Langmuir 1991;7(5):842e7.[92] D’Sa RA, Meenan BJ. Langmuir 2009;26(3):1894e903.[93] Zoulalian V, Zurcher S, Tosatti S, Textor M, Monge S, Robin JJ. Langmuir

2010;26(1):74e82.[94] Schuler M, Hamilton DW, Kunzler TP, Sprecher CM, de Wild M, Brunette DM,

et al. J Biomed Mater Res B Appl Biomater 2009;91B(2):517e27.[95] Maddikeri RR, Tosatti S, Schuler M, Chessari S, Textor M, Richards RG, et al. J

Biomed Mater Res A 2008;84A(2):425e35.[96] Tugulu S, Klok HA. Macromol Symp 2009;279:103e9.[97] Tugulu S, Klok HA. Biomacromolecules 2008;9(3):906e12.[98] Wang J, Gibson MI, Barbey R, Xiao SJ, Klok HA. Macromol Rapid Commun

2009;30(9e10):845e50.[99] Hucknall A, Simnick AJ, Hill RT, Chilkoti A, Garcia A, Johannes MS, et al.

Biointerphases 2009;4(2):FA50e7.[100] Cheng G, Zhang Z, Chen SF, Bryers JD, Jiang SY. Biomaterials 2007;28

(29):4192e9.[101] Hucknall A, Kim DH, Rangarajan S, Hill RT, Reichert WM, Chilkoti A. Adv

Mater 2009;21(19):1968e71.[102] Zhang Z, Zhang M, Chen SF, Horbetta TA, Ratner BD, Jiang SY. Biomaterials

2008;29(32):4285e91.[103] Tan J, Brash JL. J Biomed Mater Res Part A 2009;90A(1):196e204.[104] Tan J, McClungWG, Brash JL. J Biomed Mater Res Part A 2008;85A(4):873e80.[105] Tan J, Brash JL. J Appl Polym Sci 2008;108(3):1617e28.[106] Sheu MS, Hoffman AS, Feijen J. J Adhes Sci Technol 1992;6(9):995e1009.[107] Yoshimoto K, Hirase T, Nemoto S, Hatta T, Nagasaki Y. Langmuir 2008;24

(17):9623e9.[108] Hyun J, Jang H, Kim K, Na K, Tak T. J Memb Sci 2006;282(1e2):52e9.[109] Sharma S, Johnson RW, Desai TA. Langmuir 2004;20(2):348e56.[110] Dalsin JL, HuBH, Lee BP,Messersmith PB. J AmChemSoc 2003;125(14):4253e8.[111] Pidhatika B, Moller J, Vogel V, Konradi R. Chimia 2008;62(4):264e9.[112] Zhang M, Horbett TA. J Biomed Mater Res Part A 2009;89A(3):791e803.[113] Cao L, Chang M, Lee CY, Castnet DG, Sukavaneshvar S, Ratner BD, et al.

J Biomed Mater Res Part A 2007;81A(4):827e37.[114] Shen MC, Wagner MS, Castner DG, Ratner BD, Horbett TA. Langmuir 2003;19

(5):1692e9.[115] Shen M, Martinson L, Wagner MS, Castner DG, Ratner BD, Horbett TA.

J Biomater Sci Polym Ed 2002;13(4):367e90.[116] Martwiset S, Koh AE, Chen W. Langmuir 2006;22(19):8192e6.

[117] McArthur SL, McLean KM, Kingshott P, St John HAW, Chatelier RC,Griesser HJ. Colloids Surf B Biointerfaces 2000;17(1):37e48.

[118] Mrabet B, Nguyen MN, Majbri A, Mahouche S, Turmine M, Bakhrouf A, et al.Surf Sci 2009;603(16):2422e9.

[119] Ishihara K, Ishikawa E, Iwasaki Y, Nakabayashi N. J. Biomater. Sci. Polym. Ed.1999;10(10):1047e61.

[120] Amanda A, Mallapragada SK. Biotechnol Prog 2001;17(5):917e23.[121] Chapman RG, Ostuni E, Liang MN, Meluleni G, Kim E, Yan L, et al. Langmuir

2001;17(4):1225e33.[122] Vaisocherova H, Zhang Z, Yang W, Cao ZQ, Cheng G, Taylor AD, et al. Biosens

Bioelectron 2009;24(7):1924e30.[123] Zhang Z, Cheng G, Carr LR, Vaisocherova H, Chen SF, Jiang SY. Biomaterials

2008;29(36):4719e25.[124] Li GZ, Cheng G, Xue H, Chen SF, Zhang FB, Jiang SY. Biomaterials 2008;29

(35):4592e7.[125] Li GZ, Xue H, Cheng G, Chen SF, Zhang FB, Jiang SY. J Phys Chem B 2008;112

(48):15269e74.[126] Cheng G, Xite H, Zhang Z, Chen SF, Jiang SY. Angew Chem Int Ed 2008;47

(46):8831e4.[127] Ishihara K, Hasegawa T, Watanabe J, Iwasaki Y. Artif Organs 2002;26

(12):1014e9.[128] Hasegawa T, Iwasaki Y, Ishihara K. J Biomed Mater Res 2002;63

(3):333e41.[129] Chen SF, Liu LY, Jiang SY. Langmuir 2006;22(6):2418e21.[130] Yi SJ, Yuk JS, Jung SH, Zhavnerko GK, Kim YM, Ha KS. Mol Cell 2003;15

(3):333e40.

Shengfu Chen is a professor at the State KeyLaboratory of Chemical Engineering andDepartment of Chemical and Biological Engi-neering, Zhejiang University (ZJU), China. Hereceived his Ph.D. from Shanghai Institute ofApplied Physics, Chinese Academy of Sciencesin 1998, was a post-doctoral fellow at theUniversity of Washington, USA, before joiningZJU in 2007. He has authored 67 scientificpapers. His current research focuses oninvestigating interfacial phenomena andproperties of biological and chemical systemsat nanometer scale and developing newmaterials and coatings for biomedical andengineering applications, such as biocompat-ible materials, drug delivery, membraneseparation, and nanobiodevices.

Lingyan Li received her dual Ph.Ds in Chem-ical Engineering and Nanotechnology fromthe University of Washington (2004). She iscurrently a post-doctoral fellow in the RenalNanotechnology Lab at the Cleveland Clinic.Her current research focuses on the develop-ment of novel biomaterials and associatedsurface modification approaches to enhancehemocompatibility and minimize biofoulingof MEMS-based membranes for an implant-able bio-artificial kidney. She has authoredover 20 scientific papers.

Page 11: Surface hydration: Principles and applications toward low-fouling

S. Chen et al. / Polymer 51 (2010) 5283e5293 5293

Chao Zhao is currently pursuing his Ph.D.degree in the Department of Chemical andBiomolecular Engineering at the University ofAkron under the guidance of Prof. Jie Zheng.He received his B.E. in Polymer Materials andEngineering at Harbin Institute of Technology,China (2004) and M.E. in Chemical Engi-neering at Zhejiang University, China (2006),respectively. His current research focuses onthe synthesis and characterization ofbiocompatible and functionlizable polymersfor biomedical applications.

Jie Zheng is Assistant Professor of Chemicaland Biomolecular Engineering at the Univer-sity of Akron (UA). Zheng earned his B.S. andPh.D. in Chemical Engineering at ZhejiangUniversity, China (1995) and University ofWashington (2005), respectively, and thenjoined the National Cancer Institute of theNIH as a scientist. In 2007, he acceptedAssistant Professor position at the UA. Zheng'scurrent research focuses on protein (mis)folding, protein aggregation, and biomolec-ular interfaces, with particular interests inprotein interactions with biological cellmembranes and artificial surfaces towardsmolecular understanding of protein struc-ture–function relationship. He has authoredover 40 scientific papers. His work has beenrecognized by US National Science Founda-tion CAREER Award and 3 M Non-TenuredFaculty Award.