DOI:
10.1039/C005236A
(Review Article)
Mol. BioSyst., 2011,
7, 101-115
Towards proteomics-on-chip: The role of the surface
Received
28th April 2010
, Accepted 29th June 2010
First published on 12th August 2010
Abstract
Miniaturisation is revolutionary to high-throughput proteomics. These technologies have gained much interest in the past decade, as they allow for sensitive parallel analysis of small amounts of biological materials. This review describes the state of the art of proteomics-on-chip, with a particular focus on the fundamental proteomics-on-chip challenges. The important role of bio-interfacial interactions and strategies to control them are presented. Various coating methodologies for on-chip protein/peptide separation are reviewed to provide an overview of the principles of protein-resistant and protein immobilisation coatings, and their effectiveness.
 Malinda Salim | Malinda Salim received her BEng degree in Chemical Engineering from the University of Sheffield in 2004. She pursued her PhD studies in developing surface modification techniques for proteomics-on-chip application with joined collaboration between Chemical Engineering Department and Engineering Materials, under the supervision of Professor Phillip Wright and Dr Sally McArthur. Upon completing her PhD, she took up a postdoctoral research associate position with Professor Phillip Wright and Dr Seetharaman (Raman) Vaidyanathan at the University of Sheffield, applying her multi-disciplinary research background to metabolomics. She is particularly interested in the application of mass spectrometries to investigate metabolites interaction. |
 Sally L. McArthur | Sally McArthur is currently Associate Professor in Bioengineering in the Faculty of Engineering and Industrial Science at Swinburne University of Technology in Australia. Prior to that she in the Department of Engineering Materials at the University of Sheffield as Senior Lecturer in Biomedical Engineering after joining the department as a Lecturer in May 2002. Prior to moving to the UK, she was a Senior Research Fellow in the Bioengineering Department at the University of Washington, working with Professor David Castner and NESAC/Bio. |
 S. Vaidyanathan | Raman Vaidyanathan joined the Department Chemical and Process Engineering at The University of Sheffield in 2007, as a lecturer. He's a chemical graduate with a Masters degree in Biotechnology, both from India. He completed his PhD in 2001 from the University of Strathclyde in Glasgow, where he investigated the application of near infrared spectroscopy to monitor industrially relevant bioprocesses involving filamentous microorganisms. This was work done in collaboration with Eli Lilly. He has over five years of pre-doctoral industrial experience applying biotechnological solutions to environmental engineering problems. |
 Phillip C. Wright | Phillip Wright is the foundation Professor of Systems Biology and Engineering within the Department of Chemical and Process Engineering (CPE), at The University of Sheffield. He is also the Head of Department, ad Director of the ChELSI Institute at Sheffield. He started his Chemical Engineering career as a cadet at BHP Steel in Wollongong NSW Australia, and completed his BE in Chemical Engineering by a combination of both full and part time study, graduating from the University of NSW in 1991. He returned to BHP Steel as a Process Engineer for the years 1991–1993. |
1. Introduction to miniaturisation and proteomics
1.1 Miniaturisation
The first study of miniaturisation was performed in 1975 by Terry et al. on a microfabricated silicon gas chromatographic analyser.1 However, research and development of lab-on-a-chip applications boomed in the 1990's following the report by Manz et al. who proposed miniaturised devices for chemical analysis.2 Miniaturisation has demonstrated advantages over macroscale systems due to lower sample consumption, rapid high throughput analysis, and possibilities for integration and automation (including scale-out). Reduced sample consumption is desirable to save precious samples and chemicals. Miniaturisation can also shorten the time to obtain the result, since this architecture has the possibility to run multiple samples simultaneously via parallelisation, dramatically increasing throughput.3–5 Although the nature of the reaction does not usually change in miniaturised devices, the diffusion time (t) becomes significantly shorter,3,6 since t is proportional to the square of the characteristic length, d (i.e. t α d2). An immunoassay performed in the usual 96-well plate requires hours for an antibody-antigen reaction to reach optimal levels, and only minutes or seconds in a miniaturised immunoassay.5,7,8 Another change upon scaling down is the higher surface area-to-volume ratio, which accounts for the important role of surface characteristics, hence the control of biomolecules and electroosmotic flow (EOF) in miniaturised systems.4,5,9
The advantages of miniaturisation, together with an ability to integrate several experimental steps into a single micro-scale device, make microfluidics a promising tool for biological analyses.4,5,10 Recently, microfluidic devices were demonstrated as being able to mimic biological systems within the body, enabling in vivo studies to be undertaken in an in vitro environment. Key applications have included the imitation of blood vessels,11 kidney,12 and lung.13Cell culture has also been performed in microchannels to better understand effects of shear and diffusion on cell behavior in vivo.14
1.2 Proteomics
Proteomics (proteome–protein complement expressed by the genome)15 is important for pharmaceutical, biomedical, environmental and biotechnology applications.16,17 Core proteomic analytical processes usually include protein extraction from cells, protein separation, protein digestion (with a protease: usually with trypsin into peptides in what is known as “bottom up” proteomics), and analysis using mass spectrometry (MS).16,18–20 As these conventional processes can be slow, labour intensive, and introduce sample loss and degradation in each processing stage, miniaturisation and integration of these processes onto a microfluidic chip, usually coupled to a mass spectrometer are being widely investigated to improve throughput and detection.16,18–20 Two-dimensional sodium dodecyl sulfate poly(acrylamide) gel electrophoresis (2D SDS-PAGE) is a common technology used in proteomics for orthogonal protein separation based on isoelectric points and molecular weights. It usually requires several hours (often overnight) for completion,21 which reduces to minutes in miniaturised 2D SDS-PAGE apparatus.22,23
To develop a realistic proteomics-on-chip device, a diverse set of research fields need to be brought together. These can be divided into categories such as micromachining and microfluidic processes; protein–surface interactions; surface modification; proteomics workflows (protein extraction from cells, protein digestion, protein and peptide separations); coupling between unit operations; and detection strategies. Each category has its own set of intertwined challenges. For example, effective adaptations of commonly used cell lysis techniques (e.g. mechanically-based processes such as grinding in liquid nitrogen with a mortar and pestle), are difficult to achieve on chip.24 Others, such as the efficacy of proteolytic digestion on-chip, depend heavily on the success of enzyme-surface immobilisation (often reported to provide higher efficiency compared to conventional ‘in-solution’ digestion),25 preserving enzymatic activity, sample solubility, and product recovery. Packing of solid materials including beads and gels in microchannels has also been shown to be challenging.26 All of the issues require dedicated research.
Downstream, on chip devices must also ensure compatibility with (mostly electrospray ionisation (ESI) based) MS post-analysis. Readers interested in applications and methods to interface microfluidic chip-MS (ESI and MALDI) are directed to recent reviews by Jeonghoon et al.27,28 General concerns include the influence of current from the separation chamber on the electrospray current, and chip-interface design.29,30 High EOF, the bulk flow of fluid through the channel brought about by an electric field, is usually also required (e.g. using a sheathless interface).31 High EOF is, however, undesirable for many forms of protein separation.30,32 Additionally, band broadening resulting from Joule heating can detrimentally affect the separation resolution.33 Biomolecule adsorption to channel walls has also been shown to significantly affect the EOF,34,35 but many of the typically polyethylene oxide (PEO)-based low fouling coatings exhibit low EOF.35,36
Surfaces, biomolecules, and fluid movement are therefore ubiquitous and inter-dependent within any on-chip device. Controlling the extent of EOF while simultaneously controlling biointerfacial interactions through surface modification is thus critical to the development of effective on-chip proteomic devices.19,37 Different surface properties can provide different controls for EOF, prevent non-specific adsorption of biomolecules, or enable the immobilisation of biomolecules for specific capture, digestion or detection methods. Given the multidisciplinary nature of this field, there are many other issues related to on-chip proteomics that are beyond the scope of this review. A number of articles however, are recommended to readers who require more information.19,31,38–42
1.3 Scope
This review seeks to integrate fundamental proteomics-on-chip challenges to provide a readily accessible overview. The first part (section 2) of this article presents a primer on biomolecule-surface interactions, focusing on the mechanisms and impact of uncontrolled adsorption in microfluidic devices. This provides the knowledge required to understand measures that can be taken to control biointerfacial interactions within microfluidic devices for biomolecule separation using a range of surface modification strategies in section 3. Perspectives are given on the current technologies and future prospects. Finally, in section 4, a critical evaluation of surfaces that resist non-specific bioadhesion, in particular PEO-based surfaces, is included to provide an overview of the principles of protein-resistant coatings and the coating effectiveness of various common strategies.
2. Understanding protein adsorption at solid–liquid interface
2.1 Interaction between proteins and surfaces
Proteins are complex heterogeneous biological molecules that are composed of chains of amino acids (NH2–CHR–COOH) which can be hydrophobic or hydrophilic, neutral, negatively or positively charged.43 As a consequence, proteins are able to interact with each other as well as synthetic surfaces in a number of ways as shown in Fig. 1. These include hydrophobic interactions, repulsive and attractive ionic or electrostatic interactions, and short range interactions such as hydrogen bonding and van der Waals interactions.43,44 A delicate balance between a number of forces plays an important role in the stabilisation of a protein, which may be perturbed upon adsorption to or interactions with a solid surface. Since protein adsorption normally occurs at constant temperature and pressure, the thermodynamic driving forces governing them can be described by considering the Gibbs free energy for the system:
where G, H, S and T are Gibbs free energy, enthalpy, entropy and absolute temperature respectively. Generally, if the Gibbs free energy of the system is negative (ΔGads < 0), proteins will adsorb onto a surface. For a negative ΔGads, there should either be a decrease in enthalpy or an increase in entropy.43,45 Norde reported that the entropy gained from the protein unfolding upon adsorption onto surfaces may be the major driving force for protein adsorption.43,45
![Schematic representation of protein interaction (having hydrophobic, neutral hydrophilic and charged sides) with a number of synthetic surfaces [Adapted from ref. 46].](/image/article/2011/MB/c005236a/c005236a-f1.gif) |
| Fig. 1 Schematic representation of protein interaction (having hydrophobic, neutral hydrophilic and charged sides) with a number of synthetic surfaces [Adapted from ref. 46]. | |
2.2.1 Hydrophobic interactions.
Hydrophobic interaction is an additional driving force for protein adsorption at surfaces. This interaction may be short range (ca. 15–20 nm), although the magnitude of hydrophobic interaction (long or short range) is still under debate. Hydrophobic interaction minimises the free energy of the system by reducing the interfacial area between non-polar regions and aqueous solution.47 This gives rise to protein adsorption when the protein is brought into contact with a hydrophobic surface. The hydrophobic moieties residing in the protein interior may therefore be exposed to the hydrophobic surface, which cause protein unfolding. Therefore, protein adsorption onto a hydrophobic surface generally results in changes of the protein structure or denaturation (Fig. 2). The extent of denaturation depends on the degree of hydrophobicity of the protein and the surface.43,44,48,49
 |
| Fig. 2 Schematic representation of protein unfolding on a hydrophobic surface. | |
Several groups have reported the effects of surface hydrophobicity on protein denaturation. For example, Wertz et al. studied the spreading of bovine serum albumin (BSA) and human fibrinogen on hydrophobic and hydrophilic self-assembled monolayers (SAMs) using total internal reflection fluorescence.50 For both proteins, fast and extensive spreading was observed on the hydrophobic surface (greater increase of protein footprint as a function of time), whereas very little spreading was observed on the hydrophilic surface. A protein footprint is the surface area a protein occupies when adsorbed onto a surface. The ultimate surface coverage for both proteins was found to be smaller on hydrophobic compared to hydrophilic surfaces at high protein concentrations. Similarly, as reported by Veen et al., the adsorbed amount of lysozyme and α-lactalbumin at their saturation level is smaller on hydrophobic surfaces compared to hydrophilic surfaces at high protein concentrations.51 At low protein concentrations, however, the quantity of adsorbed proteins was greater on a hydrophobic surface. Although faster spreading of protein occurred on the hydrophobic surface, the proteins spread to a far lesser extent at low protein concentration. These studies contradict several general findings that greater protein surface coverage was obtained on hydrophobic surfaces, where protein spreading is usually not accounted for.52–55 The aforementioned studies also showed the dependency of protein spreading on protein supply rate; and that a balance exists between the rate of protein spreading and the rate of protein adsorption.50,51,56
2.2.2 Ionic interactions.
Ionic/electrostatic interactions arise from the attractive and repulsive interactions between the charge groups on the proteins, and those present on the surface.44 The net charges of the protein and surface are affected by the pH and the ionic strength of the solution, defined by the Debye length (between 1–1000 nm).57 These short and long range electrostatic interactions also play an important role in protein stability and function.58 A charged protein is expected to be attracted or adsorbed onto an oppositely charged surface, and be repelled when both the protein and the surface are similarly charged, if electrostatic interactions dominate. This, however, is not always the case, as adsorption of proteins occurs even on similarly charged surfaces. For example, net negatively charged proteins have been shown to adsorb onto negatively charged surfaces.43,45,48,59 Although a protein carries a net charge, it cannot be considered as a single point charge, as different amino acid residues exist on different parts of the protein surface,60 creating a charge distribution across the surface.
2.2.3 Other interactions.
The role of van der Waals forces in protein–surface interactions is of importance when the protein is in close contact with the surface, as van der Waals interactions are short ranged (typically below 2.5 nm) and always attractive.61 This interaction can be formed between polar or non-polar atoms or molecules. The balance between van der Waals and electrostatic interactions for colloidal or protein systems and stability has been interpreted in the Derjaguin-Landau-Verwey-Overbeek (DLVO) theory.62
Hydrogen bonding, which also has a short range character (ca. 0.3 nm), is another type of interaction between a protein and a solid surface.63 It is formed when a proton is shared between two electronegative atoms. Therefore, H-bonding may form between amino acids (intramolecular hydrogen bonding, which also contributes to protein stability) or between a protein and surrounding water molecules,63,64 which can stabilise native protein conformation and preserve its activity.
Understanding of the principles and factors that influence protein adsorption is important for several applications including biomaterials, pharmaceuticals, food technology, biotechnologies, filtration and biosensors.44Protein adsorption at surfaces is a complex phenomenon, and is dependent on factors such as protein properties, protein orientation on the surface, the surrounding environment, and surface properties.43–45,48,65,66 Examples of the properties of the proteins that affect adsorption are the charge, size, and structural stability. The extent of the interaction between a protein and the surface will generally be greater if the protein is large, carries zero net charge, and/or has low structural stability. A large protein, for example, human fibrinogen (340 kDa, 5 nm × 5 nm × 47 nm),54 has the ability to form multiple contact points with the surface, while smaller proteins such as chicken-egg lysozyme (14.5 kDa, 4.5 nm × 3 nm × 3 nm)67 covers a smaller area, and thus can have fewer contact points per molecule.44,45,48,49,62,68 When the protein has zero net charge (i.e. near its isoelectric point, pI), the reduced electrostatic repulsion between the individual proteins will allow more molecules to be adsorbed onto a surface. Finally, a protein that has low structural stability can readily undergo denaturation or ‘spreading’. This allows more sites on the protein to be in contact with the surface, creating greater surface interaction (see Fig. 2).43,44,48,67,68
The properties of surfaces such as surface topography, composition, charge, and heterogeneity will also determine the protein–surface interaction. A surface having rough topography will have a greater surface area available for the attachment of the protein. On surfaces with heterogeneous chemical compositions, different types of protein interaction may occur, depending on the density and distribution of functional and non-functional groups of the protein and the surface. In short, increasing the surface complexities may increase the types and numbers of interactions between the protein and the surface. Other than all these properties, the surrounding environment factors such as protein concentrations, time, pH of the media, flow conditions, ionic strengths of the buffer and temperature will also affect the protein adsorption behaviour.43,44,49,69
This section describes some of the techniques that have been developed to investigate protein adsorption at surfaces: both planar and microchannels. These can generally be divided into two classes, i.e. surface analytical and (bio) chemical (see Table 1). Although a wide range of surface characterisation methods exists, measurement of protein adsorption in closed-system microchannels is observed to be often limited to (bio)chemical analysis techniques such as radiochemical (most commonly isotopic labelling), immunological, fluorescence, and electrophoretic due to the difficulty in instrumental miniaturisation and assessing the inner surface of the microchannel. Monitoring of protein adsorption has however been shown possible in microfluidic chips integrated to surface plasmon resonance and surface plasmon resonance imaging.70,71 The information obtained and limitations of each of these techniques are presented in Table 1, where in all cases, these approaches provide minimal data on the structure of adsorbed protein layers.
Method |
Information |
Limitations |
Surface analytical |
Atomic force microscopy
|
Interaction forces, structure, kinetics |
Protein distortion due to sample-tip interaction, difficult sample preparation, tip contamination, not quantitative. |
Circular dichroism |
Structure |
Signal noise problems from interfering compounds. |
Contact angle measurement |
Wetting, hydrophilicity/hydrophobicity |
Not quantitative, limited adsorption information. |
Ellipsometry
|
Quantitative, structure, kinetics |
Surface refractive index difficult to determine. |
Reflectometry
|
Quantitative, structure, kinetics |
Surface refractive index difficult to determine. |
Secondary ion mass spectrometry with time-of-flight |
Structure, chemical |
Dehydrated state, vacuum, complex data, difficulty in identifying individual proteins, largest protein fragments are the individual amino acid immonium ions. |
Surface plasmon resonance
|
Quantitative, kinetics |
Analytes must have sufficient mass (>2000 Daltons). |
Surface-matrix assisted laser desorption/ionisation mass spectrometry |
Semi quantitative, molecular mass of adsorbed protein |
No information on the structure of adsorbed protein layer, dehydrated state, vacuum, incorporation of protein into matrix difficult. |
Total internal reflection fluorescence |
Quantitative, structure, kinetics |
Requires fluorescent labeling. |
X-ray photoelectron spectroscopy
|
Quantitative |
Little detail on molecular information, identify different proteins difficult, vacuum, dehydrated state. May cause denaturation and alter protein film thickness. |
Biochemical |
Enzyme linked immunosorbent assay |
Kinetics |
Not quantitative. Protein orientation/conformation and antibody dependent. Limited types of proteins depending on the antibodies available. |
Fluorescence spectroscopy
|
Quantitative, kinetics |
No information on the structure of adsorbed protein layer, labeling affects protein. |
SDS-PAGE
|
Quantitative |
Low sensitivity (dependent on intensity of protein bands/spots), poor reproducibility. |
Electrophoresis
|
Semi quantitative, kinetics |
No information on the structure of adsorbed protein layer, not quantitative. |
Radiochemical |
Quantitative, kinetics |
No information on the structure of adsorbed protein layer, environmental issue on disposal, labeling affects protein. |
3.
Protein adsorption and surface modification strategies in microchannel
There are thousands of peer reviewed articles relating to protein adsorption onto non-microchannel surfaces, which incorporate many adsorption variations such as different types of proteins, surfaces chemistries and structures, protein adsorption properties, and characterisation methods.84–90 Since the realisation of the importance of miniaturisation in biotechnology, reports on protein adsorption behaviour in microchannel systems have also emerged.8,83,91–95 Many publications however focus on system design and applications such as varying surface treatments for controlling microchannel properties, which will be discussed further in section 3.2.
Rossier et al. studied the adsorption isotherm of a protein and its antibody, as well as the immunosorption kinetics onto polymeric microchannels using a radiolabelling method.8 They observed fast diffusion of molecules onto the channel wall, with complete immunoreaction taking place in 1 min. However, the radiolabelling process is not desirable for studying protein adsorption as it is expensive and hazardous.96 In 2003, Lenghaus et al. reported the first attempt to elucidate enzyme adsorption onto microcapillaries of different materials under flow conditions using an enzyme assay.92 The adsorption of the enzyme onto the microcapillary surface was calculated based on the fractional enzyme loss. However, difficulties in resolving the resultant small changes in enzyme loss, especially at high initial enzyme concentrations, arose in this study. Determining the amount of adsorbed protein in a microchannel through a solution depletion method is therefore not a sensitive method of characterisation. Shortly after, Nicolau et al. modelled the adsorption of proteins onto microfluidic devices to estimate the thickness of the adsorbed protein layer at different protein concentrations, buffer ionic strengths and wall surface tensions.95 Due to the complexities of the proteins, these studies were modelled based on a generic model protein derived from an average of 26 proteins in the online database.97 Thus, these observations may not hold true for ‘non-generic’ proteins and surfaces with different characteristics. A laser induced fluorescence technique was used to probe protein adsorption onto microchannel surfaces by Lionello et al.93,94 Immunoglobulin G was adsorbed onto a polymeric microchannel under both static and dynamic conditions. The group demonstrated adsorption kinetics coupled with diffusion phenomena in microchannels, and ways of loading the protein to obtain the best coverage for immunoassay purposes. Due to bulk solution depletion upon adsorption, multiple sample loads in ‘static’ conditions were found to be necessary to obtain an optimum surface coverage. Studies of fibrinogen adsorption on glass micro-capillaries have also been performed using an enzyme-linked immunoassay (ELISA) method, where the effects of protein adsorption in static versus dynamic adsorption conditions were investigated.59 At any given time, the rate of adsorption in static conditions was always observed to be lower than the dynamic system due to bulk solution depletion. Adsorption kinetics of lysozyme on a hydrophobic cyclic olefin copolymer (COC) channel was recently studied by Alvarez, where the main focus was to demonstrate possibilities of real-time monitoring of protein adsorption using pulsed streaming potentials.83 Although a non-labeling approach was used, a more complex design of the chip setup is required. Despite varying surface activities due to different experimental conditions from all these studies, it can be concluded that the nature of protein–surface interaction is not affected by miniaturisation; and that protein adsorbs rapidly onto the microchannel due to fast diffusion: the protein reaches the surface at a higher rate than its adsorption.
3.2 Surface modification in microchannels
The desire to control surface properties in many applications has led to intensive research on surface modification techniques. Surfaces with varying physical and chemical characteristics can subsequently be obtained, which can directly influence the biological performance at the interfaces.98,99 In chip-CE for instance, higher protein separation efficiencies and migration time reproducibility were reported on coated-microchannels compared to non-coated microchannels due to the stable EOF and lower non-specific protein adsorption;100 whereas higher peak capacities and separation resolution (narrow protein bands) were observed in chip-IEF.30 Oriented antibodies on carefully controlled surfaces were also shown to improve the assay detection signal due to higher specific antigen capture efficiency.101,102 The regulation of EOF and control of protein adsorption is therefore integral to enable rapid and efficient protein separation.60,103,104
There are a number of surface modification strategies commonly used in biomedical and biotechnology applications, including wet chemical treatments such as self assembled monolayers, the Langmuir–Blodgett-Kuhn technique, polyelectrolyte multilayer deposition; and plasma modification.105–109 In this review, we will primarily focus on surface modification techniques suitable for microchannels and specifically those used in the electrophoretic separation of biomolecules.
Surface modification of microchannels can be challenging, due to the range of materials utilised to create the devices. Common construction materials in microchip fabrication include glass, fused silica, quartz, silicon and polymeric materials such as poly(dimethylsiloxane) (PDMS), poly(carbonate), poly(methylmethacrylate) (PMMA), poly(ethylene terephthalate), poly(styrene), poly(propylene) and COC.110–112 Therefore, a simple and versatile surface modification technique capable of coating a wide range of materials with excellent coating stability, desired control of EOF as well as biomolecules adsorption is of huge interest. This situation is further complicated by the requirement that coatings for electrophoresis and lab-on-chip applications are stable and retain their specific properties throughout their use.113 Surface modification in microchannels can generally be divided into two major coating categories, covalent or dynamic.112–114
3.2.1 Dynamic coating.
Dynamic coating methods are relatively simple, and require the addition of surface modifiers such as polymers or surfactants either to the running buffer (termed dynamic coating), or by rinsing the surface with modifier solutions prior to analyses (‘static’ physically adsorbed coating). These modifiers are physically adsorbed to the microchannel surfaces, usually by electrostatic interactions or H-bonding and van der Waals forces; and are most commonly used to alter EOF properties within microchannels.35,111,113–117
Different types of polymeric coatings (neutral, anionic and cationic) can be physically adsorbed on microchannel walls. Neutral polymers such as polyethylene oxide (PEO), poly(saccharides), cellulose derivatives, poly(vinyl alcohol) (PVA), poly(acrylamide), poly(vinyl pyrrolidone) (PVP) have been widely used to suppress EOF and reduce the electrostatic interactions between analytes and microchannel walls.113–118 Hydrophilic polymers such as PVA, poly(acrylamide), PEO, celluloses, and poly(saccharides) do not adsorb strongly on glass/fused silica surfaces, and the coatings can be easily removed.116,117 The stabilities of physically adsorbed polymers are often reported to be pH dependent. For example, a physically adsorbed PEO coating has been reported to degrade at pH > 8, possibly due to competitive H-bonding among silanol group, hydroxyl and PEO.32 Suppression of EOF in PVA adsorbed surfaces was also reported to be not effective at pH > 4–8 due to its poor coating stability.119
Adsorption of charged polymers on surfaces changes the surface zeta potential and thus, the EOF. These polymers can therefore be used to control the magnitude and direction of the EOF. The most commonly used charged polymers are of cationic (positively charged) nature, enabling the separation of basic analytes and reversing the direction of EOF on hydroxyl-group rich bare glass or fused silica.116 Examples of cationic polymers include poly(brene), poly(arginine), poly(ethyleneimine) (PEI), poly(l-lysine) (PLL), and poly(allylamine hydrochloride) (PAH).120–122 Anionic (negatively charged) polymers can be introduced on negatively charged surfaces via the assembly of polyelectrolyte multilayers, also known as layer-by-layer (LbL) assembly.113,116,123LbL requires the deposition of alternately charged layers of polymer. This approach has been widely applied to coat glass, silica and oxidised PDMS microchannels since its introduction in 1991 by Decher et al.124LbL of polyelectrolytes with an anionic layer as the final layer has been shown to reduce protein adsorption and provides a more stable EOF compared to bare glass or fused silica.111 On uncoated surfaces, stable and reproducible EOF is difficult to achieve without excessive rinsing and etching pre-conditioning steps with strong NaOH and/or HCl solutions, which remove hydrocarbon contamination and regenerate surface OH groups.113 For example, polyelectrolyte multilayers with a final dextran sulfate layer has been used to separate proteins at (acidic) pH 2, which is not possible with uncoated glass (pI 2–3).125
Other than the use of polymers as surface modifiers, the addition of small molecules such as surfactants, salts and various types of amine-containing molecules has also been investigated to reduce protein adsorption and changes the EOF behaviour of microchannels.126–130Biopolymer coatings such as the ELISA blocking proteins, lysozyme, and fibrinogen have been utilised by several groups.131–133 However, controlling the behaviour of these biopolymer coatings are always difficult, and specific and non-specific interactions between the coatings and the biomolecular species being separated may occur. This then leads to desorption, denaturation or exchange between the coating and the sample.
A wide range of surfactants are also used as surface modifiers within on-chip devices.116,134,135 Anionic and cationic surfactants are often used to increase and decrease or even reverse the EOF properties within a channel. These surfactants however, tend to be strongly protein denaturing, and may not be suitable for analysis where maintaining the native conformation of proteins may be required.116 The use of zwitterionic surfactants such as 3-[(3-cholamidopropyl)dimethylammonio]-1-propanesulfonate (CHAPS) has been shown to be able to neutralise the surface charge, therefore resulting in EOF suppression.136 Despite the ability of nonionic surfactants to reduce EOF,137,138 they have been generally shown to be much less effective in reducing the adsorption of proteins on surfaces compared to zwitterionic surfactants.115 Non-ionic surfactants generally have weaker interactions with proteins compared to ionic and zwitterionic surfactants.139 The concentration of the surfactants used as surface modifiers must also be carefully manipulated to avoid the formation of micelles.
While physical adsorption can be readily applied to microfluidic and on chip devices, it has limitations. It poses environmental risks, due to the solvent chemical wastes.74,111 Surfaces formed by physical adsorption are generally less stable and less reproducible, especially when they are reproduced on different types of substrates.74,111 Different surface characteristics can also be generated from the same polymer due to different coating processes (polymer solutions, pretreatment solutions and coating protocols).32,140Regeneration of the coatings is also often required with physical adsorption, as stabilities are often limited (usually after only a few EOF runs).132 This process is time consuming, and surface pretreatment steps usually require >30 min.32 The presence of surface additives in the running buffer (dynamic coating) or possible desorption of additives from the surfaces (‘static’ coating) may be detrimental towards applications requiring analysis by mass spectrometry and affinity CE due to possible ion suppression. Although being able to suppress EOF in many cases where separation resolution can be increased, these types of coatings do not effectively eliminate protein adsorption.111,116
3.2.2 Covalent coating.
Surface modification using covalent immobilisation is always preferred compared to physical or dynamic adsorption, due to the enhanced surface stability and the resulting higher resolution separations that can be achieved.114 Covalent immobilisation is often based on silane chemistries.141 The silane-coupling agent acts as an anchor between inorganic silanol containing substrates and the organic material coating. It consists of two steps: initial silane modification of the silanol surface groups (most commonly using bifunctional silanes), followed by incorporation of polymers or small molecules onto the silane surface. Silanisation processes can vary according to the different types of silane reagents, solvents, and surface pretreatment procedures, where the modified surfaces may vary accordingly.142 One of the main disadvantages of silanisation coating is the poor stability of the siloxane (Si–O–Si) bonds at high solution pH. Also, silanes are usually grafted on inorganic oxide containing surfaces such as glass and metal oxides; and it is often not applicable to polymeric materials.98,111
Similar to dynamic coating, polymers such as poly(acrylamide), PVP, PVA, PEO, PEI, PLL, proteins, and other amine and acid containing species can be covalently bound to microchannel surfaces.143–147 These can be coated using silane or non-silane based chemistry including radical polymerisation, thermal immobilisation and surface activation followed by solution based deposition.100,119,148–150 Covalently immobilised coatings have been reported to exhibit longer coating stability than physical adsorption. For example, covalent attachment of PEI on a PMMA surface exhibited a stable EOF of >50 runs;122 while >90 runs were observed for the corresponding PAH and PEI grafted glass surfaces (<60–90 runs for physically adsorbed polymers).35
The covalent coating of polymers on surfaces requires reactive functional groups to be present at the surface to initiate immobilisation, and as such only certain materials can be used. The lack of a surface ionisable group in polymeric materials such as poly(ethylene), poly(propylene), PDMS, poly(carbonate) and poly(tetrafluoroethylene) (PTFE) requires surface pre-modification to introduce reactive functional groups.151,152 This can be achieved using UV irradiation, laser activation, or plasma treatment.100,151,153,154 These treatments alone can be used to render hydrophobic surfaces hydrophilic, and can also result in an increase in EOF within channels formed in these polymeric materials due to the introduction of protonatable groups. For example, the EOF of plasma oxidised PDMS was reported to be 4 times higher than untreated PDMS.155 An EOF increase of ca. 4% was observed on a PMMA laser treated surface.154 It should, however, be noted that these surface pre-treatments are often dependent on the initial surface chemistry, subject to often rapid ageing effects including hydrophobic recovery and oxidation, and thus often show limited stability.143,155 In short, many covalent coating methods for microfluidic devices require multi-step immobilisation procedures, which may be difficult to control, and can result in irreproducible coatings.117
3.2.3 Plasma-based coating
Commonly used solvent or liquid phase coatings may also cause problems, such as clogging or poor uniformities in the microchannels, so solventless deposition (plasma and non-plasma) has also been sought as an alternative method to coat microchannels.156 Plasma surface modification can be divided into two classes, namely plasma treatment (chemically or physically altering the surface itself), and plasma polymerisation or deposition. Plasma treatments have been used to oxidise PDMS as well as other polymers, and enable the subsequent grafting of polymers or the bonding of devices together.156,157 Plasma processing has also been shown to create multi-functional microfluidics with controllable surface wetting properties, allowing fluid manipulation, which is useful for directing protein and cell growth.158 As detailed earlier, these treatments are generally effective for short periods of time.126 Plasma polymerisation is a deposition rather than a treatment process that results in the formation of a polymeric thin film from the vapour phase under the influence of a plasma discharge.159 Volatilised monomers are introduced into an evacuated reaction chamber (see Fig. 3) where, under the influence of an electric field, they undergo ionisation, generating electrons, ions, free radicals, photons, and molecules both in ground and excited states. Excitation of the monomer results in reactive species on the surface, which act as reactive sites for the covalent attachment of other species.160,161 It can therefore be viewed as a kind of graft polymerisation, although it does not alter the bulk properties of the substrate by partial penetration of the substrate, as is the case in graft polymerisation obtained by conventional means.159 Plasma processing is a complex process with repetitive fragmentation, rearrangement, crosslinking, ionisation, and polymerisation of the monomers occurring throughout the cycle. The resulting polymer thin film is deposited on the surfaces of any material exposed to the plasma glow. Plasma polymers have been utilised in biosensors,162 surface micropatterning,163 blood contacting devices,164cell culture surfaces,165 biomolecule immobilisation,166 and polymer grafting.160
![Schematic example of a radio frequency (rf) plasma polymer reactor [Reproduced from supplementary material in ref. 35].](/image/article/2011/MB/c005236a/c005236a-f3.gif) |
| Fig. 3 Schematic example of a radio frequency (rf) plasma polymer reactor [Reproduced from supplementary material in ref. 35]. | |
The polymers formed by plasma polymerisation tend to be highly branched and cross-linked with random structures incorporating various structural fragments of the monomer.159,161 However, by controlling the plasma deposition parameters, it is increasingly possible to produce more traditional polymer structures, chemistries and behaviours.167–169 Plasma polymers have a number of advantages over other surface modification techniques: good adhesion and ability to coat virtually any dry substrate; the coatings do not affect the bulk properties of the substrate; a solvent-free and one-step coating process; ultra-thin films (1–10 nm) with controllable thickness; a large number of different monomers have been shown to form films.159,160,170
In 2003, the Karube group investigated the first use of a plasma polymerisation to coat glass open-microchannels for IEF-based protein separation.171,172Hexamethyldisiloxane (HMDSO), glycidol, and acetonitrile were used as the monomers to reduce EOF. It was reported that the plasma polymerised acetonitrile coating was easily removed due to weak surface attachment; and a minimum thickness of 50 nm had to be obtained for HMDSO and glycidol coatings to remain stable at high applied voltages. Barbier et al. introduced plasma polymerised acrylic acid on PDMS microchannels and these exhibited stability over several days.173 Coatings produced on PDMS have been widely known to be susceptible to degradation due to fast hydrophobic recovery.174 General surface modification strategies and their recent development for PDMS microfluidic chips have been reviewed by several groups.175,176 It was suggested that a combination of both wet chemical treatments and solventless techniques provided the greatest promise in terms of surface stability. Incorporating additives such as carboxylic acid and PEG-based polymers to the PDMS pre-polymer prior to curing has also been shown to improve surface wettability and provide better control of EOF as well as non-specific protein adsorption.177,178 A stable and non-fouling coating was then demonstrated by Salim et al. in glass microchannels using a tetraglyme plasma polymer.179 This work has recently been adapted to produce coatings on the more challenging PDMS substrates that are stable and resist protein adsorption for up to 100 days [A. G. Pereira-Medranoet al. Rapid fabrication of Glass/PDMS hybrid μIMER for augmented high throughput membrane proteomics, submitted to Lab Chip 2010].
These early studies suggest plasma polymerisation as a versatile method to coat microchannel surfaces. To date, wet chemistries are still the preferred choice to surface modify microchannels, as solutions can be directly flowed through the ‘closed’ microfluidic chip system. Coating of the inner microchannel surfaces using plasma can be challenging; and an ‘open-system’ is usually devised, where the plasma-coated channel has to be subsequently bonded to another cover to form a seal. Consequently, miniaturised plasma sources/microplasma is being increasingly investigated to coat inner surface of microchannels and small tubings.180–182 Although several groups have shown the successful use of microplasma to produce thin film coatings to control EOF and biomolecule adsorption,183,184 this technology is still in its early development phase. Design improvements and understanding of the microplasma properties are essential to obtain controlled, uniform film properties.
4. Understanding coatings that resist and bind biomolecules
4.1 Surfaces that resist bioadhesion
Various coatings such as PEO, phospholipids, poly(saccharides), poly(acrylamide), and proteins have been presented for their ability to reduce bioadhesion in a range of biomedical and biotechnology applications.185–192 Of these, PEO was shown to offer great promise, and has been widely investigated. The structure of PEO is HO–(CH2–O–CH2)n–OH,193 where n may range from 0–100
000.194PEO-based surfaces can be used to create a non-fouling platform suitable in applications where elimination of protein adsorption is essential. By using a mixture of reactive and non-reactive grafted PEG chains, it is also possible to immobilise specific proteins and other biomolecules to a non-fouling PEG surface. PEO surfaces can be produced by several approaches, including physical adsorption,32,195 SAMs,196,197 covalent immobilisation,198,199 vapour deposition techniques,156 and plasma polymerisation of ethylene oxide monomers.163,188,200–205
In aqueous solutions, the PEO molecule is hydrophilic and highly mobile with a large exclusion volume.193 It has a strong tendency to form hydrogen bonds with water, and this extensive hydration of PEO molecule causes the van der Waals attraction to be small and no electrostatic attraction occurs.206 The highly mobile or rapid motion of the PEO molecule is also reported to give the approaching protein little time to form an interaction.206 The mechanisms for the antifouling properties of PEO are not yet fully understood, and are often related to the chain length and surface density of PEO, which gives rise to steric repulsion effects.207–210 Steric repulsion effects (osmotic and excluded volume contributions) arise from the compression of the PEO chains, where there is a loss in entropy as a protein approaches the surfaces. Adsorption of a protein onto the PEO surface is therefore thermodynamically unfavourable.74,206
Jeon et al. reported a balance between steric repulsion, van der Waals and hydrophobic interactions between protein molecules and PEO surfaces, where the net force which determines the adsorption of protein onto the PEO surface depends on the thickness of the PEO layers as well as the surface coverage.207 Sufficient layer thickness and polymer chain density can screen the protein-substrate interaction, and prevent diffusion of the protein through the steric layer to the underlying substrate.209,210 Szleifer et al. reported that surface density has a greater effect in resisting protein adsorption than chain length.211 Overall, PEO with molecular weights <10
000 Da, and high chain density are generally reported to be effective in resisting protein adsorption.151,210 High density polymers are, however, difficult to obtain with covalent attachment due to the steric exclusion volume of the attached polymer chain, which then hinders the attachment of another polymer chain.
In the radio frequency glow discharge (rfgd) plasma deposition process however, short EO containing molecules (tetraglyme, triglyme) can also produce surfaces with effective non-fouling properties. This may consist of a carpet of short length EO units –(CH2–CH2–O)n– packed in a high density manner with a reduced steric exclusion volume.200,202,204 Monomers of up to two EO units (diglyme) have also been shown to effectively resist protein adsorption.205 Long PEO chain units are therefore not necessary in rendering a surface non-fouling. Plasma polymerised surfaces of EO units are therefore postulated to be similar to the surfaces formed by the SAMs of EO terminated alkanethiols on gold.203,204 It was first demonstrated in 1991 by Prime and Whitesides, that SAMs of short chain PEO can exhibit low bioadhesion.212 The surfaces formed by SAMs, however, are dependent on the substrate materials, with noble metal surfaces (Au, Ag) required to form the alkanethiol bond.213,214
Although still under widespread investigation, the non-fouling properties of EO-based plasma polymerised surfaces are generally being correlated with the high retention of the C–O units within the plasma polymer structure. High retention of the C–O content can be achieved by lowering the plasma deposition power. By adjusting the plasma deposition power, different types of surface chemistries (i.e. extents of C–O incorporation) can be obtained from a single monomer system, producing films with varying protein adhesion properties.204,215 Based on the data from the SAM literature, it is thought that the strong tendencies for the EO units to form hydrogen bonds with water may effectively screen or form a physical barrier to the surfaces from biomolecules.205,216 Studies conducted by Shen et al. using ESCA and ToF-SIMS showed the presence of methyl-terminated short EO chain ends in the plasma deposited tetraglyme surface.203 The plasma polymerised surface formed however, may be more complex and may have disordered molecular surface structures as opposed to ordered SAMs.203 The EO units formed in the plasma polymer films probably do not have high mobilities or the conformational freedom as normally associated with grafted polymers as densely packed, short chain films are formed.202 This densely packed complete surface coverage is thought to be a factor in the ability of the coatings to resist protein adsorption.217 Therefore, conformational flexibility may not contribute to the protein resistivity in this case.218
Plasma polymerised tetraglyme surfaces have been shown to able to resist fibrinogen adsorption up to a level of lower than 10 ng/cm2, and thus resist cell adhesion.179,203 Although plasma polymerisation is capable of producing ultrathin films (1–10 nm), there also appears to be a threshold of film thickness in the plasma polymerisation process whereby an effective non-fouling surface can be obtained. Chu et al. found a minimum film thickness of about 3 nm and 6 nm for resisting BSA and fibrinogen adsorption respectively on a plasma polymerised diglyme surface.201 However, the exact surface chemistries for the different film thicknesses formed remained unclear, and may be due to a number of factors including the coverage of the plasma film and specific characteristics of plasma deposition system used.201
4.2 Surfaces for functional immobilisation of biomolecules
Proteins contain functional groups such as –COOH, –NH2, –OH and –SH, which can be physically (random) or covalently (oriented) immobilised to microchannel surfaces (Fig. 4).219 In some lab-on-chip applications such as affinity separation220,221 and enzymatic reactors,222,223 biomolecules are required to be surface-immobilised in specific orientations such that their biological activities are retained. Reactive surface chemistries commonly used for biomolecule immobilisation are amine, carboxyl, aldehyde, epoxy, isocyanate, hydroxyl, maleimide, and NHS-activated surfaces, which can be prepared by different strategies: generally wet chemical treatments.74 However, hydroxyl (alcohol monomers), amine (alkylamine monomers), carboxyl (carboxylic acid, propanoic acid monomer) and aldehyde (aldehye monomers) functionalised plasma polymers have also been shown to be successful for covalent immobilisation of biologically active molecules.224
 |
| Fig. 4 Schematic diagram of random, uncontrolled (A) and oriented (B) protein adsorption to surfaces. | |
5. Closing remarks
The central concept of the proteomics-on-chip is to integrate different types of separation and preparation tools. The aim of this is to enable fast analyses of a real-time complex sample containing up to thousands of proteins and many hundreds of thousands of peptides, present with possibility >10 orders (or even more) of magnitude in concentration range.225 Design of a universal microfluidic platform for proteomics is therefore difficult to predict due to the inherent complexities in each of the relevant unit operations, and many possible ways of integration. Most of the microfluidic chip systems built is therefore focused on specific applications such as on-line sample pre-treatment, protein digestion, protein/peptide separation, and its MS coupling. Regardless of the design, biomolecules that are present in most of the unit operations need to be effectively controlled by surface modification to obtain a globally reliable, robust, and reproducible microfluidic chip platform for complex proteome analyses. Although significant contributions have been made to control bio-interfacial interactions as already described in this article, further work is needed to effectively coat the inner surface of microchannels in a manner independent of the substrate materials used, and to develop characterisation tools for inner surface coatings and protein coverage. Integration between unit operations is also a major drawback that needs to be assessed to ensure down-stream solvent compatibility with MS instrumentation and minimal sample loss (so that low abundance proteins are still possible to measure). Addressing these challenges, together with the advancement in proteomics, microfluidics, and surface technologies, can provide the key to future growth in biotechnology, pharmaceutical, and biomedical research.
Acknowledgements
We acknowledge the EPSRC for funding (grants GR/S84347/01 and EP/E036252/1).
References
- W. C. Andrea, Lab-on-a-chip: Opportunities for chemical engineering, AIChE J., 2002, 48(8), 1590–1595 CrossRef CAS.
- A. Manz, N. Graber and H. M. Widmer, Miniaturized total chemical analysis systems: A novel concept for chemical sensing, Sens. Actuators, B, 1990, 1(1–6), 244–248 CrossRef.
- D. Janasek, J. Franzke and A. Manz, Scaling and the design of miniaturized chemical-analysis systems, Nature, 2006, 442(7101), 374–380 CrossRef CAS.
- N. Lion, F. Reymond, H. H. Girault and J. S. Rossier, Why the move to microfluidics for protein analysis?, Curr. Opin. Biotechnol., 2004, 15(1), 31–37 CrossRef CAS.
- U. Bilitewski, M. Genrich, S. Kadow and G. Mersal, Biochemical analysis with microfluidic systems, Anal. Bioanal. Chem., 2003, 377(3), 556–569 CrossRef CAS.
- M. J. Madou and R. Cubicciotti, Scaling issues in chemical and biological sensors, Proc. IEEE, 2003, 91(6), 830–838 CrossRef CAS.
- J. S. Rossier and H. H. Girault, Enzyme linked immunosorbent assay on a microchip with electrochemical detection, Lab Chip, 2001, 1(2), 153–157 RSC.
- J. S. Rossier, G. Gokulrangan, H. H. Girault, S. Svojanovsky and G. S. Wilson, Characterization of Protein Adsorption and Immunosorption Kinetics in Photoablated Polymer Microchannels, Langmuir, 2000, 16(22), 8489–8494 CrossRef CAS.
- P. S. Dittrich and A. Manz, Lab-on-a-chip: microfluidics in drug discovery, 2006, 5(3), 210–218 Search PubMed.
- Garcia, D. Carlos, L. I. U. Yan, Henry and S. Charles, Recent progress in micro-TAS for clinical applications, Analyst, 2003, 128, 1002–1008 RSC.
- C. H. Arnaud, Mimicking Biological Systems, Chemical and Engineering News, 2007, 14–19 Search PubMed.
- K.-J. Jang and K.-Y. Suh, A multi-layer microfluidic device for efficient culture and analysis of renal tubular cells, Lab Chip, 2010, 10, 36, 10.1039/b907515a.
- D. Huh, A. Mammoto, M. Montoya-Zavala and D. E. Ingber, A human breathing lung-on-a-chip, FASEB J., 2010, 24 , (1_MeetingAbstracts), 1065.22-.
- H. Yu, I. Meyvantsson, I. A. Shkel and D. J. Beebe, Diffusion dependent cell behavior in microenvironments, Lab Chip, 2005, 5(10), 1089–1095 RSC.
-
J. Roboz, Strategies, Techniques, and Applications in Cancer Biochemistry and Biology, CRC Press, 2002, p. 361 Search PubMed.
- R. E. Banks, M. J. Dunn, D. F. Hochstrasser, J.-C. Sanchez, W. Blackstock, D. J. Pappin and P. J. Selby, Proteomics: new perspectives, new biomedical opportunities, Lancet, 2000, 356(9243), 1749–1756 CrossRef CAS.
- P. Jagroop, P. L. S. Ambrosius, C. W. Phillip and A. B. Catherine, A cross-species quantitative proteomic study of salt adaptation in a halotolerant environmental isolate using 15N metabolic labelling, Proteomics, 2008, 8(11), 2266–2284 CrossRef CAS.
- J. Khandurina and A. Guttman, Bioanalysis in microfluidic devices, J. Chromatogr., A, 2002, 943(2), 159–83 CrossRef CAS.
- R. B. M. Schasfoort, Proteomics-on-a-chip: the challenge to couple lab-on-a-chip unit operations, Expert Rev. Proteomics, 2004, 1(1), 123–132 Search PubMed.
- D. Figeys and D. Pinto, Proteomics on a chip: promising developments, Electrophoresis, 2001, 22(2), 208–216 CrossRef CAS.
- S. Mouradian, Lab-on-a-chip: applications in proteomics, Curr. Opin. Chem. Biol., 2002, 6(1), 51–56 CrossRef CAS.
- J. Han and A. K. Singh, Rapid protein separations in ultra-short microchannels: microchip sodium dodecyl sulfate-polyacrylamide gel electrophoresis and isoelectric focusing, Journal of chromatography. A, 2004, 1049(1–2), 205–209 CrossRef CAS.
- Y. Li, J. S. Buch, F. Rosenberger, D. L. DeVoe and C. S. Lee, Integration of isoelectric focusing with parallel sodium dodecyl sulfate gel electrophoresis for multidimensional protein separations in a plastic microfluidic [correction of microfludic] network, Anal. Chem., 2004, 76(3), 742–748 CrossRef CAS.
- J. Kim, M. Johnson, P. Hill and B. K. Gale, Microfluidic sample preparation: cell lysis and nucleic acid purification, Integr. Biol., 2009, 1(10), 574–586 RSC.
- Y. Hiroshi, M. Masaya, H. Takeshi, B.-N. Maria Portia, A. Kazunari and M. Hideaki, Rapid and efficient proteolysis for proteomic analysis by protease-immobilized microreactor, Electrophoresis, 2009, 30(18), 3257–3264 CrossRef CAS.
- J. Carlier, K. Chuda, S. Arscott, V. Thomy, B. Verbeke, X. Coqueret, C. Jean Christophe, C. Druon and P. Tabourier, High pressure-resistant SU-8 microchannels for monolithic porous structure integration, J. Micromech. Microeng., 2006, 16, 2211–2219 CrossRef.
- L. Jeonghoon, A. S. Steven and K. M. Kermit, Microfluidic chips for mass spectrometry-based proteomics, J. Mass Spectrom., 2009, 44(5), 579–593 CrossRef CAS.
- J. Lee, S. A. Soper and K. K. Murray, Microfluidics with MALDI analysis for proteomics—A review, Anal. Chim. Acta, 2009, 649(2), 180–190 CrossRef CAS.
- R. T. Kelly, K. Tang, D. Irimia, M. Toner and R. D. Smith, Elastomeric Microchip Electrospray Emitter for Stable Cone-Jet Mode Operation in the Nanoflow Regime, Anal. Chem., 2008, 80(10), 3824–3831 CrossRef CAS.
- C. Li, Y. Yang, H. G. Craighead and K. H. Lee, Isoelectric focusing in cyclic olefin copolymer microfluidic channels coated by polyacrylamide using a UV photografting method, Electrophoresis, 2005, 26(9), 1800–1806 CrossRef CAS.
- H.-B. Javier, N. Christian, C. Alejandro and P. Matthias, On-line capillary electrophoresis-mass spectrometry for the analysis of biomolecules, Electrophoresis, 2004, 25(14), 2257–2281 CrossRef CAS.
- J. Preisler and E. S. Yeung, Characterization of Nonbonded Poly(ethylene oxide) Coating for Capillary Electrophoresis via Continuous Monitoring of Electroosmotic Flow, Anal. Chem., 1996, 68(17), 2885–2889 CrossRef CAS.
- G. Y. Tang, C. Yang, J. C. Chai and H. Q. Gong, Joule heating effect on electroosmotic flow and mass species transport in a microcapillary, Int. J. Heat Mass Transfer, 2004, 47(2), 215–227 CrossRef CAS.
- S. Ghosal, Effect of Analyte Adsorption on the Electroosmotic Flow in Microfluidic Channels, Anal. Chem., 2002, 74(4), 771–775 CrossRef CAS.
- M. Salim, P. C. Wright and S. L. McArthur, Studies of electroosmotic flow and the effects of protein adsorption in plasma-polymerized microchannel surfaces, Electrophoresis, 2009, 30(11), 1877–1887 CrossRef CAS.
- W. Hellmich, J. Regtmeier, T. T. Duong, R. Ros, D. Anselmetti and A. Ros, Poly(oxyethylene) Based Surface Coatings for Poly(dimethylsiloxane) Microchannels, Langmuir, 2005, 21(16), 7551–7557 CrossRef CAS.
-
R. E. Oosterbroek and A. v. d. Berg, Lab-on-a-chip: miniaturized systems for (bio) chemical analysis and synthesis, Elsevier B.V, 2003, p. 450 Search PubMed.
- A. J. de Mello and N. Beard, Dealing with real samples: sample pre-treatment in microfluidic systems, Lab Chip, 2003, 3(1), 11N–19 RSC.
- Y.-C. Tyan, J.-D. Liao, S.-B. Jong, P.-C. Liao, M.-H. Yang, Y.-W. Chang, R. Klauser, M. Himmelhaus and M. Grunze, Characterization of trypsin immobilized on the functionable alkylthiolate self-assembled monolayers: a preliminary application for trypsin digestion chip on protein identification using matrix-assisted laser desorption/ionization time-of-flight mass spectrometry, J. Mater. Sci.: Mater. Med., 2005, 16(2), 135–142 CrossRef CAS.
- A. Ríos, A. Escarpa, M. C. González and A. G. Crevillén, Challenges of analytical microsystems, TrAC, Trends Anal. Chem., 2006, 25(5), 467–479 CrossRef CAS.
- D. Mark, S. Haeberle, G. Roth, F. v. Stetten and R. Zengerle, Microfluidic lab-on-a-chip platforms: requirements, characteristics and applications, Chem. Soc. Rev., 2010, 39(3), 1153–1182 RSC.
- T. Nguyet Thuy, A. Ichraf, P. Antoine and T. Myriam, Recent innovations in protein separation on microchips by electrophoretic methods: An update, Electrophoresis, 2010, 31(1), 147–173 CrossRef.
- W. Norde, Adsorption of proteins from solution at the solid–liquid interface, Adv. Colloid Interface Sci., 1986, 25, 267–340 CrossRef CAS.
-
J. D. Andrade, Surface and Interfacial Aspects of Bio-medical Polymers, in Principles of protein adsorption, Plenum Press, New York, 1985, vol. 2, pp. 4–6 Search PubMed.
- C. A. Haynes and W. Norde, Globular proteins at solid/liquid interfaces, Colloids Surf., B, 1994, 2(6), 517–566 CrossRef CAS.
- J. D. Andrade, V. Hlady and A. P. Wei, Adsorption of complex proteins at interfaces, Pure Appl. Chem., 1992, 64(11), 1777–1781 CrossRef CAS.
-
K. Al-Malah, in Adsorption: Theory, Modeling, and Analysis, ed. J. Tóth, Marcel Dekker, New York, 2002, vol. 107, p. 803 Search PubMed.
- P. M. Claesson, E. Blomberg, J. C. Froberg, T. Nylander and T. Arnebrant, Protein interactions at solid surfaces, Adv. Colloid Interface Sci., 1995, 57, 161–227 CrossRef CAS.
-
K. C. Dee, R. Bizios and D. A. Puleo, Protein-surface Interactions, in Introduction to Tissue-Biomaterial Interactions, John Wiley and Sons Ltd., 1st edn, 2002, pp. 37–40 Search PubMed.
- C. F. Wertz and M. M. Santore, Effect of Surface Hydrophobicity on Adsorption and Relaxation Kinetics of Albumin and Fibrinogen: Single-Species and Competitive Behavior, Langmuir, 2001, 17(10), 3006–3016 CrossRef CAS.
- M. van der Veen, M. C. Stuart and W. Norde, Spreading of proteins and its effect on adsorption and desorption kinetics, Colloids Surf., B, 2007, 54(2), 136–142 CrossRef CAS.
- J. Buijs and V. Hlady, Adsorption Kinetics, Conformation, and Mobility of the Growth Hormone and Lysozyme on Solid Surfaces, Studied with TIRF, J. Colloid Interface Sci., 1997, 190(1), 171–181 CrossRef CAS.
- M. Malmsten, Ellipsometry studies of the effects of surface hydrophobicity on protein adsorption, Colloids Surf., B, 1995, 3(5), 297–308 CrossRef CAS.
- G. B. Sigal, M. Mrksich and G. M. Whitesides, Effect of Surface Wettability on the Adsorption of Proteins and Detergents, J. Am. Chem. Soc., 1998, 120(14), 3464–3473 CrossRef CAS.
- T. Zoungrana, G. H. Findenegg and W. Norde, Structure, Stability, and Activity of Adsorbed Enzymes, J. Colloid Interface Sci., 1997, 190(2), 437–448 CrossRef CAS.
- G. Raffaini and F. Ganazzoli, Protein Adsorption on a Hydrophobic Surface: A Molecular Dynamics Study of Lysozyme on Graphite, Langmuir, 2010, 26(8), 5679–5689 CrossRef CAS.
-
Z. Adamczyk, Irreversible adsorption of particles, in Adsorption: Theory, Modeling, and Analysis, ed. J. Tóth, Marcel Dekker, 2002, vol. 107, p. 251 Search PubMed.
- S. Kumar and R. Nussinov, Close-Range Electrostatic Interactions in Proteins, ChemBioChem, 2002, 3(7), 604–617 CrossRef CAS.
- M. Salim, B. O'Sullivan, S. L. McArthur and P. C. Wright, Characterization of fibrinogen adsorption onto glass microcapillary surfaces by ELISA, Lab Chip, 2007, 7(1), 64–70 RSC.
-
R. A. Latour, Biomaterials: Protein–Surface Interactions, in Encyclopedia of Biomaterials and Biomedical Engineering, ed. Gary L. Bowlin and G. W., Marcel Dekker, Inc., New York, 2005, vol. 2, p. 1 Search PubMed.
-
L. Hagel, Separation on the basis of chemistry, in Protein Purification Techniques: A Practical Approach, ed. S. Roe, Oxford University Press, 2nd edn, 2001, p. 155 Search PubMed.
-
K. L. Hanson, L. Filipponi and D. V. Nicolau, Biomolecules and cells on surfaces-fundamental concepts, in Microarray Technology and its Application, ed. U. R. Müller and D. V. Nicolau, Springer, New York, 2005, pp. 23–44 Search PubMed.
-
A. T. Hubbard, Encyclopedia of Surface and Colloid Science, Marcel Dekker, New York, 2002, vol. 4, p. 4383 Search PubMed.
-
T. Arakawa and J. S. Philo, Biophysical and biochemical analysis of recombinant proteins structure and analysis of proteins, in Pharmaceutical Biotechnology, D. J. A. Crommelin and R. D. Sindelar, Gutenberg Press, 2nd edn, 2002, p. 25 Search PubMed.
-
J. L. Brash and T. A. Horbett, Proteins at Interfaces: Physicochemical and Biochemical Studies, Washington DC, 1987, ACS Symposium Series, vol. 343, pp. 1–33 Search PubMed.
-
J. L. Brash and T. A. Horbett, Proteins at interfaces II: fundamentals and applications, Washington DC, 1995, ACS Symposium Series, vol. 602, pp. 1–23 Search PubMed.
- J. D. Andrade, V. Hlady, A.-P. Wei, C.-H. Ho, A. S. Lea, S. I. Jeon, Y. S. Lin and E. Stroup, Proteins at interfaces: Principles, multivariate aspects, protein resistant surfaces, and direct imaging and…, Clin. Mater., 1992, 11(1–4), 67–84 CrossRef CAS.
-
M. Malmsten, Protein Adsorption at the Solid–Liquid Interface, in Protein Architecture, ed. Y. Lvov and H. Hohwald, Marcel Dekker, 1999, p. 6 Search PubMed.
- Y. Ling, Z. Lu, Y. Gan, Y. Liu and C. M. Li, AFM study of adsorption of protein A on a poly(dimethylsiloxane) surface, Nanotechnology, 2009, 20(28), 285101 CrossRef.
- E. Ouellet, C. Lausted, T. Lin, C. W. T. Yang, L. Hood and E. T. Lagally, Parallel microfluidic surface plasmon resonance imaging arrays, Lab Chip, 2010, 10(5), 581–588 RSC.
- G. Krishnamoorthy, E. T. Carlen, H. L. deBoer, A. van den Berg and R. B. M. Schasfoort, Electrokinetic Lab-on-a-BioChip for Multi-ligand/Multi-analyte Biosensing, Anal. Chem., 2010, 82(10), 4145–4150 CrossRef CAS.
-
J. S. M. Hans Griesser, S. Matthew Wagner, G. David Castner, Peter Kingshott and M. Keith Mclean, XPS, ToF-SIMS, and MALDI-MS for Characterizing Adsorbed Protein Films, in Biopolymers at Interfaces, ed. M. Malmsten, CRC Press, 2003, vol. 110 Search PubMed.
-
H. H. Peter Kingshott, Methods of assessing protein adsorption, in Encyclopedia of Surface and Colloid Science, ed. T. A. Hubbard, CRC Press, 2002, vol. 3, p. 3343 Search PubMed.
- J. M. Goddard and J. H. Hotchkiss, Polymer surface modification for the attachment of bioactive compounds, Prog. Polym. Sci., 2007, 32(7), 698–725 CrossRef CAS.
- S. L. McGurk, R. J. Green, G. H. W. Sanders, M. C. Davies, C. J. Roberts, S. J. B. Tendler and P. M. Williams, Molecular interactions of biomolecules with surface-engineered interfaces using atomic force microscopy and surface plasmon resonance, Langmuir, 1999, 15(15), 5136–5140 CrossRef CAS.
- Y. F. Dufrene, T. G. Marchal and P. G. Rouxhet, Probing the organization of adsorbed protein layers: Complementarity of atomic force microscopy, X-ray photoelectron spectroscopy and radiolabeling, Appl. Surf. Sci., 1999, 144–145, 638–643 CrossRef CAS.
-
M. Bergkvist, Adsorption of Proteins: Methods of Orientation, in Encyclopedia of Surface and Colloid Science, ed. P. Somasundaran, CRC Press, 2006, vol. 1 Search PubMed.
- D. S. Mantus, B. D. Ratner, B. A. Carlson and J. F. Moulder, Static secondary ion mass spectrometry of adsorbed proteins, Anal. Chem., 1993, 65(10), 1431–1438 CrossRef CAS.
-
J. Marchand-Brynaert, Polymer Membranes, in Encyclopedia of Surface and Colloid Science, ed. P. Somasundaran, CRC Press, 2006, vol. 6, p. 4861 Search PubMed.
- H. Matsumura, K. Kawasaki, N. Okumura, M. Kambara and W. Norde, Characterization of the surface of protein-adsorbed dental materials by wetting and streaming potential measurements, Colloids Surf., B, 2003, 32(2), 97–103 CrossRef CAS.
-
J. W. Pollard, Two-Dimensional Polyacrylamide Gel Electrophoresis of Proteins, in Methods in Molecular Biology, ed. J. M. Walker, Humana Press Inc., New Jersey, 1984, vol. 1, p. 81 Search PubMed.
-
W. Possart, Surface Characterization of Advanced Polymers, ed. Luigia Sabbatini and Pier Giorgio, Zambonin VCH, Weinheim, New York, Basel, Cambridge, Tokyo, 1993. Acta Polymerica, 1994, 45, (2), 131 Search PubMed.
- F. Luna-Vera and J. C. Alvarez, Adsorption kinetics of proteins in plastic microfluidic channels: Real-time monitoring of lysozyme adsorption by pulsed streaming potentials, Biosens. Bioelectron., 2010, 25(6), 1539–1543 CrossRef CAS.
- F. Boumaza, P. Dejardin, F. Yan, F. Bauduin and Y. Holl, Fibrinogen adsorption on Pyrex glass tubes: A continuous kinetic study, Biophys. Chem., 1992, 42(1), 87–92 CrossRef CAS.
- J. N. B. D. Lin, A. S. Lea, P. K. Hansma and J. D. Andrade, Direct observation of Immunoglobulin adsorption dynamics using the atomic force microscope, Langmuir, 1990, 6, 509–511 CrossRef CAS.
- Y.-S. Lin and V. Hlady, Human serum albumin adsorption onto octadecyldimethylsilyl–silica gradient surface, Colloids Surf., B, 1994, 2(5), 481–491 CrossRef CAS.
- J. L. Ortega-Vinuesa, P. Tengvall and I. Lundstrom, Molecular packing of HSA, IgG, and fibrinogen adsorbed on silicon by AFM imaging, Thin Solid Films, 1998, 324(1–2), 257–273 CrossRef CAS.
- W. J. J. Dillman and I. F. Miller, On the adsorption of serum proteins on polymer membrane surfaces, J. Colloid Interface Sci., 1973, 44(2), 221–241 CrossRef CAS.
- L. B. Thompson, N. H. Mack and R. G. Nuzzo, Bifunctional polyacrylamide based polymers for the specific binding of hexahistidine tagged proteins on gold surfaces, Phys. Chem. Chem. Phys., 2010, 12(17), 4301–4308 RSC.
- N. Bonnet, D. O'Hagan and G. Hahner, Protein adsorption onto CF3-terminated oligo(ethylene glycol) containing self-assembled monolayers (SAMs): the influence of ionic strength and electrostatic forces, Phys. Chem. Chem. Phys., 2010, 12(17), 4367–4374 RSC.
- J. Jenkins, B. Prabhakarpandian, K. Lenghaus, J. Hickman and S. Sundaram, Fluidics-resolved estimation of protein adsorption kinetics in a biomicrofluidic system, Anal. Biochem., 2004, 331(2), 207–215 CrossRef CAS.
- K. Lenghaus, J. W. Dale, J. C. Henderson, D. C. Henry, E. R. Loghin and J. J. Hickman, Enzymes as Ultrasensitive Probes for Protein Adsorption in Flow Systems, Langmuir, 2003, 19(15), 5971–5974 CrossRef CAS.
- A. Lionello, J. Josserand, H. Jensen and H. H. Girault, Dynamic protein adsorption in microchannels by “stop-flow” and continuous flow, Lab Chip, 2005, 5(10), 1096–103 RSC.
- A. Lionello, J. Josserand, H. Jensen and H. H. Girault, Protein adsorption in static microsystems: effect of the surface to volume ratio, Lab Chip, 2005, 5(3), 254–60 RSC.
- D. V. Nicolau, Jr., F. Fulga and D. V. Nicolau, Impact of protein adsorption on the geometry of microfluidic devices, Biomed. Microdevices, 2003, 5(3), 227–233 CrossRef.
- W. F. Patton, Detection technologies in proteome analysis, J. Chromatogr., B: Anal. Technol. Biomed. Life Sci., 2002, 771(1–2), 3–31 CrossRef CAS.
- E. N. Vasina, E. Paszek, D. V. Nicolau, Jr. and D. V. Nicolau, The BAD project: data mining, database and prediction of protein adsorption on surfaces, Lab Chip, 2009, 9(7), 891–900 RSC.
- N. Nath, J. Hyun, H. Ma and A. Chilkoti, Surface engineering strategies for control of protein and cell interactions, Surf. Sci., 2004, 570(1–2), 98–110 CrossRef CAS.
- M. Mrksich, What can surface chemistry do for cell biology?, Curr. Opin. Chem. Biol., 2002, 6(6), 794–797 CrossRef CAS.
- J. Liu, T. Pan, A. T. Woolley and M. L. Lee, Surface-Modified Poly(methyl methacrylate) Capillary Electrophoresis
Microchips for Protein and Peptide Analysis, Anal. Chem., 2004, 76(23), 6948–6955 CrossRef CAS.
- Y. Yuan, H. Hongyan and L. J. Lee, Protein A-based antibody immobilization onto polymeric microdevices for enhanced sensitivity of enzyme-linked immunosorbent assay, Biotechnol. Bioeng., 2009, 102(3), 891–901 CrossRef CAS.
- W.-C. Sung, H.-H. Chen, H. Makamba and S.-H. Chen, Functionalized 3D-Hydrogel Plugs Covalently Patterned Inside Hydrophilic Poly(dimethylsiloxane) Microchannels for Flow-Through Immunoassays, Anal. Chem., 2009, 81(19), 7967–7973 CrossRef CAS.
- D. G. Castner and B. D. Ratner, Biomedical surface science: Foundations to frontiers, Surf. Sci., 2002, 500(1–3), 28–60 CrossRef CAS.
- W. L. W. Hau, D. W. Trau, N. J. Sucher, M. Wong and Y. Zohar, Surface-chemistry technology for microfluidics, J. Micromech. Microeng., 2003, 13(2), 272 CrossRef CAS.
- M. Tirrell, Chemical engineering of polymers: production of flexible, functional materials, Chem. Eng. Sci., 1995, 50(24), 4123–4141 CrossRef CAS.
- M. Advincula, X. Fan, J. Lemons and R. Advincula, Surface modification of surface sol–gel derived titanium oxide films by self-assembled monolayers (SAMs) and non-specific protein adsorption studies, Colloids Surf., B, 2005, 42(1), 29–43 CrossRef CAS.
- M. Herrero, R. Navarro, Y. Grohens, H. Reinecke and C. Mijangos, Controlled wet-chemical modification and bacterial adhesion on PVC-surfaces, Polym. Degrad. Stab., 2006, 91(9), 1915–1918 CrossRef CAS.
- C. Oehr, Plasma surface modification of polymers for biomedical use, Nucl. Instrum. Methods Phys. Res., Sect. B, 2003, 208, 40–47 CrossRef CAS.
- Y. L. Lee and C. Y. Chen, Surface wettability and platelet adhesion studies on Langmuir–Blodgett films, Appl. Surf. Sci., 2003, 207(1–4), 51–62 CrossRef CAS.
- K. Huikko, R. Kostiainen and T. Kotiaho, Introduction to micro-analytical systems: bioanalytical and pharmaceutical applications, Eur. J. Pharm. Sci., 2003, 20(2), 149–71 CrossRef CAS.
- A. Pallandre, B. de Lambert, R. Attia, A. M. Jonas and J.-L. Viovy, Surface treatment and characterization: perspectives to electrophoresis and lab-on-chips, Electrophoresis, 2006, 27(3), 584–610 CrossRef CAS.
- J. Liu and M. L. Lee, Permanent surface modification of polymeric capillary electrophoresis microchips for protein and peptide analysis, Electrophoresis, 2006, 27(18), 3533–3546 CrossRef CAS.
- D. Belder and M. Ludwig, Surface modification in microchip electrophoresis, Electrophoresis, 2003, 24(21), 3595–606 CrossRef CAS.
- V. Dolník, Wall coating for capillary electrophoresis on microchips, Electrophoresis, 2004, 25(21–22), 3589–3601 CrossRef CAS.
- P. G. Righetti, C. Gelfi, B. Verzola and L. Castelletti, The state of the art of dynamic coatings, Electrophoresis, 2001, 22(4), 603–611 CrossRef CAS.
- E. A. S. Doherty, R. J. Meagher, M. N. Albarghouthi and A. E. Barron, Microchannel wall coatings for protein separations by capillary and chip electrophoresis, Electrophoresis, 2003, 24(12), 34–54 CrossRef CAS.
- J. Horvath and V. Dolník, Polymer wall coatings for capillary electrophoresis, Electrophoresis, 2001, 22(4), 644–655 CrossRef CAS.
- J. Ou, T. Glawdel, C. L. Ren and J. Pawliszyn, Fabrication of a hybrid PDMS/SU-8/quartz microfluidic chip for enhancing UV absorption whole-channel imaging detection sensitivity and application for isoelectric focusing of proteins, Lab Chip, 2009, 9(13), 1926–1932 RSC.
- M. Gilges, M. H. Kleemiss and G. Schomburg, Capillary Zone Electrophoresis Separations of Basic and Acidic Proteins Using Poly(vinyl alcohol) Coatings in Fused Silica Capillaries, Anal. Chem., 1994, 66(13), 2038–2046 CrossRef CAS.
- S. L. R. Barker, M. J. Tarlov, H. Canavan, J. J. Hickman and L. E. Locascio, Plastic Microfluidic Devices Modified with Polyelectrolyte Multilayers, Anal. Chem., 2000, 72(20), 4899–4903 CrossRef CAS.
- X. Liu, D. Erickson, D. Li and U. J. Krull, Cationic polymer coatings for design of electroosmotic flow and control of DNA adsorption, Anal. Chim. Acta, 2004, 507(1), 55–62 CrossRef CAS.
- F. Kitagawa, K. Kubota, K. Sueyoshi and K. Otsuka, One-step immobilization of cationic polymer onto a poly(methyl methacrylate) microchip for high-performance, Sci. Technol. Adv. Mater., 2006, 7(6), 558–565 CrossRef CAS.
- Y. Liu, J. C. Fanguy, J. M. Bledsoe and C. S. Henry, Dynamic coating using polyelectrolyte multilayers for chemical control of electroosmotic flow in capillary electrophoresis microchips, Anal. Chem., 2000, 72(24), 5939–5944 CrossRef CAS.
- G. Decher, J. D. Hong and J. Schmitt, Buildup of ultrathin multilayer films by a self-assembly process: III. Consecutively alternating adsorption of anionic, Thin Solid Films, 1992, 210–211, 831–835 CrossRef.
- H. Katayama, Y. Ishihama and N. Asakawa, Stable Capillary Coating with Successive Multiple Ionic Polymer Layers, Anal. Chem., 1998, 70(11), 2254–2260 CrossRef CAS.
- H. Nagata, M. Tabuchi, K. Hirano and Y. Baba, Microchip electrophoretic protein separation using electroosmotic flow induced by dynamic sodium dodecyl sulfate-coating of uncoated plastic chips, Electrophoresis, 2005, 26(11), 2247–2253 CrossRef CAS.
- C. D. García, B. M. Dressen, A. Henderson and C. S. Henry, Comparison of surfactants for dynamic surface modification of poly(dimethylsiloxane) microchips, Electrophoresis, 2005, 26(3), 703–709 CrossRef CAS.
- J. S. Green and J. W. Jorgenson, Minimizing adsorption of proteins on fused silica in capillary zone electrophoresis by the addition of alkali metal salts to the buffers, Journal of Chromatography, 1989, 478, 63–70 CrossRef CAS.
- C. Gelfi, A. Vigano, M. Ripamonti, P. G. Righetti, R. Sebastiano and A. Citterio, Protein Analysis by Capillary Zone Electrophoresis Utilizing a Trifunctional Diamine for Silica Coating, Anal. Chem., 2001, 73(16), 3862–3868 CrossRef CAS.
- D. Corradini, G. Cannarsa, E. Fabbri and C. Corradini, Effects of alkylamines on electroosmotic flow and protein migration behaviour in capillary electrophoresis, J. Chromatogr., A, 1995, 709(1), 127–134 CrossRef CAS.
- P. R. Van Tassel, D. Miras, A. Hagege, M. Leroy, J.-C. Voegel and P. Schaaf, Control of Protein Adsorption in Capillary Electrophoresis via an Irreversibly Bound Protein Coating, J. Colloid Interface Sci., 1996, 183(1), 269–273 CrossRef CAS.
- N. Naruishi, Y. Tanaka, T. Higashi and S. i. Wakida, Highly efficient dynamic modification of plastic microfluidic devices using proteins in microchip capillary, J. Chromatogr., A, 2006, 1130(2), 169–174 CrossRef CAS.
- W. Schrott, Z. Slouka, P. Cervenka, J. Ston, M. Nebyla, M. Pribyl and D. Snita, Study on surface properties of PDMS microfluidic chips treated with albumin, Biomicrofluidics, 2009, 3(4), 044101–15 CrossRef.
- D. G. Carlos, M. D. Brian, H. Amber and S. H. Charles, Comparison of surfactants for dynamic surface modification of poly(dimethylsiloxane) microchips, Electrophoresis, 2005, 26(3), 703–709 CrossRef CAS.
- M. F. Mora, J. Felhofer, A. Ayon and C. D. Garcia, Surfactants as a Preferred Option to Improve Separation and Electrochemical Detection in Capillary Electrophoresis, Anal. Lett., 2008, 41(2), 312–334 CrossRef CAS.
- K. K. C. Yeung and C. A. Lucy, Suppression of Electroosmotic Flow and Prevention of Wall Adsorption in Capillary Zone Electrophoresis Using Zwitterionic Surfactants, Anal. Chem., 1997, 69(17), 3435–3441 CrossRef CAS.
- Y. Luo, B. Huang, H. Wu and R. N. Zare, Controlling electroosmotic flow in poly(dimethylsiloxane) separation channels
by means of prepolymer additives, Anal. Chem., 2006, 78(13), 4588–4592 CrossRef CAS.
- H. Wang, D. Wang, J. Wang, H. Wang, J. Gu, C. Han, Q. Jin, B. Xu, C. He, L. Cao, Y. Wang and J. Zhao, Application of poly(dimethylsiloxane)/glass microchip for fast electrophoretic separation of serum small, dense low-density lipoprotein, J. Chromatogr., A, 2009, 1216(35), 6343–6347 CrossRef CAS.
- S. Makino, J. A. Reynolds and C. Tanford, The Binding of Deoxycholate and Triton X-100 to Proteins, Journal of Biological Chemistry, 1973, 248(14), 4926–4932 CAS.
- J. Z. Xing, R. Michael and E. Richards, Determination of peptidoglycan-associated protein in Escherichia coli NCIB 8545 by capillary zone electrophoresis, J. Chromatogr., A, 1996, 740(2), 273–278 CrossRef CAS.
- S. Prakash, T. M. Long, J. C. Selby, J. S. Moore and M. A. Shannon, “Click” Modification of Silica Surfaces and Glass Microfluidic Channels, Anal. Chem., 2007, 79(4), 1661–1667 CrossRef CAS.
- E. A. S. Doherty, K. D. Berglund, B. A. Buchholz, I. V. Kourkine, T. M. Przybycien, R. D. Tilton and A. E. Barron, Critical factors for high-performance physically adsorbed (dynamic) polymeric wall coatings for capillary electrophoresis of DNA, Electrophoresis, 2002, 23(16), 2766–2776 CrossRef CAS.
- A. Papra, A. Bernard, D. Juncker, N. B. Larsen, B. Michel and E. Delamarche, Microfluidic Networks Made of Poly(dimethylsiloxane), Si, and Au Coated with Polyethylene Glycol for Patterning Proteins onto Surfaces, Langmuir, 2001, 17(13), 4090–4095 CrossRef CAS.
- S. A. Soper, A. C. Henry, B. Vaidya, M. Galloway, M. Wabuyele and R. L. McCarley, Surface modification of polymer-based microfluidic devices, Anal. Chim. Acta, 2002, 470(1), 87–99 CrossRef CAS.
- S. Hu, X. Ren, M. Bachman, C. E. Sims, G. P. Li and N. L. Allbritton, Surface-Directed, Graft Polymerization within Microfluidic Channels, Anal. Chem., 2004, 76(7), 1865–1870 CrossRef CAS.
- J. Wang, A. Muck Jr., M. P. Chatrathi, G. Chen, N. Mittal, S. D. Spillman and S. Obeidat, Bulk modification of polymeric microfluidic devices, Lab Chip, 2005, 5(2), 226–230 RSC.
- Y. Bai, C. G. Koh, M. Boreman, Y.-J. Juang, I. C. Tang, L. J. Lee and S.-T. Yang, Surface modification for enhancing antibody binding on polymer-based microfluidic device for enzyme-linked immunosorbent assay, Langmuir: The ACS Journal of Surfaces and Colloids, 2006, 22(22), 9458–9467 Search PubMed.
- D. Wu, B. Zhao, Z. Dai, J. Qin and B. Lin, Grafting epoxy-modified hydrophilic polymers onto poly(dimethylsiloxane) microfluidic chip to resist nonspecific protein adsorption, Lab Chip, 2006, 6(7), 942–947 RSC.
- B. Wang, L. Chen, Z. Abdulali-Kanji, J. H. Horton and R. D. Oleschuk, Aging Effects on Oxidized and Amine-Modified Poly(dimethylsiloxane) Surfaces Studied with Chemical Force Titrations: Effects on Electroosmotic Flow Rate in Microfluidic Channels, Langmuir, 2003, 19(23), 9792–9798 CrossRef CAS.
- Y. Bai, W.-C. Huang and S.-T. Yang, Enzyme-linked immunosorbent assay of Escherichia coli O157:H7 in surface enhanced poly(methyl methacrylate) microchannels, Biotechnol. Bioeng., 2007, 98(2), 328–339 CrossRef CAS.
- J. H. Lee, B. J. Jeong and H. B. Lee, Plasma protein adsorption and platelet adhesion onto comb-like PEO gradient surfaces, J. Biomed. Mater. Res., 1997, 34(1), 105–114 CrossRef CAS.
- H. Chen, M. A. Brook and H. Sheardown, Silicone elastomers for reduced protein adsorption, Biomaterials, 2004, 25(12), 2273–2282 CrossRef CAS.
- J. Gaudioso and H. G. Craighead, Characterizing electroosmotic flow in microfluidic devices, J. Chromatogr., A, 2002, 971(1–2), 249–253 CrossRef CAS.
- T. J. Johnson, D. Ross, M. Gaitan and L. E. Locascio, Laser Modification of Preformed Polymer Microchannels: Application To Reduce Band Broadening around Turns Subject to Electrokinetic Flow, Anal. Chem., 2001, 73(15), 3656–3661 CrossRef CAS.
- X. Ren, M. Bachman, C. Sims, G. P. Li and N. Allbritton, Electroosmotic properties of microfluidic channels composed of poly(dimethylsiloxane), J. Chromatogr., B: Biomed. Sci. Appl., 2001, 762(2), 117–125 CrossRef CAS.
- K. C. Popat and T. A. Desai, Poly(ethylene glycol) interfaces: an approach for enhanced performance of microfluidic systems, Biosens. Bioelectron., 2004, 19(9), 1037–1044 CrossRef CAS.
- R. A. D'Sa and B. J. Meenan, Chemical Grafting of Poly(ethylene glycol) Methyl Ether Methacrylate onto Polymer Surfaces by Atmospheric Pressure Plasma Processing, Langmuir, 2010, 26(3), 1894–1903 CrossRef CAS.
- K. Tsougeni, D. Papageorgiou, A. Tserepi and E. Gogolides, “Smart” polymeric microfluidics fabricated by plasma processing: controlled wetting, capillary filling and hydrophobic valving, Lab Chip, 2010, 10(4), 462–469 RSC.
-
H. Yasuda, Plasma polymerization, Academic Press, 1985, p. 432 Search PubMed.
- P. K. Chu, J. Y. Chen, L. P. Wang and N. Huang, Plasma-surface modification of biomaterials, Mater. Sci. Eng., R, 2002, 36(5–6), 143–206 CrossRef.
-
L. S. McArthur and M. K. McLean, Surface Modification, in Encyclopedia of Biomaterials and Biomedical Engineering, ed. Gary E. Wnek and G. L. B., Marcel Dekker, New York, 2004, vol. 2, p. 1421 Search PubMed.
- H. Muguruma, Plasma-polymerized films for biosensors II, TrAC, Trends Anal. Chem., 2007, 26(5), 433–443 CrossRef CAS.
- S. Bouaidat, C. Berendsen, P. Thomsen, S. G. Petersen, A. Wolff and J. Jonsmann, Micro patterning of cell and protein non-adhesive plasma polymerized coatings for biochip applications, Lab Chip, 2004, 4(6), 632–637 RSC.
- O. D. Krishna, K. Kim and Y. Byun, Covalently grafted phospholipid monolayer on silicone catheter surface for reduction in platelet adhesion, Biomaterials, 2005, 26(34), 7115–7123 CrossRef CAS.
- H. Thissen, G. Johnson, P. G. Hartley, P. Kingshott and H. J. Griesser, Two-dimensional patterning of thin coatings for the control of tissue outgrowth, Biomaterials, 2006, 27(1), 35–43 CrossRef CAS.
- B. Gupta, C. Plummer, I. Bisson, P. Frey and J. Hilborn, Plasma-induced graft polymerization of acrylic acid onto poly(ethylene terephthalate) films: characterization and human, Biomaterials, 2002, 23(3), 863–871 CrossRef CAS.
- G. P. Löpez, B. D. Ratner, C. D. Tidwell, C. L. Haycox, R. J. Rapoza and T. A. Horbett, Glow discharge plasma deposition of tetraethylene glycol dimethyl ether for fouling-resistant biomaterial surfaces, J. Biomed. Mater. Res., 1992, 26(4), 415–439 CrossRef CAS.
- G. P. López and B. D. Ratner, Substrate temperature effects on film chemistry in plasma depositions of organics. II. Polymerizable precursors, J. Polym. Sci., Part A: Polym. Chem., 1992, 30(11), 2415–2425 CrossRef CAS.
- G. J. Leggett, B. D. Ratner and J. C. Vickerman, Characterization of plasma-deposited styrene films by XPS and static SIMS, Surf. Interface Anal., 1995, 23(1), 22–28 CAS.
- C. M. Chan, T. M. Ko and H. Hiraoka, Polymer surface modification by plasmas and photons, Surf. Sci. Rep., 1996, 24(1–2), 1–54 CrossRef CAS.
- M. Loughran, S. W. Tsai, K. Yokoyama and I. Karube, Simultaneous iso-electric focusing of proteins in a micro-fabricated capillary coated with hydrophobic and hydrophilic, Curr. Appl. Phys., 2003, 3(6), 495–499 CrossRef.
- S.-W. Tsai, M. Loughran, A. Hiratsuka, K. Yano and I. Karube, Application of plasma-polymerized films for isoelectric focusing of proteins in a capillary electrophoresis chip, Analyst, 2003, 128(3), 237–244 RSC.
- V. Barbier, M. Tatoulian, H. Li, F. Arefi-Khonsari, A. Ajdari and P. Tabeling, Stable modification of PDMS surface properties by plasma polymerization: application to the formation of double emulsions in microfluidic systems, Langmuir: The ACS Journal of Surfaces and Colloids, 2006, 22(12), 5230–5232 Search PubMed.
- M. Meincken, T. A. Berhane and P. E. Mallon, Tracking the hydrophobicity recovery of PDMS compounds using the adhesive force determined by AFM force distance, Polymer, 2005, 46(1), 203–208 CrossRef CAS.
- Z. Jinwen, E. Amanda Vera and V. Nicolas Hans, Recent developments in PDMS surface modification for microfluidic devices, Electrophoresis, 2009, 31(1), 2–16.
- I. Wong and C.-M. Ho, Surface molecular property modifications for poly(dimethylsiloxane) (PDMS) based microfluidic devices, Microfluid. Nanofluid., 2009, 7(3), 291–306 CrossRef CAS.
- J. Zhou, H. Yan, K. Ren, W. Dai and H. Wu, Convenient Method for Modifying Poly(dimethylsiloxane) with Poly(ethylene glycol) in Microfluidics, Anal. Chem., 2009, 81(16), 6627–6632 CrossRef CAS.
- X. Yan, Y. Xiao-Dong, X. Jing-Juan and C. Hong-Yuan, Bulk modification of PDMS microchips by an amphiphilic copolymer, Electrophoresis, 2007, 28(18), 3302–3307 CrossRef CAS.
- M. Salim, G. Mishra, G. J. S. Fowler, B. O'Sullivan, P. C. Wright and S. L. McArthur, Non-fouling microfluidic chip produced by radio frequency tetraglyme plasma deposition, Lab Chip, 2007, 7(4), 523–525 RSC.
- M. Kadowaki, H. Yoshizawa, S. Mori and M. Suzuki, Plasma CVD on the inner surface of a microchannel, Thin Solid Films, 2006, 506–507, 123–127 CrossRef CAS.
- H. Yoshiki, A. Oki, H. Ogawa and Y. Horiike, Inner wall modification of a poly(ethylene terephthalate) (PET) capillary by 13.56 MHz capacitively coupled microplasma, Thin Solid Films, 2002, 407(1–2), 156–162 CrossRef CAS.
- H. Yoshiki, Thin film coatings on capillary inner walls by microplasma, Vacuum, 2009, 84(5), 559–563 CrossRef CAS.
- J. L. Lauer, J. L. Shohet, R. M. Albrecht, C. Pratoomtong, R. Murugesan, S. Esnault, J. S. Malter, U. H. von Andrian, R. D. Bathke and S. B. Shohet, Reduced adhesion of human blood platelets to polyethylene tubing by microplasma surface modification, J. Appl. Phys., 2004, 96(8), 4539–4546 CrossRef CAS.
- H. Yoshiki, A. Oki, H. Ogawa and Y. Horiike, Generation of a capacitively coupled microplasma and its application to the inner-wall modification of a poly(ethylene terephthalate) capillary, J. Vac. Sci. Technol., A, 2002, 20(1), 24–29 CrossRef CAS.
- S. Chen, J. Zheng, L. Li and S. Jiang, Strong Resistance of Phosphorylcholine Self-Assembled Monolayers to Protein Adsorption: Insights into Nonfouling Properties of Zwitterionic Materials, J. Am. Chem. Soc., 2005, 127(41), 14473–14478 CrossRef CAS.
- S. L. McArthur, K. M. McLean, P. Kingshott, H. A. W. St John, R. C. Chatelier and H. J. Griesser, Effect of polysaccharide structure on protein adsorption, Colloids Surf., B, 2000, 17, 37–48 CrossRef CAS.
- P. Kingshott and H. J. Griesser, Surfaces that resist bioadhesion, Curr. Opin. Solid State Mater. Sci., 1999, 4(4), 403–412 CrossRef.
- Z. Ademovic, B. Holst, R. Kahn, I. Jørring, T. Brevig, J. Wei, X. Hou, B. Winter-Jensen and P. Kingshott, The method of surface PEGylation influences leukocyte adhesion and activation, J. Mater. Sci.: Mater. Med., 2006, 17(3), 203–211 CrossRef CAS.
- S. B. Brueggemeier, D. Wu, S. J. Kron and S. P. Palecek, Protein–acrylamide copolymer hydrogels for array-based detection of tyrosine kinase activity from cell lysates, Biomacromolecules, 2005, 6(5), 2765–2775 CrossRef CAS.
- S. Pillai, A. Arpanaei, R. L. Meyer, V. Birkedal, L. Gram, F. Besenbacher and P. Kingshott, Preventing Protein Adsorption from a Range of Surfaces Using an Aqueous Fish Protein Extract, Biomacromolecules, 2009, 10(10), 2759–2766 CrossRef CAS.
- Y. Xu, M. Takai and K. Ishihara, Phospholipid Polymer Biointerfaces for Lab-on-a-Chip Devices, Ann. Biomed. Eng., 2010, 38(6), 1938–1953 CrossRef.
- K. Jang, K. Sato, K. Mawatari, T. Konno, K. Ishihara and T. Kitamori, Surface modification by 2-methacryloyloxyethyl phosphorylcholine coupled to a photolabile linker for cell micropatterning, Biomaterials, 2009, 30(7), 1413–1420 CrossRef CAS.
-
J. M. Harris, Introduction to Biotechnical and Biomedical Applications of Poly(Ethylene Glycol), in Poly(ethylene glycol) Chemistry, ed. J. M. Harris, Plenum press, 1992, p. 1 Search PubMed.
-
N. B. Graham and M. E. McNeill, Poly(ethylene oxide): Applications in Drug Delivery, in Polymeric Materials Encyclopedia, ed. J. C. Salamone, CRC Press, New York, 1996, vol. 8, p. 6042 Search PubMed.
- N. Iki and E. S. Yeung, Non-bonded poly(ethylene oxide) polymer-coated column for protein separation by capillary electrophoresis, J. Chromatogr., A, 1996, 731(1–2), 273–282 CrossRef CAS.
- L. Li, S. Chen, J. Zheng, B. D. Ratner and S. Jiang, Protein Adsorption on Oligo(ethylene glycol)-Terminated Alkanethiolate Self-Assembled Monolayers: The Molecular Basis for Nonfouling Behavior, J. Phys. Chem. B, 2005, 109(7), 2934–2941 CrossRef CAS.
- P. Harder, M. Grunze, R. Dahint, G. M. Whitesides and P. E. Laibinis, Molecular Conformation in Oligo(ethylene glycol)-Terminated Self-Assembled Monolayers on Gold and Silver Surfaces Determines Their Ability To Resist Protein Adsorption, J. Phys. Chem. B, 1998, 102(2), 426–436 CrossRef.
- P. Kingshott, J. Wei, D. Bagge-Ravn, N. Gadegaard and L. Gram, Covalent Attachment of Poly(ethylene glycol) to Surfaces, Critical for Reducing Bacterial Adhesion, Langmuir, 2003, 19(17), 6912–6921 CrossRef CAS.
- P. Kingshott, H. Thissen and H. J. Griesser, Effects of cloud-point grafting, chain length, and density of PEG layers on competitive adsorption of ocular proteins, Biomaterials, 2002, 23(9), 2043–2056 CrossRef CAS.
- D. Beyer, W. Knoll, H. Ringsdorf, J. H. Wang, R. B. Timmons and P. Sluka, Reduced protein adsorption on plastics via direct plasma deposition of triethylene glycol monoallyl ether, J. Biomed. Mater. Res., 1997, 36(2), 181–189 CrossRef CAS.
- L. Q. Chu, W. Knoll and R. Forch, Pulsed Plasma Polymerized Di(ethylene glycol) Monovinyl Ether Coatings for Nonfouling Surfaces, Chem. Mater., 2006, 18(20), 4840–4844 CrossRef CAS.
- Y. V. Pan, T. C. McDevitt, T. K. Kim, D. Leach-Scampavia, P. S. Stayton, D. D. Denton and B. D. Ratner, Micro-Scale Cell Patterning on Nonfouling Plasma Polymerized Tetraglyme Coatings by Protein Microcontact Printing, Plasmas Polym., 2002, 7(2), 171–183 CrossRef CAS.
- M. Shen, Y. V. Pan, M. S. Wagner, K. D. Hauch, D. G. Castner, B. D. Ratner and T. A. Horbett, Inhibition of monocyte adhesion and fibrinogen adsorption on glow discharge plasma deposited tetraethylene glycol dimethyl ether, J. Biomater. Sci., Polym. Ed., 2001, 12(9), 961–978 CrossRef CAS.
- M. Shen, M. S. Wagner, D. G. Castner, B. D. Ratner and T. A. Horbett, Multivariate Surface Analysis of Plasma-Deposited Tetraglyme for Reduction of Protein Adsorption and Monocyte
Adhesion, Langmuir, 2003, 19(5), 1692–1699 CrossRef CAS.
- Y. J. Wu, R. B. Timmons, J. S. Jen and F. E. Molock, Non-fouling surfaces produced by gas phase pulsed plasma polymerization of an ultra low molecular weight ethylene oxide, Colloids Surf., B, 2000, 18(3–4), 235–248 CrossRef CAS.
-
C.-G. Golander, J. N. H., Kap Lim, P. Claesson, P. Stenius and J. D. Andrade, Properties of Immobilized PEG Films and the Interaction with Proteins: Experiments and Modeling, in Poly(ethylene glycol) Chemistry, ed. J. M. Harris, Plenum Press, New York, 1992, p. 221 Search PubMed.
- S. I. Jeon, J. H. Lee, J. D. Andrade and P. G. De Gennes, Protein–surface interactions in the presence of polyethylene oxide: I. Simplified theory, J. Colloid Interface Sci., 1991, 142(1), 149–158 CrossRef CAS.
- Y. S. Lin, V. Hlady and C. G. Golander, The surface density gradient of grafted poly(ethylene glycol): Preparation, characterization and protein adsorption, Colloids Surf., B, 1994, 3(1–2), 49–62 CrossRef CAS.
- K. Holmberg, F. Tiberg, M. Malmsten and C. Brink, Grafting with hydrophilic polymer chains to prepare protein-resistant surfaces, Colloids Surf., A, 1997, 123–124, 297–306 CrossRef CAS.
- S. Pasche, J. Voros, H. J. Griesser, N. D. Spencer and M. Textor, Effects of Ionic Strength and Surface Charge on Protein Adsorption at PEGylated Surfaces, J. Phys. Chem. B, 2005, 109(37), 17545–17552 CrossRef CAS.
- I. Szleifer, Polymers and proteins: interactions at interfaces, Curr. Opin. Solid State Mater. Sci., 1997, 2(3), 337–344 CrossRef CAS.
- K. L. Prime and G. M. Whitesides, Self-assembled organic monolayers: model systems for studying adsorption of proteins at surfaces, Science, 1991, 252(5009), 1164–1167 CrossRef CAS.
-
K. Holmberg and G. Quash, Protein Adsorption in Solid-Phase Diagnostics, in Biopolymers at Interfaces, ed. M. Malmsten, Marcel Dekker, 2003, vol. 110, p. 750 Search PubMed.
- S. W. Lee and P. E. Laibinis, Protein-resistant coatings for glass and metal oxide surfaces derived from oligo(ethylene glycol)-terminated alkyltrichlorosilanes, Biomaterials, 1998, 19, 1669–1675 CrossRef CAS.
- F. Bretagnol, M. Lejeune, A. Papadopoulou-Bouraoui, M. Hasiwa, H. Rauscher, G. Ceccone, P. Colpo and F. Rossi, Fouling and non-fouling surfaces produced by plasma polymerization of ethylene oxide monomer, Acta Biomater., 2006, 2(2), 165–172 CrossRef.
- J. Zheng, L. Li, H.-K. Tsao, Y.-J. Sheng, S. Chen and S. Jiang, Strong Repulsive Forces between Protein and Oligo (Ethylene Glycol) Self-Assembled Monolayers: A Molecular Simulation Study, Biophys. J., 2005, 89(1), 158–166 CrossRef.
- K. L. Prime and G. M. Whitesides, Adsorption of proteins onto surfaces containing end-attached oligo(ethylene oxide): a model system using self-assembled monolayers, J. Am. Chem. Soc., 1993, 115(23), 10714–10721 CrossRef CAS.
- R. S. Kane, P. Deschatelets and G. M. Whitesides, Kosmotropes Form the Basis of Protein-Resistant Surfaces, Langmuir, 2003, 19(6), 2388–2391 CrossRef CAS.
-
M. Schaferling and D. Kambhampati, Protein microarray surface chemistry and coupling schemes, in Protein microarray technology, ed. D. Kamchampati, Wiley, Weinheim, 2003, pp. 11–36 Search PubMed.
- W.-C. Lee and K. H. Lee, Applications of affinity chromatography in proteomics, Anal. Biochem., 2004, 324(1), 1–10 CrossRef CAS.
- C. You, M. Bhagawati, A. Brecht and J. Piehler, Affinity capturing for targeting proteins into micro and nanostructures, Anal. Bioanal. Chem., 2009, 393(6–7), 1563–1570 CrossRef CAS.
- K. Jana, rcaron, enková and F. Frantiscaronek, Immobilized microfluidic enzymatic reactors, Electrophoresis, 2004, 25(21–22), 3550–3563 CrossRef CAS.
- B. Lucas, F. M. Maria, L. do L. Claudimir, A. Arturo and D. G. Carlos, Lab-on-a-Chip Biosensor for Glucose Based on a Packed Immobilized Enzyme Reactor, Electroanalysis, 2007, 19(23), 2451–2456 CrossRef.
-
S. Kim Shyong, Britcher, Leane, Kumar, Sunil, Griesser and J. Hans, Plasma methods for the generation of chemically reactive surfaces for biomolecule immobilization and cell colonization: A review, Wiley-VCH, Weinheim, Allemagne, 2006, pp. 392–418 Search PubMed.
- N. L. Anderson and N. G. Anderson, The human plasma proteome: history, character, and diagnostic prospects, Molecular & cellular proteomics: MCP, 2002, 1(11), 845–867 Search PubMed.
|
This journal is © The Royal Society of Chemistry 2011 |
Click here to see how this site uses Cookies. View our privacy policy here.