Hostname: page-component-8448b6f56d-t5pn6 Total loading time: 0 Render date: 2024-04-23T07:54:05.587Z Has data issue: false hasContentIssue false

Modeling and simulation of protein–surface interactions: achievements and challenges

Published online by Cambridge University Press:  29 January 2016

Musa Ozboyaci*
Affiliation:
Heidelberg Institute for Theoretical Studies (HITS), Schloss-Wolfsbrunnenweg 35, 69118 Heidelberg, Germany Heidelberg Graduate School of Mathematical and Computational Methods for the Sciences (HGS MathComp), Heidelberg University, Im Neuenheimer Feld 368, 69120 Heidelberg, Germany
Daria B. Kokh
Affiliation:
Heidelberg Institute for Theoretical Studies (HITS), Schloss-Wolfsbrunnenweg 35, 69118 Heidelberg, Germany
Stefano Corni
Affiliation:
Centro S3, CNR Instituto Nanoscienze, via Campi 213/a, 41125 Modena, Italy
Rebecca C. Wade*
Affiliation:
Heidelberg Institute for Theoretical Studies (HITS), Schloss-Wolfsbrunnenweg 35, 69118 Heidelberg, Germany Zentrum für Molekulare Biologie der Universität Heidelberg, DKFZ-ZMBH Allianz, Im Neuenheimer Feld 282, 69120 Heidelberg, Germany Interdisciplinary Center for Scientific Computing (IWR), Heidelberg University, 69120 Heidelberg, Germany
*
*Authors for correspondence: Musa Ozboyaci, Heidelberg Institute for Theoretical Studies (HITS), Schloss-Wolfsbrunnenweg 35, 69118 Heidelberg, Germany; Heidelberg Graduate School of Mathematical and Computational Methods for the Sciences (HGS MathComp), Heidelberg University, Im Neuenheimer Feld 368, 69120 Heidelberg, Germany & Rebecca C. Wade, Heidelberg Institute for Theoretical Studies (HITS), Schloss-Wolfsbrunnenweg 35, 69118 Heidelberg, Germany; Zentrum für Molekulare Biologie der Universität Heidelberg, DKFZ-ZMBH Allianz, Im Neuenheimer Feld 282, 69120 Heidelberg, Germany; Interdisciplinary Center for Scientific Computing (IWR), Heidelberg University, 69120 Heidelberg, Germany. Tel.:+49-6221-533-247; Emails: musa.oezboyaci@h-its.org, rebecca.wade@h-its.org
*Authors for correspondence: Musa Ozboyaci, Heidelberg Institute for Theoretical Studies (HITS), Schloss-Wolfsbrunnenweg 35, 69118 Heidelberg, Germany; Heidelberg Graduate School of Mathematical and Computational Methods for the Sciences (HGS MathComp), Heidelberg University, Im Neuenheimer Feld 368, 69120 Heidelberg, Germany & Rebecca C. Wade, Heidelberg Institute for Theoretical Studies (HITS), Schloss-Wolfsbrunnenweg 35, 69118 Heidelberg, Germany; Zentrum für Molekulare Biologie der Universität Heidelberg, DKFZ-ZMBH Allianz, Im Neuenheimer Feld 282, 69120 Heidelberg, Germany; Interdisciplinary Center for Scientific Computing (IWR), Heidelberg University, 69120 Heidelberg, Germany. Tel.:+49-6221-533-247; Emails: musa.oezboyaci@h-its.org, rebecca.wade@h-its.org
Rights & Permissions [Opens in a new window]

Abstract

Understanding protein–inorganic surface interactions is central to the rational design of new tools in biomaterial sciences, nanobiotechnology and nanomedicine. Although a significant amount of experimental research on protein adsorption onto solid substrates has been reported, many aspects of the recognition and interaction mechanisms of biomolecules and inorganic surfaces are still unclear. Theoretical modeling and simulations provide complementary approaches for experimental studies, and they have been applied for exploring protein–surface binding mechanisms, the determinants of binding specificity towards different surfaces, as well as the thermodynamics and kinetics of adsorption. Although the general computational approaches employed to study the dynamics of proteins and materials are similar, the models and force-fields (FFs) used for describing the physical properties and interactions of material surfaces and biological molecules differ. In particular, FF and water models designed for use in biomolecular simulations are often not directly transferable to surface simulations and vice versa. The adsorption events span a wide range of time- and length-scales that vary from nanoseconds to days, and from nanometers to micrometers, respectively, rendering the use of multi-scale approaches unavoidable. Further, changes in the atomic structure of material surfaces that can lead to surface reconstruction, and in the structure of proteins that can result in complete denaturation of the adsorbed molecules, can create many intermediate structural and energetic states that complicate sampling. In this review, we address the challenges posed to theoretical and computational methods in achieving accurate descriptions of the physical, chemical and mechanical properties of protein-surface systems. In this context, we discuss the applicability of different modeling and simulation techniques ranging from quantum mechanics through all-atom molecular mechanics to coarse-grained approaches. We examine uses of different sampling methods, as well as free energy calculations. Furthermore, we review computational studies of protein–surface interactions and discuss the successes and limitations of current approaches.

Type
Review
Copyright
Copyright © Cambridge University Press 2016 

1. Introduction

Protein–inorganic surface interactions have gained increasing attention owing to their widespread occurrence in nature, and their broad range of applications in nanobiotechnology (Choi et al. Reference Choi, Kim, Kim and Min2009; Hill et al. Reference Hill, Vega and Mirkin2007; Hu et al. Reference Hu, Das, Hecht and Scoles2005; Jackson et al. Reference Jackson, Omanovic and Roscoe2000; Laera et al. Reference Laera, Ceccone, Rossi, Gilliland, Hussain, Siligardi and Calzolai2011; Manecka et al. Reference Manecka, Labrash, Rouxel, Dubot, Lalevée, Andaloussi, Renard, Langlois and Versace2014; Millo et al. Reference Millo, Pandelia, Utesch, Wisitruangsakul, Mroginski, Lubitz, Hildebrandt and Zebger2009; Park et al. Reference Park, Lytton-Jean, Lee, Weigand, Schatz and Mirkin2008; Qin et al. Reference Qin, Banholzer, Millstone and Mirkin2007a, Reference Qin, Banholzer, Xu, Huang and Mirkinb; Slocik et al. Reference Slocik, Govorov and Naik2011; Xu et al. Reference Xu, Cao, Svec and Fréchet2010). Adhesion of proteins on solid substrates is utilized by many organisms, e.g. sea urchins make use of the adsorption of matrix proteins to specific crystal patches on endoskeletal calcite surfaces (Wilt, Reference Wilt1999), and has even been evolutionarily adapted to enable some organisms to live in specific habitats, e.g. for the adhesion of mussels to mineral rocks (Yu et al. Reference Yu, Wei, Danner, Ashley, Israelachvili and Waite2011). Humans have long used inorganic materials that make direct interactions with proteins. For example, gold crowns as dental prosthetics date back to the ancient Etruscan civilization (Demann et al. Reference Demann, Stein and Haubenreich2005), and man-made nanoparticles were used as pigments in ointments by the ancient Romans (Casals et al. Reference Casals, Bastus, Vázquez, Varon, Comenge and Puntes2008). However, it is only quite recently that advances in science and technology have enabled the production of completely new or engineered surfaces and hence allowed new applications. For example, the remarkable structural and mechanical properties of graphene, isolated in 2004 (Novoselov et al. Reference Novoselov2004), have drawn increasing attention to carbon allotropes and catalyzed further research into their interactions with proteins, motivated by various biotechnological applications, including efficient biosensors (Alava et al. Reference Alava, Mann, Théodore, Benitez, Dichtel, Parpia and Craighead2013).

Computational modeling and simulation of biomolecules can help scientists to unravel the mechanisms of molecular-level events and predict the behavior of complex systems at a level of detail that cannot be directly measured in experiments. Since the development of the first modeling and simulation methods for complex molecules, computational research in the field has expanded enormously. Their importance is shown by the Nobel Prize in Chemistry being awarded in 2013 to Martin Karplus, Michael Levitt and Arieh Warshel “for the development of multiscale models for complex chemical systems”. Proteins, nucleic acids, lipids and their interactions in aqueous environments have been widely studied computationally by means of molecular mechanics (MM) force-fields (FFs) (Brooks et al. Reference Brooks, Bruccoleri, Olafson, States, Swaminathan and Karplus1983; Cornell et al. Reference Cornell, Cieplak, Bayly, Gould, Merz, Ferguson, Spellmeyer, Fox, Caldwell and Kollman1995; Oostenbrink et al. Reference Oostenbrink, Villa, Mark and Van Gunsteren2004) developed and tailored specifically for these types of molecules. However, many of these FFs fall short in reproducing the properties of protein-inorganic surface systems. To alleviate this problem, many useful models (Heinz et al. Reference Heinz, Jha, Luettmer-Strathmann, Farmer and Naik2011; Iori & Corni, Reference Iori and Corni2008; Kokh et al. Reference Kokh, Corni, Winn, Hoefling, Gottschalk and Wade2010) and FF parameters (De la Torre et al. Reference De la Torre, Hernández Cifre, Ortega, Schmidt, Fernandes, Pérez Sánchez and Pamies2009; Heinz et al. Reference Heinz, Lin, Kishore Mishra and Emami2013; Iori et al. Reference Iori, Di Felice, Molinari and Corni2009; Schneider & Colombi Ciacchi, Reference Schneider and Colombi Ciacchi2010; Wright et al. Reference Wright, Rodger, Corni and Walsh2013a) for material surfaces have been introduced that have been designed to be compatible with FFs for biomolecular systems. These FFs are still rather young and their improvement is an area of active research.

A number of reviews addressing different aspects of protein–surface interaction studies have been published previously. Several of these provide a general overview of the adsorption of proteins at solid surfaces (Cohavi et al. Reference Cohavi, Reichmann, Abramovich, Tesler, Bellapadrona, Kokh, Wade, Vaskevich, Rubinstein and Schreiber2010; Costa et al. Reference Costa, Garrain and Baaden2013; Horbett & Brash, Reference Horbett and Brash1995; Rabe et al. Reference Rabe, Verdes and Seeger2011; Qu et al. Reference Qu, He, Lin, Sha, Shi, Lu and Xu2013), whereas others focus on more specific aspects such as the determination of the adsorption kinetics of protein–surface binding by weakly bound mobile precursor states (Garland et al. Reference Garland, Shen and Zhu2012), and adsorption on various different surface types, such as metallic surfaces (Tomba et al. Reference Tomba, Colombi Ciacchi and De Vita2009; Vallee et al. Reference Vallee, Humblot and Pradier2010), polymer surfaces (Hahm, Reference Hahm2014; Wei et al. Reference Wei, Becherer, Angioletti-Uberti, Dzubiella, Wischke, Neffe, Lendlein, Ballauff and Haag2014) and protein repellent surfaces (Szott & Horbett, Reference Szott and Horbett2011). The physico-chemical properties of nanomaterials, and their applications in medicine, biology and biotechnology, have also been reviewed in several papers (see, e.g. Ansari & Husain, Reference Ansari and Husain2012; Dufort et al. Reference Dufort, Sancey and Coll2012; Khlebtsov & Dykman, Reference Khlebtsov and Dykman2010; Mahmoudi et al. Reference Mahmoudi, Lynch, Ejtehadi, Monopoli, Bombelli and Laurent2011; Mahon et al. Reference Mahon, Salvati, Baldelli Bombelli, Lynch and Dawson2012; Mandal et al. Reference Mandal, Nasrolahi Shirazi and Parang2014; Salata, Reference Salata2004; Saptarshi et al. Reference Saptarshi, Duschl and Lopata2013; Shemetov et al. Reference Shemetov, Nabiev and Sukhanova2012). Various aspects of the computational methods employed in modeling and simulation of protein–surface interactions are addressed in other reviews, including issues in computational modeling of peptide–surface interactions (Di Felice & Corni, Reference Di Felice and Corni2011), problems with these simulation techniques (Latour, Reference Latour2008) and approaches to multiscale modeling of soft matter that are transferable to protein-surface systems (Praprotnik et al. Reference Praprotnik, Site and Kremer2008).

This review provides a discussion of the computational models and simulation techniques that have been used in studies of protein–surface interactions. Due to the broad range of models used in these studies, only models of protein–surface interactions based on chemical structures are discussed in this review and, therefore, more abstract models to describe these interactions, such as that developed by Oberle et al. (Reference Oberle, Yigit, Angioletti-Uberti, Dzubiella and Ballauff2015) for the description of competitive adsorption of proteins to nanoparticles, are not discussed further. A brief introduction to various types of material surfaces is provided and some of the properties that need attention from a modeling point of view are discussed. We further give an overview of the different modeling, sampling and free energy calculation techniques employed in recent studies. We discuss the properties that can be computed by these methods and how they can assist and complement experiments. Some of the important findings from applications of these methods are reviewed, and drawbacks and shortcomings of the available techniques are discussed. The paper concludes with a discussion of the general limitations and future directions of the field.

2. Which types of surfaces can be modeled?

The interactions of proteins with inorganic materials are determined not only by the properties of the proteins, but also by the chemical composition, molecular structure, size and shape of the material. Inorganic surfaces possess distinct physico-chemical properties, such as reactivity towards different compounds, material stability and specific adsorption characteristics for different adsorbents. These properties allow different types of surfaces to be employed for different applications, e.g. implantation or chromatography. Understanding the basis of these physico-chemical properties usually requires atomic-level investigation of the surfaces. The types of material surface modeled commonly in computational studies of protein–surface interactions are elemental metals and alloys, metal oxides and minerals, self-assembled monolayers (SAMs), polymers and carbon allotrope surfaces, see Fig. 1.

Fig. 1. Simulations of proteins with different types of surface: (a) lysozyme on a polyethylene surface (reprinted with permission from (Wei et al. Reference Wei, Carignano and Szleifer2011). Copyright (2011) American Chemical Society), (b) the MRKDV peptide on a bare silver surface (adapted with permission from (Aliaga et al. Reference Aliaga, Ahumada, Sepúlveda, Gomez-Jeria, Garrido, Weiss-López and Campos-Vallette2011). Copyright (2011) American Chemical Society), (c) RAD16II on a rutile surface (reprinted with permission from (Monti, Reference Monti2007). Copyright (2007) American Chemical Society), and (d) NiFe hydrogenase on a SAM surface (reprinted with permission from (Utesch et al. Reference Utesch, Millo, Castro, Hildebrandt, Zebger and Mroginski2013). Copyright (2013) American Chemical Society).

2.1 Elemental metals and alloys

Protein–metal surface interactions can be studied with experimental techniques such as atomic force microscopy (AFM) (Binnig et al., Reference Binnig, Quate and Gerber1986), surface plasmon resonance (SPR) (Jönsson et al. Reference Jönsson, Fägerstam, Ivarsson, Johnsson, Karlsson, Lundh, Löfås, Persson, Roos and Rönnberg1991) and localized SPR (Stuart et al. Reference Stuart, Haes, Yonzon, Hicks and Van Duyne2005). Due to their chemical inertness and unique optical properties (Jain et al. Reference Jain, Huang, El-Sayed and El-Sayed2008), the noble metals, gold and silver, are the most commonly used metals employed as probes or sensors in these techniques. Along with well-known applications of metal surfaces, such as biosensors and implants (Liu et al. Reference Liu, Chu and Ding2004), metal surfaces are also used in bioelectronics as electrodes as they allow controlled exchange of electrons with metalloproteins immobilized on them (Alessandrini et al. Reference Alessandrini, Salerno, Frabboni and Facci2005; Andolfi et al. Reference Andolfi, Bruce, Cannistraro, Canters, Davis, Hill, Crozier, Verbeet, Wrathmell and Astier2004). The interactions of proteins with bare metal surfaces have been the subject of many computational studies, which cover a wide range of different elemental and alloy surfaces, such as Cu(100) (Chen & Wang, Reference Chen, Wang, Huang and Gao2010), Au(111) (Bizzarri, Reference Bizzarri2006; Hoefling et al. Reference Hoefling, Monti, Corni and Gottschalk2011; Siwko & Corni, Reference Siwko and Corni2013; Venkat et al. Reference Venkat, Corni and Di Felice2007; Zanetti-Polzi et al. Reference Zanetti-Polzi, Daidone, Bortolotti and Corni2014), Au(100) (Hagiwara et al. Reference Hagiwara, Sakiyama and Watanabe2009), Au nanoparticle (Todorova et al. Reference Todorova, Chiappini, Mager, Simona, Patel, Stevens and Yarovsky2014), Fe (Zhang et al. Reference Zhang, Chen, Feng, Ding, Zhou and Jia2009b), Ni (Yang & Zhao, Reference Yang and Zhao2006), Pd (Coppage et al. Reference Coppage, Slocik, Briggs, Frenkel, Heinz, Naik and Knecht2011), Pt (Kantarci et al. Reference Kantarci, Tamerler, Sarikaya, Haliloglu and Doruker2005), Ag (Aliaga et al. Reference Aliaga, Ahumada, Sepúlveda, Gomez-Jeria, Garrido, Weiss-López and Campos-Vallette2011; Ghosh et al. Reference Ghosh, Jana and Guchhait2012) and steel (Imamura et al. Reference Imamura, Kawasaki, Awadzu, Sakiyama and Nakanishi2003).

2.2 Oxides and minerals

Metal surfaces (excluding noble metals such as gold, platinum and palladium) are oxidized when exposed to water or air, forming metal oxides that are very common in the Earth's crust. Due to their good mechanical stability, catalytic properties and biocompatibility (Andreescu et al. Reference Andreescu, Ornatska, Erlichman, Estevez, Leiter and Matijević2012), metal oxides and minerals are used in a wide range of applications that include fabrication of biomaterials (Whaley et al. Reference Whaley, English, Hu, Barbara and Belcher2000), cellular delivery of drugs and biomolecules (Kievit & Zhang, Reference Kievit and Zhang2011; Xu et al. Reference Xu, Zeng, Lu and Yu2006), tissue engineering (Shin et al. Reference Shin, Jo and Mikos2003) and proteomics (Sugiyama et al. Reference Sugiyama, Masuda, Shinoda, Nakamura, Tomita and Ishihama2007). Computational studies to investigate the interactions of oxide and mineral surfaces with proteins or peptides have mostly been carried out for different forms of titanium dioxide, such as rutile and anatase (Carravetta et al. Reference Carravetta, Monti and Zhang2009; Kang et al. Reference Kang, Li, Tu, Wang and Ågren2010; Köppen et al. Reference Köppen, Bronkalla and Langel2008; Monti, Reference Monti2007; Monti et al. Reference Monti, Carravetta, Battocchio, Iucci and Polzonetti2008; Sun et al. Reference Sun, Han, Lindgren, Shen and Laaksonen2014a; Wu et al. Reference Wu, Chen, Skelton, Cummings and Zheng2013), silicon dioxides (Chen et al. Reference Chen, Su, Neoh and Choe2009a; Nonella & Seeger, Reference Nonella and Seeger2008; Patwardhan et al. Reference Patwardhan, Emami, Berry, Jones, Naik, Deschaume, Heinz and Perry2012; Rimola et al. Reference Rimola, Sodupe and Ugliengo2009; Tosaka et al. Reference Tosaka, Yamamoto, Ohdomari and Watanabe2010), calcite (Wierzbicki et al. Reference Wierzbicki, Sikes, Madura and Drake1994) and mica (Kang et al. Reference Kang, Huynh, Xia, Zhang, Fang, Wei and Zhou2013).

2.3 Self-assembled monolayers

SAMs are thin films that coat surfaces by spontaneous adsorption of organic molecules that form ordered molecular assemblies. Typically, in a SAM, molecules are chemisorbed onto a surface substrate through their reactive head groups and are, therefore, very stable. SAMs can be categorized into two groups according to their head-group type: thiol-based and silate-based (Schreiber, Reference Schreiber2004). The head-group is attached to a tail group that, by following a fast adsorption phase (seconds), undergoes a slow reorganization (hours), in which interactions with other tail groups increase and the packing is improved (Love et al. Reference Love, Estroff, Kriebel, Nuzzo and Whitesides2005). The tail group can be functionalized with small chemical groups or large molecules, such as peptides. These, together with the length of the tail group, allow the physico-chemical interfacial properties, in particular, hydrophobicity and ionization to be adjusted according to the desired application. Alkanethiols are the most common type of SAM. They have the chemical formula S-(CH2)n-R, where R stands for a functional group, such as -CH3, -COOH, -NH2 or -OH.

Bare metals and oxide surfaces are prone to non-specific adsorption of proteins and other organic molecules. These adsorption processes may result in undesirable agglomeration of adsorbates on the surfaces. The self-assembly of organic molecules on a metal surface in a SAM creates a physical barrier between the surface and the adsorbates, acting as an electrical insulator and passivating the surface atoms (Love et al. Reference Love, Estroff, Kriebel, Nuzzo and Whitesides2005). The properties of SAMs have been reviewed in (Chaki & Vijayamohanan, Reference Chaki and Vijayamohanan2002; Gooding et al. Reference Gooding, Mearns, Yang and Liu2003; Love et al. Reference Love, Estroff, Kriebel, Nuzzo and Whitesides2005; Senaratne et al. Reference Senaratne, Andruzzi and Ober2005). Modeling of protein–SAM interactions has been reported, mostly for alkanethiol SAMs, in (Alvarez-Paggi et al. Reference Alvarez-Paggi, Martín, DeBiase, Hildebrandt, Martí and Murgida2010; Hsu et al. Reference Hsu, Sheu and Tsay2008; O'Mahony et al. Reference O'Mahony, O'Dwyer, Nijhuis, Greer, Quinn and Thompson2013; Sun et al. Reference Sun, Welsh and Latour2005; Utesch et al. Reference Utesch, Millo, Castro, Hildebrandt, Zebger and Mroginski2013; Wang et al. Reference Wang, Zhao, Zhao, Wang, Yang, Yu and Zheng2010a, Reference Wang, Zhao, Yu, Zhao, Li and Zhengb; Xie et al. Reference Xie, Liu and Zhou2012) and peptide–SAM interactions have been modeled by Nowinski et al. (Reference Nowinski, Sun, White, Keefe and Jiang2012).

2.4 Polymers

Polymer surfaces have attracted much attention in the nanotechnology field due to their mechanical stability, low cost and their wide applicability (Nie & Kumacheva, Reference Nie and Kumacheva2008).

Particularly synthetic polymer-based biomaterials, due to their non-fouling properties, are currently being investigated intensively for applications in controlled drug delivery (Hoffman, Reference Hoffman2008), in highly sensitive biosensors (Anker et al. Reference Anker, Hall, Lyandres, Shah, Zhao and Van Duyne2008), and in bioelectronics (Senaratne et al. Reference Senaratne, Andruzzi and Ober2005). Nanostructured polymer materials used for bio-related and medicinal research include electrostatic polymer brushes, micelles, layer-by-layer deposition and thin films (Stuart et al. Reference Stuart, Huck, Genzer, Müller, Ober, Stamm, Sukhorukov, Szleifer, Tsukruk, Urban, Winnik, Zauscher, Luzinov and Minko2010). Polymer brushes are surface modifiers that share many properties in common with SAMs (Senaratne et al. Reference Senaratne, Andruzzi and Ober2005). They are prepared by grafting polymers of the same or varying kinds on surfaces forming homopolymer and mixed brushes, respectively. Polymeric micelles are formed through self-assembly of amphiphilic copolymers and typically have a diameter of size 30–50 nm (Otsuka et al. Reference Otsuka, Nagasaki and Kataoka2003; Stuart et al. Reference Stuart, Huck, Genzer, Müller, Ober, Stamm, Sukhorukov, Szleifer, Tsukruk, Urban, Winnik, Zauscher, Luzinov and Minko2010).

These micelles are particularly important due to their lower critical micelle concentration (CMC), higher stability and slower rate of dissociation than surfactant micelles. These properties have allowed polymeric micelles to act as effective cancer treatment tools with high drug deposition at the target site (Otsuka et al. Reference Otsuka, Nagasaki and Kataoka2003). Polymer surfaces and their applications have been reviewed by (Barbey et al. Reference Barbey, Lavanant, Paripovic, Schüwer, Sugnaux, Tugulu and Klok2009; Kim et al. Reference Kim, Khang and Lee2008; Nie et al. Reference Nie, Petukhova and Kumacheva2010; Otsuka et al. Reference Otsuka, Nagasaki and Kataoka2003; Senaratne et al. Reference Senaratne, Andruzzi and Ober2005; Stuart et al. Reference Stuart, Huck, Genzer, Müller, Ober, Stamm, Sukhorukov, Szleifer, Tsukruk, Urban, Winnik, Zauscher, Luzinov and Minko2010). The limited number of computational studies of protein-polymer surfaces to date have been carried out for polymer types including polystyrene, polyethylene and polydimethylsiloxane (Boughton et al. Reference Boughton, Andricioaei and Chen2010; Jeyachandran et al. Reference Jeyachandran, Mielczarski, Rai and Mielczarski2009; Liu et al. Reference Liu, Wu, Feng, Shao and Cai2012; Lu et al. Reference Lu, Lee and Park1992; O'Brien et al. Reference O'Brien, Stuart, Bruce and Latour2008; Raffaini & Ganazzoli, Reference Raffaini and Ganazzoli2007; Zhang et al. Reference Zhang, Zhao and Sun2009a; Wei et al. Reference Wei, Carignano and Szleifer2011).

2.5 Carbon allotropes

Pure carbon may exist in a number of different allotropes. Owing to their unique thermal, electrical, chemical and mechanical properties, carbon-based nanomaterials have been the subject of numerous applications in analytical chemistry (Scida et al. Reference Scida, Stege, Haby, Messina and García2011). These applications mostly focus on carbon nanomaterials with sp2-carbon bonding, such as fullerenes, carbon nanotubes (CNT), graphene and graphite. This is due to the extremely high surface areas of fullerenes and CNTs relative to their size, which makes them suitable for design as highly efficient drug carriers. Furthermore, the excellent electrical properties of graphene and CNTs make them suitable for biosensor applications (Liu & Liang, Reference Liu and Liang2012). In all the four materials, the carbon atoms make three chemical bonds with other carbons in the surface-plane with delocalized π electron clouds in the direction perpendicular to the surface (Scida et al. Reference Scida, Stege, Haby, Messina and García2011). This configuration makes the mutual van der Waals interactions between CNTs very strong and hence leads them to be very hydrophobic (Guldi et al. Reference Guldi, Rahman, Sgobba and Ehli2006). To alter the hydrophobicity, surface modifications with surface defects and polar groups have been suggested, but these affect the stability of the materials as well as their mechanical and electrical properties (Scida et al. Reference Scida, Stege, Haby, Messina and García2011). Computational studies of protein–carbon surface interactions have mostly focused on graphene/graphite (Mereghetti & Wade, Reference Mereghetti and Wade2011; Mücksch & Urbassek, Reference Mücksch and Urbassek2011; Raffaini & Ganazzoli, Reference Raffaini and Ganazzoli2010; Kang et al. Reference Kang, Huynh, Xia, Zhang, Fang, Wei and Zhou2013; Sun et al. Reference Sun, Feng, Hou and Li2014b; Yu et al. Reference Yu, Wang, Lin, Zhao, Zhao and Zheng2012b), CNT (Balamurugan et al. Reference Balamurugan, Gopalakrishnan, Raman and Subramanian2010; Chen et al. Reference Chen, Wang, Liu, Wu, Kang, Moore and Gubbins2009b; Tallury & Pasquinelli, Reference Tallury and Pasquinelli2010; Wang et al. Reference Wang, Humphreys, Chung, Delduco, Lustig, Wang, Parker, Rizzo, Subramoney, Chiang and Jagota2003) and fullerenes (Durdagi et al. Reference Durdagi, Mavromoustakos, Chronakis and Papadopoulos2008; Kraszewski et al. Reference Kraszewski, Tarek, Treptow and Ramseyer2010; Noon et al. Reference Noon, Kong and Ma2002).

3 Which surface properties need consideration in modeling?

The modeling of protein-water-solid surface interfaces poses problems because a variety of distinct physical and chemical properties may be associated with different surface types. Factors such as the size and shape of nanoparticles, the crystal packing of a surface, the presence of surface defects, the density of SAM molecules, the change of the chemical state of the surface, such as protonation, oxidation, or hydroxylation, due to the presence of water and the environmental conditions, such as the presence of surfactants, all affect the interaction of biomolecules and the solid substrate. Therefore, it is very important to choose the level of microscopic details to be included in the computational model carefully when modeling protein–surface interactions, which must adequately describe the physical and chemical properties of the studied system under experimental conditions.

Several important characteristics of the surfaces that should be considered in modeling protein adsorption are discussed in the next sections: ionization and hydration, polarization, surface reconstruction upon binding, as well as the topography and morphology of surfaces and nanoparticles.

3.1 Ionization and hydration

SAMs and metal oxide surfaces may be protonated and/or hydroxylated to varying degrees depending on the environmental conditions and on the material itself, e.g. pH, material shape and size. Many theoretical studies on SAMs do not take into account the ionization states of the functional groups of SAMs. However, a recent study (Utesch et al. Reference Utesch, Millo, Castro, Hildebrandt, Zebger and Mroginski2013) reporting a titration curve of an amino-terminated alkanethiol SAM showed that the level of ionization was very sensitive at pH values around 6 (with 16 ± 5% protonation at pH 7 and 52 ± 9% at pH 6). A positive correlation between the ionization level of a SAM and the strength of adsorption of charged proteins was found (Zhou et al. Reference Zhou, Zheng and Jiang2004; Utesch et al. Reference Utesch, Millo, Castro, Hildebrandt, Zebger and Mroginski2013). Therefore, it is crucial to compute and model the surface ionization to obtain reliable results compatible with experiments. The protonation states of SAMs in modeling studies are mostly represented by random assignment of protonated and deprotonated groups (Alvarez-Paggi et al. Reference Alvarez-Paggi, Castro, Tórtora, Castro, Radi and Murgida2013; Utesch et al. Reference Utesch, Sezer, Weidinger and Mroginski2012; Zhou et al. Reference Zhou, Zheng and Jiang2004). In several studies, they are represented either by a uniform distribution of small partial charges (Sun et al. Reference Sun, Welsh and Latour2005) or by large partial charges being assigned to functional groups in neutral surfaces (Wang et al. Reference Wang, Zhao, Yu, Zhao, Li and Zheng2010b).

As with SAMs, SiO2 and TiO2 surfaces have different levels of ionization of the surface groups depending on the pH of the environment, and this has been shown to determine selectivity of the adsorption of proteins (Patwardhan et al. Reference Patwardhan, Emami, Berry, Jones, Naik, Deschaume, Heinz and Perry2012). Studies have shown that the concentration of silanol groups (Si-OH) and the degree of their ionization define the hydrophobicity of silicon dioxide (silica) surfaces and govern their adsorption properties and thus also, the behavior of silica-based materials in processes such as biomolecular adhesion and biomineralization (Sumper & Brunner, Reference Sumper and Brunner2008; Voskerician et al. Reference Voskerician, Shive, Shawgo, von Recum, Anderson, Cima and Langer2003). In computational studies of protein–oxide surface interactions, ionization may be represented explicitly (Friedrichs et al. Reference Friedrichs, Köppen and Langel2013; Köppen & Langel, Reference Köppen and Langel2010; Patwardhan et al. Reference Patwardhan, Emami, Berry, Jones, Naik, Deschaume, Heinz and Perry2012; Tosaka et al. Reference Tosaka, Yamamoto, Ohdomari and Watanabe2010), or by assigning uniform partial charges to selected surface atoms (Kubiak-Ossowska & Mulheran, Reference Kubiak-Ossowska and Mulheran2012). Köppen & Langel (Reference Köppen and Langel2010) showed that the adhesion energies of peptides on titanium dioxide surfaces are sensitive to the values of the partial charges of the surface hydroxyl groups. Therefore, care must be taken to ensure a reliable parameterization of the ionization charges.

Conventional simulations of proteins usually neglect changes in the protonation states of ionizable groups as a function of time. However, an accurate simulation of a protein-surface system may require a more sophisticated approach, such as constant pH simulation, to treat the variation of the protonation states of residues in the interfacial region, as well as of the surface interaction sites. Although constant pH simulation techniques have been applied to various molecular systems, including small chemotherapeutic drug molecules binding to nanodiamonds (Adnan et al. Reference Adnan, Lam, Chen, Lee, Schaffer, Barnard, Schatz, Ho and Liu2011), they have not, as far as we are aware, been applied to simulations of peptide/protein–surface interactions.

The most important issue for modeling the adsorption of biomolecules on oxide surfaces is the treatment of the properties of the surface hydration shell. Due to dissociation, the hydroxyl groups and hydrogen atoms form bonds with unsaturated surface metal (e.g. Ti) and O atoms, respectively. The degree of water dissociation defines the physical properties of the surface and strongly affects the binding properties of biological molecules. Kang et al. (Reference Kang, Li, Tu, Wang and Ågren2010) investigated the role of the water in the adsorption process on rutile surfaces and observed that human serum albumin (HSA) adsorbs on a rutile surface modified with -OH groups more strongly than on an unmodified rutile surface. Hydrogen bond analysis in the same study showed that the bonds formed between the structured hydroxyl groups on the modified surface reduced the possibility of hydrogen bond formation between the surface and the water molecules, hence making it easier for the protein to adsorb onto the modified surface.

It should be noted that water dissociation is reversible on all oxides (Henderson, Reference Henderson2002). Therefore, the concentration and position of hydroxyl groups may change with time. Even though relying on the average distribution of hydroxyl groups on a particular surface may be sufficient, it may not always give accurate results, especially if the protein adsorption kinetics are determined by surface hydration kinetics. This poses a problem for conventional modeling techniques for biomolecules, which ideally have to capture the dynamics of dissociative adsorption and associative desorption of water molecules on oxide surfaces.

3.2 Polarization

The electrostatic potential of solutes and solvent molecules induces an attractive polarization of metal surfaces. Metal polarization is negligible for neutral adsorbates that do not have large dipoles and, since proteins usually have a relatively small total charge, it is often neglected in simulations (Braun et al. Reference Braun, Sarikaya and Schulten2002). Early studies of metal surfaces also showed that the effect of pure water on polarization of the metal is often negligible (Barabino & Marchesi, Reference Barabino, Cesare-Gavotti and Marchesi1984; Shelley et al. Reference Shelley, Patey, Bérard and Torrie1997; Spohr, Reference Spohr1995) and that charge-induced polarization does not cause any change in the structure of the water on the surface (Feng et al. Reference Feng, Pandey, Berry, Farmer, Naik and Heinz2011). However, for the adsorption of biomolecules with considerable dipole moments, including some peptides and proteins, the surface polarization contributes to the binding energy and influences the binding mode. Although studies have shown that, for water-metal surface systems, the energy due to polarization is less than 10% of the total binding energy (Feng et al. Reference Feng, Pandey, Berry, Farmer, Naik and Heinz2011; Neves et al. Reference Neves, Motheo, Fartaria and Silva Fernandes2007; Siepmann & Sprik, Reference Siepmann and Sprik1995; Vila Verde et al. Reference Vertegel, Siegel and Dordick2009, Reference Vigmond, Iwakura, Mizutani and Katsura2011), for proteins, in particular, on Au(111) surfaces, the contribution from polarization has been estimated to be about 10–20% of the total binding energy (Heinz et al. Reference Heinz, Jha, Luettmer-Strathmann, Farmer and Naik2011). Further, on surfaces such as Au(100), where the van der Waals attraction is weaker, polarization was found to tune the adsorption of proteins (Heinz et al. Reference Heinz, Farmer, Pandey, Slocik, Patnaik, Pachter and Naik2009) and act as a major contributor to the adsorption of highly charged peptides (Heinz et al. Reference Heinz, Jha, Luettmer-Strathmann, Farmer and Naik2011). For the simulation of amino acid adsorption on Au(111), polarization effects were also found to be important for reproducing experimental binding propensities (Hoefling et al. Reference Hoefling, Monti, Corni and Gottschalk2011).

The polarization of surfaces of other types can also play an important role in the interactions of surfaces with their environments. Schyman & Jorgensen (Reference Schyman and Jorgensen2013) showed that while a non-polarizable FF is adequate for the description of interactions between water and small hydrocarbons, such as benzene (C6H6) and coronene (C24H12), a polarizable FF is required for CNTs and fullerenes in order to reproduce interaction energy values obtained from density functional theory (DFT) calculations. Therefore, the effect of polarization should be carefully considered in computer simulations of biomolecules, in particular, of charged proteins, with metal and carbon surfaces.

3.3 Reconstruction

The atomic structure of the surface of a material generally differs from that of its interior because of differences in the forces acting on the atoms in the vicinity of the surface. The type and the degree of reconstruction of a surface is determined by environmental conditions, such as temperature and pressure (Somorjai & Li, Reference Somorjai and Li2011), as well as the structure of the material and may be affected by adsorbing molecules. Ideally, reconstruction of surfaces has to be taken into account in the simulation of adsorption (Ghiringhelli et al. Reference Ghiringhelli, Hess, van der Vegt and Delle Site2008): an enormous task in most cases. On the other hand, it has been reported in several studies that the reconstruction of some surfaces in certain conditions is negligible (Feng et al. Reference Feng, Pandey, Berry, Farmer, Naik and Heinz2011; Iori et al. Reference Iori, Corni and Di Felice2008; Raffaini & Ganazzoli, Reference Raffaini and Ganazzoli2012; Wright et al. Reference Wright, Rodger, Corni and Walsh2013a). However, other experimental and computational studies show that large scale surface reconstruction may take place after adsorption of small molecules (Eralp et al. Reference Eralp, Shavorskiy and Held2011; Gibbs et al. Reference Gibbs, Ocko, Zehner and Mochrie1990; Lal et al. Reference Lal, Plummer, Richmond and Smith2004, Reference Lal, Plummer and Smith2006). Therefore, care must be taken in modeling if a major reconstruction of the surface takes place upon adsorption.

3.4 Topography

The topographic characteristics of a surface or nanoparticle at the micro- to nanometer-scale are important determinants of protein adsorption properties, such as binding affinities and surface saturation values (Fenoglio et al. Reference Fenoglio, Fubini, Ghibaudi and Turci2011; Gagner et al. Reference Gagner, Lopez, Dordick and Siegel2011, Reference Gagner, Shrivastava, Qian, Dordick and Siegel2012; Roach et al. Reference Roach, Farrar and Perry2006). The topography of a surface can be characterized by its exposed crystal planes, its roughness and its defects (due to locally varying chemical composition or the crystalline structure of the surface), as well as kinks, edges and steps that occur during the growth of a crystal. Studies have shown that the same type of protein may have different adsorption energies, varying from highly favorable to highly unfavorable, for surfaces with different lattice structures but the same chemical composition (Heinz et al. Reference Heinz, Farmer, Pandey, Slocik, Patnaik, Pachter and Naik2009; Oren et al. Reference Oren, Tamerler and Sarikaya2005). Although modeling of the intrinsic crystal structure is straightforward, it is important to consider that different crystal planes of a material (which can be kinetically or thermodynamically favored) may be exposed during a surface adsorption process. For example, during the growth of a nanoparticle, different surface planes of the same particle can display different structures at the same time, e.g. (100) and (111) (Korzeniewski et al. Reference Korzeniewski, Climent, Feliu, Bard and Zoski2011).

Similarly, protein binding properties depend on the material structure. Rechendorff et al. (Reference Rechendorff, Hovgaard, Foss, Zhdanov and Besenbacher2006) showed that adsorption of fibrinogen on a tantalum surface can be induced by up to 70% by increasing the surface (root mean squared) roughness from 2·0 to 32·9 nm. The increase was much greater than the increase in the surface area due to the surface roughness of around 20%. On the other hand, Rechendorff et al. (Reference Rechendorff, Hovgaard, Foss, Zhdanov and Besenbacher2006) found that the adsorption of the more globular protein, bovine serum albumin, to tantalum induced by the surface roughness was similar to the increase in the surface area, thus demonstrating a selective effect of the material structure on the adsorption processes of different proteins. Finally, using molecular dynamics (MD) simulations, Nada (Reference Nada2014) investigated the interactions of an aspartic acid with step edges and kinks, as well as flat regions of a calcite crystal surface. They showed that aspartic acid binds preferentially to an acute step edge and not to an obtuse one, due to the ordered structure of water near the obtuse step edge that prevented the aspartic acid from binding strongly (Nada, Reference Nada2014). In summary, these studies demonstrate that the topography of a surface may determine its binding characteristics to proteins and peptides. Although, many topographic features can be neglected in modeling studies when comparing to experiments with well-characterized surfaces, the characteristics of the surface topography should be taken into account to realistically model and simulate the interactions of proteins with the materials used in real-life applications.

3.5 Morphology

The size of a nanoparticle, and therefore the curvature of its surface, has a strong impact on both the physico-chemical characteristics of the nanoparticle itself, such as surface polarizability, surface charge (Chiu et al. Reference Chiu, Moore, Shinoda and Nielsen2009) and isoelectric point (Suttiponparnit et al. Reference Suttiponparnit, Jiang, Sahu, Suvachittanont, Charinpanitkul and Biswas2010), and the properties of the layer of coating molecules which can determine its acidity (Wang et al. Reference Wang, Nap, Lagzi, Kowalczyk, Han, Grzybowski and Szleifer2011). The curvature of a nanoparticle may change the properties of its hydration shell as well. The solvation free energy of a nanoparticle becomes more favorable as the surface curvature decreases (and the size of the particle increases) owing to the increased water–particle interaction energy, which in turn determines its polarity (Chiu et al. Reference Chiu, Moore, Shinoda and Nielsen2009).

The effect of surface curvature on protein adsorption properties, i.e. the kinetics, thermodynamics and structural stability of adsorbed proteins, increases with decreasing particle size (Lacerda et al. Reference Lacerda, Park, Meuse, Pristinski, Becker, Karim and Douglas2010), but also depends on the adsorbate. Moreover, a strong difference in binding to spherical nanoparticles and nanorods is often observed (Gagner et al. Reference Gagner, Lopez, Dordick and Siegel2011). A structural adaption to the curvature of the nanoparticle surface upon adsorption may lead to a loss of enzymatic activity of some proteins (Wu & Narsimhan, Reference Wu and Narsimhan2008), or it may lead to significant changes in secondary or tertiary structures of self-assembling peptides (Shaw et al. Reference Shaw, Middleton, Volk and Lévy2012) or proteins (Tavanti et al. Reference Tavanti, Pedone and Menziani2015; Yang et al. Reference Yang, Johnson, Wu, Woods, George and Murphy2013) adsorbing onto the surfaces. It has been shown that while surface curvature may help to retain the tertiary structure of some proteins with globular structures adsorbed on small nanoparticles (Vertegel et al. Reference Vila Verde, Acres and Maranas2004; Lundqvist et al. Reference Lundqvist, Sethson and Jonsson2004), it can also cause significant loss in the secondary structure of a protein upon adsorption (Gagner et al. Reference Gagner, Lopez, Dordick and Siegel2011).

The effects of the surface curvature of nanoparticles can be neglected in computational studies when the binding of relatively small proteins to large nanoparticles is studied. However, the size and curvature of a nanoparticle may play a major role in the adsorption patterns of proteins not only due to geometric adaptation of the protein to a nanoparticle of similar size, but also due to changes in the physico-chemical properties of the nanoparticle itself. DFT calculations of amino acid adsorption on a single-walled carbon nanotube (SWCNT) showed that glycine adsorbs more strongly on a nanotube (3,3) than on a flat graphite surface, whereas phenylalanine adsorbs more strongly on a flat graphite surface, and the amino acids cysteine and histidine, showed no significant change in their adsorption energies (Roman et al. Reference Roman, Diño, Nakanishi and Kasai2006). Binding free energy calculations of amyloidogenic apoC-II(60–70) peptide on fullerene, CNT and graphene using MD simulations showed that the binding affinity was weakest for the fullerene and strongest for the graphene due to reduced efficiency of π-stacking interactions between the aromatic side chains of the peptide and the fullerene and CNT arising from the increased surface curvature (Todorova et al. Reference Todorova, Makarucha, Hine, Mostofi and Yarovsky2013). Xie et al. (Reference Xie, Luo, Lin, Xi, Yang and Wei2014) calculated the adsorption free energies of Alzheimer's β-amyloid peptide fragments (Aβ) to two different fullerene nanoparticle systems, C180 and three C60 (3C60). They found tighter binding of the peptides on the larger C180.

Raffaini & Ganazzoli (Reference Raffaini and Ganazzoli2013) showed that the binding strength of a protein to a carbon nanotube depends on the morphology and that the interaction energy between the protein and the nanoparticle was larger for the concave interior surface of the nanotube than for the convex outer surface. In a similar study, Chen et al. (Reference Chen, Wang, Liu, Wu, Kang, Moore and Gubbins2009b) showed that the length and the diameter of CNTs affect the energetics of the interactions with a peptide drug. While longer CNTs provide more space to trap the peptide inside the tube, a smaller diameter increases the interaction energy.

The surface morphology not only alters the binding characteristics of a peptide or protein to a surface but also the interactions between biomolecules near a surface or a nanoparticle, thus leading to intermolecular structural changes. Li et al. (Reference Li, Luo, Derreumaux and Wei2011) performed 3 separate sets of simulations of Aβ(16–22) peptides that abnormally self-assemble into β-rich aggregates: amorphous peptide in solution, amorphous peptide with SWCNT, and prefibrillar peptide with SWCNT. They observed that without the CNTs, the amorphous peptides form β-sheet structures. On the other hand, simulations with SWCNTs showed that the amorphous peptides tend to form disordered coils, whereas the β-sheet structures formed by the prefibrillar peptides were destabilized due to the interactions of the peptides with the CNTs.

Finally, noteworthy to mention is that a change in the morphology of a surface/particle will lead to changes in its physico-chemical properties. In a study by Emami et al. (Reference Emami, Puddu, Berry, Varshney, Patwardhan, Perry and Heinz2014b), it was shown that the size of the silica nanoparticles determines their surface ionization levels. The larger nanoparticles have higher ionization and therefore bind more strongly to peptides with high net positive charge. Further, Baier et al. (Reference Baier, Blumenstein, Preusker, Jeurgens, Welzel, Do, Pleiss and Bill2014) recently investigated the binding free energy profile of 12-mer peptides on polar (001) and non-polar (100) ZnO surfaces. Employing an enhanced sampling approach (see Section 10.2), the authors showed that there is a positive correlation between the ZnO particle sizes obtained in experiments in the presence of the peptides and the calculated affinities of the peptides for the ZnO surface. Their result showed that the selective adsorption of a peptide can impact the growth of certain nanocrystals. Therefore, one should consider modeling, not only the size and curvature of the material, but also other properties that depend on the morphology, in particular for nanoparticles.

4. Which modeling and simulation techniques are applicable to protein–surface interactions?

The approaches used for modeling protein-inorganic surface systems cover a broad range of scales from sub-atomic quantum mechanical (QM) through classical atomic levels to mesoscopic descriptions and further to continuum descriptions on a macroscopic scale. Figure 2 shows the most common techniques to model and to simulate molecular interactions. Simulations at the QM level are applied to systems of a few hundreds of atoms at most and they do not reach the nanosecond scale, hence they are directly applicable only to small nanoparticle-ligand systems (Mahmoudi et al. Reference Mahmoudi, Lynch, Ejtehadi, Monopoli, Bombelli and Laurent2011).

Fig. 2. Typical time and length scales of different simulation techniques: quantum mechanics (QM), including coupled cluster (CC) and DFT methods (inset adapted with permission from (Iori et al. Reference Iori, Corni and Di Felice2008). Copyright (2008) American Chemical Society); molecular mechanics (MM) including all-atom molecular dynamics (AA-MD) simulations, implicit solvent and coarse grained MD (IS-MD and CG-MD), and the Brownian dynamics (BD) technique; and continuum mechanics (CM). The ranges of time and length scales are approximate.

Many physical processes at the protein-solid surface interface are driven by physisorption, i.e. proceed without formation of chemical bonds between the adsorbate and the solid surface. In contrast to the relatively well-understood chemisorption, the nature and behavior of non-covalent adsorption is often unclear since multiple factors, which depend strongly on the surface type, influence the interactions that govern the adsorption process. Even if chemical binding takes place, physical adsorption drives the first stages of molecular recognition and induces long-time scale structural adaptations of a protein to a solid surface. Non-covalent binding processes can be described within a MM framework, which drastically reduces computational costs compared with QM, and thus enables the simulation of the dynamics of systems that consist of millions of atoms for up to microseconds with solvent molecules modeled explicitly. All-atom MM simulations are important, in particular, for investigation of the dynamic and thermodynamic properties of protein adsorption (Mahmoudi et al. Reference Mahmoudi, Lynch, Ejtehadi, Monopoli, Bombelli and Laurent2011; Utesch et al. Reference Utesch, Daminelli and Mroginski2011). The nano length scale is usually appropriate for studying protein–surface interactions at the molecular level, and hence, all-atom simulations are common methods of choice (Gagner et al. Reference Gagner, Shrivastava, Qian, Dordick and Siegel2012).

In experiments, adsorption events typically take place over time periods from milliseconds to hours which are far from the accessible time scales of all-atom simulations (Mücksch & Urbassek, Reference Mücksch and Urbassek2011). The initial stages of protein adsorption to surfaces occur on a sub-second time scale. These may be followed by a slow stage in which large secondary structural changes occur, such as the transition from α-helix towards β-sheet of lysozyme on a SAM surface (Sethuraman & Belfort, Reference Sethuraman and Belfort2005), and may entail denaturation over periods lasting up to several hours, or even days (Gray, Reference Gray2004; Pan et al. Reference Pan, Qin, Meng, Cao and Wang2012). Capturing the time and length scales of a complete adsorption process therefore necessitates employing mesoscopic, coarse-grained (CG) and multiscale approaches (Gray, Reference Gray2004; Wei et al. Reference Wei, Carignano and Szleifer2011). Furthermore, hybrid approaches such as QM/MM to bridge typical time and length scales of conventional approaches, and enhanced sampling simulation techniques to accelerate adsorption and desorption events have been proposed and successfully applied to study protein–inorganic surface interactions (Euston et al. Reference Euston, Hughes, Naser and Westacott2008; Utesch et al. Reference Utesch, Daminelli and Mroginski2011; Zhang & Sun, Reference Zhang and Sun2010).

5 Quantum mechanics studies of protein–surface interactions

QM-based methods are those where the quantum nature of electrons is explicitly taken into account while the much heavier nuclei are usually considered as classical particles moving in the field generated by the electrons and the other nuclei. Several approaches belong to this class, ranging from those that are relatively fast but rich in adjustable parameters, such as tightbinding and semi-empirical methods, to very accurate but also computationally very expensive, parameter-free calculations, such as coupled-cluster. The QM method that has been applied most extensively so far for studying biomolecules at surfaces is DFT, as it represents the best compromise between accuracy and computational feasibility.

The fundamental idea behind DFT (Martin, Reference Martin2004), based on the Hohenberg and Kohn theorems (Hohenberg, Reference Hohenberg and Kohn1964), is that, for non-degenerate systems, the ground state electron density, n(r), alone determines the entire behavior of the system, and that such n(r) minimizes the energy of the system. The most useful approximation of such energy as a function of n(r) has been proposed by Kohn & Sham (Reference Kohn and Sham1965). It translates the minimization problem to a non-linear, independent particle approach akin to the Hartree–Fock one. A central quantity within this approach is the so-called exchange correlation functional (f xc ), i.e. the n(r)-dependent energy contribution due to quantum-mechanical and many-body deviations from a mean-field description of the electrons. So far, there is neither an exact expression for f xc nor a one-fits-all approximation. An important and often not-obvious choice to be made for any DFT calculation is therefore which f xc is the best for the system under study.

Given the geometry of the nuclei in space, DFT allows the forces acting on them to be calculated. This relation can be used to find the nuclei geometries that provide local and global minima of the system energy (i.e. the optimal or relaxed geometry). These are obviously the most important structures, since they are the structures in whose neighborhood the system fluctuates. In the following, we shall refer to this kind of DFT calculation as static. In fact, forces can also be used to simulate the dynamics of the nuclei, and therefore, in principle, the thermodynamics as well, via different propagation algorithms, such as Born-Oppenheimer and Car-Parrinello (Marx & Hutter, Reference Marx, Hutter and Grotendorst2000). DFT ab initio molecular dynamics (AIMD) is often used to collectively refer to this kind of simulations. Several software packages are available and maintained for performing static DFT and/or AIMD calculations. They differ in the numerical approaches implemented to solve the Kohn–Sham equations (e.g. using plane-waves or localized atom-centered functions), in the available f xc , in the level of parallelism (i.e. suitable for highly parallel computers or for a few-core, high memory workstation), in the pre- and post-processing tools offered, and finally in the distribution policy (e.g. open-source versus proprietary, free versus commercial).

Considering proteins on surfaces, static DFT calculations of single amino acids or even di- and tri-peptides, (Lee et al. Reference Lee, Kim and Lee2014; Muir et al. Reference Muir, Costa and Idriss2014) interacting with a surface, with no solvent present, are now affordable and widely used (Arrouvel et al. Reference Arrouvel, Diawara, Costa and Marcus2007; Di Felice et al. Reference Di Felice, Selloni and Molinari2003; Di Felice & Selloni, Reference Di Felice and Selloni2004; Ghiringhelli et al. Reference Ghiringhelli, Schravendijk and Delle Site2006; Iori et al. Reference Iori, Corni and Di Felice2008; Rimola et al. Reference Rimola, Sodupe and Ugliengo2009). They may seem rather distant from biophysically relevant systems, but they can actually provide important information by themselves, or be preliminary to other approaches (e.g. provide the basis for a classical FF parameterization). Among the DFT calculations that directly provide useful information, we mention those aimed at understanding whether a covalent bond is established between an amino acid and an inorganic surface. In fact, when a covalent bond is present (chemisorption), it is unlikely that the protein and solvent environment missing in the calculations dramatically change its nature, although they do modify the details. Notable examples of calculations of this kind are works on Cys–Au(111) interactions (Buimaga-Iarinca & Calborean, Reference Buimaga-Iarinca and Calborean2012; Di Felice et al. Reference Di Felice, Selloni and Molinari2003; Di Felice & Selloni, Reference Di Felice and Selloni2004; Fajín et al. Reference Fajín, Gomes and Cordeiro2013; Nazmutdinov et al. Reference Nazmutdinov, Manyurov, Zinkicheva, Jang and Ulstrup2007), often used for protein immobilization (Vigmond et al. Reference Vila Verde, Beltramo and Maranas1994). Static DFT calculations on amino acids or simpler molecules representative of chemical groups in natural amino acids on metals (Hong et al. Reference Hong, Heinz, Naik, Farmer and Pachter2009; Iori et al. Reference Iori, Corni and Di Felice2008, Reference Iori, Di Felice, Molinari and Corni2009) have also highlighted the peculiar nature of the amino acid–metal interaction. Depending on the partners, such an interaction ranges from clear non-bonding (e.g. alkyl side chains on Au(111)) to clear chemisorption (Cys on Au(111)). It also encompasses border-line cases where the interaction has some typical covalent bond characteristics, such as electron sharing and directionality, and yet it is only marginally stronger than a non-bonded interaction (e.g. imidazole on Au(111), see Fig. 3).

Fig. 3. Isosurface plots for density functional theory (DFT) single electron states at the imidazole/Au(111) interface. (a) Bonding orbital with σ-like shape; (b) antibonding orbital with σ-like shape. The atomic p-like character of the orbital on the imidazole N is visible in both panels as density within the ring (red circle). Color scheme: the orbital density isosurface is represented in magenta; Au: orange, N: grey; C: yellow; H: cyan. Adapted with permission from (Iori et al. Reference Iori, Corni and Di Felice2008). Copyright (2008) American Chemical Society.

DFT calculations have also been used to elucidate amino acid adsorption on silica (Rimola et al. Reference Rimola, Costa, Sodupe, Lambert and Ugliengo2013), hydroxyapatite (Jimenez-Izal et al. Reference Jimenez-Izal, Chiatti, Corno, Rimola and Ugliengo2012), alumina (Arrouvel et al. Reference Arrouvel, Diawara, Costa and Marcus2007) and titania (Carravetta et al. Reference Carravetta, Monti and Zhang2009; Koch et al. Reference Koch, Lipton, Filipek and Renugopalakrishnan2011) surfaces where electrostatic interactions and hydrogen bonds are important, and where the formation of chemical bonds often implies complex reaction mechanisms. The interaction of quartz and aluminosilicate structures with phospholipids was also studied by DFT to understand why the enzyme phospholipase A2 digests phospholipids faster in the presence of quartz than aluminosilicates (Snyder & Madura, Reference Snyder and Madura2008). The interaction of amino-acids with graphene has also been the subject of recent DFT studies (Akdim et al. Reference Akdim, Pachter, Kim, Naik, Walsh, Trohalaki, Hong, Kuang and Farmer2013; Wang et al. Reference Wang, Guo, Wang, Zhang, Huang, Lu, Wang, Zhang and Leng2014).

Static DFT simulations have been performed to parameterize classical FFs for protein–surface interactions in water (Bellucci et al. Reference Bellucci, Brancolini, Calzolari, Carrillo Parramon, Corni, Di Felice, Horbett, Brash and Norde2012; Carravetta & Monti, Reference Carravetta and Monti2006; Carravetta et al. Reference Carravetta, Monti and Zhang2009; Di Felice & Corni, Reference Di Felice and Corni2011; Ghiringhelli et al. Reference Ghiringhelli, Schravendijk and Delle Site2006; Iori et al. Reference Iori, Di Felice, Molinari and Corni2009; Schneider & Colombi Ciacchi, Reference Schneider and Colombi Ciacchi2010; Wright et al. Reference Wright, Rodger, Corni and Walsh2013a, Reference Wright, Rodger, Walsh and Corni2013b). The general strategy here is to perform energy calculations on the optimized geometries and/or along specific coordinates, for either entire amino acids or simpler amino acid analogues, on the target surface, and to use the results as a training set. Classical parameters are thus adjusted so as to reproduce as closely as possible the QM interaction energies and geometries in the classical calculations. DFT calculations have also been used to generate partial atomic charges for the surface atoms to be used for Coulombic interactions in the classical models.

AIMD is intrinsically much more expensive than static DFT simulations and, for this reason, its application in the protein-surface field has been less popular. Recently, Motta et al. (Reference Motta, Gaigeot and Costa2012) investigated the adsorption of glycine on a stepped boehmite (AlO(OH)) surface in water, identifying inner-sphere adsorption, i.e. displacement of surface hydroxyl groups by glycine molecules, as the most favorable. Wright & Walsh (Reference Wright and Walsh2012) focused on ammonium and acetate ions, which are analogues of common chemical groups in amino acids, at the water/quartz interface. Colombi Ciacchi and colleagues used AIMD to build a model of a native silicon oxide surface (Colombi Ciacchi & Payne, Reference Cicero, Calzolari, Corni and Catellani2005; Cole et al. Reference Collier, Vellore, Yancey, Stuart and Latour2007) that was afterwards used for studying peptide–silica interactions (Schneider & Colombi Ciacchi, Reference Schneider and Colombi Ciacchi2012).

The current limits of static and AIMD DFT calculations in the field of protein–surface interactions are well illustrated by some recent examples shown in Fig. 4. (Rimola et al. Reference Rimola, Aschi, Orlando and Ugliengo2012) exploited static DFT to study the adsorption of an entire dodecapeptide on a hydroxyapatite surface, including some key water molecules (the system was composed of around 500 atoms). They discussed the driving force for the adsorption and the role of the surface in determining the peptide folding. A system of similar size (a dodecapeptide on a graphene sheet) has also been investigated by Akdim et al. (Reference Akdim, Pachter, Kim, Naik, Walsh, Trohalaki, Hong, Kuang and Farmer2013) to confirm by DFT the adsorbed peptide geometries obtained with a classical force-field.

Fig. 4. Examples of systems studied in large scale static density functional theory (DFT) calculations and ab initio molecular dynamics (AIMD) simulations. (a) Static DFT: dodecapeptides adsorbed on a hydroxyapatite (0001) surface, after DFT geometry optimization, color scheme: O peptide: red, O water: cyan, N: blue, C: dark yellow, H: light gray, Ca: green; P: yellow. Adapted with permission from (Rimola et al. Reference Rimola, Aschi, Orlando and Ugliengo2012). Copyright (2012) American Chemical Society; (b) AIMD: Side view of a polyserine β-sheet on an Au(111) slab in liquid water. Color scheme: O: red, N: blue, C: gray, H: white, Au: yellow. Adapted with permission from (Calzolari et al. Reference Calzolari, Cicero, Cavazzoni, Di Felice, Catellani and Corni2010). Copyright (2010) American Chemical Society. The dashed lines indicate the periodically repeated cell.

Calzolari et al. (Reference Calzolari, Cicero, Cavazzoni, Di Felice, Catellani and Corni2010) considered a model polyserine β-sheet, periodically replicated, simulated by AIMD on Au(111) in liquid water. This system was composed of approximately 500 atoms and its time evolution could be simulated for 20 ps. Such a simulation time and the composition of the system allowed the investigation of some specific questions, such as the nature of the local β-sheet/Au interactions, the competition between water and the serine side chain for gold, as well as the nature of the β-sheet/water interface. However, it is apparent that several other important questions are inaccessible with this kind of approach. Today, larger systems and longer AIMD simulations are affordable but the most expensive AIMD simulations are still confined to a few thousand atoms and a few hundred ps.

Beside the current limitations related to the computational cost of DFT simulations, one notorious drawback of the f xc functionals used so far is worth discussion here. It is the inability to account for long-range dispersion (London) interactions, sometimes referred to as van der Waals’ interactions, that results in an underestimated interaction strength, and even no solute-surface binding when such interactions are the only relevant ones (e.g. for inert metal surfaces and saturated, non-polar molecules). Various corrections have been proposed to solve this problem (Tkatchenko et al. Reference Tkatchenko, Romaner, Hofmann, Zojer, Ambrosch-Draxl and Scheffler2010), and some of them have also been tested in the framework of protein–surface interactions with encouraging results. In particular, the DFT-Dn methods (Grimme, Reference Grimme2004; Grimme et al. Reference Grimme, Antony, Ehrlich and Krieg2010; Wu et al. Reference Wu, Vargas, Nayak, Lotrich and Scoles2001) (that add to the DFT energy, empirical atom-atom d−6 terms, where d is the pairwise interatomic distance, suitably damped for small d) have been used for amino acid and peptide adsorption on mineral surfaces (Folliet et al. Reference Folliet, Gervais, Costa, Laurent, Babonneau, Stievano, Lambert and Tielens2013; Rimola et al. Reference Rimola, Sodupe and Ugliengo2009, Reference Rosa, Corni and Di Felice2012) and on graphene (Akdim et al. Reference Akdim, Pachter, Kim, Naik, Walsh, Trohalaki, Hong, Kuang and Farmer2013). The f xc functional, vdW-DF, which does not contain empirical terms (Dion et al. Reference Dion, Rydberg, Schröder, Langreth and Lundqvist2004), has been tested against experimental desorption energy data for some small molecules on Au(111) (Wright et al. Reference Wright, Rodger, Corni and Walsh2013a). It was used to provide the main data (stable geometries and the related energies for amino acid analogues on Au(111) and Au(100)) needed to parameterize the GolP-CHARMM FF (Rosa et al. Reference Rosa, Corni and Di Felice2014b; Wright et al. Reference Wright, Rodger, Corni and Walsh2013a). Tests confirmed the reliability of the vdW-DF adsorption energies, within a few kJ/mol of the experimental values (Fig. 5), and pointed to an already documented tendency of vdW-DF to provide contact distances and Au lattice parameters that are slightly too large by 0·1–0·2 Å (Lee et al. Reference Lee, Murray, Kong, Lundqvist and Langreth2010). Other functionals akin to vdW-DF have been proposed to correct this deficiency (Lee et al. Reference Lee, Murray, Kong, Lundqvist and Langreth2010; Klimeš et al. Reference Klimeš, Bowler and Michaelides2010, Reference Klimeš, Bowler and Michaelides2011), and are awaiting testing and validation in the field of molecular adsorption, and specifically for protein–surface interactions. The computational overhead in using these functionals is modest, and they have also been applied to other biomolecules adsorbed on gold, such as nucleic acids (Rosa et al. Reference Rosa, Corni and Di Felice2012, Reference Rosa, Corni and Di Felice2014a, Reference Rosa, Corni and Di Feliceb). AIMD is also possible with these functionals, as exemplified by a recent study of the liquid water/gold interface (Nadler & Sanz, Reference Nadler and Sanz2012). In this case, dispersion interactions do not change the picture provided by conventional functionals (Cicero et al. Reference Clementi, Nymeyer and Onuchic2011). In summary, the DFT limitations connected with the lack of dispersion interactions are currently being overcome by recent methodological developments.

Fig. 5. Comparison between vdW-DF and experimental adsorption energies to an Au(111) surface for a set of molecular adsorbates. Reprinted with permission from (Wright et al. Reference Wright and Walsh2013). Copyright (2013) American Chemical Society.

In the future, approaches other than DFT, such as Quantum Monte Carlo (QMC) (Austin et al. Reference Austin, Zubarev and Lester2012), may also become popular for investigating biomolecule–surface interactions. QMC is based on polyelectronic wavefunctions and naturally accounts for long–range dispersion interactions. Such wavefunctions are determined by Monte Carlo (MC) algorithms, which are highly parallelizable. While QMC calculations are currently too expensive to routinely study biomolecule–surface interactions, its intrinsic high parallelism combined with the constantly growing size of modern supercomputers, make its application to this field likely in the years to come.

To conclude, despite their size and time-scale limitations, DFT-based approaches are playing an important role in revealing the physico-chemical basis of protein–surface interactions. They are used either to provide a detailed picture of the local amino acid–surface interactions or as a basis for developing classical atomistic models, i.e. as a source of benchmark data to train classical FFs or of model structures for complex materials such as amorphous silica.

6. Challenges in applying biomolecular molecular mechanics force fields to protein–surface interactions

Modeling and simulation of protein–surface interactions brings along the challenges associated not only with modeling the surfaces and proteins separately, but also with modeling the system as a whole (see Fig. 6 for a depiction of protein–surface interactions in aqueous solvent). The FFs routinely used in modeling and simulation studies of proteins are parameterized for interactions between biomolecular fragments or small chemical compounds in aqueous solution. Although the FFs developed for protein simulations may provide a good approximation for modeling the interactions between a protein and a surface in some cases, in general, the force field parameters to be used have to be derived and calibrated for the systems of interest to obtain high quality results.

Fig. 6. Schematic illustration of protein–surface interactions in aqueous solvent. The main interaction interfaces can be categorized as: protein–surface, protein–solvent, solvent–surface and protein–solvent–surface. The protein-surface interface (depicted in the left circle) includes direct interactions. The interactions can be non-specific such as van der Waals and electrostatic interactions (represented with dashed lines in the figure), or specific such as strong histidine–gold interactions (shown with a continuous line) and even stronger chemisorption interactions. At the protein-solvent interface (depicted in the top circle), the structural and physical properties of the protein and the solvent deviate from those inside the protein and in the bulk solvent, respectively. In particular, water forms layers around the polar and charged residues as depicted by the two spheres in the figure. At the interface, the relative dielectric permittivity of water and of the protein is lower than that of their bulk counterparts. At the solvent-surface interface (depicted in the right circle), the solvent may form structured layers or be completely disordered. On a gold surface, for instance, water forms two ordered layers that are separated by high energy barriers and have a lowered relative dielectric permittivity in the direction normal to the surface. At the protein-solvent-surface interface (depicted in the bottom circle), the interactions involve a complex interplay between the constituents. The protein may make strong indirect interactions with the surface through a stable network of hydrogen bonds (represented by dashed lines) in the adsorption region.

6.1 Interaction potentials

The classical potential energy functions employed in the all-atom molecular mechanics force fields (MM FFs) for biomolecules, such as AMBER (Cornell et al. Reference Cornell, Cieplak, Bayly, Gould, Merz, Ferguson, Spellmeyer, Fox, Caldwell and Kollman1995), CHARMM (Brooks et al. Reference Brooks, Bruccoleri, Olafson, States, Swaminathan and Karplus1983), GROMOS (Oostenbrink et al. Reference Oostenbrink, Villa, Mark and Van Gunsteren2004), and OPLS-AA (Jorgensen et al. Reference Jorgensen, Maxwell and Tirado-Rives1996), are widely used and thoroughly evaluated for simulation of biomolecules in aqueous solution. Most commonly used biomolecular FFs are expressed as a sum of pairwise interaction terms that represent changes of the (i) chemical bond lengths and bond angles of a molecule as harmonic spring functions; and (ii) torsions as periodic functions (dihedral angles, or torsional rotation of atoms around a central bond); and (iii) non-bonded electrostatic and van der Waals intra-  and inter-molecular interactions:

$${E_{{\rm Total}}} = {E_{{\rm bond}}} + {E_{{\rm bond}}}_{\_{\rm angle}} + {E_{{\rm torsion}}} + {E_{{\rm electrostatic}}} + {E_{{\rm vdW}}}$$

Electrostatic interactions between the atoms in the system are approximated by Coulomb's law with fixed point charges assigned to each atom, whereas van der Waals interactions are typically described by the Lennard–Jones (12–6) potential:

$${U_{{\rm LJ}}} = 4{\varepsilon _{ij}}\left[ {{{\left( {\displaystyle{{{\sigma _{ij}}} \over r}} \right)}^{12}} - {{\left( {\displaystyle{{{\sigma _{ij}}} \over r}} \right)}^6}} \right]$$

where the two terms represent repulsive and attractive interactions, respectively, and the parameters, and σ are expressed as a combination of parameters of atoms i and j. If this form of the FF is suitable for the solid surface to be studied, it can also be applied for simulations of protein adsorption.

The FFs derived specifically for simulation of inorganic materials are generally aimed at reproducing the structural and physico-chemical properties of the bulk or surface of a material in vacuum and often have a functional form that is different from that of conventional biomolecular force-fields. Usually, monatomic systems with close-packed structures can be reasonably well described by two-body interaction potentials, whereas a three-body potential must be used for semiconductors (Vashishta et al. Reference Vashishta, Kalia, Rino and Ebbsjö1990). Additional angle-dependent terms can be employed to preserve the directionality of bonds, e.g. in the crystal packing arrangement of a bulk solid (Cruz-Chu et al. Reference Cruz-Chu, Aksimentiev and Schulten2006). Furthermore, a Buckingham potential of the form Ae Br −(C/r 6), where A, B and C are force field parameters, is often used instead of a Lennard–Jones potential to better describe the repulsion between ions in inorganic materials (see, for example, Van Beest & Kramer, Reference Van Beest, Kramer and van Santen1990). Moreover, a balanced parameterization of a simple pairwise interaction energy model commonly used in standard biomolecular FFs is not straightforward and is not always possible for the water-material interface. For instance, a combination of two-body and three-body non-bonded potentials was employed in a silica force-field to preserve the tetrahedral structure of the silica glass and to reproduce accurately silica surfaces, pores and surface wettability (Cruz-Chu et al. Reference Cruz-Chu, Aksimentiev and Schulten2006). Another limitation of a simple pair-wise function is that it is based on a fixed point charge approximation for Coulomb's interaction, which omits the interactions between higher order multipoles and polarization effects arising from the electrostatic field of the environment (a non-polarizable force-field). The partial atomic charges are usually assigned for a molecule in an aqueous environment and, thus, the polarization effect of the aqueous surroundings is implicitly accounted for. Non-polarizable FFs have been extensively validated over the last decades and currently offer a robust description of the equilibrium properties of biomolecules in aqueous solution. However, polarization effects of the surface (such as metal or carbon surfaces, as discussed earlier) may contribute notably to the binding energy of an adsorbate, as well as to the structure of the surface itself.

Polarizable FFs for biological molecules are currently under active development and validation (Halgren & Damm, Reference Halgren and Damm2001). However, apart from their notably higher computational cost, polarizable FFs for biomolecules are not designed to reproduce the polarization properties of inorganic surfaces. In Section 7, we will consider several models adapted for simulation of surface polarization effects.

Despite all the limitations mentioned above, the early simulations of protein and peptides on solid surfaces have typically been performed using standard biomolecular FFs. In the last decade, however, much effort has been invested to evaluate and adapt biomolecular FFs to the simulation of interfaces between biomolecules in water and specific inorganic materials, and several new approaches for developing general bio-inorganic FFs have been reported. For example, a dual-scale FF that includes two sets of partial atomic charges, optimized for each phase separately, has been developed (Biswas et al. Reference Biswas, Vellore, Yancey, Kucukkal, Collier, Brooks, Stuart and Latour2012). In the recently reported INTERFACE FF (Heinz et al. Reference Heinz, Lin, Kishore Mishra and Emami2013), partial atomic charges and van der Waals parameters have been optimized using a non-polarizable fixed-charge model for a large number of different materials based on the properties of the solid surfaces, e.g. crystal unit parameters, surface tension, water contact angle and interfacial tension, and adsorption energy of selected peptides. Some applications of these new models for specific surfaces will be discussed in Section 7.

6.2 Solvation models

The properties of the hydration shell of a solid substrate, in particular variations of physical properties, such as the density, free energy and dielectric constant of water in a hydration shell, and chemical properties, such as surface reconstruction and ionization, often govern adsorption processes. Therefore, a rigorous treatment of water, whether explicit or implicit, is required to account for the effects of water on adsorption. FF parameters for water molecules are usually parameterized for the bulk water state and thus do not take into account the specific features of the surface or protein hydration shell. Moreover, the properties of a solid surface hydration shell differ from those of small solutes and biomolecules in that, generally, more water–water hydrogen bonds are broken and the water is forced to build an extended ordered surface layer. In general, there is still uncertainty as to whether standard explicit water models are able to adequately reproduce adsorption phenomena for proteins and surfaces. Studies of water properties on a wide range of surfaces have been reviewed by Henderson (Reference Henderson2002).

The microscopic properties of the water molecules on a surface can be parameterized using DFT ab initio computations, which have been widely used for predicting the adsorption energies of water molecules as well as water dissociation properties on different surfaces. The AIMD technique mentioned earlier is a powerful tool for exploring water properties in a monolayer or even in several layers (e.g. on gold, see (Velasco-Velez et al. Reference Velasco-Velez, Wu, Pascal, Wan, Guo, Prendergast and Salmeron2014)). However, it was shown that the spatial and temporal limitations of DFT simulations may prevent accurate modeling of water structural transitions from the hydration shell to bulk, which is important for the modeling of protein adsorption (Große Holthaus et al. Reference Große Holthaus, Köppen, Frauenheim and Colombi Ciacchi2012). Accordingly, an all-atom MM representation of water has to be used for exploring the adsorption processes of molecules in solution, which usually occur on a much longer time-scale. Likewise, an all-atom MM model of the solvent can be replaced by a continuum solvent model to enable longer simulations of large proteins or protein solutions for investigating their adsorption properties.

Implicit solvation is a common way to reduce the computational cost of simulations that arises in large part from calculating the interactions between the explicit water molecules. In this approach, the contribution of the solvent to the binding energies between the solutes is represented with a mean field model, and is usually decomposed into two major contributions: electrostatic contributions due to charge interactions and the change of dielectric polarization of their surroundings, and non-polar contributions due to the change in entropy and the dispersion energy upon removing water from the solute surface. The non-polar desolvation energy is commonly estimated as a linear function of the solvent accessible surface area. The most common approaches for computing the continuum electrostatic contribution to the binding free energy is to solve the Poisson–Boltzmann (PB) equation or the Generalized Born (GB) model, which provides an approximation to the PB method. The PB method is often used with a rigid molecule approximation because it is relatively computationally expensive to solve the PB equation for every macromolecular configuration during a simulation. For applications of these methods to free energy calculations of protein-surface systems, see Section 10.

The implicit solvent models traditionally used for molecules in solution do not, per se, account for the unique properties of the hydration shell and the interactions of the water in direct contact with the inorganic surface. However, they can be parameterized to account for the hydration shell characteristics of individual materials. An example of such a continuum solvent model is ProMetCS, which was derived for protein-gold (111) interfaces (Kokh et al. Reference Kokh, Corni, Winn, Hoefling, Gottschalk and Wade2010). In ProMetCS, the effects of the hydration shell distortion at the protein-surface interface were accounted for by the addition of an analytical function parameterized to reproduce the potential of mean force (PMF) of a probe ion/atom on the gold surface obtained from explicit water MD simulations. The effects arising from the partial replacement of the metal hydration shell by a protein adsorption site are described by a free energy penalty, proportional to the protein-surface contact area, compensating the Lennard–Jones attraction to a large extent. The PMF of an atom on the gold surface computed using the ProMetCS FF reproduces the profile of the PMF function obtained in explicit solvent MD simulations (Fig. 7).

Fig. 7. Modeling of protein-Au(111) interactions using a continuum solvent model parameterized by comparison with explicit solvent MD simulations. Solid line: The potential of mean force (PMF) for a test atom as a function of atom-gold surface distance, as obtained from MD simulations in explicit water solvent; squares: the corresponding LJ potential; dashed line: their difference, associated with the desolvation energy; dotted line: PMF computed using the protein-metal continuum solvent ProMetCS model (which includes both LJ and metal hydrophobic desolvation energies, see text). Reprinted with permission from (Kokh et al. Reference Kokh, Corni, Winn, Hoefling, Gottschalk and Wade2010). Copyright (2010) American Chemical Society.

7. All-atom molecular mechanics studies of protein–surface interactions

Some of the factors that are particularly important to take into account when developing or applying MM FFs for simulation of protein-surface adsorption are: (i) compatibility with a standard MM FF for biomolecules; (ii) the ability to provide dynamic and conformational properties of biomolecules in solution as well as on the surface; (ii) the ability to provide physical and chemical properties of the surface such as structure and polarization; (iii) a correct reproduction of the hydration shell properties for a particular surface; and (iv) the ability to model changes in the surface layer upon protein adsorption or surface reconstruction, for example, due to water dissociation or adsorption of ions.

In this section, we will focus on the FFs developed for describing protein–surface, protein–water and surface–water interactions in atomic detail. Further, we will briefly introduce the techniques employed to tackle some of the challenges posed by different surface types. In the following sub-sections, the most commonly studied surface types, namely elemental metal, titanium oxide, silicon oxide, mineral, self-assembled monolayer and sp2-carbon surfaces, are reviewed.

7.1 Metal surfaces

The metal polarization in the presence of the charges of a solute, as discussed earlier, has a significant effect on adsorption and, therefore, efforts have been put into integrating polarization effects into FF models. The simplest way to introduce the induced polarization of the metal surface in a computational model is using a classical image-charge approximation for a charge and a zero-potential metal surface in a continuum dielectric medium with an interaction energy expressed by Coulomb's law between the charge and its image of the opposite sign (see Fig. 8a ). This approximation was shown to give an accurate description for the interaction energy of an Al(111) surface with a charge when the separation distance was more than 2·5 Å (Finnis & Finnis, Reference Finnis1991), while at closer distances to the surface, the image charge model overestimated the interaction potential (Finnis & Finnis, Reference Finnis1991; Smith et al. Reference Smith, Chen and Weinert1989). The choice of the image plane position was discussed by Heinz et al. (Reference Heinz, Jha, Luettmer-Strathmann, Farmer and Naik 2011 ) who suggested that the image plane may deviate from the jellium edge (which corresponds to a plane half a lattice spacing above the first layer of surface atoms). The position of this image plane is further away from the surface atoms than the jellium edge and its deviation is represented by a system dependent offset, δ, which is determined by the response of the electron density of the metal (Smith et al. Reference Smith, Chen and Weinert1989). Heinz et al. (Reference Heinz, Jha, Luettmer-Strathmann, Farmer and Naik2011) exploited this scheme to estimate polarization effects a posteriori from non-polarizable fully atomistic MD simulations of a water-Au surface, as well as a water-peptide-gold system.

Fig. 8. Models of polarizable gold surfaces. (a) The method of image charges. Charges q i and q j induce polarization charges inside the metal shown as −q i and −q j respectively. (b) Classical polarization models. Each gold atom of the surface is assigned a dipole with either variable (Drude oscillator) (left) or fixed (rigid rod that is free to rotate) (right) moment. Reproduced with permission from (Iori & Corni, Reference Iori and Corni2008). Copyright (2008) John Wiley and Sons.

Although implementation of the image-charge approximation in an all-atom MD simulation is straightforward, it is impractical for large systems due to the increased computational load that scales with the number of interacting particles. In an alternative approach, which can be incorporated into any commonly used MD energy function, virtual dipoles or rods that can adjust their position in response to the external electrostatic field are introduced on all the surface atoms. Each virtual dipole is constrained at one end to a real surface atom and, depending on the model, can either change the magnitude of its dipole moment (Drude oscillator) or its orientation (rigid rods, see Fig. 8b ) and, hereby screen the external electrostatic field. A model with virtual rigid rods for simulation of the polarization of a metal surface was proposed and implemented by Iori & Corni (Reference Iori and Corni2008) and used in the GolP family of FFs optimized for Au surfaces (Iori et al. Reference Iori, Di Felice, Molinari and Corni2009; Wright et al. Reference Wright, Rodger, Corni and Walsh2013a, Reference Wright, Rodger, Walsh and Cornib), as well as for Ag and graphene.

The first optimized sets of Lennard–Jones parameters to include metal–water and metal–amino acid interactions was developed by Ghiringhelli et al. (Reference Ghiringhelli, Hess, van der Vegt and Delle Site2008) on the basis of DFT calculations, and by Heinz et al. (Reference Heinz, Vaia, Farmer and Naik2008) and Vila Verde et al. (Reference Vertegel, Siegel and Dordick2009) to reproduce Au(111) hydrophilicity. In the GolP FF (Iori et al. Reference Iori, Di Felice, Molinari and Corni2009), which is based on the OPLS FF, additional parameters describing the interactions of biomolecular groups with the gold surface are parameterized from QM calculations and experimental studies of adsorption energy. To reproduce the binding energy and orientation of small molecular fragments, the authors included and optimized a set of additional parameters that describe the van der Waals interactions with gold as well as stronger, chemical-like bonds between aromatic groups and gold atoms in the form of a Lennard–Jones function. This parameterization can be directly used in standard MM force-fields. The GolP FF has been applied in MD simulations of the adsorption of amino acids, as well as several proteins, on gold (Brancolini et al. Reference Brancolini, Kokh, Calzolai, Wade and Corni2012; Cohavi et al. Reference Cole, Payne, Csányi, Mark Spearing and Colombi Ciacchi2011; Hoefling et al. Reference Hoefling, Iori, Corni and Gottschalk2010a, Reference Hoefling, Iori, Corni and Gottschalkb, Reference Hoefling, Monti, Corni and Gottschalk2011; Kokh et al. Reference Kokh, Corni, Winn, Hoefling, Gottschalk and Wade2010). Furthermore, the ProMetCS (Kokh et al. Reference Kokh, Corni, Winn, Hoefling, Gottschalk and Wade2010) continuum solvent model describes protein–surface interactions in atomic detail using the GolP Lennard–Jones interaction parameters together with an image charge model for protein-metal surface electrostatic interactions.

Following the same strategy, a new force-field, GolP-CHARMM (Wright et al. Reference Wright, Rodger, Corni and Walsh2013a, Reference Wright, Rodger, Walsh and Corni2013b), has recently been developed on the basis of the CHARMM FF for studying the interaction of proteins with Au(111) and Au(100) surfaces. In GolP-CHARMM, special attention has been paid to the correct description of the properties of the water molecules adsorbed on gold, with the properties being compared with ab initio MD simulations (Cicero et al. Reference Clementi, Nymeyer and Onuchic2011). The new FF improves the quality of the simulation of the hydration shell, reproducing the increased tendency of interfacial water molecules to donate hydrogen bonds to other water molecules compared with those in the bulk water and reproducing the energetics of water molecules on a gold surface.

Heinz et al. (Reference Heinz, Vaia, Farmer and Naik2008) developed Lennard–Jones parameters for face-centered cubic metals based on experimentally determined densities and surface tensions at 298 K under atmospheric pressure for use in modeling of mixed interfaces. They provide two sets of parameters for metal atoms (Ag, Al, Au, Cu, Ni, Pb, Pd, Pt); one set suited for 12–6 Lennard–Jones functions to be used in AMBER, CHARMM, CVFF and OPLS-AA force-fields, and another for 9–6 Lennard–Jones functions to be used in the COMPASS (Sun, Reference Sun1998) and PCFF force-fields. Penna et al. (Reference Penna, Mijajlovic and Biggs2014) used surface models and the Lennard–Jones parameters of Heinz et al. (Reference Heinz, Vaia, Farmer and Naik2008) with a CHARMM FF (MacKerell et al. Reference MacKerell, Bashford, Bellott, Dunbrack, Evanseck, Field, Fischer, Gao, Guo, Ha, Joseph-McCarthy, Kuchnir, Kuczera, Lau, Mattos, Michnick, Ngo, Nguyen, Prodhom, Reiher, Roux, Schlenkrich, Smith, Stote, Straub, Watanabe, Wiórkiewicz-Kuczera, Yin and Karplus1998) to investigate the interactions of peptides with Au and Pt surfaces.

In addition to polarization, surface charges should be taken into account when modeling a surface. Bizzarri (Reference Bizzarri2006) investigated the dynamics and electron transfer (ET) properties of the bacterial ET protein, azurin, anchored to neutral, positively and negatively charged gold surfaces using classical MD simulations. They calculated the ET rate between the azurin copper atom and the sulfur atom of the cysteine surface anchor using classical Marcus theory (Marcus & Sutin, Reference Marcus and Sutin1985). The computed ET rate was highest on the positively charged surface and lowest on the neutral surface. However, it was not clear to what extent the relatively larger structural changes observed in the molecule anchored on the positively charged surface play a role in this trend.

7.2 Titanium oxide surfaces

Because of its high stability and resistance to corrosion, titanium is the material of choice for many medical and technical applications and it is, therefore, one of the most widely used materials in computational modeling. Crystals of TiO2 (titania) appear in three forms: rutile, anatase and brookite. The rutile form, in particular its interaction with water, has been the most studied by computer modeling. An oxide layer, predominantly of amorphous structure, is formed at the titanium-water interface (Schneider & Colombi Ciacchi, Reference Schneider and Colombi Ciacchi2010). The properties of this layer depend strongly on the physico-chemical conditions and the temperature regime of adsorption.

There is controversy between theory and experiment regarding the extent to which water dissociates on TiOx; simulations often show a larger degree of dissociation than experiment (Diebold, Reference Diebold2003; Huang et al. Reference Huang, Gubbins, Li and Lu2014; Sun et al. Reference Sun, Liu, Selloni, Lu and Smith2010). It is believed that for rutile (110), dissociation takes place predominantly on defects, while water can also dissociate to some extent on a defect-free (100) surface, and effectively dissociates on a (001) surface. Accordingly, the density distribution of water strongly varies for different TiO2 surfaces from two sharp maxima on a rutile surface to a smooth distribution on hydrophobic titania (Huang et al. Reference Huang, Gubbins, Li and Lu2014). The mixture of hydroxyl groups and molecularly adsorbed water may form a partially ordered solvation layer that influences interactions of peptides with the surface (Li et al. Reference Li, Monti and Carravetta2012).

The behavior of water on different types of TiO2 surfaces, including partial dissociation of water, was investigated using the reactive FF ReaxFF (Huang et al. Reference Huang, Gubbins, Li and Lu2014; Kim et al. Reference Kim, Kumar, Persson, Sofo, van Duin and Kubicki2013) and AIMD (Carravetta et al. Reference Carravetta, Monti and Zhang2009; Schneider & Colombi Ciacchi, Reference Schneider and Colombi Ciacchi2011). More details on theoretical studies of titania–water interactions can be found in the review by Sun et al. (Reference Sun, Liu, Selloni, Lu and Smith2010).

As mentioned before, water dissociation on oxide surfaces is reversible and cannot be simply described by classical MM force-fields. Bandura & Kubicki (Reference Bandura and Kubicki2003) pointed out that the two-body interaction potential is not suitable for capturing the correct directionality of hydroxyl groups on a TiO2 surface and that this problem can be solved by using a polarizable FF or by introducing bond stretching and angle bending terms for Ti-O-H. They used the Matsui & Akaogi (Reference Matsui and Akaogi1991) model which incorporated additional bond-stretching and angle-bending terms for O-Ti-H into the Buckingham potential for the atoms of the (110) surface of rutile. The FF was later refined by Předota et al. (Reference Předota, Bandura, Cummings, Kubicki, Wesolowski, Chialvo and Machesky2004) and then revised for anatase (101), (112) and rutile (110) surfaces. In the FF developed by Borodin et al. (Reference Borodin, Smith, Bandyopadhyaya and Byutner2003), an exponential term for modeling short-range repulsion was added to the standard Lennard–Jones function together with an additional polarization term, and the FF was optimized against the results of QM calculations for polyethylenoxide on small TiO5H9 clusters and then applied to TiO2 surfaces.

Since the use of AIMD and reactive FFs is too time-consuming for simulations of large systems and/or long time-scales, TiOx parameters have been incorporated into standard MM FFs employed in biomolecular simulations. A classical FF with a Finnis–Sinclairtype many-body potential for the surface coupled to a Buckingham potential for the short-range repulsive interactions was developed by Schneider and Colombi Ciacchi and applied to simulating peptide interactions on a Ti/TiOx interface (Schneider & Colombi Ciacchi, Reference Schneider and Colombi Ciacchi2010). The parameters of the AMBER FF were optimized for TiO2 interactions with collagen peptides in a study by Köppen & Langel (Reference Köppen and Langel2010). The ratio of protonated, hydroxylated and stoichiometric units on the surface was adapted to the physiological pH value and retained during MD simulations. The charges in the bulk of TiO2 were fixed, but different charges on the surface oxygen and hydrogen atoms were tested. The binding of Glu and Lys to TiO2 was found to be mediated by hydroxyl groups on the surface, with the adsorption energy strongly dependent on the charge model used.

Friedrichs & Langel (Reference Friedrichs and Langel2014) re-parameterized the Matsui–Akaogi potential to the standard Lennard–Jones form for a description of the interactions between the rutile surface atoms and used their model to simulate adsorption of peptides on the rutile surface. Although this enabled standard MD software to be used, the simplicity of the form of the potential gave rise to some deviation of physical properties, such as crystal structure parameters and permittivity, from those observed in experiments and simulations with the original potential.

The adsorption of different peptides onto a rutile surface, TiO2(110), was simulated in a series of studies by Monti and colleagues of the Ala-Glu and Ala-Lys dipeptides (Carravetta & Monti, Reference Carravetta and Monti2006) and (Monti et al. Reference Monti, Carravetta, Battocchio, Iucci and Polzonetti2008), and β-sheet conformations of the EAK16 and RAD16 oligopeptides (Carravetta et al. Reference Carravetta, Monti and Zhang2009). The FF for biomolecule-surface and water–surface interactions used in these studies was optimized against data from QM calculations for a small titania cluster and experimental data (Carravetta & Monti, Reference Carravetta and Monti2006). The TIP3P water model and a combination of a Buckingham potential for the interaction between the surface and surface-water atoms and a Lennard–Jones potential for most of the rest of the atoms were employed. Peptide binding to the surface was found to preferentially occur through the backbone atoms; particularly carboxyl oxygens and the C-terminal carboxy group on the peptides interacted with exposed Ti atoms. Water molecules were also observed to mediate peptide-surface binding through hydrogen bonds. Since the dissociation of water was not taken into account, the contribution of hydroxyl groups to peptide binding was not analyzed. Finally, the adsorption of glycine (Li et al. Reference Li, Monti and Carravetta2012), and cysteine (Li et al. Reference Li, Monti, Ågren and Carravetta2014) onto TiO2 was explored using the reactive force-field, Reaxff (Van Duin et al. Reference Van Duin, Dasgupta, Lorant and Goddard2001), extended to treat the interaction of solid inorganic substrates with biomolecules. It is interesting that these adsorbing molecules, especially glycine, tend to form self-assembled clusters and that only weak adsorption was observed. Additionally, the proton transfer reactivity of the amino acid is observed to be enhanced by the presence of the surface.

It should be noted that MM simulations of proteins on reactive surfaces like titanium oxide should be performed with caution, because the presence of defects may change the surface properties dramatically, and thus affect the adsorption behavior. In a study by Utesch et al. (Reference Utesch, Daminelli and Mroginski2011), the BMP-2 protein interacted unexpectedly weakly with the TiO2 surface, because it was unable to penetrate the highly-ordered water layer formed on the defectless surface.

7.3 Silicon oxide surfaces

A number of FFs have been developed for MD simulation of bulk silicon oxides and their interfaces with water due to the wide range of technological applications of silica-based materials (see reviews by Butenuth et al. (Reference Butenuth, Moras, Schneider, Koleini, Köppen, Meißner, Wright, Walsh and Colombi Ciacchi2012) and by Rimola et al. (Reference Rimola, Costa, Sodupe, Lambert and Ugliengo2013)). There are, however, some challenges in building a FF for silicon oxide surfaces. In particular, for modeling the tetragonal bulk structure of SiO2 (silica), many-body potentials are often used, which are generally not compatible with common biomolecular force-fields. At the SiOx-water interface, water molecules dissociate and saturate dangling Si and O groups, which leads to formation of different types of exposed functional groups: hydrophilic silanols (Si-OH), and hydrophobic siloxanes (with oxygen buried, Si-O-Si) and silanes (with silica atoms saturated by hydrogen). Silanol groups can make hydrogen bonds to each other or two silanol hydroxyl groups can bind to one Si atom. Since they are hydrophilic in character, these groups form hydrogen bonds with surface water molecules. Crystalline silica has a dense ordered network of geminal hydrogen-bonded silanol groups leading to ordered water layers (Notman & Walsh, Reference Notman and Walsh2009) (see Fig. 9), while amorphous silica has isolated geminal and hydrogen-bonded silanol groups resulting in only locally ordered regions of water molecules and hydroxyl groups (Aarts et al. Reference Aarts, Pipino, Hoefnagels, Kessels and van de Sanden2005).

Fig. 9. Lateral density profiles of the first layer of structured water on the (101-0), (0001) and (011-1) surfaces of silica. Reprinted with permission from (Notman & Walsh, Reference Notman and Walsh2009). Copyright (2009) American Chemical Society. Positions of the water molecules are dominated by the positions of the silanol hydroxyl groups and to a lesser extent by the underlying crystal structure.

Silica has an isoelectric point of about 2–3 (Parks, Reference Parks1965), and, at neutral pH, between 5% and 20% of silanol groups are deprotonated, leading to a negative surface charge. Experimental studies of peptide adsorption on silica nanoparticles in aqueous solution (Patwardhan et al. Reference Patwardhan, Emami, Berry, Jones, Naik, Deschaume, Heinz and Perry2012; Puddu & Perry, Reference Puddu and Perry2012) showed that binding was dominated by electrostatic interactions between negatively charged siloxide groups and ammonium groups at N-termini, and Lys and Arg containing peptides and, to a lesser extent, hydrogen bonds between negatively charged silanol and siloxide groups and polar groups in Ser, His and Asp-containing peptides. Binding of hydrophobic peptides was observed at low pH (Puddu & Perry, Reference Puddu and Perry2012). A FF for MD simulations of a deprotonated negatively charged SiO2 surface in solution at neutral pH was reported by Butenuth et al. (Reference Butenuth, Moras, Schneider, Koleini, Köppen, Meißner, Wright, Walsh and Colombi Ciacchi2012).

Recently, several FFs for silica compatible with biomolecular FFs have been developed. A FF based on the CHARMM FF was parameterized by Lopes et al. (Reference Lopes, Murashov, Tazi, Demchuk and MacKerell2006) to obtain a structural and dynamic representation of water in the vicinity of neutral quartz crystalline surfaces using MD simulations. This FF was used for a set of peptides known to be strong and weak binders to the quartz (100) surface and showed that Pro, Trp and Leu were the main residues forming close contacts with the surface (Oren et al. Reference Oren, Notman, Kim, Evans, Walsh, Samudrala, Tamerler and Sarikaya2010).

The GLASSFF_2·01 FF is compatible with CHARMM and with TIP3P water parameters and was developed by Cruz-Chu et al. (Reference Cruz-Chu, Aksimentiev and Schulten2006) for simulation of amorphous silica surfaces and nanopores. This FF employs standard two-body Lennard–Jones and Coulomb potentials as well as a three-body directional term for modeling the tetrahedral arrangement of bulk silica. To construct amorphous silica, the authors applied annealing cycles in MD simulations, which initiated surface reconstruction from crystal to amorphous silica accompanied by formation of dangling atoms (oxygens with less than two bonds and silicons with less than four bonds) that were required for modeling the wetting properties of silica correctly.

A comparison of FFs for the prediction of the physical and chemical properties of bulk silica and its aqueous interface such as atomic charge, bond length and angles, density of silanol groups on the surface and degree of ionization of silanol groups can be found in the papers by Skelton et al. (Reference Skelton, Fenter, Kubicki, Wesolowski and Cummings2011) and by Emami et al. (Reference Emami, Puddu, Berry, Varshney, Patwardhan, Perry and Heinz2014a). In the latter study, a data set with more than 20 different silica surface models that cover most important types of silica surface chemistry at different surface ionization states was assembled. Then, atomic parameters for the different chemical groups in silica were derived to reproduce a large set of chemical and physical properties of bulk and surface silica as well as the water contact angle, heat of water immersion and water adsorption isotherms observed in experiments. The parameters were integrated into several FFs (AMBER, CHARMM, COMPASS, INTERFACE (Heinz et al. Reference Heinz, Lin, Kishore Mishra and Emami2013)) compatible with the TIP3P and SPC water models. In particular, the CHARMM-INTERFACE FF (Heinz et al. Reference Heinz, Lin, Kishore Mishra and Emami2013) was applied to simulation of peptide adsorption on silica nanoparticles of different size at different pH values (Emami et al. Reference Emami, Puddu, Berry, Varshney, Patwardhan, Perry and Heinz2014b), showing good agreement of nanoparticle coverage with experimental data (Puddu & Perry, Reference Puddu and Perry2012).

The Dual-FF  (Biswas et al. Reference Biswas, Vellore, Yancey, Kucukkal, Collier, Brooks, Stuart and Latour2012), (discussed in more detail in Sec.7·5) was optimized to reproduce the experimental binding energies of TGTG-X-GTGT peptides, with X = N, D, G, K, F,T, W and V, on hydroxylated quartz (100) and glass surfaces in MD simulations (Snyder et al. Reference Snyder, Abramyan, Yancey, Thyparambil, Wei, Stuart and Latour2012). In the Dual-FF, CHARMM parameters were used for biomolecules, while interfacial parameters, such as partial charges and Lennard–Jones parameters for silica–peptide and silica–water interactions, were tuned to obtain the best agreement with peptide adsorption energies.

7.4 Mineral surfaces

The main difficulty in simulating mineral-biomolecule interfaces in aqueous solution is that, while both minerals and biomolecules may have atom types in common, e.g. oxygen, the FF for bonded terms are different in minerals and biomolecules. Specifically, minerals are ionic systems and are best represented by ionic bonds based on Coulomb electrostatics and Pauli repulsion terms (Hauptmann et al. Reference Hauptmann, Dufner, Brickmann, Kast and Berry2003). In contrast, biomolecules are covalently bonded. As for oxide surfaces, water may dissociate on a mineral surface leaving adsorbed hydroxyl groups at charged group sites on the surface with a partial occupancy of about 50% of the charged sites (De Leeuw, Reference De Leeuw2010). This process cannot be modeled in a standard MM FF, but the presence of hydroxyl groups/hydrogens on the surface may strongly affect the adsorption properties of biomolecules.

A number of FFs for modeling the bulk and surface properties of minerals have been developed (see (Mishra et al. Reference Mishra, Flatt and Heinz2013) and references therein), and several non-reactive FFs for the simulation of mineral-water interfaces have been reported. A general FF, CLAYFF, for modeling the interface of a hydrated multicomponent mineral with aqueous solution was proposed by Cygan et al. (Reference Cygan, Liang and Kalinichev2004), where charges on oxygen atoms and hydroxyl groups vary according to their structural environment. This FF includes a flexible SPC-based model for water and hydroxyl groups and treats most interatomic interactions as non-bonded with Lennard–Jones and Coulomb terms.

A methodology for the generation of a general bio-compatible FF for minerals was proposed by Freeman et al. (Reference Freeman, Harding, Cooke, Elliott, Lardge and Duffy2007). In this work, existing FFs and models were used for the various components of the system (including the TIP3P model for water), while new parameters were derived for cross-species interaction terms, for example, for the interaction of a hydroxyl group with a Ca2+ ion. This methodology was used for simulation of small organic molecules on calcite and magnesite surfaces (Freeman et al. Reference Freeman, Asteriadis, Yang and Harding2009) by combining the AMBER FF for organic molecules with inter-species terms proposed by Freeman et al. (Reference Freeman, Harding, Cooke, Elliott, Lardge and Duffy2007); and for simulation of the binding of ovocleidin-17 protein to calcium carbonate nanoparticles (Freeman et al. Reference Freeman, Harding, Quigley and Rodger2010) and surfaces (Freeman et al. Reference Freeman, Harding, Quigley and Rodger2011). Furthermore, Katti et al. reported parameters for modeling clay minerals (NaSi16(Al6FeMg)O40(OH)8) in the CHARMM FF (Katti et al. Reference Katti, Schmidt, Ghosh and Katti2005b) and amino-acid adsorption to these clay minerals (Katti et al. Reference Katti, Ghosh, Schmidt and Katti2005a).

Hydroxyapatite (HAP, Ca10 (PO4)6 (OH)2) is the main mineral component of bones and teeth and thus a promising material for application in bone replacement. A FF for HAP has been developed with a Born–Mayer–Huggins potential with an exponential repulsion term similar to that of Buckingham potential instead of the standard Lennard–Jones plus Coulomb expression for the non-bonded interactions. The FF parameters were derived by fitting the QM electrostatic field in a 3D crystal environment and experimental crystal parameters (Hauptmann et al. Reference Hauptmann, Dufner, Brickmann, Kast and Berry2003). This FF was subsequently re-parameterized for monoclinic hydroxyapatite using an energy function consistent with common biomolecular FFs (Bhowmik et al. Reference Bhowmik, Katti and Katti2007). Dong et al. (Reference Dong, Wang, Wu and Pan2007) adapted these HAP parameters for the standard biomolecule FF for modeling the adsorption dynamics of the BMP-2 protein on a HAP (001) surface. In other studies, the Lennard–Jones and Coulomb parameters proposed by Hauptmann et al. (Reference Hauptmann, Dufner, Brickmann, Kast and Berry2003) were combined with CHARMM parameters for the protein using the standard mixing rule in order to study the adsorption of fibronectin on HAP (001) (Shen et al. Reference Shen, Wu, Wang and Pan2008) and with OPLS-AA parameters for simulating the HAP-water interface (Zahn & Hochrein, Reference Zahn and Hochrein2003).

7.5 Self-assembled monolayer surfaces

The applicability of several commonly used biomolecular FFs for the simulation of peptide adsorption on hydrophobic and hydrophilic SAMs has been systematically evaluated by Latour and colleagues for CHARMM19 (Sun & Latour, Reference Sun and Latour2006; Sun et al. Reference Sun, Dominy and Latour2007; Vellore et al. Reference Vellore, Yancey, Collier, Latour and Stuart2010), and CHARMM22, OPLS-AA, and AMBER94 (Collier et al. Reference Colombi Ciacchi and Payne2012) FFs. In the latter study by Collier et al. (Reference Colombi Ciacchi and Payne2012), the FF parameters and partial charges of the SAMs (-OH and -COOH) were assigned by analogy to amino acids with similar functional groups. The authors compared experimental findings with their computed free energies of adsorption and qualitative behavior of peptides on different SAM surfaces such as the change in conformation upon adsorption and the orientation on the surface. This study (Collier et al. Reference Colombi Ciacchi and Payne2012) demonstrated that although some FFs perform reasonably well (the best match to experiment was obtained in simulations using CHARMM22 and AMBER94 (Cornell et al. Reference Cornell, Cieplak, Bayly, Gould, Merz, Ferguson, Spellmeyer, Fox, Caldwell and Kollman1995)), none of the FFs capture the specific interaction properties of the SAM-water interface. In particular, systematic overestimation of the binding strength of hydrophobic peptides and underestimation for negatively charged peptides was observed. On the other hand, Collier et al. (Reference Colombi Ciacchi and Payne2012) also noted that altering FF parameters to reproduce the properties of adsorbed peptides unavoidably led to alteration of peptide behavior in solution. These results suggested that for accurate simulations of peptide adsorption a new FF strategy was required.

A new approach, the dual-FF, was proposed by Biswas et al. (Reference Biswas, Vellore, Yancey, Kucukkal, Collier, Brooks, Stuart and Latour2012). In this FF, different sets of non-bonded parameters, i.e. atomic partial charges and parameters of the Lennard–Jones potential, were used to represent intra-phase (i.e. between peptides or between water molecules) and inter-phase (peptide–SAM) interactions simultaneously in simulations. Furthermore, the non-polarizable TIP3P (Jorgensen et al. Reference Jorgensen, Chandrasekhar, Madura, Impey and Klein1983) water model was reparameterized for water-SAM interfaces based on the electrostatic properties of water molecules in the interface computed using the polarizable TIP4P-FQ (Rick et al. Reference Rick, Stuart and Berne1994) FF. In addition, extended experimental benchmarks of binding free energies of peptides with different sequences on various functionalized SAM surface interfaces by Wei & Latour (Reference Wei and Latour2009, Reference Wei and Latour2010) were used for optimization of the parameters for amino-acids.

Utesch et al. (Reference Utesch, Sezer, Weidinger and Mroginski2012) studied the adsorption of sulfite oxidase (SO) on a functionalized SAM surface under different ionic strength conditions using the CHARMM32 FF (MacKerell et al. Reference MacKerell, Bashford, Bellott, Dunbrack, Evanseck, Field, Fischer, Gao, Guo, Ha, Joseph-McCarthy, Kuchnir, Kuczera, Lau, Mattos, Michnick, Ngo, Nguyen, Prodhom, Reiher, Roux, Schlenkrich, Smith, Stote, Straub, Watanabe, Wiórkiewicz-Kuczera, Yin and Karplus1998) for parameters of both the protein and the SAM surface. The catalytic activity of SO was earlier experimentally observed to be dependent on the flexibility of the tether connecting the molybdenum cofactor (Moco) domain and the cytochrome b5 domain of SO. In line with experimental findings, their results indicated that the adsorption of the enzyme's cytochrome b5 domain is inhibited at high ionic strength (750 mM), whereas under much lower ionic strength conditions (100 mM), stable interactions with the surface take place, leading to a loss of flexibility of the tether.

Effects of water on adsorption on SAMs have been reported in several studies. Residence times (τ F) and self-diffusion coefficients (DF) of interface water molecules on several different functionalized SAM surfaces of type S(CH2)-F were inspected by Wang et al. (Reference Wang, Zhao, Zhao, Wang, Yang, Yu and Zheng2010a, Reference Wang, Zhao, Yu, Zhao, Li and Zhengb) and found to be in the order: τ COOH > τ NH2 > τ OH > τ CH3 > τ bulk and DCOOH < DOH < DNH2 < DCH3 < Dbulk. This ranking was found to be directly correlated with the number of hydrogen bonds formed between the interfacial water and the modified SAM surfaces. Further, it was shown that the structure and the dynamics of the water determine the adsorption behavior of amyloid-β peptide on the modified SAM surfaces and that the energy barrier for the adsorption of the peptide is lower on hydrophobic CH3-SAM than on the hydrophilic OH-SAM.

The mobility of interfacial water on the zwitterionic sulfobetaine (SBT)-SAM surface was investigated in a separate study by Xie et al. (Reference Xie, Liu and Zhou2012). They found that the lower density packing and the higher flexibility of the SBT-SAM allowed water molecules to penetrate into the surface, thereby increasing the number of interactions of SBT-SAM with the water molecules and decreasing the mobility of the interface water molecules significantly compared with that of pure bulk water (DSBT < Dbulk/20). The mean force-distance profile showed that the affinity of the SBT-SAM surface towards neuromedin-B peptide was smaller than for the other two SAM surfaces (CH3-, OH-SAM). The observations support the experimental results showing the antifouling properties of SBT-SAM surfaces (Xie et al. Reference Xie, Liu and Zhou2012).

7.6 sp2-Carbon surfaces

The ability of a biomolecular FF to predict the interaction properties of a single water molecule described by a standard TIP3P or TIP4P model, and several ions on CNT and fullerene surfaces was explored by Schyman & Jorgensen (Reference Schyman and Jorgensen2013). Adsorption energies, as well as the binding site on the surface and the orientation of the water molecule obtained from the simulations using MM FFs, were compared with those of DFT calculations. The non-polarizable OPLS-AA FF was found to be adequate for the description of water properties on benzene (C6H6) and coronene (C24H12). However, with increasing surface size above (C54H18) and especially for C60, the simulations became inaccurate, indicating that surface polarization played an important role. Indeed, a polarizable biomolecular FF with induced dipoles on all non-hydrogen atoms (OPLS-AAP) was shown to yield reasonably good agreement with QM calculations for the interaction properties of water molecules and ions.

Another polarizable FF, AMOEBA (Ren & Ponder, Reference Ren and Ponder2002, Reference Ren and Ponder2003), has been extended for the description of the non-bonded interactions between peptides and CNT (De Miranda Tomasio & Walsh, Reference De Miranda Tomásio and Walsh2007) and graphite (Walsh, Reference Walsh2008), and exploited for simulations of peptide binding on CNT (Tomásio et al. Reference Tomásio and Walsh2009). These studies demonstrated that surface polarization caused a strong interaction between the carbon surface and aromatic rings of the side chains of the residues, particularly of tryptophan, placed in the surface plane, which caused strong binding of peptides rich in aromatic residues. The binding of peptides on CNT was also shown to be affected by the interplay between the ability of the aromatic rings to align on a non-flat surface and the total peptide-surface contact area that is increased due to surface curvature. The importance of π-stacking between aromatic side chains and carbon surfaces was also shown in a study of peptide adsorption on a single-layer graphene substrate by Akdim et al. (Reference Akdim, Pachter, Kim, Naik, Walsh, Trohalaki, Hong, Kuang and Farmer2013) using the polarizable AMOEBABIO-09 FF implemented in TINKER 5·1. (Saint Louis, Washington. Software Tools for Molecular Design.). Although these studies demonstrate that polarizable MM FFs are able to provide reasonably accurate descriptions of the interaction properties of peptides on carbon substrates, such simulations are still computationally very demanding for the use of protein adsorption simulations. A viable approach is the inclusion of polarizability in graphene simulations via the rigid rod model, which was employed by Hughes et al. (Reference Hughes, Tomásio and Walsh2014) to calculate the adsorption free energies of single amino acids on a aqueous graphene surface.

The role of the water in peptide binding to a graphene surface was explored by Camden et al. (Reference Camden, Barr and Berry2013), who simulated GXG tripeptides for 20 amino acid residues (X) with an explicit water model. In the TEAM FF used in these simulations, the carbon atoms of the graphene surface were not polarizable, and thus important polarization-driven interactions between aromatic groups and the surface were missing. However, this study demonstrated that desolvation effects may play a very important role in peptide binding to a graphene surface. The dynamics of water on carbon-based hydrophilic graphite-COOH and hydrophobic graphite-CH3 surfaces was studied by Li et al. (Reference Li, Liu, Li, Ye, Chen, Fang, Wu and Zhou2005) using the standard OPLS-AA FF. They observed that the diffusion of water was slowed dramatically in the vicinity of the surfaces, and this effect extended up to 15 Å from the surface. The water diffusion coefficient within a 3 Å distance from a hydrophilic surface (-COOH) was 4 times smaller than that of a hydrophobic surface (-CH3).

Implicit solvent models offer an alternative way to explore biomolecule-carbon interfaces in aqueous solution. It can be expected that the hydration shell on carbon surfaces has a much less defined structure than on hydrophilic surfaces and thus, is better suited for a continuum model. Indeed, no strong influence of structural water on the conformation of a test peptide was observed in evaluation of implicit and explicit water models (Walther et al. Reference Walther, Jaffe, Halicioglu and Koumoutsakos2001). In an implicit solvent model, however, one has to account for hydrophobic desolvation effects, which may promote binding of biomolecules to carbon surfaces, explicitly. This was demonstrated by Mereghetti & Wade (Reference Mereghetti and Wade2011) in a study of hydrophobin adsorption on graphite.

Raffaini and Ganazzoli reported simulations of protein adsorption on graphite surfaces in a series of studies using the consistent valence FF (CVFF) (Dauber-Osguthorpe et al. Reference Dauber-Osguthorpe, Roberts, Osguthorpe, Wolff, Genest and Hagler1988) with a Morse potential applied to bonded interactions (Raffaini & Ganazzoli, Reference Raffaini and Ganazzoli2003, Reference Raffaini and Ganazzoli2004a, Reference Raffaini and Ganazzolib, Reference Raffaini and Ganazzoli2006, Reference Raffaini and Ganazzoli2007, Reference Raffaini and Ganazzoli2010). In their simulations, many proteins were observed to denature partially or completely upon adsorption to the hydrophobic surface.

Raffaini & Ganazzoli (Reference Raffaini and Ganazzoli2010) also carried out kinetic calculations for lysozyme spreading on a graphite surface by fitting the time evolution of the interaction energies, protein-surface center of mass distance and components of the radius of gyration of the protein to a function consisting of an exponential term and a stretched exponential term (Kohlrausch function):where γ stands for an arbitrary quantity that decreases with time (t), t 0 and t 1 are relaxation times, and A, B, C and δ are fitting parameters. In one simulation, they found that, upon adsorption, lysozyme underwent a liquid-like spreading on the surface suggested by a very fast initial decrease of distance between the center of mass of the protein and the surface with t 0 = 28 ps and by loss of most of the secondary structure. The initial kinetic step was followed by a longer stage with a t 1 ~ 1 ns in which all the secondary structure was lost. The authors point out that the fast kinetics of spreading, observed in the study may be attributable to the implicit solvent model used in the simulations.

8. Coarse-Grained molecular mechanics modeling of protein–surface interactions

Protein adhesion on a surface is often a combination of hierarchical processes occurring on a wide range of different time-scales. Moreover, an understanding of adhesion processes in real life applications often requires considering the interplay between protein–protein and surface-protein interactions. Thus, increasing the time- and length-scales of simulations beyond the all-atom MD limits, while keeping the atomic details of the protein-water-surface interface in the simulation model, is required. Coarse-graining methodologies offer a trade-off between computational speed and accuracy (Tozzini, Reference Tozzini2005). However, apart from the established limitations of the CG models developed for biological molecules, there are other fundamental problems in developing and applying a CG FF for adsorption processes. One is the reliable modeling of the hydration shell on a surface that depends strongly on the chemistry of the surface and the surrounding solution. Taking into account that even standard all-atom FFs have notable deficiencies for modeling the surface-water interface, developing a CG water model to satisfactorily describe all accessible water states, and yet reduce the number of degrees of freedom, is challenging.

CG simulations are often employed in studies of adsorption of model homopolymer chains on surfaces. Due to the greater structural and dynamic complexities of peptides and proteins, the physical characteristics obtained from these polymer simulations cannot be directly extrapolated to peptide adsorption and spreading on a surface. To build a simplified description for proteins, CG models of protein-like AB polymers, in which each of the monomers is represented with types A (with hydrophilic properties) or type B (with hydrophobic properties) beads, have been introduced and applied to study adsorption (Liu & Chakrabarti, Reference Liu and Chakrabarti1999).

Studies of CG models with improved system descriptions have been reported for protein adsorption on flat surfaces and on nanoparticles. The binding of the negatively charged proline rich salivary protein, PRP-1, to a negatively charged surface was studied by MC simulation using a model in which each residue was described by one bead that was either positively or negatively charged, or uncharged (Skepö et al. Reference Skepö, Linse and Arnebrant2006). The protein changes conformation depending on salt concentration, and therefore, an implicit solvent model was used in which ions were treated explicitly. The effects of salt on protein conformation and consequently on adsorption properties were investigated in the simulations. Similar united residue models for peptide chains were used in several other papers (Evers et al. Reference Evers, Andersson, Lund and Skepö2012; Knotts et al. Reference Knotts, Rathore and de Pablo2005, Reference Knotts, Rathore and de Pablo2008; Skepö, Reference Skepö2008; Xie et al. Reference Xie, Zhou and Jiang2010). In two of these studies, short-range attraction potentials were added to model the binding of statherin protein on pure hydrophobic, charged, and mixed surfaces (Skepö, Reference Skepö2008), and the binding of β-casein on hydrophobic and charged surfaces (Evers et al. Reference Evers, Andersson, Lund and Skepö2012). Further, MC simulations of bovine serum albumin (BSA) and HSA on silver nanoparticles using a CG model elucidated conformational changes of the protein upon adsorption, which were found to be in agreement with measured tryptophan adsorption spectra (Voicescu et al. Reference Voicescu, Ionescu and Angelescu2012).

Tavanti et al. applied a Gō-type model, which takes only the interactions that are present in a molecule in its natively-folded structure into account, to study corona formation by ubiquitin molecules on gold nanoparticles of various sizes (10, 16, 20 and 24 nm diameter) in the presence of both implicit and explicit citrate (Tavanti et al. Reference Tavanti, Pedone and Menziani2015). They used a FF and parameters developed originally for protein and RNA folding described by (Clementi et al. Reference Cohavi, Corni, De Rienzo, Di Felice, Gottschalk, Hoefling, Kokh, Molinari, Schreiber, Vaskevich and Wade2000; Pincus et al. Reference Pincus, Cho, Hyeon, Thirumalai and Conn2008) for ubiquitin, and the parameters described by Ding et al. (Reference Ding, Radic, Chen, Chen, Geitner, Brown and Ke2013) for citrate. With this setup, they showed that the aggregation characteristics of the proteins depend on the size of the nanoparticle and that a loss in secondary structure of the proteins is more prominent in nanoparticles of smaller size, i.e. 10 and 16 nm diameter. Further, they observed that the protein corona starts forming together with a slow phase of protein reorientation on the nanoparticle surface to optimize interactions on the nanoparticle surface.

In several studies, the MARTINI FF (Marrink et al. Reference Marrink, Risselada, Yefimov, Tieleman and de Vries2007; Seo et al. Reference Seo, Rauscher, Pomès and Tieleman2012), in which each residue is described by several beads, was employed to investigate the adsorption of proteins to surfaces (Griepernau et al. Reference Griepernau, Hanke, Santen, B, Hansmann, Meinke, Mohanty, Nadler, Hansmann, Meinke, Mohanty, Nadler and Zimmermann2008; Liang et al. Reference Liang, Fieg, Keil and Jakobtorweihen2012) and to nanoparticles (Hung et al. Reference Hung, Mwenifumbo, Mager, Kuna, Stellacci, Yarovsky and Stevens2011). MARTINI, which was originally developed for lipid bilayers, has been extended for simulation of proteins in solvent by applying an elastic network model to keep the protein near the native protein structure.

Most of the studies mentioned above employed CG FFs evaluated for biomolecules or surfaces alone, and hence their accuracy for modeling the protein-solvent-surface interface was not evaluated. Moreover, none of these CG models represents all the effects of the surface hydration shell that can be explored in atomic detail MD simulations. A step towards the inclusion of such solvent effects was performed by Carrillo-Parramon et al. (Reference Carrillo-Parramon, Brancolini and Corni2013), who exploited the surface desolvation terms as defined in the ProMetCS implicit solvent FF in an essential dynamics CG model of ubiquitin interacting with a gold surface. Their model reproduced fluctuation characteristics of the protein obtained with classical atomistic MD.

Models developed specifically for protein structures near interfaces have been proposed in several studies. Bhirde et al. (Reference Bhirde, Hassan, Harr and Chen2014) modeled the aggregation of gold nanoparticles in an albumin solution using a rigid-body CG model of inter–nanoparticle, inter- and intra-protein and nanoparticle–protein interactions consisting of a 12–6 Lennard–Jones term for all three types of interactions and an electrostatic term for protein–protein interactions based on the screened Coulomb potential implicit solvent model (SCPISM) (Cardone et al. Reference Cardone, Pant and Hassan2013; Hassan & Steinbach, Reference Hassan and Steinbach2011). They investigated the role of nanoparticle size and coating of its surface in adsorption of albumin by performing a number of MC simulations with different resolutions of coarse-graining and different values of the Lennard–Jones parameters (σ and ε) and the nanoparticle concentration. This study showed that the albumin molecules formed aggregates with the nanoparticles, and the stability of these aggregates depended on the surface coating characteristics and the nanoparticle concentration. Wu and Narsimhan proposed an implicit solvent CG model in which 1–3 beads were used per protein residue, and applied it to the unfolding of lysozyme on a neutral silica surface, investigating the dependence on temperature and ionic strength (Wu & Narsimhan, Reference Wu and Narsimhan2009). The authors evaluated their approach by comparing with all-atom MD simulations of protein dynamics in solution.

Ravichandran & Talbot (Reference Ravichandran and Talbot2000) investigated the kinetics of adsorption and the structure of the adsorbed layer of hen egg white lysozyme (HEWL) on a solid surface. The HEWL molecules were represented as spheres with a uniform charge distribution on each of the molecules. The authors pointed out that the simplified description of the molecules would not work with proteins that are not globular, or that have an anisotropic charge distribution. In a separate study, Ravichandran et al. (Reference Ravichandran, Madura and Talbot2001) investigated the early stages of the adsorption process and the binding orientations of HEWL on a positively charged surface with a full atomic detail representation of the protein. Although HEWL has an overall positive charge, it could bind to the charged surface due to the non-uniform distribution of the charges on the protein. Their result shows the necessity of less CG representations in certain systems.

9. Applications of sampling methods to protein–surface interactions

Simulations of biological molecules give many insights into the molecular mechanisms of interactions and association, as well as their thermodynamic properties and kinetics. To perform a simulation, descriptions of the structures present in the system (e.g. a set of coordinates of the atoms for all-atom models), a reliable FF and a sampling method are required. The sampling technique used in a simulation should be chosen according to the properties of the system to be investigated. In this section, we will discuss advantages and disadvantages of conventional and enhanced sampling techniques in the context of their application to protein–surface interactions.

9.1 Molecular dynamics

Among the different sampling methods available, classical MD is by far the most widely used technique to investigate the characteristics of protein–surface interactions. Examples include simulations for structural refinement of docked complexes (Aliaga et al. Reference Aliaga, Ahumada, Sepúlveda, Gomez-Jeria, Garrido, Weiss-López and Campos-Vallette2011; Alvarez-Paggi et al. Reference Alvarez-Paggi, Castro, Tórtora, Castro, Radi and Murgida2013; Brancolini et al. Reference Brancolini, Kokh, Calzolai, Wade and Corni2012, Reference Brancolini, Corazza, Vuano, Fogolari, Mimmi, Bellotti, Stoppini, Corni and Esposito2015; Imamura et al. Reference Imamura, Kawasaki, Nagayasu, Sakiyama and Nakanishi2007), and for the investigation of the binding orientations of proteins on surfaces (Alvarez-Paggi et al. Reference Alvarez-Paggi, Martín, DeBiase, Hildebrandt, Martí and Murgida2010; Boughton et al. Reference Boughton, Andricioaei and Chen2010; Coppage et al. Reference Coppage, Slocik, Ramezani-Dakhel, Bedford, Heinz, Naik and Knecht2013); the kinetic mechanisms of adsorption (Raffaini & Ganazzoli, Reference Raffaini and Ganazzoli2010); the ET pathways and properties of adsorbed proteins (Bizzarri, Reference Bizzarri2006; Siwko & Corni, Reference Siwko and Corni2013; Utesch et al. Reference Utesch, Sezer, Weidinger and Mroginski2012; Zanetti-Polzi et al. Reference Zanetti-Polzi, Daidone, Bortolotti and Corni2014; Zhou et al. Reference Zhou, Zheng and Jiang2004); the effects of pH (Emami et al. Reference Emami, Puddu, Berry, Varshney, Patwardhan, Perry and Heinz2014b; Imamura et al. Reference Imamura, Kawasaki, Awadzu, Sakiyama and Nakanishi2003; Tosaka et al. Reference Tosaka, Yamamoto, Ohdomari and Watanabe2010; Utesch et al. Reference Utesch, Millo, Castro, Hildebrandt, Zebger and Mroginski2013) and ionic strength (Bizzarri, Reference Bizzarri2006) on adsorption; the role of ions in mediating adsorption (Wu et al. Reference Wu, Chen, Skelton, Cummings and Zheng2013); and structural and energetic aspects of adsorption of proteins on surfaces (Apicella et al. Reference Apicella, Soncini, Deriu, Natalello, Bonanomi, Dellasega, Tortora, Regonesi and Casari2013; Jose & Sengupta, Reference Jose and Sengupta2013; Hoefling et al. Reference Hoefling, Monti, Corni and Gottschalk2011; Hung et al. Reference Hung, Mwenifumbo, Mager, Kuna, Stellacci, Yarovsky and Stevens2011; Kubiak-Ossowska & Mulheran, Reference Kubiak-Ossowska and Mulheran2010a, Reference Kubiak-Ossowska and Mulheranb; O'Mahony et al. Reference O'Mahony, O'Dwyer, Nijhuis, Greer, Quinn and Thompson2013; Vila Verde et al. Reference Vertegel, Siegel and Dordick2009, Reference Vigmond, Iwakura, Mizutani and Katsura2011; Wang et al. Reference Wang, Zhao, Zhao, Wang, Yang, Yu and Zheng2010a, Reference Wang, Zhao, Yu, Zhao, Li and Zhengb; Yu et al. Reference Yu, Becker and Carri2012a; Steckbeck et al. Reference Steckbeck, Schneider, Wittig, Rischka, Grunwald and Colombi Ciacchi2014; Sun et al. Reference Sun, Han, Lindgren, Shen and Laaksonen2014a; Sun et al. Reference Sun, Feng, Hou and Li2014b). Further, conventional MD can be used to simulate physical perturbations, such as mechanical or electrical forces exerted on molecules in experiments. Examples include voltage-biased MD simulations of peptides on a gold surface to validate the signals resulting from a surface enhanced Raman spectroscopy (SERS) experiment (Chen et al. Reference Chen, Cruz-Chu, Woodard, Gartia, Schulten and Liu2012). Mimicking the experimental conditions helps to understand how experiments work and how their results should be interpreted.

Classical MD simulations are very helpful in understanding many molecular phenomena at atomic detail. However, the capabilities and limitations of the MD method, as of other sampling methods, must be known in order to set up a simulation and interpret the results obtained properly. When comparing MD results with the thermodynamic and kinetic properties of the same system obtained from experiments, it should be remembered that a trend observed in an MD simulation may not always agree with the expected average properties. A system kinetically trapped in an energy well, may stay in its metastable state for the rest of the simulation and appear as if in equilibrium (Wei et al. Reference Wei, Carignano and Szleifer2011). In fact, a single simulation corresponds to a single trajectory of a system, and it may not always ensure an adequate sampling. Drawing conclusions from a statistically inadequate sampling will thus often lead to incorrect interpretations. To improve the statistical significance of their sampling, Penna et al. (Reference Penna, Mijajlovic and Biggs2014) chose a direct approach and performed more than 240 explicit solvent MD simulations for the investigation of adsorption mechanisms of a platinum-binding peptide and a gold-binding peptide on neutral Pt(111) and Au(111) surfaces, respectively. From these simulations, they obtained statistics on many binding characteristics, such as anchoring events between each of the peptide residues and the surfaces and the distribution of transition times from the anchoring phase to the initial lockdown phase.

Another important sampling problem is the simulation time required to reproduce an adsorption event on a molecular scale. Even though a simulation that captures the dynamics of a system for several nanoseconds is helpful for structural refinement of molecules or for understanding the initial stages of a process, many molecular events take place over much longer time scales. In their paper, Wei et al. (Reference Wei, Carignano and Szleifer2011) stressed the necessity of very long simulations to be able to investigate a complete adsorption process on a surface in detail. To bypass the energy barriers that normally require long simulation times to overcome, Emami et al. (Reference Emami, Puddu, Berry, Varshney, Patwardhan, Perry and Heinz2014b) employed 1 ns-long annealing simulations at 500 K prior to their production simulations at 298 K. From these classical MD simulations, they investigated the adsorption of three peptides differing in net charge to silica surfaces under four different pH values (with changing ionization levels of the silica surface depending on the pH value). The authors showed that the times spent by each of the peptides on the silica surfaces in the simulations could be used as a relative measure of adsorption strength as a function of pH (see Fig. 10). Even though the trends in adsorption level with pH for each of the peptides from the simulations and the experiments are in good agreement, the time spent on the surface by each of the peptides fails to fully reproduce the difference in relative adsorption levels between the negatively charged and overall neutral peptides measured experimentally.

Fig. 10. Adsorption of three different peptides (positively charged, neutral and negatively charged) on silica nanoparticles as a function of pH. (a) Amounts of adsorbed peptides measured in experiments (Puddu & Perry, Reference Puddu and Perry2012). (b) Times spent by the peptides on the surface calculated from the MD simulations. Reprinted with permission from (Emami et al. Reference Emami, Puddu, Berry, Varshney, Patwardhan, Perry and Heinz2014b). Copyright (2014) American Chemical Society.

To improve the sampling characteristics of conventional MD simulations, replica exchange MD (REMD) (Sugita et al. Reference Sugita and Okamoto1999) was developed based on a MC simulation method called parallel tempering (Hansmann, Reference Hansmann and Okamoto1997) (see the following section). In conventional MD simulations, structures may become trapped in local energy minima on the potential energy landscape, hence, sampling of configurational phase space is limited. To overcome high energy barriers, the REMD method makes use of sampling of the configurational space with independent MD simulations (i.e. replicas) of a system at different replica conditions, e.g. at different temperatures or with a different Hamiltonian. By exchanging the replicas, configurations that are trapped in local energy minima can explore other parts of the energy landscape and an improved Boltzmann-weighted sampling can be achieved. REMD methods are useful to sample a large set of configurations in different potential wells separated by high energy barriers, which otherwise is not possible in a classical MD simulation. Indeed, temperature-based REMD (T-REMD) simulations were used in several studies to accelerate the sampling of peptide–surface interactions (Corni et al. Reference Corni, Hnilova, Tamerler and Sarikaya2013; Li et al. Reference Li, Luo, Derreumaux and Wei2011; Oren et al. Reference Oren, Notman, Kim, Evans, Walsh, Samudrala, Tamerler and Sarikaya2010) and of interactions of basic fibroblast growth factor (bFGF), a small protein, with a hydroxyapatite (001) surface (Liao & Zhou, Reference Liao and Zhou2014). Liao and Zhou observed that while the protein displaces the surface hydration shell and binds tightly to the surface in the T-REMD simulations (five replicas in the range from 310 K to 2500 K), bFGF did not contact the surface directly in classical MD simulations at 310 K (Liao & Zhou, Reference Liao and Zhou2014). However, the T-REMD method is usually not applicable to large systems, e.g. a large protein and a surface with explicit solvent, due to the number of replicas required which scales as O(f 1/2), with f being the system's total number of degrees of freedom (Fukunishi et al. Reference Fukunishi, Watanabe and Takada2002; Wright & Walsh, Reference Wright and Walsh2013).

Typically, simulation of a protein-surface system in atomic detail requires solvation in tens of thousands of water molecules. The large number of water molecules becomes problematic because the water–water interaction energy term (E ww) dominates the energy terms due to interactions between the solute molecules and between the solute and water molecules, thereby demanding more replicas in a temperature-based REMD simulation (Huang et al. Reference Huang, Hagen, Kim, Friesner, Zhou and Berne2007). To overcome the poor scaling of the T-REMD method with the size of the system, replica exchange with solute tempering (REST) was proposed (Liu et al. Reference Liu, Kim, Friesner and Berne2005). In this approach, the potential energy function is tailored such that the E ww term is eliminated from the acceptance probability.

This increases the chances of acceptance of exchanges between replicas compared with regular TREMD simulations. Wright & Walsh (Reference Wright and Walsh2013) investigated the transferability of the REST method to peptide–surface interactions by using a quartz binding protein and a fully hydroxylated α-quartz system as a benchmark for the REST and the T-REMD simulations. The results showed that the REST approach was 50% less expensive in terms of CPU time than the T-REMD method using the same simulation parameters. Furthermore, up to about 80% of the CPU time could be saved with the REST method, owing to the smaller number of replicas used in the simulations. Using the REST parameters optimized for peptide–surface interaction simulations (Wright & Walsh, Reference Wright and Walsh2013), Tang et al. (Reference Tang, Palafox-Hernandez, Law, E. Hughes, Swihart, Prasad, Knecht and Walsh2013) performed REST simulations for numerous gold binding peptides to obtain an accurate ensemble of peptide conformations adsorbed on a gold surface.

The REMD and REST methods sample the phase space with a canonical distribution (fixed number of atoms, volume and temperature) and therefore fall short in exploring the very low probability states. Although investigating the dynamics of a system does not necessarily require an extensive exploration of these states, it may lead to erroneous results, particularly when calculating free energies. This problem can be greater for systems with peptides strongly interacting with surfaces (Wright & Walsh, Reference Wright and Walsh2013). To obtain more exhaustive sampling of the phase space, replica exchange methods can be used in conjunction with methods employing biasing potentials such as metadynamics (Bussi et al. Reference Bussi, Gervasio, Laio and Parrinello2006a).

9.2 Monte Carlo methods

MC methods are stochastic techniques used to solve problems that obey certain probability functions. MC methods have long been used in molecular simulations as an alternative to MD, owing to their efficiency in sampling of conformational ensembles with a Boltzmann distribution. From these ensembles, the geometric averages and thermodynamic properties of a system can be approximated (Schlick, Reference Schlick2002). The MC methods applied to biomolecular systems have been reviewed (see Hansmann & Okamoto, Reference Hansmann1999; Taylor et al. Reference Taylor, Jewsbury and Essex2002), and information on MC methods can be found elsewhere (Landau & Binder, Reference Landau and Binder2009).

Commonly employed for atomistic and CG simulations of protein-surface systems, Metropolis sampling (Metropolis et al. Reference Metropolis, Rosenbluth, Rosenbluth, Teller and Teller1953) is a simple and yet very efficient Markov chain Monte Carlo (MCMC) scheme. The Metropolis scheme is used to create a random walk on a set of points with an importance-weight based on a probability distribution (Frenkel & Smit, Reference Frenkel and Smit2001). MC simulations with the Metropolis algorithm have been used to explore the adsorbent conformations and protein denaturation on the surfaces (Anand et al. Reference Anand, Sharma, Kumar and Belfort2009; Castells et al. Reference Castells, Yang and Van Tassel2002; Euston & Naser, Reference Euston and Naser2005; Skepö, Reference Skepö2008; Skepö et al. Reference Skepö, Linse and Arnebrant2006; Zhdanov & Kasemo, Reference Zhdanov and Kasemo2000a), to investigate ordering of the proteins adsorbed on the surface (Zhdanov & Kasemo, Reference Zhdanov and Kasemo2000b), and to calculate the binding free energy of adsorption (Lund et al. Reference Lund, Åkesson and Jönsson2005).

Al-Mekhnaqi et al. (Reference Al-Mekhnaqi, Mayeed and Newaz2009) applied MC simulations with an energy minimization scheme to predict the conformations of a HSA subdomain A on a graphite surface with transition probabilities determined using the Metropolis method. The protein structure was modeled with a united atom approximation, and the structures obtained during MC simulations were minimized using a direct search scheme, which is a technique used for solving optimization problems without constraints. The simulations started with the peptide chain fully extended. In each step, one dihedral angle was chosen randomly and given a random value. After an energy minimization step, a new step was performed and this conformation was accepted with the Metropolis criterion. Simulations were ended after a predefined number of perturbations was achieved. Using this method, they were able to reproduce very similar conformations of the albumin subdomain adsorbed to a graphite surface to those obtained by all-atom molecular simulations (Raffaini & Ganazzoli, Reference Raffaini and Ganazzoli2003).

The Metropolis MC method can fail in sampling the configuration space of a system if the system becomes trapped in a low-energy state at low temperatures. Therefore, typically, many independent Metropolis MC simulations of a system are carried out. Alternatively, an energy-based MC method, parallel tempering Monte Carlo (PTMC), (Swendsen & Wang, Reference Swendsen and Wang1986) can be used. It is the basis for the REMD method. Every time a certain number of MC moves have been performed, the configurations are exchanged between a replica and its neighboring replicas that have the closest temperature values. The exchange operation is ruled by the Boltzmann distribution and done with the Metropolis acceptance probability. The PTMC method has been applied to protein–surface interactions in several studies, including prediction of the binding orientations of proteins (Xie et al. Reference Xie, Zhou and Jiang2010), and investigation of thermal and mechanical properties of proteins tethered on surfaces using a hybrid PTMC method with MD (Knotts et al. Reference Knotts, Rathore and de Pablo2005).

Xie et al. (Reference Xie, Zhou and Jiang2010) compared the performance of the PTMC method with that of conventional serial Metropolis MC simulations of lysozyme on charged surfaces using the same parameters. Figure 11a shows the distribution of the orientations of the lysozyme on a negatively charged surface obtained from the different MC methods. A comparison of these distributions with the adsorption energy landscape (Fig. 11b ) shows that with the PTMC method, the two lowest energy minima were found. In the four serial MC simulations, on the other hand, each simulation revealed a different energy minimum: two local energy minima in addition to the two lower minima found with PTMC.

Fig. 11. (a) Distribution of orientations of lysozyme on a negatively charged surface obtained by a single parallel tempering MC (labeled as p0) simulation and by four separate conventional serial MC (labeled as s1-s4) simulations. Orientations are represented by the cosine of angle θ, the angle between the unit vector along the dipole moment of lysozyme and the unit vector normal to the surface. (b) Energy landscape of interaction of lysozyme with the surface. Energy minima corresponding to the most visited orientations in the MC simulations are indicated with arrows. Reprinted with permission from (Xie et al. Reference Xie, Zhou and Jiang2010). Copyright (2010), AIP Publishing LLC.

In two other studies, the same group applied the PTMC method to obtain the most probable orientations of fibronectin on a HAP surface (Liao et al. Reference Liao, Xie and Zhou2014) and of protein G on -COOH and -NH2 functionalized SAMs (Liu et al. Reference Liu, Liao and Zhou2013). These orientations were used as starting orientations for MD simulations to investigate the adsorption mechanisms of the proteins.

Finally, basin-hopping Monte Carlo (BHMC) (Li & Scheraga, Reference Li and Scheraga1987; Wales & Doye, Reference Wales and Doye1997) is worth mentioning, since it is one of the most reliable MC methods for searching the lowest energy configurations of molecules (Iwamatsu & Okabe, Reference Iwamatsu and Okabe2004). Although the BHMC method has been successfully applied to conformational sampling of peptides in implicit solvent (Strodel & Wales, Reference Strodel and Wales2008), to the best of our knowledge, this method has yet to be employed for peptides or proteins near inorganic surfaces.

9.3 Brownian dynamics

The diffusive dynamics of a Brownian particle suspended in a solution can be modeled by the BD technique. BD is commonly applied for simulation of diffusion-driven binding processes leading to the formation of ‘diffusional encounter complexes’. Further information on details of BD methods, and their applications can be found in other reviews (Allison et al. Reference Allison, Northrup and McCammon1986; Gabdoulline & Wade, Reference Gabdoulline and Wade2002; Madura et al. Reference Madura, Davist, Gilson, Wade, Luty, McCammon, Lipkowitz and Boyd1994).

BD has been used in many studies of adsorption of proteins on surfaces with CG (Ravichandran & Talbot, Reference Ravichandran and Talbot2000) and all-atom (Ravichandran et al. Reference Ravichandran, Madura and Talbot2001) representations of the proteins. Simplified BD simulation methods designed for CG models have been proposed (De la Torre et al. Reference De la Torre, Hernández Cifre, Ortega, Schmidt, Fernandes, Pérez Sánchez and Pamies2009; Gorba & Helms, Reference Gorba and Helms2003) and one (Gorba & Helms, Reference Gorba and Helms2003) was tested for the diffusional dynamics of cytochrome c on a charged surface. As an alternative to CG models, several BD studies of protein adsorption to different surfaces have been performed with rigid, atomic detail models of the solutes (Brancolini et al. Reference Brancolini, Kokh, Calzolai, Wade and Corni2012; Mereghetti & Wade, Reference Mereghetti and Wade2011; Kokh et al. Reference Kokh, Corni, Winn, Hoefling, Gottschalk and Wade2010). Mereghetti and Wade (Mereghetti & Wade, Reference Mereghetti and Wade2011) applied BD to the simulation of hydrophobin proteins on a graphite surface. They explained the high affinity of the protein for the graphite surface from the average proximity and orientation of hydrophobin on the surface with its hydrophobic face towards the surface obtained from the simulations.

Brancolini and coworkers employed BD simulations for docking of ubiquitin (Brancolini et al. Reference Brancolini, Kokh, Calzolai, Wade and Corni2012) and of human β2-microglobulin (Brancolini et al. Reference Brancolini, Toroz and Corni2014, Reference Brancolini, Corazza, Vuano, Fogolari, Mimmi, Bellotti, Stoppini, Corni and Esposito2015; Brancolini et al. Reference Brancolini, Corazza, Vuano, Fogolari, Mimmi, Bellotti, Stoppini, Corni and Esposito2015) on bare, citrate covered and thiol-protected gold nanoparticles using the ProMetCS model (Kokh et al. Reference Kokh, Corni, Winn, Hoefling, Gottschalk and Wade2010). The most abundant orientations of ubiquitin on the gold obtained from BD simulations were used as starting orientations for fully flexible MD simulations. By using a combination of BD and MD methods, they were able to reduce the number of possible initial orientations for MD simulations, evaluate the stability of each orientation with MD simulations and observe the dynamics of adsorption in detail.

Overall, BD is an effective method for simulation of biomolecules as well as interactions of molecules with surfaces. An implicit representation of the solvent and a rigid-body treatment of the molecules help to reduce the time required for calculations, allowing thousands of simulations to be performed within a few hours with current computer technology. However, while the rigid-body model may be appropriate for globular proteins, it is less suitable for peptides and unstructured proteins. For these, flexible CG models of the solute can be employed in BD simulations.

10. Applications of free energy calculation methods to protein–surface interactions

The computation of binding free energies is used for determining binding affinities between proteins and surfaces, as well as the kinetics and transition free energies of adsorption processes. Free energy calculations are, therefore, important for bridging between theoretical and experimental studies. To date, many methods have been developed, which differ in their accuracy and complexity, for use in molecular simulation (Swanson et al. Reference Swanson, Henchman and McCammon2004). Most of these methods employ MD or MC techniques for sampling of the phase space. The free energy calculation methods can be classified as end-point methods and pathway methods. In end-point methods, only the reference unbound and the final bound states are sampled to obtain the free energy difference between the states. A commonly used end-point method is the MM Poisson–Boltzmann Surface Area (MM-PBSA) method (Kollman et al. Reference Kollman, Massova, Reyes, Kuhn, Huo, Chong, Lee, Lee, Duan, Wang, Donini, Cieplak, Srinivasan, Case and Cheatham2000; Srinivasan et al. Reference Srinivasan, Cheatham, Cieplak, Kollman and Case1998). The MM-PBSA method has been applied to a horse heart cytochrome c-COOH-SAM surface system to compute the adsorption free energy (Alvarez-Paggi et al. Reference Alvarez-Paggi, Martín, DeBiase, Hildebrandt, Martí and Murgida2010). Although easy to implement and perform, this method was shown to have a relatively large error range (Singh & Warshel, Reference Singh and Warshel2010) and variable performance (Hou et al. Reference Hou, Wang, Li and Wang2011) for protein–ligand interactions, and to require long MD simulations and extensive PB calculations.

A computationally faster, though less accurate, alternative to MM-PBSA, is MM Generalized-Born Surface Area (MM-GBSA) (Kollman et al. Reference Kollman, Massova, Reyes, Kuhn, Huo, Chong, Lee, Lee, Duan, Wang, Donini, Cieplak, Srinivasan, Case and Cheatham2000; Srinivasan et al. Reference Srinivasan, Cheatham, Cieplak, Kollman and Case1998). This method is based on solving the GB equation, which provides the solvation free energy of each individual atom and, therefore, is suitable for modeling the interactions of flexible solutes. Guo et al. (Reference Guo, Yao, Ning, Wang and Liu2014) computed binding free energies of three different proteins on a graphene surface from classical MD simulations using MM-GBSA. For each of the protein-graphene systems, they performed two separate simulations, each with a different protein orientation with respect to the surface. The results showed that the calculated free energy values for the two simulations may vary up to 65 kcal mol−1 with standard error values around 11 kcal mol−1. Although these calculations give insights into the adsorption strength of a binding mode from a trajectory, a better sampling of configuration phase space is required for MD simulations. To reduce the effect of the insufficient sampling on the energy calculations, Xie et al. (Reference Xie, Luo, Lin, Xi, Yang and Wei2014) used the REMD method for sampling the adsorption of the Aβ peptide to fullerenes. After clustering of the conformers, the MM-GBSA method was employed to obtain binding free energies for the largest clusters.

The end-point methods provide good approximations to free energy differences. Pathway methods are formally exact. Based on a predefined reaction coordinate, pathway methods employ either non-physical (alchemical) or physical intermediates. A commonly used method is free energy perturbation (FEP) (Mori et al. Reference Mori, Hamers, Pedersen and Cui2013). Although FEP has been applied to study binding of a formate anion (HCOO) to a rutile surface (Mori et al. Reference Mori, Hamers, Pedersen and Cui2013), to the best of our knowledge, this method has not been applied to study protein–surface interactions. Other pathway methods based on physical reaction coordinates and applied to protein–surface interactions are discussed below. These methods can be categorized into two groups depending on how the system is sampled: equilibrium and non-equilibrium simulation methods.

10.1 Equilibrium methods

A commonly used equilibrium sampling technique to obtain a free energy profile of a system is umbrella sampling (Torrie & Valleau, Reference Torrie and Valleau1974, Reference Torrie and Valleau1977). In this method, a reaction coordinate is defined that connects two thermodynamic states for which the free energy difference is computed. The reaction coordinate used is usually selected based on a geometric entity, e.g. distance, angle, etc. The reaction coordinate is then divided into distinct windows in which a bias (umbrella) potential is applied to keep the system near that coordinate point by means of restraints. For each window, a separate simulation of the system is performed for sampling around the corresponding coordinate point. Then, the simulations are combined with a reweighting procedure since they are performed in biased ensembles. Two reweighting procedures are weighted histogram analysis method (WHAM) (Kumar et al. Reference Kumar, Rosenberg, Bouzida, Swendsen and Kollman1992; Souaille & Roux, Reference Souaille and Roux2001) and umbrella integration (Kästner & Thiel, Reference Kästner and Thiel2005). Using the final probability distribution of the configurations, the PMF and hence, the free energy profile of the system, is obtained.

The umbrella sampling technique has been successfully employed for the calculation of the free energy of adsorption of a model surfactant on hydrophobic and hydrophilic silica surfaces (Xu et al. Reference Xu, Yang and Yang2008), nanoparticles on phospholipid membranes (Li & Gu, Reference Li and Gu2010), ions on hydrophobic surfaces (Horinek et al. Reference Horinek, Serr, Bonthuis, Boström, Kunz and Netz2008) and amino acids on a ZnO surface (Nawrocki & Cieplak, Reference Nawrocki and Cieplak2013). Umbrella sampling combined with WHAM has also been applied to peptide–surface interactions. Examples include adsorption of various tripeptides to a CH3-SAM (Sun et al. Reference Sun, Dominy and Latour2007), and an RGD peptide to a titanium oxide surface (Schneider & Colombi Ciacchi, Reference Schneider and Colombi Ciacchi2010). Although, as in these two examples, the distance of the peptide from the surface is typically used as the reaction coordinate for sampling, other definitions can also be employed. For instance, Boughton et al. (Reference Boughton, Andricioaei and Chen2010) used the angular orientation of the α-helix of magainin 2, an α-helical amphiphilic peptide, on a polystyrene surface as a reaction coordinate to compute its adsorption free energy by umbrella sampling combined with WHAM.

Umbrella sampling, even though it restricts the spatial sampling necessary to calculate free energy differences, can still be subject to conformational sampling problems. Wang et al. (Reference Wang, Stuart and Latour2008) therefore combined umbrella sampling with a biased REMD technique. In this method, an initial umbrella sampling of the system is employed to obtain a rough estimate of the PMF profile of the system. The negative of the PMF is then used as a biasing potential for REMD simulations. A final PMF profile is constructed from a probability-ratio method analysis for the REMD simulations. Comparison of the estimated PMF profile of a Na+ ion on a carboxylic acid functionalized SAM surface with the theoretical profile, showed that the biased REMD gives much better agreement with the theoretical profile compared with those obtained from conventional REMD and umbrella sampling simulations. The method was applied in separate studies to adsorption of the G4-K-G4 peptide on a polylactic acid surface for the investigation of binding free energy profiles (O'Brien et al. Reference O'Brien, Stuart, Bruce and Latour2008), for the assessment of the protein FF parameters to be used in protein-surface simulations (Vellore et al. Reference Vellore, Yancey, Collier, Latour and Stuart2010), and for investigation of the effect of pressure on the calculation of protein-surface binding free energies from MD simulations (Yancey et al. Reference Yancey, Vellore, Collier, Stuart and Latour2010).

Another common approach used in molecular simulations to calculate the free energy landscape as well as investigate new reaction pathways is metadynamics (Bussi et al. Reference Bussi, Laio and Parrinello2006b; Laio & Gervasio, Reference Laio and Gervasio2008; Laio & Parrinello, Reference Laio and Parrinello2002). In this approach, the dynamics of the system is steered by thermodynamic forces and compensated by a history-dependent biased potential (Laio & Parrinello, Reference Laio and Parrinello2002). Owing to the bias potential, the frequency at which a state is visited during a simulation typically decreases linearly with its free energy (Bussi et al. Reference Bussi, Gervasio, Laio and Parrinello2006a). Accurate free energy profiles can be obtained from metadynamics with an appropriate reweighting scheme (Bonomi et al. Reference Bonomi, Barducci and Parrinello2009) to recover an unbiased probability distribution (Meissner et al. Reference Meißner, Schneider, Schiffels and Colombi Ciacchi2014). To date, metadynamics has been successfully applied for calculation of the free energy of binding of a formate anion to a TiO2 surface (Mori et al. Reference Mori, Hamers, Pedersen and Cui2013) and of amino acids to silver and gold surfaces (Palafox-Hernandez et al. Reference Palafox-Hernandez, Tang, Hughes, Li, Swihart, Prasad, Walsh and Knecht2014), and for calculation of the free energy landscapes of alanine dipeptide in its free and gold surface bound states from 2μs-long simulations (Bellucci & Corni, Reference Bellucci and Corni2014).

The combination of metadynamics with other enhanced sampling techniques allows a better exploration of low probability states, hence increasing the accuracy of the free energy calculations (Bussi et al. Reference Bussi, Gervasio, Laio and Parrinello2006a). Qin & Buehler (Reference Qin and Buehler2014) investigated the adsorption free energy of mussel adhesion proteins onto a silica surface using metadynamics starting from configurations obtained from REMD simulations.

Schneider & Colombi Ciacchi (Reference Schneider and Colombi Ciacchi2012) have successfully applied a hybrid REST and metadynamics to study the adsorption of small peptides on Si and Ti surfaces. With only four replicas at temperatures of 300, 350, 400 and 450 K, they were able to obtain agreement with the experimental adsorption free energy of the RKLPDA peptide on a Ti surface (experimental: 38·0 ± 8 kJ mol−1, computed: 38·6 ± 3·9 kJ mol−1). The same approach was employed recently by Meissner et al. (Reference Meißner, Schneider, Schiffels and Colombi Ciacchi2014) to estimate the circular dichroism (CD) spectra of a peptide in its free state and bound to a silica surface. CD spectroscopy is a useful technique to monitor secondary structure content in biomolecules, in particular, α-helix content in peptide structures. Meissner et al. calculated the CD ellipticity values of conformations from simulation snapshots and used these values as external collective values in their reweighting procedure. The estimated fractional helicity values of the free and adsorbed peptides obtained from the simulations show good agreement with experiments. Even though these results are very promising, application of this method to protein adsorption requires careful selection of parameters, e.g. adequate number of replicas, appropriate selection of temperatures.

Thermodynamic integration is another approach used in free energy calculations of protein–surface interactions. This method, like umbrella sampling, requires a predefined reaction path between the initial and the final states. Juffer et al. (Reference Juffer, Argos and de Vlieg1996) applied thermodynamic integration to compute the adsorption free energies of the enzyme cutinase and its variants to a charged surface in the presence of explicit ions. Hoefling et al. (Reference Hoefling, Iori, Corni and Gottschalk2010a) applied the method to obtain the PMF profiles of amino acids adsorbing on an Au(111) surface. Further, Schneider & Colombi Ciacchi (Reference Schneider and Colombi Ciacchi2010) employed thermodynamic integration, along with the umbrella sampling/WHAM method, to compare the free energy profiles of adsorption of RGD peptides on oxidized Ti obtained by the two methods. Their results showed a perfect agreement between the results of the two methods. Finally, Friddle et al. (Reference Friddle, Battle, Trubetskoy, Tao, Salter, Moradian-Oldak, De Yoreo and Wierzbicki2011) used the adaptive biasing force simulation method, which is a technique based on thermodynamic integration to obtain free energy profiles using a biasing force, for the adsorption of a 12-mer C-terminal fragment of the amelogenin protein on various different crystal terminations of two different hydroxyapatite surfaces ((100) and (001)). By complementing AFM measurements with the free energy calculations, Friddle et al. (Reference Friddle, Battle, Trubetskoy, Tao, Salter, Moradian-Oldak, De Yoreo and Wierzbicki2011) were able to predict the crystal termination of the hydroxyapatite surface used in experiments and identify the interactions governing the adsorption of the amelogenin peptide to it.

10.2 Non-equilibrium methods

Observing adsorption/desorption events of large molecules using conventional MD methods requires very long simulation times. These limitations can be overcome by steered molecular dynamics (SMD) simulations (Utesch et al. Reference Utesch, Daminelli and Mroginski2011), which is the most commonly used non-equilibrium simulation technique for biomolecular systems. In a typical SMD simulation, a protein or a peptide is pulled with a non-physical external force along a predefined reaction coordinate with a constant velocity or with a constant force, thereby accelerating adsorption or desorption events. Further, SMD allows simulation of systems under mechanical stress such as stretching, shearing and bending, and can hence be used to predict the effects of external disturbances on protein-surface systems (Hamdi et al. Reference Hamdi, Ferreira, Sharma and Mavroidis2008). These disturbances may lead to uncertainties in experimental measurements. Particularly, SMD simulations conducted to investigate the effect of the tip of an AFM showed that pushing a molecule towards a surface with an AFM tip will bias the results by increasing adhesion, as the force exerted on the molecule leads it to spread more on the surface (Horinek et al. Reference Horinek, Serr, Bonthuis, Boström, Kunz and Netz2008; Mücksch & Urbassek, Reference Mücksch and Urbassek2011). With SMD simulations, the adsorption mechanisms of peptides and proteins on different types of surfaces have been investigated (Alvarez-Paggi et al. Reference Alvarez-Paggi, Martín, DeBiase, Hildebrandt, Martí and Murgida2010; Dong et al. Reference Dong, Wang, Wu and Pan2007; Emami et al. Reference Emami, Puddu, Berry, Varshney, Patwardhan, Perry and Heinz2014b; Friedrichs et al. Reference Friedrichs, Köppen and Langel2013; Hamdi et al. Reference Hamdi, Ferreira, Sharma and Mavroidis2008; Shen et al. Reference Shen, Wu, Wang and Pan2008; Utesch et al. Reference Utesch, Daminelli and Mroginski2011; Yang & Zhao, Reference Yang and Zhao2007) and compared with experimental measurements (Schneider & Colombi Ciacchi, Reference Schneider and Colombi Ciacchi2010, Reference Schneider and Colombi Ciacchi2012). In addition to accelerating the sampling of certain events, the SMD method is used to calculate free energy differences. It was shown by Jarzynski (Reference Jarzynski1997) that a free energy difference can be obtained from a non-equilibrium process following the equation: e βΔF  = 〈e βW 〉 where β = 1/(k B T) (k B is the Boltzmann constant and T is the absolute temperature) and ∆F and W stand for the free energy difference and the non-reversible work done along a reaction coordinate, respectively. The average is taken on all (in principle) the realizations of the non-equilibrium process that connect the initial and final states of interest. Based on this equation, a free energy difference can be obtained using a population of SMD simulations that start from an equilibrium ensemble. Binding free energies of peptides and proteins to various different surfaces computed using the Jarzynski equation have been reported in a number of papers (Chen et al. Reference Chen, Wang, Liu, Wu, Kang, Moore and Gubbins2009b; Kang et al. Reference Kang, Liu, Wang, Shen, Wu and Guan2009; Mijajlovic et al. Reference Mijajlovic, Penna and Biggs2013). Baier et al. (Reference Baier, Blumenstein, Preusker, Jeurgens, Welzel, Do, Pleiss and Bill2014) investigated the adsorption free energies of ZnO-binding peptides on ZnO surfaces. To this end, they applied a hybrid steered MD with umbrella sampling method. In this approach, the peptides were steered towards the surface with a constant force applied to the peptides. After an equilibration period without any external forces applied upon adsorption, the peptides were pulled back from the ZnO surface. Selected conformers were then simulated for another 5 ns with a harmonic potential applied to their centers of mass. The PMF profiles were obtained using the WHAM method. Combined SMD simulations with umbrella sampling were also used in a study to calculate the adsorption free energy of a cationic peptide on a silica surface (Emami et al. Reference Emami, Puddu, Berry, Varshney, Patwardhan, Perry and Heinz2014b).

11. Outlook and future directions

The modeling and simulation studies reviewed in this paper suggest that the adsorption of aqueous peptides and/or proteins to surfaces is governed by a number of properties that determine the strength and specificity of the interactions. These properties can be summarized as: pH, solvated ion types and surfactants, and ionic strength of the solution; ionization levels, physical character (i.e. polar, non-polar, charged, etc.), size, shape, thickness, structural and compositional homogeneity/heterogeneity, chemical modifications and molecular structure of the surface; and flexibility, physical character and intra–peptide interactions determined by the sequence and affinity of the peptide/protein. To derive design rules for proteins and surfaces with desired binding characteristics, the studies often oversimplify adsorption mechanisms, focusing on only one or a few of these properties as determinants of the specificity or the affinity of the proteins for the surfaces. The transferability of the simple design rules derived in these studies to systems with different surface/protein types under different solution conditions is, hence, to be considered with caution. To be able to draw a complete picture of these interactions and thus to draw universal design rules, one has to take all of these properties into account and investigate the significance of each of them in the system of interest thoroughly.

The current FFs used for biomolecular interactions were developed and optimized specifically for their interactions in aqueous environment. The major FFs used for simulations of biomolecule–inorganic surface interactions, on the other hand, are based on mixing the parameters of separately parameterized biomolecular and inorganic FFs. Although their applicability to inorganic interfaces has been tested and validated to a certain extent, many more studies are needed for an extensive calibration of the parameter sets. The number of experimental studies providing structural information on protein–surface interactions is currently limited. However, advances in experimental capabilities for application to these interactions are very promising, for instance probing the 3D structures of peptides adsorbed on metal oxides (Mirau et al. Reference Mirau, Naik and Gehring2011) and identification of sites on ubiquitin engaged in binding gold nanoparticles (Calzolai et al. Reference Calzolai, Franchini, Gilliland and Rossi2010) by means of nuclear magnetic resonance (NMR) techniques.

A bottleneck in the simulation of protein-surface systems is sampling. Undoubtedly, the most commonly used simulation method at present is classical MD. However, with classical MD simulations, complete sampling of the phase space and, therefore, the adsorption dynamics is not possible. Enhanced sampling methods (e.g. replica exchange methods) are therefore invaluable tools for capturing the less probable states of the system that might otherwise require dozens of classical MD simulations. Even though simulations of peptide–surface interactions by means of enhanced sampling methods have been reported in a number of studies (discussed in previous sections), simulations of protein–surface interactions with these methods are still often not feasible due to the large size and complexity of proteins. Further studies are necessary to develop simulation protocols, evaluate them and optimize suitable parameters.

Advances in the simulation of protein–surface interactions are tied to the general advances in simulation methods. For instance, we discussed the significance of the representation of a change in ionization states of SAM and oxide surfaces and hence the need for constant pH simulations earlier in this review. However, many constant pH simulations suffer from convergence issues and pose even more issues with explicit solvation (Mongan & Case, Reference Mongan and Case2005). Therefore, improved simulation methods are required for accurate modeling and simulation of protein-inorganic surface systems.

Finally, none of the simulation techniques covered in this review is able to provide an accurate picture of protein adsorption events that take place on a large range of time and length scales by itself. Therefore, appropriate multi-scale modeling and simulation approaches should be developed and employed in a concerted manner.

Acknowledgements

We thank Dr. Neil J. Bruce for his careful reading of the manuscript.

Financial support

M. O. acknowledges the support of Heidelberg Graduate School of Mathematical and Computational Methods for the Sciences (HGS MathComp), Heidelberg University. S. C. acknowledges funding from MIUR through PRIN 2012A7LMS3003. Our work is supported by the Klaus Tschira Foundation.

References

Aarts, I. M. P., Pipino, A. C. R., Hoefnagels, J. P. M., Kessels, W. M. M. & van de Sanden, M. C. M. (2005) Quasi-ice monolayer on atomically smooth amorphous SiO2 at room temperature observed with a high-finesse optical resonator. Physical Review Letters 95, 166104.Google Scholar
Adnan, A., Lam, R., Chen, H., Lee, J., Schaffer, D. J., Barnard, A. S., Schatz, G. C., Ho, D. & Liu, W. K. (2011) Atomistic simulation and measurement of pH dependent cancer therapeutic interactions with nanodiamond carrier. Molecular Pharmaceutics 8, 368374.Google Scholar
Akdim, B., Pachter, R., Kim, S. S., Naik, R. R., Walsh, T. R., Trohalaki, S., Hong, G., Kuang, Z. & Farmer, B. L. (2013) Electronic properties of a graphene device with peptide adsorption: insight from simulation. ACS Applied Materials & Interfaces 5, 74707477.CrossRefGoogle ScholarPubMed
Alava, T., Mann, J. A., Théodore, C., Benitez, J. J., Dichtel, W. R., Parpia, J. M. & Craighead, H. G. (2013) Control of the graphene–protein interface is required to preserve adsorbed protein function. Analytical Chemistry 85, 27542759.CrossRefGoogle ScholarPubMed
Alessandrini, A., Salerno, M., Frabboni, S. & Facci, P. (2005) Single-metalloprotein wet biotransistor. Applied Physics Letters 86, 133902.CrossRefGoogle Scholar
Aliaga, A. E., Ahumada, H., Sepúlveda, K., Gomez-Jeria, J. S., Garrido, C., Weiss-López, B. E. & Campos-Vallette, M. M. (2011) SERS, Molecular dynamics and molecular orbital studies of the MRKDV peptide on silver and membrane surfaces. The Journal of Physical Chemistry C 115, 39823989.CrossRefGoogle Scholar
Allison, S. A., Northrup, S. H. & McCammon, J. A. (1986) Simulation of biomolecular diffusion and complex formation. Biophysical Journal 49, 167175.CrossRefGoogle ScholarPubMed
Al-Mekhnaqi, A. M., Mayeed, M. S. & Newaz, G. M. (2009) Prediction of protein conformation in water and on surfaces by Monte Carlo simulations using united-atom method. Molecular Simulation 35, 292300.CrossRefGoogle Scholar
Alvarez-Paggi, D., Castro, M. A., Tórtora, V., Castro, L., Radi, R. & Murgida, D. H. (2013) Electrostatically driven second-sphere ligand switch between high and low reorganization energy forms of native cytochrome c. Journal of the American Chemical Society 135, 43894397.Google Scholar
Alvarez-Paggi, D., Martín, D. F., DeBiase, P. M., Hildebrandt, P., Martí, M. A. & Murgida, D. H. (2010) Molecular basis of coupled protein and electron transfer dynamics of cytochrome c in biomimetic complexes. Journal of the American Chemical Society 132, 57695778.Google Scholar
Anand, G., Sharma, S., Kumar, S. K. & Belfort, G. (2009) Stability of tethered proteins. Langmuir 25, 49985005.CrossRefGoogle ScholarPubMed
Andolfi, L., Bruce, D., Cannistraro, S., Canters, G. W., Davis, J. J., Hill, H. A. O., Crozier, J., Verbeet, M. P., Wrathmell, C. L. & Astier, Y. (2004) The electrochemical characteristics of blue copper protein monolayers on gold. Journal of Electroanalytical Chemistry 565, 2128.Google Scholar
Andreescu, S., Ornatska, M., Erlichman, J. S., Estevez, A. & Leiter, J. C. (2012) Biomedical applications of metal oxide nanoparticles. In Fine Particles in Medicine and Pharmacy (ed Matijević, E.), pp. 57100. Springer US, Boston, MA, USA.Google Scholar
Anker, J. N., Hall, W. P., Lyandres, O., Shah, N. C., Zhao, J. & Van Duyne, R. P. (2008) Biosensing with plasmonic nanosensors. Nature Materials 7, 442453.Google Scholar
Ansari, S. A. & Husain, Q. (2012) Potential applications of enzymes immobilized on/in nano materials: a review. Biotechnology Advances 30, 512523.Google Scholar
Apicella, A., Soncini, M., Deriu, M. A., Natalello, A., Bonanomi, M., Dellasega, D., Tortora, P., Regonesi, M. E. & Casari, C. S. (2013) A hydrophobic gold surface triggers misfolding and aggregation of the amyloidogenic josephin domain in monomeric form, while leaving the oligomers unaffected. PLoS ONE 8, e58794.Google Scholar
Arrouvel, C., Diawara, B., Costa, D. & Marcus, P. (2007) DFT periodic study of the adsorption of glycine on the anhydrous and hydroxylated (0001) surfaces of α-alumina. The Journal of Physical Chemistry C 111, 1816418173.Google Scholar
Austin, B. M., Zubarev, D. Y. & Lester, W. A. (2012) Quantum Monte Carlo and related approaches. Chemical Reviews 112, 263288.Google Scholar
Baier, J., Blumenstein, N. J., Preusker, J., Jeurgens, L. P. H., Welzel, U., Do, T. A., Pleiss, J. & Bill, J. (2014) The influence of ZnO-binding 12-mer peptides on bio-inspired ZnO formation. CrystEngComm 16, 53015307.Google Scholar
Balamurugan, K., Gopalakrishnan, R., Raman, S. S. & Subramanian, V. (2010) Exploring the changes in the structure of α-helical peptides adsorbed onto a single walled carbon nanotube using classical molecular dynamics simulation. The Journal of Physical Chemistry B 114, 1404814058.Google Scholar
Bandura, A. V. & Kubicki, J. D. (2003) Derivation of force field parameters for TiO2−H2O systems from ab initio calculations. The Journal of Physical Chemistry B 107, 1107211081.CrossRefGoogle Scholar
Barabino, G., Cesare-Gavotti, & Marchesi, M. (1984) Molecular dynamics simulation of water near walls using an improved wall—water interaction potential. Chemical Physics Letters 104, 478484.CrossRefGoogle Scholar
Barbey, R., Lavanant, L., Paripovic, D., Schüwer, N., Sugnaux, C., Tugulu, S. & Klok, H.-A. (2009) Polymer brushes via surface-initiated controlled radical polymerization: synthesis, characterization, properties, and applications. Chemical Reviews 109, 54375527.Google Scholar
Bellucci, L., Brancolini, G., Calzolari, A., Carrillo Parramon, O., Corni, S. & Di Felice, R. (2012) Proteins and peptides at gold surfaces: insights from atomistic simulations. In Proteins at Interfaces III State of the Art (eds Horbett, T., Brash, J. L. & Norde, W.), pp. 229250. American Chemical Society, Washington, DC, USA.Google Scholar
Bellucci, L. & Corni, S. (2014) Interaction with a gold surface reshapes the free energy landscape of alanine dipeptide. The Journal of Physical Chemistry C 118, 1135711364.Google Scholar
Bhirde, A. A., Hassan, S. A., Harr, E. & Chen, X. (2014) Role of albumin in the formation and stabilization of nanoparticle aggregates in serum studied by continuous photon correlation spectroscopy and multiscale computer simulations. The Journal of Physical Chemistry C 118, 1619916208.Google Scholar
Bhowmik, R., Katti, K. S. & Katti, D. (2007) Molecular dynamics simulation of hydroxyapatite–polyacrylic acid interfaces. Polymer 48, 664674.CrossRefGoogle Scholar
Binnig, G., Quate, C. F. & Gerber, C. (1986) Atomic force microscope. Physical Review Letters 56, 930933.Google Scholar
Biswas, P. K., Vellore, N. A., Yancey, J. A., Kucukkal, T. G., Collier, G., Brooks, B. R., Stuart, S. J. & Latour, R. A. (2012) Simulation of multiphase systems utilizing independent force fields to control intraphase and interphase behavior. Journal of Computational Chemistry 33, 14581466.Google Scholar
Bizzarri, A. R. (2006) Topological and dynamical properties of Azurin anchored to a gold substrate as investigated by molecular dynamics simulation. Biophysical Chemistry 122, 206214.CrossRefGoogle Scholar
Bonomi, M., Barducci, A. & Parrinello, M. (2009) Reconstructing the equilibrium Boltzmann distribution from well-tempered metadynamics. Journal of Computational Chemistry 30, 16151621.CrossRefGoogle ScholarPubMed
Borodin, O., Smith, G. D., Bandyopadhyaya, R. & Byutner, O. (2003) Molecular dynamics study of the influence of solid interfaces on poly(ethylene oxide) structure and dynamics. Macromolecules 36, 78737883.Google Scholar
Boughton, A. P., Andricioaei, I. & Chen, Z. (2010) Surface orientation of magainin 2: molecular dynamics simulation and sum frequency generation vibrational spectroscopic studies. Langmuir 26, 1603116036.Google Scholar
Brancolini, G., Corazza, A., Vuano, M., Fogolari, F., Mimmi, M. C., Bellotti, V., Stoppini, M., Corni, S. & Esposito, G. (2015) Probing the influence of citrate-capped gold nanoparticles on an amyloidogenic protein. ACS Nano 9, 26002613.Google Scholar
Brancolini, G., Kokh, D. B., Calzolai, L., Wade, R. C. & Corni, S. (2012) Docking of ubiquitin to gold nanoparticles. ACS Nano 6, 98639878.Google Scholar
Brancolini, G., Toroz, D. & Corni, S. (2014) Can small hydrophobic gold nanoparticles inhibit β2-microglobulin fibrillation? Nanoscale 6, 79037911.CrossRefGoogle ScholarPubMed
Braun, R., Sarikaya, M. & Schulten, K. (2002) Genetically engineered gold-binding polypeptides: structure prediction and molecular dynamics. Journal of Biomaterials Science, Polymer Edition 13, 747757.Google Scholar
Brooks, B. R., Bruccoleri, R. E., Olafson, B. D., States, D. J., Swaminathan, S. & Karplus, M. (1983) CHARMM: a program for macromolecular energy, minimization, and dynamics calculations. Journal of Computational Chemistry 4, 187217.CrossRefGoogle Scholar
Buimaga-Iarinca, L. & Calborean, A. (2012) Electronic structure of the ll-cysteine dimers adsorbed on Au(111): a density functional theory study. Physica Scripta 86, 035707.Google Scholar
Bussi, G., Gervasio, F. L., Laio, A. & Parrinello, M. (2006a) Free-energy landscape for β hairpin folding from combined parallel tempering and metadynamics. Journal of the American Chemical Society 128, 1343513441.CrossRefGoogle ScholarPubMed
Bussi, G., Laio, A. & Parrinello, M. (2006b) Equilibrium free energies from nonequilibrium metadynamics. Physical Review Letters 96, 090601.CrossRefGoogle ScholarPubMed
Butenuth, A., Moras, G., Schneider, J., Koleini, M., Köppen, S., Meißner, R., Wright, L. B., Walsh, T. R. & Colombi Ciacchi, L. (2012) Ab initio derived force-field parameters for molecular dynamics simulations of deprotonated amorphous-SiO2/water interfaces. Physica Status Solidi (B) 249, 292305.CrossRefGoogle Scholar
Calzolai, L., Franchini, F., Gilliland, D. & Rossi, F. (2010) Protein−nanoparticle interaction: identification of the ubiquitin−gold nanoparticle interaction site. Nano Letters 10, 31013105.Google Scholar
Calzolari, A., Cicero, G., Cavazzoni, C., Di Felice, R., Catellani, A. & Corni, S. (2010) Hydroxyl-rich β-sheet adhesion to the gold surface in water by first-principle simulations. Journal of the American Chemical Society 132, 47904795.Google Scholar
Camden, A. N., Barr, S. A. & Berry, R. J. (2013) Simulations of peptide-graphene interactions in explicit water. The Journal of Physical Chemistry B 117, 1069110697.Google Scholar
Cardone, A., Pant, H. & Hassan, S. A. (2013) Specific and non-specific protein association in solution: computation of solvent effects and prediction of first-encounter modes for efficient configurational bias Monte Carlo simulations. The Journal of Physical Chemistry B 117, 1236012374.CrossRefGoogle ScholarPubMed
Carravetta, V. & Monti, S. (2006) Peptide−TiO2 surface interaction in solution by ab initio and molecular dynamics simulations. The Journal of Physical Chemistry B 110, 61606169.Google Scholar
Carravetta, V., Monti, S. & Zhang, W. (2009) Interaction of biomolecular systems with titanium-based materials: computational investigations. Theoretical Chemistry Accounts 123, 299309.Google Scholar
Carrillo-Parramon, O., Brancolini, G. & Corni, S. (2013) Simulation of protein–surface interactions by a coarse-grained method. BioNanoScience 3, 1220.Google Scholar
Casals, E., Bastus, N., Vázquez, S., Varon, M., Comenge, J. & Puntes, V. (2008) Inorganic nanoparticles and biology. Contributions to Science 4, 171176.Google Scholar
Castells, V., Yang, S. & Van Tassel, P. R. (2002) Surface-induced conformational changes in lattice model proteins by Monte Carlo simulation. Physical Review E 65, 031912.CrossRefGoogle ScholarPubMed
Chaki, N. K. & Vijayamohanan, K. (2002) Self-assembled monolayers as a tunable platform for biosensor applications. Biosensors and Bioelectronics 17, 112.CrossRefGoogle ScholarPubMed
Chen, H., Su, X., Neoh, K.-G. & Choe, W.-S. (2009a) Context-dependent adsorption behavior of cyclic and linear peptides on metal oxide surfaces. Langmuir 25, 15881593.Google Scholar
Chen, Q., Wang, Q., Liu, Y.-C., Wu, T., Kang, Y., Moore, J. D. & Gubbins, K. E. (2009b) Energetics investigation on encapsulation of protein/peptide drugs in carbon nanotubes. The Journal of Chemical Physics 131, 015101.CrossRefGoogle ScholarPubMed
Chen, S.-K., Wang, B.-C., Huang, W.-Z. & Gao, S. (2010) Molecular dynamics simulation of the interaction between Maps and Cu(100) surface. In proceedings of The 2nd Conference on Environmental Science and Information Application Technology pp. 533536. IEEE.Google Scholar
Chen, Y., Cruz-Chu, E. R., Woodard, J. C., Gartia, M. R., Schulten, K. & Liu, L. (2012) Electrically induced conformational change of peptides on metallic nanosurfaces. ACS Nano 6, 88478856.CrossRefGoogle ScholarPubMed
Chiu, C.-C., Moore, P. B., Shinoda, W. & Nielsen, S. O. (2009) Size-dependent hydrophobic to hydrophilic transition for nanoparticles: a molecular dynamics study. The Journal of Chemical Physics 131, 244706.Google Scholar
Choi, J.-W., Kim, J. S., Kim, S.-U. & Min, J. (2009) Charge storage in redox-active azurin monolayer on 11-MUA modified gold surface. Biochip Journal 3, 157163.Google Scholar
Cicero, G., Calzolari, A., Corni, S. & Catellani, A. (2011) Anomalous wetting layer at the Au(111) surface. The Journal of Physical Chemistry Letters 2, 25822586.Google Scholar
Clementi, C., Nymeyer, H. & Onuchic, J. N. (2000) Topological and energetic factors: what determines the structural details of the transition state ensemble and ‘en-route’ intermediates for protein folding? An investigation for small globular proteins. Journal of Molecular Biology 298, 937953.Google Scholar
Cohavi, O., Corni, S., De Rienzo, F., Di Felice, R., Gottschalk, K. E., Hoefling, M., Kokh, D., Molinari, E., Schreiber, G., Vaskevich, A. & Wade, R. C. (2010) Protein-surface interactions: challenging experiments and computations. Journal of Molecular Recognition 23, 259262.Google Scholar
Cohavi, O., Reichmann, D., Abramovich, R., Tesler, A. B., Bellapadrona, G., Kokh, D. B., Wade, R. C., Vaskevich, A., Rubinstein, I. & Schreiber, G. (2011) A quantitative, real-time assessment of binding of peptides and proteins to gold surfaces. Chemistry - A European Journal 17, 13271336.Google Scholar
Cole, D. J., Payne, M. C., Csányi, G., Mark Spearing, S. & Colombi Ciacchi, L. (2007) Development of a classical force field for the oxidized Si surface: application to hydrophilic wafer bonding. The Journal of Chemical Physics 127, 204704. CrossRefGoogle ScholarPubMed
Collier, G., Vellore, N. A., Yancey, J. A., Stuart, S. J. & Latour, R. A. (2012) Comparison between empirical protein force fields for the simulation of the adsorption behavior of structured LK peptides on functionalized surfaces. Biointerphases 7, 24.Google Scholar
Colombi Ciacchi, L. & Payne, M. C. (2005) First-principles molecular-dynamics study of native oxide growth on Si(001). Physical Review Letters 95, 196101.Google Scholar
Coppage, R., Slocik, J. M., Briggs, B. D., Frenkel, A. I., Heinz, H., Naik, R. R. & Knecht, M. R. (2011) Crystallographic recognition controls peptide binding for bio-based nanomaterials. Journal of the American Chemical Society 133, 1234612349.CrossRefGoogle ScholarPubMed
Coppage, R., Slocik, J. M., Ramezani-Dakhel, H., Bedford, N. M., Heinz, H., Naik, R. R. & Knecht, M. R. (2013) Exploiting localized surface binding effects to enhance the catalytic reactivity of peptide-capped nanoparticles. Journal of the American Chemical Society 135, 1104811054.Google Scholar
Cornell, W. D., Cieplak, P., Bayly, C. I., Gould, I. R., Merz, K. M., Ferguson, D. M., Spellmeyer, D. C., Fox, T., Caldwell, J. W. & Kollman, P. A. (1995) A second generation force field for the simulation of proteins, nucleic acids, and organic molecules. Journal of the American Chemical Society 117, 51795197.Google Scholar
Corni, S., Hnilova, M., Tamerler, C. & Sarikaya, M. (2013) Conformational behavior of genetically-engineered dodecapeptides as a determinant of binding affinity for gold. The Journal of Physical Chemistry C 117, 1699017003.Google Scholar
Costa, D., Garrain, P.-A. & Baaden, M. (2013) Understanding small biomolecule-biomaterial interactions: a review of fundamental theoretical and experimental approaches for biomolecule interactions with inorganic surfaces. Journal of Biomedical Materials Research Part A 101A, 12101222.Google Scholar
Cruz-Chu, E. R., Aksimentiev, A. & Schulten, K. (2006) Water−silica force field for simulating nanodevices. The Journal of Physical Chemistry B 110, 2149721508.Google Scholar
Cygan, R. T., Liang, J.-J. & Kalinichev, A. G. (2004) Molecular models of hydroxide, oxyhydroxide, and clay phases and the development of a general force field. The Journal of Physical Chemistry B 108, 12551266.Google Scholar
Dauber-Osguthorpe, P., Roberts, V. A., Osguthorpe, D. J., Wolff, J., Genest, M. & Hagler, A. T. (1988) Structure and energetics of ligand binding to proteins:Escherichia coli dihydrofolate reductase-trimethoprim, a drug-receptor system. Proteins: Structure, Function, and Genetics 4, 3147.Google Scholar
De la Torre, J. G., Hernández Cifre, J. G., Ortega, A., Schmidt, R. R., Fernandes, M. X., Pérez Sánchez, H. E. & Pamies, R. (2009) SIMUFLEX: Algorithms and tools for simulation of the conformation and dynamics of flexible molecules and nanoparticles in dilute solution. Journal of Chemical Theory and Computation 5, 26062618.CrossRefGoogle ScholarPubMed
De Leeuw, N. H. (2010) Computer simulations of structures and properties of the biomaterial hydroxyapatite. Journal of Materials Chemistry 20, 53765389.Google Scholar
De Miranda Tomásio, S. & Walsh, T. R. (2007) Atomistic modelling of the interaction between peptides and carbon nanotubes. Molecular Physics 105, 221229.Google Scholar
Demann, E. T. K., Stein, P. S. & Haubenreich, J. E. (2005) Gold as an implant in medicine and dentistry. Journal of Long-Term Effects of Medical Implants 15, 687698.Google Scholar
Di Felice, R. & Corni, S. (2011) Simulation of peptide–surface recognition. The Journal of Physical Chemistry Letters 2, 15101519.Google Scholar
Di Felice, R. & Selloni, A. (2004) Adsorption modes of cysteine on Au(111): thiolate, amino-thiolate, disulfide. The Journal of Chemical Physics 120, 49064914.Google Scholar
Di Felice, R., Selloni, A. & Molinari, E. (2003) DFT study of cysteine adsorption on Au(111). The Journal of Physical Chemistry B 107, 11511156.Google Scholar
Diebold, U. (2003) The surface science of titanium dioxide. Surface Science Reports 48, 53229.Google Scholar
Ding, F., Radic, S., Chen, R., Chen, P., Geitner, N. K., Brown, J. M. & Ke, P. C. (2013) Direct observation of a single nanoparticle–ubiquitin corona formation. Nanoscale 5, 91629169.Google Scholar
Dion, M., Rydberg, H., Schröder, E., Langreth, D. C. & Lundqvist, B. I. (2004) Van der Waals density functional for general geometries. Physical Review Letters 92, 246401.CrossRefGoogle Scholar
Dong, X., Wang, Q., Wu, T. & Pan, H. (2007) Understanding adsorption-desorption dynamics of BMP-2 on hydroxyapatite (001) surface. Biophysical Journal 93, 750759.Google Scholar
Dufort, S., Sancey, L. & Coll, J.-L. (2012) Physico-chemical parameters that govern nanoparticles fate also dictate rules for their molecular evolution. Advanced Drug Delivery Reviews 64, 179189.Google Scholar
Durdagi, S., Mavromoustakos, T., Chronakis, N. & Papadopoulos, M. G. (2008) Computational design of novel fullerene analogues as potential HIV-1 PR inhibitors: analysis of the binding interactions between fullerene inhibitors and HIV-1 PR residues using 3D QSAR, molecular docking and molecular dynamics simulations. Bioorganic & Medicinal Chemistry 16, 99579974.CrossRefGoogle ScholarPubMed
Emami, F. S., Puddu, V., Berry, R. J., Varshney, V., Patwardhan, S. V., Perry, C. C. & Heinz, H. (2014a) Force field and a surface model database for silica to simulate interfacial properties in atomic resolution. Chemistry of Materials 26, 26472658.CrossRefGoogle Scholar
Emami, F. S., Puddu, V., Berry, R. J., Varshney, V., Patwardhan, S. V., Perry, C. C. & Heinz, H. (2014b) Prediction of specific biomolecule adsorption on silica surfaces as a function of pH and particle size. Chemistry of Materials 26, 57255734.Google Scholar
Eralp, T., Shavorskiy, A. & Held, G. (2011) The adsorption geometry and chemical state of lysine on Cu{110}. Surface Science 605, 468472.Google Scholar
Euston, S. R., Hughes, P., Naser, M. A. & Westacott, R. E. (2008) Comparison of the adsorbed conformation of barley lipid transfer protein at the decane−water and vacuum−water interface: a molecular dynamics simulation. Biomacromolecules 9, 14431453.Google Scholar
Euston, S. R. & Naser, M. A. (2005) Simulating the equation of state of model globular proteins adsorbed at a surface. Langmuir 21, 42274235.CrossRefGoogle ScholarPubMed
Evers, C. H. J., Andersson, T., Lund, M. & Skepö, M. (2012) Adsorption of unstructured protein β-casein to hydrophobic and charged surfaces. Langmuir 28, 1184311849.Google Scholar
Fajín, J. L. C., Gomes, J. R. B. & Cordeiro, M. N. D. S. (2013) DFT study of the adsorption of D-(L-)cysteine on flat and chiral stepped gold surfaces. Langmuir 29, 88568864.Google Scholar
Feng, J., Pandey, R. B., Berry, R. J., Farmer, B. L., Naik, R. R. & Heinz, H. (2011) Adsorption mechanism of single amino acid and surfactant molecules to Au {111} surfaces in aqueous solution: design rules for metal-binding molecules. Soft Matter 7, 2113.Google Scholar
Fenoglio, I., Fubini, B., Ghibaudi, E. M. & Turci, F. (2011) Multiple aspects of the interaction of biomacromolecules with inorganic surfaces. Advanced Drug Delivery Reviews 63, 11861209.Google Scholar
Finnis, M. W. (1991) The interaction of a point charge with an aluminium (111) surface. Surface Science 241, 6172.Google Scholar
Folliet, N., Gervais, C., Costa, D., Laurent, G., Babonneau, F., Stievano, L., Lambert, J.-F. & Tielens, F. (2013) A molecular picture of the adsorption of glycine in mesoporous silica through NMR experiments combined with DFT-D calculations. The Journal of Physical Chemistry C 117, 41044114.Google Scholar
Freeman, C. L., Asteriadis, I., Yang, M. & Harding, J. H. (2009) Interactions of organic molecules with calcite and magnesite surfaces. The Journal of Physical Chemistry C 113, 36663673.Google Scholar
Freeman, C. L., Harding, J. H., Cooke, D. J., Elliott, J. A., Lardge, J. S. & Duffy, D. M. (2007) New forcefields for modeling biomineralization processes. The Journal of Physical Chemistry C 111, 1194311951.Google Scholar
Freeman, C. L., Harding, J. H., Quigley, D. & Rodger, P. M. (2010) Structural control of crystal nuclei by an eggshell protein. Angewandte Chemie International Edition 49, 51355137.CrossRefGoogle ScholarPubMed
Freeman, C. L., Harding, J. H., Quigley, D. & Rodger, P. M. (2011) Simulations of ovocleidin-17 binding to calcite surfaces and its implications for eggshell formation. The Journal of Physical Chemistry C 115, 81758183.Google Scholar
Frenkel, D. & Smit, B. (2001) Understanding Molecular Simulation: From Algorithms to Applications, 2nd edition. Academic Press, Orlando, FL, USA.Google Scholar
Friddle, R. W., Battle, K., Trubetskoy, V., Tao, J., Salter, E. A., Moradian-Oldak, J., De Yoreo, J. J. & Wierzbicki, A. (2011) Single-molecule determination of the face-specific adsorption of amelogenin's C-terminus on hydroxyapatite. Angewandte Chemie International Edition 50, 75417545.Google Scholar
Friedrichs, W., Köppen, S. & Langel, W. (2013) Titanium binding dodecapeptides and the impact of water structure. Surface Science 617, 4252.Google Scholar
Friedrichs, W. & Langel, W. (2014) Atomistic modeling of peptide adsorption on rutile (100) in the presence of water and of contamination by low molecular weight alcohols. Biointerphases 9, 031006.Google Scholar
Fukunishi, H., Watanabe, O. & Takada, S. (2002) On the Hamiltonian replica exchange method for efficient sampling of biomolecular systems: Application to protein structure prediction. The Journal of Chemical Physics 116, 90589067.Google Scholar
Gabdoulline, R. R. & Wade, R. C. (2002) Biomolecular diffusional association. Current Opinion in Structural Biology 12, 204213.Google Scholar
Gagner, J. E., Lopez, M. D., Dordick, J. S. & Siegel, R. W. (2011) Effect of gold nanoparticle morphology on adsorbed protein structure and function. Biomaterials 32, 72417252.Google Scholar
Gagner, J. E., Shrivastava, S., Qian, X., Dordick, J. S. & Siegel, R. W. (2012) Engineering nanomaterials for biomedical applications requires understanding the nano-bio interface: a perspective. The Journal of Physical Chemistry Letters 3, 31493158.Google Scholar
Garland, A., Shen, L. & Zhu, X. (2012) Mobile precursor mediated protein adsorption on solid surfaces. Progress in Surface Science 87, 122 Google Scholar
Ghiringhelli, L. M., Hess, B., van der Vegt, N. F. A. & Delle Site, L. (2008) Competing adsorption between hydrated peptides and water onto metal surfaces: from electronic to conformational properties. Journal of the American Chemical Society 130, 1346013464.Google Scholar
Ghiringhelli, L. M., Schravendijk, P. & Delle Site, L. (2006) Adsorption of alanine on a Ni(111) surface: a multiscale modeling oriented density functional study. Physical Review B 74, 035437.Google Scholar
Ghosh, S., Jana, S. & Guchhait, N. (2012) Domain specific association of small fluorescent probe trans -3-(4-monomethylaminophenyl)-acrylonitrile (MMAPA) with bovine serum albumin (BSA) and its dissociation from protein binding sites by Ag nanoparticles: spectroscopic and molecular docking study. The Journal of Physical Chemistry B 116, 11551163.Google Scholar
Gibbs, D., Ocko, B. M., Zehner, D. M. & Mochrie, S. G. J. (1990) Structure and phases of the Au(001) surface: in-plane structure. Physical Review B 42, 73307344.CrossRefGoogle ScholarPubMed
Gooding, J. J., Mearns, F., Yang, W. & Liu, J. (2003) Self-assembled monolayers into the 21st century: recent advances and applications. Electroanalysis 15, 8196.Google Scholar
Gorba, C. & Helms, V. (2003) Diffusional dynamics of cytochrome c molecules in the presence of a charged surface. Soft Materials 1, 185202.Google Scholar
Gray, J. J. (2004) The interaction of proteins with solid surfaces. Current Opinion in Structural Biology 14, 110115.Google Scholar
Griepernau, B., Hanke, C., Santen, L., B, R. A. Hansmann, U. H. E., Meinke, J. H., Mohanty, S. & Nadler, W. (2008) Coarse-grained simulations of protein adsorption on solid surfaces. In From Computational Biophysics to Systems Biology (CBSB08), Proceedings of the NIC Workshop (eds Hansmann, U. H. E., Meinke, J. H., Mohanty, S., Nadler, W. & Zimmermann, O.), pp. 223226. John von Neumann Institute for Computing, Julich, Germany.Google Scholar
Grimme, S. (2004) Accurate description of van der Waals complexes by density functional theory including empirical corrections. Journal of Computational Chemistry 25, 14631473.Google Scholar
Grimme, S., Antony, J., Ehrlich, S. & Krieg, H. (2010) A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. The Journal of Chemical Physics 132, 154104.Google Scholar
Große Holthaus, S., Köppen, S., Frauenheim, T. & Colombi Ciacchi, L. (2012) Atomistic simulations of the  ZnO(12-10)/water interface: a comparison between first-principles, tight-binding, and empirical methods. Journal of Chemical Theory and Computation 8, 45174526.Google Scholar
Guldi, D. M., Rahman, G. M. A., Sgobba, V. & Ehli, C. (2006) Multifunctional molecular carbon materials—from fullerenes to carbon nanotubes. Chemical Society Reviews 35, 471487.CrossRefGoogle ScholarPubMed
Guo, J., Yao, X., Ning, L., Wang, Q. & Liu, H. (2014) The adsorption mechanism and induced conformational changes of three typical proteins with different secondary structural features on graphene. RSC Advances 4, 9953.Google Scholar
Hagiwara, T., Sakiyama, T. & Watanabe, H. (2009) Molecular simulation of bovine β-lactoglobulin adsorbed onto a positively charged solid surface. Langmuir 25, 226234.Google Scholar
Hahm, J. (2014) Fundamentals of nanoscale polymer–protein interactions and potential contributions to solid-state nanobioarrays. Langmuir 30, 98919904.Google Scholar
Halgren, T. A. & Damm, W. (2001) Polarizable force fields. Current Opinion in Structural Biology 11, 236242.Google Scholar
Hamdi, M., Ferreira, A., Sharma, G. & Mavroidis, C. (2008) Prototyping bio-nanorobots using molecular dynamics simulation and virtual reality. Microelectronics Journal 39, 190201.Google Scholar
Hansmann, U. H. E. (1997) Parallel tempering algorithm for conformational studies of biological molecules. Chemical Physics Letters 281, 140150.Google Scholar
Hansmann, U. H. E. & Okamoto, Y. (1999) New Monte Carlo algorithms for protein folding. Current Opinion in Structural Biology 9, 177183.Google Scholar
Hassan, S. A. & Steinbach, P. J. (2011) Water-exclusion and liquid-structure forces in implicit solvation. The Journal of Physical Chemistry B 115, 1466814682.Google Scholar
Hauptmann, S., Dufner, H., Brickmann, J., Kast, S. M. & Berry, R. S. (2003) Potential energy function for apatites. Physical Chemistry Chemical Physics 5, 635639.Google Scholar
Heinz, H., Farmer, B. L., Pandey, R. B., Slocik, J. M., Patnaik, S. S., Pachter, R. & Naik, R. R. (2009) Nature of molecular interactions of peptides with gold, palladium, and Pd−Au bimetal surfaces in aqueous solution. Journal of the American Chemical Society 131, 97049714.Google Scholar
Heinz, H., Jha, K. C., Luettmer-Strathmann, J., Farmer, B. L. & Naik, R. R. (2011) Polarization at metal-biomolecular interfaces in solution. Journal of the Royal Society Interface 8, 220232.Google Scholar
Heinz, H., Lin, T.-J., Kishore Mishra, R. & Emami, F. S. (2013) Thermodynamically consistent force fields for the assembly of inorganic, organic, and biological nanostructures: the INTERFACE force field. Langmuir 29, 17541765.Google Scholar
Heinz, H., Vaia, R. A., Farmer, B. L. & Naik, R. R. (2008) Accurate simulation of surfaces and interfaces of face-centered cubic metals using 12–6 and 9–6 Lennard-Jones potentials. The Journal of Physical Chemistry C 112, 1728117290.Google Scholar
Henderson, M. (2002) The interaction of water with solid surfaces: fundamental aspects revisited. Surface Science Reports 46, 1308.CrossRefGoogle Scholar
Hill, H. D., Vega, R. A. & Mirkin, C. A. (2007) Nonenzymatic detection of bacterial genomic DNA using the bio bar code assay. Analytical Chemistry 79, 92189223.Google Scholar
Hoefling, M., Iori, F., Corni, S. & Gottschalk, K.-E. (2010a) Interaction of amino acids with the Au(111) surface: adsorption free energies from molecular dynamics simulations. Langmuir 26, 83478351.Google Scholar
Hoefling, M., Iori, F., Corni, S. & Gottschalk, K.-E. (2010b) The conformations of amino acids on a gold(111) surface. ChemPhysChem 11, 17631767.Google Scholar
Hoefling, M., Monti, S., Corni, S. & Gottschalk, K. E. (2011) Interaction of β-sheet folds with a gold surface. PLoS ONE 6, e20925.Google Scholar
Hoffman, A. S. (2008) The origins and evolution of ‘controlled’ drug delivery systems. Journal of Controlled Release 132, 153163.Google Scholar
Hohenberg, P. & Kohn, W. (1964) Inhomogeneous electron gas. Physical Review 136, B864B871.Google Scholar
Hong, G., Heinz, H., Naik, R. R., Farmer, B. L. & Pachter, R. (2009) Toward understanding amino acid adsorption at metallic interfaces: a density functional theory study. ACS Applied Materials & Interfaces 1, 388392.Google Scholar
Horbett, T. A. & Brash, J. L. (1995) Proteins at Interfaces II. American Chemical Society, Washington, DC, USA.Google Scholar
Horinek, D., Serr, A., Bonthuis, D. J., Boström, M., Kunz, W. & Netz, R. R. (2008) Molecular hydrophobic attraction and ion-specific effects studied by molecular dynamics. Langmuir 24, 12711283.Google Scholar
Hou, T., Wang, J., Li, Y. & Wang, W. (2011) Assessing the performance of the MM/PBSA and MM/GBSA methods. 1. The accuracy of binding free energy calculations based on molecular dynamics simulations. Journal of Chemical Information and Modeling 51, 6982.Google Scholar
Hsu, H.-J., Sheu, S.-Y. & Tsay, R.-Y. (2008) Preferred orientation of albumin adsorption on a hydrophilic surface from molecular simulation. Colloids and Surfaces B: Biointerfaces 67, 183191.CrossRefGoogle ScholarPubMed
Hu, Y., Das, A., Hecht, M. H. & Scoles, G. (2005) Nanografting de novo proteins onto gold surfaces. Langmuir 21, 91039109.Google Scholar
Huang, L., Gubbins, K. E., Li, L. & Lu, X. (2014) Water on titanium dioxide surface: a revisiting by reactive molecular dynamics simulations. Langmuir 30, 1483214840.Google Scholar
Huang, X., Hagen, M., Kim, B., Friesner, R. A., Zhou, R. & Berne, B. J. (2007) Replica exchange with solute tempering: efficiency in large scale systems. The Journal of Physical Chemistry B 111, 54055410.Google Scholar
Hughes, Z. E., Tomásio, S. M. & Walsh, T. R. (2014) Efficient simulations of the aqueous bio-interface of graphitic nanostructures with a polarisable model. Nanoscale 6, 54385448.Google Scholar
Hung, A., Mwenifumbo, S., Mager, M., Kuna, J. J., Stellacci, F., Yarovsky, I. & Stevens, M. M. (2011) Ordering surfaces on the nanoscale: implications for protein adsorption. Journal of the American Chemical Society 133, 14381450.CrossRefGoogle ScholarPubMed
Imamura, K., Kawasaki, Y., Awadzu, T., Sakiyama, T. & Nakanishi, K. (2003) Contribution of acidic amino residues to the adsorption of peptides onto a stainless steel surface. Journal of Colloid and Interface Science 267, 294301.Google Scholar
Imamura, K., Kawasaki, Y., Nagayasu, T., Sakiyama, T. & Nakanishi, K. (2007) Adsorption characteristics of oligopeptides composed of acidic and basic amino acids on titanium surface. Journal of Bioscience and Bioengineering 103, 712.Google Scholar
Iori, F. & Corni, S. (2008) Including image charge effects in the molecular dynamics simulations of molecules on metal surfaces. Journal of Computational Chemistry 29, 16561666.Google Scholar
Iori, F., Corni, S. & Di Felice, R. (2008) Unraveling the interaction between histidine side chain and the Au(111) surface: a DFT study. The Journal of Physical Chemistry C 112, 1354013545.Google Scholar
Iori, F., Di Felice, R., Molinari, E. & Corni, S. (2009) GolP: An atomistic force-field to describe the interaction of proteins with Au(111) surfaces in water. Journal of Computational Chemistry 30, 14651476.Google Scholar
Iwamatsu, M. & Okabe, Y. (2004) Basin hopping with occasional jumping. Chemical Physics Letters 399, 396400.Google Scholar
Jackson, D. R., Omanovic, S. & Roscoe, S. G. (2000) Electrochemical studies of the adsorption behavior of serum proteins on titanium. Langmuir 16, 54495457.Google Scholar
Jain, P. K., Huang, X., El-Sayed, I. H. & El-Sayed, M. A. (2008) Noble metals on the nanoscale: optical and photothermal properties and some applications in imaging, sensing, biology, and medicine. Accounts of Chemical Research 41, 15781586.Google Scholar
Jarzynski, C. (1997) Nonequilibrium equality for free energy differences. Physical Review Letters 78, 26902693.Google Scholar
Jeyachandran, Y. L., Mielczarski, E., Rai, B. & Mielczarski, J. A. (2009) Quantitative and qualitative evaluation of adsorption/desorption of bovine serum albumin on hydrophilic and hydrophobic surfaces. Langmuir 25, 1161411620.Google Scholar
Jimenez-Izal, E., Chiatti, F., Corno, M., Rimola, A. & Ugliengo, P. (2012) Glycine adsorption at nonstoichiometric (010) hydroxyapatite surfaces: a B3LYP study. The Journal of Physical Chemistry C 116, 1456114567.Google Scholar
Jönsson, U., Fägerstam, L., Ivarsson, B., Johnsson, B., Karlsson, R., Lundh, K., Löfås, S., Persson, B., Roos, H. & Rönnberg, I. (1991) Real-time biospecific interaction analysis using surface plasmon resonance and a sensor chip technology. BioTechniques 11, 620627.Google Scholar
Jorgensen, W. L., Chandrasekhar, J., Madura, J. D., Impey, R. W. & Klein, M. L. (1983) Comparison of simple potential functions for simulating liquid water. The Journal of Chemical Physics 79, 926.Google Scholar
Jorgensen, W. L., Maxwell, D. S. & Tirado-Rives, J. (1996) Development and testing of the OPLS all-atom force field on conformational energetics and properties of organic liquids. Journal of the American Chemical Society 118, 1122511236.Google Scholar
Jose, J. C. & Sengupta, N. (2013) Molecular dynamics simulation studies of the structural response of an isolated Aβ1–42 monomer localized in the vicinity of the hydrophilic TiO2 surface. European Biophysics Journal 42, 487494.Google Scholar
Juffer, A. H., Argos, P. & de Vlieg, J. (1996) Adsorption of proteins onto charged surfaces: A Monte Carlo approach with explicit ions. Journal of Computational Chemistry 17, 17831803.Google Scholar
Kang, S., Huynh, T., Xia, Z., Zhang, Y., Fang, H., Wei, G. & Zhou, R. (2013) Hydrophobic interaction drives surface-assisted epitaxial assembly of amyloid-like peptides. Journal of the American Chemical Society 135, 31503157.Google Scholar
Kang, Y., Li, X., Tu, Y., Wang, Q. & Ågren, H. (2010) On the mechanism of protein adsorption onto hydroxylated and nonhydroxylated TiO2 surfaces. The Journal of Physical Chemistry C 114, 1449614502.Google Scholar
Kang, Y., Liu, Y.-C., Wang, Q., Shen, J.-W., Wu, T. & Guan, W.-J. (2009) On the spontaneous encapsulation of proteins in carbon nanotubes. Biomaterials 30, 28072815.Google Scholar
Kantarci, N., Tamerler, C., Sarikaya, M., Haliloglu, T. & Doruker, P. (2005) Molecular dynamics simulations on constraint metal binding peptides. Polymer 46, 43074313.Google Scholar
Kästner, J. & Thiel, W. (2005) Bridging the gap between thermodynamic integration and umbrella sampling provides a novel analysis method: ‘Umbrella integration’. The Journal of Chemical Physics 123, 144104.Google Scholar
Katti, D. R., Ghosh, P., Schmidt, S. & Katti, K. S. (2005a) Mechanical properties of the sodium montmorillonite interlayer intercalated with amino acids. Biomacromolecules 6, 32763282.Google Scholar
Katti, D. R., Schmidt, S. R., Ghosh, P. & Katti, K. S. (2005b) Modeling the response of pyrophyllite interlayer to applied stress using steered molecular dynamics. Clays and Clay Minerals 53, 171178.Google Scholar
Khlebtsov, N. G. & Dykman, L. A. (2010) Optical properties and biomedical applications of plasmonic nanoparticles. Journal of Quantitative Spectroscopy and Radiative Transfer 111, 135.Google Scholar
Kievit, F. M. & Zhang, M. (2011) Surface engineering of iron oxide nanoparticles for targeted cancer therapy. Accounts of Chemical Research 44, 853862.Google Scholar
Kim, M. S., Khang, G. & Lee, H. B. (2008) Gradient polymer surfaces for biomedical applications. Progress in Polymer Science 33, 138164.Google Scholar
Kim, S.-Y., Kumar, N., Persson, P., Sofo, J., van Duin, A. C. T. & Kubicki, J. D. (2013) Development of a ReaxFF reactive force field for titanium dioxide/water systems. Langmuir 29, 78387846.Google Scholar
Klimeš, J., Bowler, D. R. & Michaelides, A. (2010) Chemical accuracy for the van der Waals density functional. Journal of Physics: Condensed Matter 22, 022201.Google Scholar
Klimeš, J., Bowler, D. R. & Michaelides, A. (2011) Van der Waals density functionals applied to solids. Physical Review B 83, 195131.Google Scholar
Knotts, T. A., Rathore, N. & de Pablo, J. J. (2005) Structure and stability of a model three-helix-bundle protein on tailored surfaces. Proteins: Structure, Function, and Bioinformatics 61, 385397.Google Scholar
Knotts, T. A., Rathore, N. & de Pablo, J. J. (2008) An entropic perspective of protein stability on surfaces. Biophysical Journal 94, 44734483.Google Scholar
Koch, R., Lipton, A. S., Filipek, S. & Renugopalakrishnan, V. (2011) Arginine interactions with anatase TiO2 (100) surface and the perturbation of 49Ti NMR chemical shifts – a DFT investigation: relevance to Renu-Seeram bio solar cell. Journal of Molecular Modeling 17, 14671472.Google Scholar
Kohn, W. & Sham, L. J. (1965) Self-consistent equations including exchange and correlation effects. Physical Review 140, A1133A1138.CrossRefGoogle Scholar
Kokh, D. B., Corni, S., Winn, P. J., Hoefling, M., Gottschalk, K. E. & Wade, R. C. (2010) ProMetCS: An atomistic force field for modeling protein−metal surface interactions in a continuum aqueous solvent. Journal of Chemical Theory and Computation 6, 17531768.Google Scholar
Kollman, P. A., Massova, I., Reyes, C., Kuhn, B., Huo, S., Chong, L., Lee, M., Lee, T., Duan, Y., Wang, W., Donini, O., Cieplak, P., Srinivasan, J., Case, D. A. & Cheatham, T. E. (2000) Calculating structures and free energies of complex molecules: combining molecular mechanics and continuum models. Accounts of Chemical Research 33, 889897.Google Scholar
Köppen, S., Bronkalla, O. & Langel, W. (2008) Adsorption configurations and energies of amino acids on anatase and rutile surfaces. The Journal of Physical Chemistry C 112, 1360013606.Google Scholar
Köppen, S. & Langel, W. (2010) Simulation of adhesion forces and energies of peptides on titanium dioxide surfaces. Langmuir 26, 1524815256.Google Scholar
Korzeniewski, C., Climent, V. & Feliu, J. M. (2011) Electrochemistry at platinum single crystal electrodes. In Electroanalytical Chemistry (eds Bard, C. & Zoski, A. J.), pp. 75170. CRC Press, Boca Raton, FL, USA.Google Scholar
Kraszewski, S., Tarek, M., Treptow, W. & Ramseyer, C. (2010) Affinity of C60 neat fullerenes with membrane proteins: a computational study on potassium channels. ACS Nano 4, 41584164.Google Scholar
Kubiak-Ossowska, K. & Mulheran, P. A. (2010a) What governs protein adsorption and immobilization at a charged solid surface? Langmuir 26, 76907694.Google Scholar
Kubiak-Ossowska, K. & Mulheran, P. A. (2010b) Mechanism of hen egg white lysozyme adsorption on a charged solid surface. Langmuir 26, 1595415965.Google Scholar
Kubiak-Ossowska, K. & Mulheran, P. A. (2012) Protein diffusion and long-term adsorption states at charged solid surfaces. Langmuir 28, 1557715585.Google Scholar
Kumar, S., Rosenberg, J. M., Bouzida, D., Swendsen, R. H. & Kollman, P. A. (1992) The weighted histogram analysis method for free-energy calculations on biomolecules. I. The method. Journal of Computational Chemistry 13, 10111021.Google Scholar
Lacerda, S. H. D. P., Park, J. J., Meuse, C., Pristinski, D., Becker, M. L., Karim, A. & Douglas, J. F. (2010) Interaction of gold nanoparticles with common human blood proteins. ACS Nano 4, 365379.Google Scholar
Laera, S., Ceccone, G., Rossi, F., Gilliland, D., Hussain, R., Siligardi, G. & Calzolai, L. (2011) Measuring protein structure and stability of protein–nanoparticle systems with synchrotron radiation circular dichroism. Nano Letters 11, 44804484.Google Scholar
Laio, A. & Gervasio, F. L. (2008) Metadynamics: a method to simulate rare events and reconstruct the free energy in biophysics, chemistry and material science. Reports on Progress in Physics 71, 126601.Google Scholar
Laio, A. & Parrinello, M. (2002) Escaping free-energy minima. Proceedings of the National Academy of Sciences 99, 1256212566.Google Scholar
Lal, M., Plummer, M., Richmond, N. J. & Smith, W. (2004) Solvation of metal nanoparticles in a subcritical — supercritical fluid: a computer simulation study. The Journal of Physical Chemistry B 108, 60526061.Google Scholar
Lal, M., Plummer, M. & Smith, W. (2006) Solvent density effects on the solvation behavior and configurational structure of bare and passivated 38-atom gold nanoparticle in supercritical ethane. The Journal of Physical Chemistry B 110, 2087920888.Google Scholar
Landau, D. P. & Binder, K. (2009) A Guide to Monte Carlo Simulations in Statistical Physics. Cambridge University Press, Cambridge, UK.CrossRefGoogle Scholar
Latour, R. A. (2008) Molecular simulation of protein-surface interactions: benefits, problems, solutions, and future directions (Review). Biointerphases 3, FC2FC12.Google Scholar
Lee, K., Murray, É.D., Kong, L., Lundqvist, B. I. & Langreth, D. C. (2010) Higher-accuracy van der Waals density functional. Physical Review B 82, 081101.Google Scholar
Lee, S.-S., Kim, B. & Lee, S. (2014) Structures and bonding properties of gold–Arg-Cys complexes: DFT study of simple peptide-coated metal. The Journal of Physical Chemistry C 118, 2084020847.Google Scholar
Li, C., Monti, S., Ågren, H. & Carravetta, V. (2014) Cysteine on TiO2 (110): a theoretical study by reactive dynamics and photoemission spectra simulation. Langmuir 30, 88198828.Google Scholar
Li, C., Monti, S. & Carravetta, V. (2012) Journey toward the surface: how glycine adsorbs on titania in water solution. The Journal of Physical Chemistry C 116, 1831818326.Google Scholar
Li, H., Luo, Y., Derreumaux, P. & Wei, G. (2011) Carbon nanotube inhibits the formation of β-sheet-rich oligomers of the Alzheimer's amyloid-β(16-22) peptide. Biophysical Journal 101, 22672276.Google Scholar
Li, J., Liu, T., Li, X., Ye, L., Chen, H., Fang, H., Wu, Z. & Zhou, R. (2005) Hydration and dewetting near graphite−CH3 and graphite−COOH plates. The Journal of Physical Chemistry B 109, 1363913648.Google Scholar
Li, Y. & Gu, N. (2010) Thermodynamics of charged nanoparticle adsorption on charge-neutral membranes: a simulation study. The Journal of Physical Chemistry B 114, 27492754.Google Scholar
Li, Z. & Scheraga, H. (1987) Monte Carlo-minimization approach to the multiple-minima problem in protein folding. Proceedings of the National Academy of Sciences 84, 66116615.Google Scholar
Liang, J., Fieg, G., Keil, F. J. & Jakobtorweihen, S. (2012) Adsorption of proteins onto ion-exchange chromatographic media: a molecular dynamics study. Industrial & Engineering Chemistry Research 51, 1604916058.Google Scholar
Liao, C., Xie, Y. & Zhou, J. (2014) Computer simulations of fibronectin adsorption on hydroxyapatite surfaces. RSC Advances 4, 15759.Google Scholar
Liao, C. & Zhou, J. (2014) Replica-exchange molecular dynamics simulation of basic fibroblast growth factor adsorption on hydroxyapatite. The Journal of Physical Chemistry B 118, 58435852.Google Scholar
Liu, H. & Chakrabarti, A. (1999) Molecular dynamics study of adsorption and spreading of a polymer chain onto a flat surface. Polymer 40, 72857293.Google Scholar
Liu, J., Liao, C. & Zhou, J. (2013) Multiscale simulations of protein G B1 adsorbed on charged self-assembled monolayers. Langmuir 29, 1136611374.Google Scholar
Liu, P., Kim, B., Friesner, R. A. & Berne, B. J. (2005) Replica exchange with solute tempering: A method for sampling biological systems in explicit water. Proceedings of the National Academy of Sciences 102, 1374913754.Google Scholar
Liu, X., Chu, P. & Ding, C. (2004) Surface modification of titanium, titanium alloys, and related materials for biomedical applications. Materials Science and Engineering: R: Reports 47, 49121.Google Scholar
Liu, Y., Wu, M., Feng, X., Shao, X. & Cai, W. (2012) Adsorption behavior of hydrophobin proteins on polydimethylsiloxane substrates. The Journal of Physical Chemistry B 116, 1222712234.Google Scholar
Liu, Z. & Liang, X.-J. (2012) Nano-carbons as theranostics. Theranostics 2, 235237.Google Scholar
Lopes, P. E. M., Murashov, V., Tazi, M., Demchuk, E. & MacKerell, A. D. (2006) Development of an empirical force field for silica. Application to the quartz−water interface. The Journal of Physical Chemistry B 110, 27822792.Google Scholar
Love, J. C., Estroff, L. A., Kriebel, J. K., Nuzzo, R. G. & Whitesides, G. M. (2005) Self-assembled monolayers of thiolates on metals as a form of nanotechnology. Chemical Reviews 105, 11031170.Google Scholar
Lu, D. R., Lee, S. J. & Park, K. (1992) Calculation of solvation interaction energies for protein adsorption on polymer surfaces. Journal of Biomaterials Science, Polymer Edition 3, 127147.Google Scholar
Lund, M., Åkesson, T. & Jönsson, B. (2005) Enhanced protein adsorption due to charge regulation. Langmuir 21, 83858388.Google Scholar
Lundqvist, M., Sethson, I. & Jonsson, B.-H. (2004) Protein adsorption onto silica nanoparticles: conformational changes depend on the particles’ curvature and the protein stability. Langmuir 20, 1063910647.Google Scholar
MacKerell, A. D., Bashford, D., Bellott, M., Dunbrack, R. L., Evanseck, J. D., Field, M. J., Fischer, S., Gao, J., Guo, H., Ha, S., Joseph-McCarthy, D., Kuchnir, L., Kuczera, K., Lau, F. T. K., Mattos, C., Michnick, S., Ngo, T., Nguyen, D. T., Prodhom, B., Reiher, W. E., Roux, B., Schlenkrich, M., Smith, J. C., Stote, R., Straub, J., Watanabe, M., Wiórkiewicz-Kuczera, J., Yin, D. & Karplus, M. (1998) All-atom empirical potential for molecular modeling and dynamics studies of proteins. The Journal of Physical Chemistry B 102, 35863616.Google Scholar
Madura, J. D., Davist, M. E., Gilson, M. K., Wade, R. C., Luty, B. A. & McCammon, J. A. (1994) Biological applications of electrostatic calculations and Brownian dynamics simulations. In Reviews in Computational Chemistry, Volume 5 (eds Lipkowitz, K. B. & Boyd, D. B.), pp. 229267. John Wiley & Sons, Inc., New York, NY, USA.Google Scholar
Mahmoudi, M., Lynch, I., Ejtehadi, M. R., Monopoli, M. P., Bombelli, F. B. & Laurent, S. (2011) Protein−nanoparticle interactions: opportunities and challenges. Chemical Reviews 111, 56105637.Google Scholar
Mahon, E., Salvati, A., Baldelli Bombelli, F., Lynch, I. & Dawson, K. A. (2012) Designing the nanoparticle–biomolecule interface for ‘targeting and therapeutic delivery’. Journal of Controlled Release 161, 164174.Google Scholar
Mandal, D., Nasrolahi Shirazi, A. & Parang, K. (2014) Self-assembly of peptides to nanostructures. Organic & Biomolecular Chemistry 12, 35443561.Google Scholar
Manecka, G. M., Labrash, J., Rouxel, O., Dubot, P., Lalevée, J., Andaloussi, S. A., Renard, E., Langlois, V. & Versace, D. L. (2014) Green photoinduced modification of natural poly(3-hydroxybutyrate- co -3-hydroxyvalerate) surface for antibacterial applications. ACS Sustainable Chemistry & Engineering 2, 9961006.Google Scholar
Marcus, R. A. & Sutin, N. (1985) Electron transfers in chemistry and biology. Biochimica et Biophysica Acta (BBA) - Reviews on Bioenergetics 811, 265322.Google Scholar
Marrink, S. J., Risselada, H. J., Yefimov, S., Tieleman, D. P. & de Vries, A. H. (2007) The MARTINI force field: coarse grained model for biomolecular simulations. The Journal of Physical Chemistry B 111, 78127824.Google Scholar
Martin, R. M. (2004) Electronic structure: basic theory and practical methods. Cambridge University Press, Cambridge, UK.Google Scholar
Marx, D. & Hutter, J. (2000) Ab initio molecular dynamics: theory and implementation. In Modern Methods and Algorithms of Quantum Chemistry (ed Grotendorst, J.), pp. 301449. Forschungszentrum Jülich, Jülich, Germany.Google Scholar
Matsui, M. & Akaogi, M. (1991) Molecular dynamics simulation of the structural and physical properties of the four polymorphs of TiO2. Molecular Simulation 6, 239244.Google Scholar
Meißner, R. H., Schneider, J., Schiffels, P. & Colombi Ciacchi, L. (2014) Computational prediction of circular dichroism spectra and quantification of helicity loss upon peptide adsorption on silica. Langmuir 30, 34873494.Google Scholar
Mereghetti, P. & Wade, R. C. (2011) Diffusion of hydrophobin proteins in solution and interactions with a graphite surface. BMC Biophysics 4, 2.Google Scholar
Metropolis, N., Rosenbluth, A. W., Rosenbluth, M. N., Teller, A. H. & Teller, E. (1953) Equation of state calculations by fast computing machines. The Journal of Chemical Physics 21, 10871092.Google Scholar
Mijajlovic, M., Penna, M. J. & Biggs, M. J. (2013) Free energy of adsorption for a peptide at a liquid/solid interface via nonequilibrium molecular dynamics. Langmuir 29, 29192926.Google Scholar
Millo, D., Pandelia, M.-E., Utesch, T., Wisitruangsakul, N., Mroginski, M. A., Lubitz, W., Hildebrandt, P. & Zebger, I. (2009) Spectroelectrochemical study of the [NiFe] hydrogenase from Desulfovibrio vulgaris Miyazaki F in solution and immobilized on biocompatible gold surfaces. The Journal of Physical Chemistry B 113, 1534415351.Google Scholar
Mirau, P. A., Naik, R. R. & Gehring, P. (2011) Structure of peptides on metal oxide surfaces probed by NMR. Journal of the American Chemical Society 133, 1824318248.Google Scholar
Mishra, R. K., Flatt, R. J. & Heinz, H. (2013) Force field for tricalcium silicate and insight into nanoscale properties: cleavage, initial hydration, and adsorption of organic molecules. The Journal of Physical Chemistry C 117, 1041710432.Google Scholar
Mongan, J. & Case, D. A. (2005) Biomolecular simulations at constant pH. Current Opinion in Structural Biology 15, 157163.Google Scholar
Monti, S. (2007) RAD16II β-sheet filaments onto titanium dioxide: dynamics and adsorption properties. Journal of Physical Chemistry C 111, 1696216973.Google Scholar
Monti, S., Carravetta, V., Battocchio, C., Iucci, G. & Polzonetti, G. (2008) Peptide/TiO2 surface interaction: a theoretical and experimental study on the structure of adsorbed ALA-GLU and ALA-LYS. Langmuir 24, 32053214.Google Scholar
Mori, T., Hamers, R. J., Pedersen, J. A. & Cui, Q. (2013) An explicit consideration of desolvation is critical to binding free energy calculations of charged molecules at ionic surfaces. Journal of Chemical Theory and Computation 9, 50595069.Google Scholar
Motta, A., Gaigeot, M.-P. & Costa, D. (2012) AIMD evidence of inner sphere adsorption of glycine on a stepped (101) boehmite AlOOH surface. The Journal of Physical Chemistry C 116, 2341823427.Google Scholar
Mücksch, C. & Urbassek, H. M. (2011) Molecular dynamics simulation of free and forced BSA adsorption on a hydrophobic graphite surface. Langmuir 27, 1293812943.Google Scholar
Muir, J. M. R., Costa, D. & Idriss, H. (2014) DFT computational study of the RGD peptide interaction with the rutile TiO2 (110) surface. Surface Science 624, 814.CrossRefGoogle Scholar
Nada, H. (2014) Difference in the conformation and dynamics of aspartic acid on the flat regions, step edges, and kinks of a calcite surface: a molecular dynamics study. The Journal of Physical Chemistry C 118, 1433514345.Google Scholar
Nadler, R. & Sanz, J. F. (2012) Effect of dispersion correction on the Au(1 1 1)-H2O interface: a first-principles study. The Journal of Chemical Physics 137, 114709.Google Scholar
Nawrocki, G. & Cieplak, M. (2013) Amino acids and proteins at ZnO–water interfaces in molecular dynamics simulations. Physical Chemistry Chemical Physics 15, 1362813636.CrossRefGoogle ScholarPubMed
Nazmutdinov, R. R., Manyurov, I. R., Zinkicheva, T. T., Jang, J. & Ulstrup, J. (2007) Cysteine adsorption on the Au(111) surface and the electron transfer in configuration of a scanning tunneling microscope: a quantum-chemical approach. Russian Journal of Electrochemistry 43, 328341.Google Scholar
Neves, R. S., Motheo, A. J., Fartaria, R. P. S. & Silva Fernandes, F. M. S. (2007) Modelling water adsorption on Au(210) surfaces. I. A force field for water–Au interactions by DFT. Journal of Electroanalytical Chemistry 609, 140146.Google Scholar
Nie, Z. & Kumacheva, E. (2008) Patterning surfaces with functional polymers. Nature Materials 7, 277290.Google Scholar
Nie, Z., Petukhova, A. & Kumacheva, E. (2010) Properties and emerging applications of self-assembled structures made from inorganic nanoparticles. Nature Nanotechnology 5, 1525.Google Scholar
Nonella, M. & Seeger, S. (2008) Investigating alanine–silica interaction by means of first-principles molecular-dynamics simulations. ChemPhysChem 9, 414421.Google Scholar
Noon, W. H., Kong, Y. & Ma, J. (2002) Molecular dynamics analysis of a buckyball-antibody complex. Proceedings of the National Academy of Sciences 99, 64666470.Google Scholar
Notman, R. & Walsh, T. R. (2009) Molecular dynamics studies of the interactions of water and amino acid analogues with quartz surfaces. Langmuir 25, 16381644.Google Scholar
Novoselov, K. S. (2004) Electric field effect in atomically thin carbon films. Science 306, 666669.Google Scholar
Nowinski, A. K., Sun, F., White, A. D., Keefe, A. J. & Jiang, S. (2012) Sequence, structure, and function of peptide self-assembled monolayers. Journal of the American Chemical Society 134, 60006005.Google Scholar
O'Brien, C. P., Stuart, S. J., Bruce, D. A. & Latour, R. A. (2008) Modeling of peptide adsorption interactions with a poly(lactic acid) surface. Langmuir 24, 1411514124.Google Scholar
O'Mahony, S., O'Dwyer, C., Nijhuis, C. A., Greer, J. C., Quinn, A. J. & Thompson, D. (2013) Nanoscale dynamics and protein adhesivity of alkylamine self-assembled monolayers on graphene. Langmuir 29, 72717282.Google Scholar
Oberle, M., Yigit, C., Angioletti-Uberti, S., Dzubiella, J. & Ballauff, M. (2015) Competitive protein adsorption to soft polymeric layers: binary mixtures and comparison to theory. The Journal of Physical Chemistry B 119, 32503258.Google Scholar
Oostenbrink, C., Villa, A., Mark, A. E. & Van Gunsteren, W. F. (2004) A biomolecular force field based on the free enthalpy of hydration and solvation: the GROMOS force-field parameter sets 53A5 and 53A6. Journal of Computational Chemistry 25, 16561676.Google Scholar
Oren, E. E., Notman, R., Kim, I. W., Evans, J. S., Walsh, T. R., Samudrala, R., Tamerler, C. & Sarikaya, M. (2010) Probing the molecular mechanisms of quartz-binding peptides. Langmuir 26, 1100311009.Google Scholar
Oren, E. E., Tamerler, C. & Sarikaya, M. (2005) Metal recognition of septapeptides via polypod molecular architecture. Nano Letters 5, 415419.Google Scholar
Otsuka, H., Nagasaki, Y. & Kataoka, K. (2003) PEGylated nanoparticles for biological and pharmaceutical applications. Advanced Drug Delivery Reviews 55, 403419.Google Scholar
Palafox-Hernandez, J. P., Tang, Z., Hughes, Z. E., Li, Y., Swihart, M. T., Prasad, P. N., Walsh, T. R. & Knecht, M. R. (2014) Comparative study of materials-binding peptide interactions with gold and silver surfaces and nanostructures: a thermodynamic basis for biological selectivity of inorganic materials. Chemistry of Materials 26, 49604969.Google Scholar
Pan, H., Qin, M., Meng, W., Cao, Y. & Wang, W. (2012) How do proteins unfold upon adsorption on nanoparticle surfaces? Langmuir 28, 1277912787.Google Scholar
Park, S. Y., Lytton-Jean, A. K. R., Lee, B., Weigand, S., Schatz, G. C. & Mirkin, C. A. (2008) DNA-programmable nanoparticle crystallization. Nature 451, 553556.Google Scholar
Parks, G. A. (1965) The isoelectric points of solid oxides, solid hydroxides, and aqueous hydroxo complex systems. Chemical Reviews 65, 177198.Google Scholar
Patwardhan, S. V., Emami, F. S., Berry, R. J., Jones, S. E., Naik, R. R., Deschaume, O., Heinz, H. & Perry, C. C. (2012) Chemistry of aqueous silica nanoparticle surfaces and the mechanism of selective peptide adsorption. Journal of the American Chemical Society 134, 62446256.Google Scholar
Penna, M. J., Mijajlovic, M. & Biggs, M. J. (2014) Molecular-level understanding of protein adsorption at the interface between water and a strongly interacting uncharged solid surface. Journal of the American Chemical Society 136, 53235331.Google Scholar
Pincus, D. L., Cho, S. S., Hyeon, C. & Thirumalai, D. (2008) Minimal models for proteins and RNA: from folding to function. In Progress in molecular biology and translational science (ed Conn, P. M.), pp. 203250. Academic Press, San Diego, CA, USA.Google Scholar
Praprotnik, M., Site, L. D. & Kremer, K. (2008) Multiscale simulation of soft matter: from scale bridging to adaptive resolution. Annual Review of Physical Chemistry 59, 545571.Google Scholar
Předota, M., Bandura, A. V., Cummings, P. T., Kubicki, J. D., Wesolowski, D. J., Chialvo, A. A. & Machesky, M. L. (2004) Electric double layer at the rutile (110) surface. 1. Structure of surfaces and interfacial water from molecular dynamics by use of ab initio potentials. The Journal of Physical Chemistry B 108, 1204912060.Google Scholar
Puddu, V. & Perry, C. C. (2012) Peptide adsorption on silica nanoparticles: evidence of hydrophobic interactions. ACS Nano 6, 63566363.Google Scholar
Qin, L., Banholzer, M. J., Millstone, J. E. & Mirkin, C. A. (2007a) Nanodisk codes. Nano Letters 7, 38493853.Google Scholar
Qin, L., Banholzer, M. J., Xu, X., Huang, L. & Mirkin, C. A. (2007b) Rational design and synthesis of catalytically driven nanorotors. Journal of the American Chemical Society 129, 1487014871.Google Scholar
Qin, Z. & Buehler, M. J. (2014) Molecular mechanics of mussel adhesion proteins. Journal of the Mechanics and Physics of Solids 62, 1930.Google Scholar
Qu, Z. G., He, X. C., Lin, M., Sha, B. Y., Shi, X. H., Lu, T. J. & Xu, F. (2013) Advances in the understanding of nanomaterial–biomembrane interactions and their mathematical and numerical modeling. Nanomedicine 8, 9951011.Google Scholar
Rabe, M., Verdes, D. & Seeger, S. (2011) Understanding protein adsorption phenomena at solid surfaces. Advances in Colloid and Interface Science 162, 87106.Google Scholar
Raffaini, G. & Ganazzoli, F. (2003) Simulation study of the interaction of some albumin subdomains with a flat graphite surface. Langmuir 19, 34033412.Google Scholar
Raffaini, G. & Ganazzoli, F. (2004a) Molecular dynamics simulation of the adsorption of a fibronectin module on a graphite surface. Langmuir 20, 33713378.Google Scholar
Raffaini, G. & Ganazzoli, F. (2004b) Surface ordering of proteins adsorbed on graphite. The Journal of Physical Chemistry B 108, 1385013854.Google Scholar
Raffaini, G. & Ganazzoli, F. (2006) Adsorption of charged albumin subdomains on a graphite surface. Journal of Biomedical Materials Research Part A 76A, 638645.Google Scholar
Raffaini, G. & Ganazzoli, F. (2007) Understanding the performance of biomaterials through molecular modeling: crossing the bridge between their intrinsic properties and the surface adsorption of proteins. Macromolecular Bioscience 7, 552566.Google Scholar
Raffaini, G. & Ganazzoli, F. (2010) Protein adsorption on a hydrophobic surface: a molecular dynamics study of lysozyme on graphite. Langmuir 26, 56795689.Google Scholar
Raffaini, G. & Ganazzoli, F. (2012) Molecular modelling of protein adsorption on the surface of titanium dioxide polymorphs. Philosophical Transactions of the Royal Society A 370, 14441462.Google Scholar
Raffaini, G. & Ganazzoli, F. (2013) Surface topography effects in protein adsorption on nanostructured carbon allotropes. Langmuir 29, 48834893.Google Scholar
Ravichandran, S., Madura, J. D. & Talbot, J. (2001) A Brownian dynamics study of the initial stages of hen egg-white lysozyme adsorption at a solid interface. The Journal of Physical Chemistry B 105, 36103613.Google Scholar
Ravichandran, S. & Talbot, J. (2000) Mobility of adsorbed proteins: a Brownian dynamics study. Biophysical Journal 78, 110120.Google Scholar
Rechendorff, K., Hovgaard, M. B., Foss, M., Zhdanov, V. P. & Besenbacher, F. (2006) Enhancement of protein adsorption induced by surface roughness. Langmuir 22, 1088510888.Google Scholar
Ren, P. & Ponder, J. W. (2002) Consistent treatment of inter- and intramolecular polarization in molecular mechanics calculations. Journal of Computational Chemistry 23, 14971506.Google Scholar
Ren, P. & Ponder, J. W. (2003) Polarizable atomic multipole water model for molecular mechanics simulation. The Journal of Physical Chemistry B 107, 59335947.Google Scholar
Rick, S. W., Stuart, S. J. & Berne, B. J. (1994) Dynamical fluctuating charge force fields: application to liquid water. The Journal of Chemical Physics 101, 6141.Google Scholar
Rimola, A., Aschi, M., Orlando, R. & Ugliengo, P. (2012) Does adsorption at hydroxyapatite surfaces induce peptide folding? Insights from large-scale B3LYP calculations. Journal of the American Chemical Society 134, 1089910910.Google Scholar
Rimola, A., Costa, D., Sodupe, M., Lambert, J.-F. & Ugliengo, P. (2013) Silica surface features and their role in the adsorption of biomolecules: computational modeling and experiments. Chemical Reviews 113, 42164313.Google Scholar
Rimola, A., Sodupe, M. & Ugliengo, P. (2009) Affinity scale for the interaction of amino acids with silica surfaces. The Journal of Physical Chemistry C 113, 57415750.Google Scholar
Roach, P., Farrar, D. & Perry, C. C. (2006) Surface tailoring for controlled protein adsorption: effect of topography at the nanometer scale and chemistry. Journal of the American Chemical Society 128, 39393945.Google Scholar
Roman, T., Diño, W. A., Nakanishi, H. & Kasai, H. (2006) Amino acid adsorption on single-walled carbon nanotubes. The European Physical Journal D 38, 117120.Google Scholar
Rosa, M., Corni, S. & Di Felice, R. (2012) A density functional theory study of cytosine on Au(111). The Journal of Physical Chemistry C 116, 2136621373.Google Scholar
Rosa, M., Corni, S. & Di Felice, R. (2014a) Enthalpy–entropy tuning in the adsorption of nucleobases at the Au(111) surface. Journal of Chemical Theory and Computation 10, 17071716.Google Scholar
Rosa, M., Corni, S. & Di Felice, R. (2014b) van der Waals effects at molecule-metal interfaces. Physical Review B 90, 125448.Google Scholar
Salata, O. (2004) Applications of nanoparticles in biology and medicine. Journal of Nanobiotechnology 2, 3.Google Scholar
Saptarshi, S. R., Duschl, A. & Lopata, A. L. (2013) Interaction of nanoparticles with proteins: relation to bio-reactivity of the nanoparticle. Journal of Nanobiotechnology 11, 26.Google Scholar
Schlick, T. (2002) Molecular Modeling and Simulation. Springer, New York, NY, USA.Google Scholar
Schneider, J. & Colombi Ciacchi, L. (2010) First principles and classical modeling of the oxidized titanium (0001) surface. Surface Science 604, 11051115.Google Scholar
Schneider, J. & Colombi Ciacchi, L. (2011) A classical potential to model the adsorption of biological molecules on oxidized titanium surfaces. Journal of Chemical Theory and Computation 7, 473484.Google Scholar
Schneider, J. & Colombi Ciacchi, L. (2012) Specific material recognition by small peptides mediated by the interfacial solvent structure. Journal of the American Chemical Society 134, 24072413.Google Scholar
Schreiber, F. (2004) Self-assembled monolayers: from simple model systems to biofunctionalized interfaces. Journal of Physics: Condensed Matter 16, R881R900.Google Scholar
Schyman, P. & Jorgensen, W. L. (2013) Exploring adsorption of water and ions on carbon surfaces using a polarizable force field. The Journal of Physical Chemistry Letters 4, 468474.Google Scholar
Scida, K., Stege, P. W., Haby, G., Messina, G. A. & García, C. D. (2011) Recent applications of carbon-based nanomaterials in analytical chemistry: critical review. Analytica Chimica Acta 691, 617.Google Scholar
Senaratne, W., Andruzzi, L. & Ober, C. K. (2005) Self-assembled monolayers and polymer brushes in biotechnology: current applications and future perspectives. Biomacromolecules 6, 24272448.Google Scholar
Seo, M., Rauscher, S., Pomès, R. & Tieleman, D. P. (2012) Improving internal peptide dynamics in the coarse-grained MARTINI model: toward large-scale simulations of amyloid- and elastin-like peptides. Journal of Chemical Theory and Computation 8, 17741785.Google Scholar
Sethuraman, A. & Belfort, G. (2005) Protein structural perturbation and aggregation on homogeneous surfaces. Biophysical Journal 88, 13221333.Google Scholar
Shaw, C. P., Middleton, D. A., Volk, M. & Lévy, R. (2012) Amyloid-derived peptide forms self-assembled monolayers on gold nanoparticle with a curvature-dependent β-sheet structure. ACS Nano 6, 14161426.Google Scholar
Shelley, J. C., Patey, G. N., Bérard, D. R. & Torrie, G. M. (1997) Modeling and structure of mercury-water interfaces. The Journal of Chemical Physics 107, 2122.Google Scholar
Shemetov, A. A., Nabiev, I. & Sukhanova, A. (2012) Molecular interaction of proteins and peptides with nanoparticles. ACS Nano 6, 45854602.Google Scholar
Shen, J., Wu, T., Wang, Q. & Pan, H. (2008) Molecular simulation of protein adsorption and desorption on hydroxyapatite surfaces. Biomaterials 29, 513532.Google Scholar
Shin, H., Jo, S. & Mikos, A. G. (2003) Biomimetic materials for tissue engineering. Biomaterials 24, 43534364.Google Scholar
Siepmann, J. I. & Sprik, M. (1995) Influence of surface topology and electrostatic potential on water/electrode systems. The Journal of Chemical Physics 102, 511.Google Scholar
Singh, N. & Warshel, A. (2010) Absolute binding free energy calculations: on the accuracy of computational scoring of protein-ligand interactions. Proteins: Structure, Function, and Bioinformatics 78, 17051723.Google Scholar
Siwko, M. E. & Corni, S. (2013) Cytochrome c on a gold surface: investigating structural relaxations and their role in protein–surface electron transfer by molecular dynamics simulations. Physical Chemistry Chemical Physics 15, 59455956.Google Scholar
Skelton, A. A., Fenter, P., Kubicki, J. D., Wesolowski, D. J. & Cummings, P. T. (2011) Simulations of the quartz(1011)/water interface: a comparison of classical force fields, ab initio molecular dynamics, and x-ray reflectivity experiments. The Journal of Physical Chemistry C 115, 20762088.Google Scholar
Skepö, M. (2008) Model simulations of the adsorption of statherin to solid surfaces: effects of surface charge and hydrophobicity. The Journal of Chemical Physics 129, 185101185111.Google Scholar
Skepö, M., Linse, P. & Arnebrant, T. (2006) Coarse-grained modeling of proline rich protein 1 (PRP-1) in bulk solution and adsorbed to a negatively charged surface. The Journal of Physical Chemistry B 110, 1214112148.Google Scholar
Slocik, J. M., Govorov, A. O. & Naik, R. R. (2011) Plasmonic circular dichroism of peptide-functionalized gold nanoparticles. Nano Letters 11, 701705.Google Scholar
Smith, N. V., Chen, C. T. & Weinert, M. (1989) Distance of the image plane from metal surfaces. Physical Review B 40, 75657573.Google Scholar
Snyder, J. A., Abramyan, T., Yancey, J. A., Thyparambil, A. A., Wei, Y., Stuart, S. J. & Latour, R. A. (2012) Development of a tuned interfacial force field parameter set for the simulation of protein adsorption to silica glass. Biointerphases 7, 56.Google Scholar
Snyder, J. A. & Madura, J. D. (2008) Interaction of the phospholipid head group with representative quartz and aluminosilicate structures: an ab initio study. The Journal of Physical Chemistry B 112, 70957103.Google Scholar
Somorjai, G. A. & Li, Y. (2011) Impact of surface chemistry. Proceedings of the National Academy of Sciences 108, 917924.Google Scholar
Souaille, M. & Roux, B. (2001) Extension to the weighted histogram analysis method: combining umbrella sampling with free energy calculations. Computer Physics Communications 135, 4057.Google Scholar
Spohr, E. (1995) Computer modeling of interfaces between aqueous and metallic phases. Acta Chemica Scandinavica 49, 189202.Google Scholar
Srinivasan, J., Cheatham, T. E., Cieplak, P., Kollman, P. A. & Case, D. A. (1998) Continuum solvent studies of the stability of DNA, RNA, and phosphoramidate−DNA helices. Journal of the American Chemical Society 120, 94019409.Google Scholar
Steckbeck, S., Schneider, J., Wittig, L., Rischka, K., Grunwald, I. & Colombi Ciacchi, L. (2014) Identification of materials’ binding peptide sequences guided by a MALDI-ToF MS depletion assay. Analytical Methods 6, 15011509.Google Scholar
Strodel, B. & Wales, D. J. (2008) Implicit solvent models and the energy landscape for aggregation of the amyloidogenic KFFE peptide. Journal of Chemical Theory and Computation 4, 657672.Google Scholar
Stuart, D. A., Haes, A. J., Yonzon, C. R., Hicks, E. M. & Van Duyne, R. P. (2005) Biological applications of localised surface plasmonic phenomenae. IEE Proceedings - Nanobiotechnology 152, 1332.Google Scholar
Stuart, M. A. C., Huck, W. T. S., Genzer, J., Müller, M., Ober, C., Stamm, M., Sukhorukov, G. B., Szleifer, I., Tsukruk, V. V, Urban, M., Winnik, F., Zauscher, S., Luzinov, I. & Minko, S. (2010) Emerging applications of stimuli-responsive polymer materials. Nature Materials 9, 101113.CrossRefGoogle ScholarPubMed
Sugita, Y. & Okamoto, Y. (1999) Replica-exchange molecular dynamics method for protein folding. Chemical Physics Letters 314, 141151.Google Scholar
Sugiyama, N., Masuda, T., Shinoda, K., Nakamura, A., Tomita, M. & Ishihama, Y. (2007) Phosphopeptide enrichment by aliphatic hydroxy acid-modified metal oxide chromatography for nano-LC-MS/MS in proteomics applications. Molecular & Cellular Proteomics 6, 11031109.Google Scholar
Sumper, M. & Brunner, E. (2008) Silica biomineralisation in diatoms: the model organism Thalassiosira pseudonana. ChemBioChem 9, 11871194.Google Scholar
Sun, C., Liu, L.-M., Selloni, A., Lu, G. Q. (Max) & Smith, S. C. (2010) Titania-water interactions: a review of theoretical studies. Journal of Materials Chemistry 20, 10319.Google Scholar
Sun, H. (1998) COMPASS: An ab initio force-field optimized for condensed-phase applications–overview with details on alkane and benzene compounds. The Journal of Physical Chemistry B 102, 73387364.Google Scholar
Sun, T., Han, G., Lindgren, M., Shen, Z. & Laaksonen, A. (2014a) Adhesion of lactoferrin and bone morphogenetic protein-2 to a rutile surface: dependence on the surface hydrophobicity. Biomaterials Science 2, 10901099.Google Scholar
Sun, X., Feng, Z., Hou, T. & Li, Y. (2014b) Mechanism of graphene oxide as an enzyme inhibitor from molecular dynamics simulations. ACS Applied Materials & Interfaces 6, 71537163.Google Scholar
Sun, Y., Dominy, B. N. & Latour, R. A. (2007) Comparison of solvation-effect methods for the simulation of peptide interactions with a hydrophobic surface. Journal of Computational Chemistry 28, 18831892.Google Scholar
Sun, Y. & Latour, R. A. (2006) Comparison of implicit solvent models for the simulation of protein–surface interactions. Journal of Computational Chemistry 27, 19081922.Google Scholar
Sun, Y., Welsh, W. J. & Latour, R. A. (2005) Prediction of the orientations of adsorbed protein using an empirical energy function with implicit solvation. Langmuir 21, 56165626.Google Scholar
Suttiponparnit, K., Jiang, J., Sahu, M., Suvachittanont, S., Charinpanitkul, T. & Biswas, P. (2010) Role of surface area, primary particle size, and crystal phase on titanium dioxide nanoparticle dispersion properties. Nanoscale Research Letters 6, 27.Google Scholar
Swanson, J. M. J., Henchman, R. H. & McCammon, J. A. (2004) Revisiting free energy calculations: a theoretical connection to MM/PBSA and direct calculation of the association free energy. Biophysical Journal 86, 6774.Google Scholar
Swendsen, R. H. & Wang, J.-S. (1986) Replica Monte Carlo simulation of spin-glasses. Physical Review Letters 57, 26072609.Google Scholar
Szott, L. M. & Horbett, T. A. (2011) Protein interactions with surfaces: computational approaches and repellency. Current Opinion in Chemical Biology 15, 683689.Google Scholar
Tallury, S. S. & Pasquinelli, M. A. (2010) Molecular dynamics simulations of flexible polymer chains wrapping single-walled carbon nanotubes. The Journal of Physical Chemistry B 114, 41224129.Google Scholar
Tang, Z., Palafox-Hernandez, J. P., Law, W.-C., E. Hughes, Z., Swihart, M. T., Prasad, P. N., Knecht, M. R. & Walsh, T. R. (2013) Biomolecular recognition principles for bionanocombinatorics: an integrated approach to elucidate enthalpic and entropic factors. ACS Nano 7, 96329646.Google Scholar
Tavanti, F., Pedone, A. & Menziani, M. C. (2015) A closer look into the ubiquitin corona on gold nanoparticles by computational studies. New Journal of Chemistry 39, 24742482.Google Scholar
Taylor, R. D., Jewsbury, P. J. & Essex, J. W. (2002) A review of protein-small molecule docking methods. Journal of Computer-Aided Molecular Design 16, 151166.Google Scholar
Tkatchenko, A., Romaner, L., Hofmann, O. T., Zojer, E., Ambrosch-Draxl, C. & Scheffler, M. (2010) Van der Waals interactions between organic adsorbates and at organic/inorganic interfaces. MRS Bulletin 35, 435442.Google Scholar
Todorova, N., Chiappini, C., Mager, M., Simona, B., Patel, I. I., Stevens, M. M. & Yarovsky, I. (2014) Surface presentation of functional peptides in solution determines cell internalization efficiency of TAT conjugated nanoparticles. Nano Letters 14, 52295237.Google Scholar
Todorova, N., Makarucha, A. J., Hine, N. D. M., Mostofi, A. A. & Yarovsky, I. (2013) Dimensionality of carbon nanomaterials determines the binding and dynamics of amyloidogenic peptides: multiscale theoretical simulations. PLoS Computational Biology 9, e1003360.Google Scholar
Tomásio, S. M. & Walsh, T. R. (2009) Modeling the binding affinity of peptides for graphitic surfaces. Influences of aromatic content and interfacial shape. The Journal of Physical Chemistry C 113, 87788785.Google Scholar
Tomba, G., Colombi Ciacchi, L. & De Vita, A. (2009) Atomic-level studies of molecular self-assembly on metallic surfaces. Advanced Materials 21, 10551066.Google Scholar
Torrie, G. M. & Valleau, J. P. (1974) Monte Carlo free energy estimates using non-Boltzmann sampling: application to the sub-critical Lennard-Jones fluid. Chemical Physics Letters 28, 578581.Google Scholar
Torrie, G. M. & Valleau, J. P. (1977) Nonphysical sampling distributions in Monte Carlo free-energy estimation: umbrella sampling. Journal of Computational Physics 23, 187199.Google Scholar
Tosaka, R., Yamamoto, H., Ohdomari, I. & Watanabe, T. (2010) Adsorption mechanism of ribosomal protein L2 onto a silica surface: a molecular dynamics simulation study. Langmuir 26, 99509955.Google Scholar
Tozzini, V. (2005) Coarse-grained models for proteins. Current Opinion in Structural Biology 15, 144150.Google Scholar
Utesch, T., Daminelli, G. & Mroginski, M. A. (2011) Molecular dynamics simulations of the adsorption of bone morphogenetic protein-2 on surfaces with medical relevance. Langmuir 27, 1314413153.Google Scholar
Utesch, T., Millo, D., Castro, M. A., Hildebrandt, P., Zebger, I. & Mroginski, M. A. (2013) Effect of the protonation degree of a self-assembled monolayer on the immobilization dynamics of a [NiFe] hydrogenase. Langmuir 29, 673682.Google Scholar
Utesch, T., Sezer, M., Weidinger, I. M. & Mroginski, M. A. (2012) Adsorption of sulfite oxidase on self-assembled monolayers from molecular dynamics simulations. Langmuir 28, 57615769.Google Scholar
Vallee, A., Humblot, V. & Pradier, C.-M. (2010) Peptide interactions with metal and oxide surfaces. Accounts of Chemical Research 43, 12971306.Google Scholar
Van Beest, B. W. H., Kramer, G. J. & van Santen, R. A. (1990) Force fields for silicas and aluminophosphates based on ab initio calculations. Physical Review Letters 64, 19551958.Google Scholar
Van Duin, A. C. T., Dasgupta, S., Lorant, F. & Goddard, W. A. (2001) ReaxFF: a reactive force field for hydrocarbons. The Journal of Physical Chemistry A 105, 93969409.Google Scholar
Vashishta, P., Kalia, R. K., Rino, J. P. & Ebbsjö, I. (1990) Interaction potential for SiO2: A molecular-dynamics study of structural correlations. Physical Review B 41, 1219712209.Google Scholar
Velasco-Velez, J.-J., Wu, C. H., Pascal, T. A., Wan, L. F., Guo, J., Prendergast, D. & Salmeron, M. (2014) The structure of interfacial water on gold electrodes studied by x-ray absorption spectroscopy. Science 346, 831834.Google Scholar
Vellore, N. A., Yancey, J. A., Collier, G., Latour, R. A. & Stuart, S. J. (2010) Assessment of the transferability of a protein force field for the simulation of peptide-surface interactions. Langmuir 26, 73967404.Google Scholar
Venkat, A. S., Corni, S. & Di Felice, R. (2007) Electronic coupling between azurin and gold at different protein/substrate orientations. Small 3, 14311437.Google Scholar
Vertegel, A. A., Siegel, R. W. & Dordick, J. S. (2004) Silica nanoparticle size influences the structure and enzymatic activity of adsorbed lysozyme. Langmuir 20, 68006807.Google Scholar
Vigmond, S. J., Iwakura, M., Mizutani, F. & Katsura, T. (1994) Site-specific immobilization of molecularly engineered dihydrofolate reductase to gold surfaces. Langmuir 10, 28602862.Google Scholar
Vila Verde, A., Acres, J. M. & Maranas, J. K. (2009) Investigating the specificity of peptide adsorption on gold using molecular dynamics simulations. Biomacromolecules 10, 21182128.Google Scholar
Vila Verde, A., Beltramo, P. J. & Maranas, J. K. (2011) Adsorption of homopolypeptides on gold investigated using atomistic molecular dynamics. Langmuir 27, 59185926.Google Scholar
Voicescu, M., Ionescu, S. & Angelescu, D. G. (2012) Spectroscopic and coarse-grained simulation studies of the BSA and HSA protein adsorption on silver nanoparticles. Journal of Nanoparticle Research 14, 1174.Google Scholar
Voskerician, G., Shive, M. S., Shawgo, R. S., von Recum, H., Anderson, J. M., Cima, M. J. & Langer, R. (2003) Biocompatibility and biofouling of MEMS drug delivery devices. Biomaterials 24, 19591967.Google Scholar
Wales, D. J. & Doye, J. P. K. (1997) Global optimization by basin-hopping and the lowest energy structures of Lennard-Jones clusters containing up to 110 atoms. The Journal of Physical Chemistry A 101, 51115116.Google Scholar
Walsh, T. R. (2008) Modelling the nanoscale patterning of nucleic acid base pairs deposited on graphite. Molecular Physics 106, 16131619.Google Scholar
Walther, J. H., Jaffe, R., Halicioglu, T. & Koumoutsakos, P. (2001) Carbon nanotubes in water: structural characteristics and energetics. The Journal of Physical Chemistry B 105, 99809987.Google Scholar
Wang, D., Nap, R. J., Lagzi, I., Kowalczyk, B., Han, S., Grzybowski, B. A. & Szleifer, I. (2011) How and why nanoparticle's curvature regulates the apparent pKa of the coating ligands. Journal of the American Chemical Society 133, 21922197.Google Scholar
Wang, F., Stuart, S. J. & Latour, R. A. (2008) Calculation of adsorption free energy for solute-surface interactions using biased replica-exchange molecular dynamics. Biointerphases 3, 918.Google Scholar
Wang, M., Guo, Y., Wang, Q., Zhang, X., Huang, J., Lu, X., Wang, K., Zhang, H. & Leng, Y. (2014) Density functional theory study of interactions between glycine and TiO2/graphene nanocomposites. Chemical Physics Letters 599, 8691.Google Scholar
Wang, Q., Zhao, C., Zhao, J., Wang, J., Yang, J.-C., Yu, X. & Zheng, J. (2010a) Comparative molecular dynamics study of Aβ adsorption on the self-assembled monolayers. Langmuir 26, 33083316.Google Scholar
Wang, Q., Zhao, J., Yu, X., Zhao, C., Li, L. & Zheng, J. (2010b) Alzheimer Aβ 1–42 monomer adsorbed on the self-assembled monolayers. Langmuir 26, 1272212732.Google Scholar
Wang, S., Humphreys, E. S., Chung, S.-Y., Delduco, D. F., Lustig, S. R., Wang, H., Parker, K. N., Rizzo, N. W., Subramoney, S., Chiang, Y.-M. & Jagota, A. (2003) Peptides with selective affinity for carbon nanotubes. Nature Materials 2, 196200.Google Scholar
Wei, Q., Becherer, T., Angioletti-Uberti, S., Dzubiella, J., Wischke, C., Neffe, A. T., Lendlein, A., Ballauff, M. & Haag, R. (2014) Protein interactions with polymer coatings and biomaterials. Angewandte Chemie International Edition 53, 80048031.Google Scholar
Wei, T., Carignano, M. A. & Szleifer, I. (2011) Lysozyme adsorption on polyethylene surfaces: why are long simulations needed? Langmuir 27, 1207412081.Google Scholar
Wei, Y. & Latour, R. A. (2010) Correlation between desorption force measured by atomic force microscopy and adsorption free energy measured by surface plasmon resonance spectroscopy for peptide − surface interactions. Langmuir 26, 1885218861.Google Scholar
Wei, Y. & Latour, R. A. (2009) Benchmark experimental data set and assessment of adsorption free energy for peptide−surface interactions. Langmuir 25, 56375646.Google Scholar
Whaley, S. R., English, D. S., Hu, E. L., Barbara, P. F. & Belcher, A. M. (2000) Selection of peptides with semiconductor binding specificity for directed nanocrystal assembly. Nature 405, 665668.Google Scholar
Wierzbicki, A., Sikes, C. S., Madura, J. D. & Drake, B. (1994) Atomic force microscopy and molecular modeling of protein and peptide binding to calcite. Calcified Tissue International 54, 133141.Google Scholar
Wilt, F. H. (1999) Matrix and mineral in the sea urchin larval skeleton. Journal of Structural Biology 126, 216226.Google Scholar
Wright, L. B., Rodger, P. M., Corni, S. & Walsh, T. R. (2013a) GolP-CHARMM: First-principles based force fields for the interaction of proteins with Au(111) and Au(100). Journal of Chemical Theory and Computation 9, 16161630.Google Scholar
Wright, L. B., Rodger, P. M., Walsh, T. R. & Corni, S. (2013b) First-principles-based force field for the interaction of proteins with Au(100)(5×1): an extension of GolP-CHARMM. The Journal of Physical Chemistry C 117, 2429224306.Google Scholar
Wright, L. B. & Walsh, T. R. (2012) First-principles molecular dynamics simulations of NH4(+) and CH3COO(−) adsorption at the aqueous quartz interface. The Journal of Chemical Physics 137, 224702.Google Scholar
Wright, L. B. & Walsh, T. R. (2013) Efficient conformational sampling of peptides adsorbed onto inorganic surfaces: insights from a quartz binding peptide. Physical Chemistry Chemical Physics 15, 47154726.Google Scholar
Wu, C., Chen, M., Skelton, A. A., Cummings, P. T. & Zheng, T. (2013) Adsorption of arginine-glycine-aspartate tripeptide onto negatively charged rutile (110) mediated by cations: the effect of surface hydroxylation. ACS Applied Materials & Interfaces 5, 25672579.Google Scholar
Wu, X. & Narsimhan, G. (2008) Effect of surface concentration on secondary and tertiary conformational changes of lysozyme adsorbed on silica nanoparticles. Biochimica et Biophysica Acta 1784, 16941701.Google Scholar
Wu, X. & Narsimhan, G. (2009) Coarse grain molecular dynamics simulation for the prediction of tertiary conformation of lysozyme adsorbed on silica surface. Molecular Simulation 35, 974985.Google Scholar
Wu, X., Vargas, M. C., Nayak, S., Lotrich, V. & Scoles, G. (2001) Towards extending the applicability of density functional theory to weakly bound systems. The Journal of Chemical Physics 115, 8748.Google Scholar
Xie, L., Luo, Y., Lin, D., Xi, W., Yang, X. & Wei, G. (2014) The molecular mechanism of fullerene-inhibited aggregation of Alzheimer's β-amyloid peptide fragment. Nanoscale 6, 97529762.Google Scholar
Xie, Y., Liu, M. & Zhou, J. (2012) Molecular dynamics simulations of peptide adsorption on self-assembled monolayers. Applied Surface Science 258, 81538159.Google Scholar
Xie, Y., Zhou, J. & Jiang, S. (2010) Parallel tempering Monte Carlo simulations of lysozyme orientation on charged surfaces. The Journal of Chemical Physics 132, 065101.Google Scholar
Xu, Y., Cao, Q., Svec, F. & Fréchet, J. M. J. (2010) Porous polymer monolithic column with surface-bound gold nanoparticles for the capture and separation of cysteine-containing peptides. Analytical Chemistry 82, 33523358.Google Scholar
Xu, Z., Yang, X. & Yang, Z. (2008) On the mechanism of surfactant adsorption on solid surfaces: free-energy investigations. The Journal of Physical Chemistry B 112, 1380213811.Google Scholar
Xu, Z. P., Zeng, Q. H., Lu, G. Q. & Yu, A. B. (2006) Inorganic nanoparticles as carriers for efficient cellular delivery. Chemical Engineering Science 61, 10271040.Google Scholar
Yancey, J. A., Vellore, N. A., Collier, G., Stuart, S. J. & Latour, R. A. (2010) Development of molecular simulation methods to accurately represent protein-surface interactions: the effect of pressure and its determination for a system with constrained atoms. Biointerphases 5, 8595.Google Scholar
Yang, J. A., Johnson, B. J., Wu, S., Woods, W. S., George, J. M. & Murphy, C. J. (2013) Study of wild-type α-synuclein binding and orientation on gold nanoparticles. Langmuir 29, 46034615.Google Scholar
Yang, Z. & Zhao, Y.-P. (2006) QM/MM and classical molecular dynamics simulation of histidine-tagged peptide immobilization on nickel surface. Materials Science and Engineering: A 423, 8491.Google Scholar
Yang, Z. & Zhao, Y.-P. (2007) Adsorption of His-tagged peptide to Ni, Cu and Au (100) surfaces: molecular dynamics simulation. Engineering Analysis with Boundary Elements 31, 402409.Google Scholar
Yu, J., Becker, M. L. & Carri, G. A. (2012a) The influence of amino acid sequence and functionality on the binding process of peptides onto gold surfaces. Langmuir 28, 14081417.Google Scholar
Yu, J., Wei, W., Danner, E., Ashley, R. K., Israelachvili, J. N. & Waite, J. H. (2011) Mussel protein adhesion depends on interprotein thiol-mediated redox modulation. Nature Chemical Biology 7, 588590.Google Scholar
Yu, X., Wang, Q., Lin, Y., Zhao, J., Zhao, C. & Zheng, J. (2012b) Structure, orientation, and surface interaction of Alzheimer amyloid-β peptides on the graphite. Langmuir 28, 65956605.Google Scholar
Zahn, D. & Hochrein, O. (2003) Computational study of interfaces between hydroxyapatite and water. Physical Chemistry Chemical Physics 5, 40044007.Google Scholar
Zanetti-Polzi, L., Daidone, I., Bortolotti, C. A. & Corni, S. (2014) Surface packing determines the redox potential shift of cytochrome c adsorbed on gold. Journal of the American Chemical Society 136, 1292912937.Google Scholar
Zhang, L. & Sun, Y. (2010) Molecular simulation of adsorption and its implications to protein chromatography: a review. Biochemical Engineering Journal 48, 408415.Google Scholar
Zhang, L., Zhao, G. & Sun, Y. (2009a) Molecular insight into protein conformational transition in hydrophobic charge induction chromatography: a molecular dynamics simulation. The Journal of Physical Chemistry B 113, 68736880.Google Scholar
Zhang, Z., Chen, S., Feng, Y., Ding, Y., Zhou, J. & Jia, H. (2009b) Electrochemical and molecular simulation studies on the corrosion inhibition of L-glutamine monolayers on an iron surface. Journal of the Serbian Chemical Society 74, 407415.Google Scholar
Zhdanov, V. P. & Kasemo, B. (2000a) Monte Carlo simulation of diffusion of adsorbed proteins. Proteins: Structure, Function, and Genetics 39, 7681.Google Scholar
Zhdanov, V. P. & Kasemo, B. (2000b) Ordering of adsorbed proteins. Proteins: Structure, Function, and Genetics 40, 539542.Google Scholar
Zhou, J., Zheng, J. & Jiang, S. (2004) Molecular simulation studies of the orientation and conformation of cytochrome c adsorbed on self-assembled monolayers. The Journal of Physical Chemistry B 108, 1741817424.Google Scholar
Figure 0

Fig. 1. Simulations of proteins with different types of surface: (a) lysozyme on a polyethylene surface (reprinted with permission from (Wei et al. 2011). Copyright (2011) American Chemical Society), (b) the MRKDV peptide on a bare silver surface (adapted with permission from (Aliaga et al. 2011). Copyright (2011) American Chemical Society), (c) RAD16II on a rutile surface (reprinted with permission from (Monti, 2007). Copyright (2007) American Chemical Society), and (d) NiFe hydrogenase on a SAM surface (reprinted with permission from (Utesch et al. 2013). Copyright (2013) American Chemical Society).

Figure 1

Fig. 2. Typical time and length scales of different simulation techniques: quantum mechanics (QM), including coupled cluster (CC) and DFT methods (inset adapted with permission from (Iori et al. 2008). Copyright (2008) American Chemical Society); molecular mechanics (MM) including all-atom molecular dynamics (AA-MD) simulations, implicit solvent and coarse grained MD (IS-MD and CG-MD), and the Brownian dynamics (BD) technique; and continuum mechanics (CM). The ranges of time and length scales are approximate.

Figure 2

Fig. 3. Isosurface plots for density functional theory (DFT) single electron states at the imidazole/Au(111) interface. (a) Bonding orbital with σ-like shape; (b) antibonding orbital with σ-like shape. The atomic p-like character of the orbital on the imidazole N is visible in both panels as density within the ring (red circle). Color scheme: the orbital density isosurface is represented in magenta; Au: orange, N: grey; C: yellow; H: cyan. Adapted with permission from (Iori et al. 2008). Copyright (2008) American Chemical Society.

Figure 3

Fig. 4. Examples of systems studied in large scale static density functional theory (DFT) calculations and ab initio molecular dynamics (AIMD) simulations. (a) Static DFT: dodecapeptides adsorbed on a hydroxyapatite (0001) surface, after DFT geometry optimization, color scheme: O peptide: red, O water: cyan, N: blue, C: dark yellow, H: light gray, Ca: green; P: yellow. Adapted with permission from (Rimola et al. 2012). Copyright (2012) American Chemical Society; (b) AIMD: Side view of a polyserine β-sheet on an Au(111) slab in liquid water. Color scheme: O: red, N: blue, C: gray, H: white, Au: yellow. Adapted with permission from (Calzolari et al. 2010). Copyright (2010) American Chemical Society. The dashed lines indicate the periodically repeated cell.

Figure 4

Fig. 5. Comparison between vdW-DF and experimental adsorption energies to an Au(111) surface for a set of molecular adsorbates. Reprinted with permission from (Wright et al. 2013). Copyright (2013) American Chemical Society.

Figure 5

Fig. 6. Schematic illustration of protein–surface interactions in aqueous solvent. The main interaction interfaces can be categorized as: protein–surface, protein–solvent, solvent–surface and protein–solvent–surface. The protein-surface interface (depicted in the left circle) includes direct interactions. The interactions can be non-specific such as van der Waals and electrostatic interactions (represented with dashed lines in the figure), or specific such as strong histidine–gold interactions (shown with a continuous line) and even stronger chemisorption interactions. At the protein-solvent interface (depicted in the top circle), the structural and physical properties of the protein and the solvent deviate from those inside the protein and in the bulk solvent, respectively. In particular, water forms layers around the polar and charged residues as depicted by the two spheres in the figure. At the interface, the relative dielectric permittivity of water and of the protein is lower than that of their bulk counterparts. At the solvent-surface interface (depicted in the right circle), the solvent may form structured layers or be completely disordered. On a gold surface, for instance, water forms two ordered layers that are separated by high energy barriers and have a lowered relative dielectric permittivity in the direction normal to the surface. At the protein-solvent-surface interface (depicted in the bottom circle), the interactions involve a complex interplay between the constituents. The protein may make strong indirect interactions with the surface through a stable network of hydrogen bonds (represented by dashed lines) in the adsorption region.

Figure 6

Fig. 7. Modeling of protein-Au(111) interactions using a continuum solvent model parameterized by comparison with explicit solvent MD simulations. Solid line: The potential of mean force (PMF) for a test atom as a function of atom-gold surface distance, as obtained from MD simulations in explicit water solvent; squares: the corresponding LJ potential; dashed line: their difference, associated with the desolvation energy; dotted line: PMF computed using the protein-metal continuum solvent ProMetCS model (which includes both LJ and metal hydrophobic desolvation energies, see text). Reprinted with permission from (Kokh et al. 2010). Copyright (2010) American Chemical Society.

Figure 7

Fig. 8. Models of polarizable gold surfaces. (a) The method of image charges. Charges qi and qj induce polarization charges inside the metal shown as −qi and −qj respectively. (b) Classical polarization models. Each gold atom of the surface is assigned a dipole with either variable (Drude oscillator) (left) or fixed (rigid rod that is free to rotate) (right) moment. Reproduced with permission from (Iori & Corni, 2008). Copyright (2008) John Wiley and Sons.

Figure 8

Fig. 9. Lateral density profiles of the first layer of structured water on the (101-0), (0001) and (011-1) surfaces of silica. Reprinted with permission from (Notman & Walsh, 2009). Copyright (2009) American Chemical Society. Positions of the water molecules are dominated by the positions of the silanol hydroxyl groups and to a lesser extent by the underlying crystal structure.

Figure 9

Fig. 10. Adsorption of three different peptides (positively charged, neutral and negatively charged) on silica nanoparticles as a function of pH. (a) Amounts of adsorbed peptides measured in experiments (Puddu & Perry, 2012). (b) Times spent by the peptides on the surface calculated from the MD simulations. Reprinted with permission from (Emami et al. 2014b). Copyright (2014) American Chemical Society.

Figure 10

Fig. 11. (a) Distribution of orientations of lysozyme on a negatively charged surface obtained by a single parallel tempering MC (labeled as p0) simulation and by four separate conventional serial MC (labeled as s1-s4) simulations. Orientations are represented by the cosine of angle θ, the angle between the unit vector along the dipole moment of lysozyme and the unit vector normal to the surface. (b) Energy landscape of interaction of lysozyme with the surface. Energy minima corresponding to the most visited orientations in the MC simulations are indicated with arrows. Reprinted with permission from (Xie et al. 2010). Copyright (2010), AIP Publishing LLC.