Aging of Plasma-Polymerized Allylamine Nanofilms and the

Oct 30, 2014 - ... Nanofilms and the Maintenance of Their Cell Adhesion Capacity ... of bioactive layers is the precondition for the application of im...
1 downloads 0 Views 8MB Size
Article pubs.acs.org/Langmuir

Aging of Plasma-Polymerized Allylamine Nanofilms and the Maintenance of Their Cell Adhesion Capacity Birgit Finke,*,† Henrike Rebl,‡ Frank Hempel,† Jan Schaf̈ er,† Klaus Liefeith,§ Klaus-Dieter Weltmann,† and J. Barbara Nebe‡ †

Leibniz-Institute for Plasma Science and Technology (INP), Felix-Hausdorff-Straße 2, D-17489 Greifswald, Germany Department of Cell Biology, University Medical Center Rostock, Schillingallee 69, D-18057 Rostock, Germany § Department of Biomaterials, Institute for Bioprocessing and Analytical Measurement Techniques e.V., Rosenhof, D-37308 Heilbad Heiligenstadt, Germany ‡

ABSTRACT: The long-term stability and γ-sterilisability of bioactive layers is the precondition for the application of implants. Thus, aging processes of a microwave deposited, plasma polymerized allylamine nanofilm (PPAAm) with positively charged amino groups were evaluated concerning physicochemical characteristics and cell adhesion capacity over the course of one year. XPS, FT-IR, surface free energy, and water contact angle measurements elucidated not only the oxidation of the PPAAm film due to atmospheric oxygen reacting with surface free radicals but also the influence of atmospheric moisture during sample storage in ambient air. Surprisingly, within 7 days 70% of the primary amino groups are lost and mostly converted into amides. A positive zeta-potential was verified for half a year and longer. Increasing polar surface groups and a water contact angle shift from 60° to 40° are further indications of altered surface properties. Nevertheless, MG-63 human osteoblastic cells adhered and spread out considerably on aged and additionally γ-sterilized PPAAm layers deposited on polished titanium alloys (Ti-6Al-4V_P). These cell-relevant characteristics were highly significant over the whole period of one year and may not be related to the existence of primary amino groups. Rather, the oxidation products, the chemical amide group, that is, seem to support the attachment of osteoblasts at all times up to one year. coupling possibilities.13 Gas discharge plasma techniques are very well suited to this purpose. Plasma processes can not only be used for the enhancement of bonding locations (hydroxyl groups), but also for the deposition of polymeric-like thin coatings on metals with the desired reactive functional groups like amino- or carboxyl groups on the surface.1,7 Simultaneously, a greater versatility for binding biomolecules using different immobilization chemistries via various linkers is possible.7,14,15 Klee et al.16 modified titanium by means of chemical vapor deposition (CVD)− polymerization of amino functionalized (2,2) paracyclophane. Friedrich et al.17 reported on polymer surface modifications with allylamine under chemical structure retention of the monomer and a relatively high amino group density (NH2/C = 18%) during pulsed plasma polymerization by radio frequency (RF) excitation. Kylian et al.18 followed a new strategy by depositing amino-rich thin films (NH2/C = 18%) using RF magnetron sputtering of nylon-6.6 in Ar/N2 or N 2/H2 discharges. A possible problem of plasma polymers is swelling and dissolution of unbound material in aqueous environ-

1. INTRODUCTION Titanium and its alloys serve as the material of choice for loadbearing, bone-contacting orthopedic and dental implants. To tailor interactions of the bone−titanium interface, a lot of topographical and chemical strategies are under experimental observation.1−3 In particular, biochemical surface modifications with biomimetic motifs promote biocompatibility, cell physiology, and the bonding of implants with the surrounding bone.4−6 The immediate occupation of an implant surface by osteoblasts is vitally important for the application in the host. A key step in this process is the cell attachment via adhesion receptors, named integrins. Therefore, immobilized proteins and peptides similar to the extracellular matrix (ECM; e.g., type I collagen, RGD peptides) that function as ligands for integrins are commonly used to design a cell attractive interface.7−10 The covalent but also the electrostatic bonding of biomolecules to the implant surface is the preferred approach. Conventionally, it is possible to apply the wet-chemical silanechemistry. The coupling of the amino group carrying silanes, e.g., 3-amino-propyltriethoxysilane (APTES), occurs via surface hydroxyl groups of the native titanium oxide layer.11,12 But properties of this silane-linker such as their deliquescence and their tendency to polymerize and to form island-like domains affect adherence adversely and give rise to the search for other © 2014 American Chemical Society

Received: May 21, 2014 Revised: October 27, 2014 Published: October 30, 2014 13914

dx.doi.org/10.1021/la5019778 | Langmuir 2014, 30, 13914−13924

Langmuir

Article

ments.17−19 But the selection of appropriate plasma parameters can lead to a good compromise between a sufficient number of functional groups and insolubility in water. But which plasma polymer is especially qualified for use as a cell adherent surface coating on titanium? Findings in cell biology concerning the net negative charge of eukaryotic cells open up new surface functionalization strategies. The matrix substance hyaluronan (HA) was found to be expressed by chondrocytes20−22 and osteoblasts14,22 and is characterized by a negative charge due to carboxyl groups of the glucuronic acid. Corresponding to the outermost oxide layer, clean titanium surfaces are slightly negatively charged.14,23 Therefore, a positively charged implant surface at physiological pH-value with NH2-groups24−26 as charge carriers should be applied to stimulate the electrostatic immobilization of HA, as our former studies involving plasma polymerized allylamine (PPAAm) pointed out.14,26−31 The following chemical equation describes the electrostatic relation between the two partners:

diameter) were used for the chemical functionalization, physicochemical analysis, and in vitro cell culture. The surface roughness (Ra) was determined with a surface profiler Dektak3ST (Veeco, USA). The radius of the standard diamond stylus was 2.5 μm, the stylus force 30 mg, and the scan length 8 mm. The measurements were carried out 10 times in different directions giving an average value of Ra = 0.19 ± 0.02 μm. 2.2. Deposition of Plasma-Polymerized Allylamine (PPAAm) Thin Films. A commercial microwave plasma reactor (2.45 GHz; V55G, Plasma Finish, Germany) was used for the deposition of PPAAm films with a thickness of about 50 nm.14 Substrates were positioned downstream, 9 cm under the coupling window. First the substrates were cleaned and activated by a continuous wave (cw) oxygen plasma (500 W, 50 Pa, 100 sccm O2/25 sccm Ar, 60 s) and, without breaking the vacuum, subsequently coated with a 50−100 nm thin layer of PPAAm using a pulsed low pressure microwave discharge plasma (500 W, 50 Pa, duty cycle DC 0.15, total pulse length 2 s, excitation pulse length 300 ms) with the monomer allylamine (H2C CHCH2NH2) (VWR International GmbH, Germany).14 A liquid handling system allowed exact dosing of allylamine by a calibrated needle valve (0.125 ± 0.009 mL/min). Argon was applied as a carrier gas (50 sccm). The samples were stored without sterilization or γsterilized (Co 60, minimal irradiation dose 25 kGy, Synergy Health Radeberg GmbH, Germany) at room temperature in closed petridishes in the dark and ambient air. The samples for all investigations were stored for 0, 1, 7, 30, 90, 180, and 360 d and for XPS analyses additionally for 600 d. 2.3. Surface Characterization. 2.3.1. X-ray Photoelectron Spectroscopy (XPS). The elemental chemical surface composition and chemical binding properties of the PPAAm thin films were determined by XPS at different time points up to 600 days. The Axis Ultra DLD electron spectrometer (Kratos, UK) runs with monochromatic Al Kα radiation (1486 eV; 150 W) and implemented charge neutralization. The spot size was 250 μm in diameter. Spectra were recorded at a pass energy of 80 eV for the determination of wide scans and chemical elemental composition and of 10 eV for highly resolved C 1s and N 1s peaks. Three spots in different positions on each sample were analyzed and averaged. In all cases the takeoff angle was 90°. Data acquisition and processing was performed with the “vision 2.1.3” software (operating software Kratos). Peak fitting was processed using the “Casa XPS” software version 2.14.dev29 (Casa Software Ltd., UK), with Gauss-Lorentz (30% Lorentz) distribution, Shirley baseline, and a fixed full width at half-maximum (fwhm) between 0.9 and 1.2 eV for the C 1s peak and 1.2−1.5 eV for the N 1s peak, respectively. The CC/CH component of the aliphatic C 1s peak was set to 285.0 eV. The other components of the C 1s peak were fixed to known values: amines (CNH) and secondary carboxyls (OCO) at 285.7 ± 0.1 eV; hydroxyls, ethers, imines, nitriles (CO, CN, CN) at 286.6 ± 0.2 eV; aldehydes, ketones, amides (CO, NCO) at 288.0 ± 0.3 eV; esters and carboxyls (CC(O)OX) at 289.2 ± 0.2 eV) and for labeled primary amino groups CF3 at 292.7 ± 0.2 eV.36,44 Chemical derivatization with trifluorobenzaldehyde (TFBA) (Alfa Aesar, Germany) at 40°C for 2 h over a saturated gas phase was used for the labeling of primary amino groups on the desired storage date. The density of the amino groups NH2/C was determined from the fluorine elemental fraction.37 2.3.2. Fourier Transform Infrared Reflection Absorption Spectroscopy (FT-IRRAS). The chemical composition and molecular structure of PPAAm thin films were analyzed by means of FTIRRAS (FT-IR Type: Spectrum One, PerkinElmer, Germany).38 The IRRAS unit of the FT-IR spectrometer in the sample compartment used parallel-polarized light at an incidence angle of 75°. To improve the sensitivity of the FT-IR measurement, Au films were deposited on a glass substrate as reference and then coated with PPAAm. In this way FT-IRRAS measurements of minimal PPAAm film thicknesses (∼30 nm) were possible. The spectra were obtained at the spectral resolution of 4 cm−1, with the number of scans at 32, in the wavenumber region 4000 to 600 cm−1. 2.3.3. Surface Free Energy, Contact Angle. The polar and disperse part of surface free energy was calculated from measurements of

PPAAm coatings on titanium prepared by microwave plasma have favorable physicochemical properties. Typically they are very thin (∼50 nm), adhere well, are homogeneous, are strongly cross-linked, are resistant to delamination, are mechanically stable, and are equipped with a sufficient density of positively charged amino groups.29,14 In more comprehensive studies different cell biological aspects associated with PPAAm on titanium surfaces, like the mobility of the cytoskeletally associated protein vinculin,27 the metabolic activity of osteoblasts, and their long-time adhesion on PPAAm28 as well as the relation between topography and chemistry of PPAAm coated and uncoated titanium surfaces with diverse surface roughnesses of their adjusted topography, were examined.32 Intramuscular implantation of PPAAmcoated test samples in rats revealed a reduced inflammation reaction compared to uncoated titanium in vivo.30 For the implementation of plasma polymer films (PPAAm) as a bioactive implant surface design it is indispensable to know more about the long term characteristics of the physicochemical layer properties.33−35 The alterations of the film composition (as detected by X-ray photoelectron spectroscopy (XPS)) and the characterization of the surface functional groups (by Fourier transform infrared reflection absorption spectroscopy (FT-IRRAS)) as well as the variation of surface free energy, water contact angle, zeta potential, and γ-sterilisability is of relevance. Therefore, the focus of this study is to identify (i) whether general compositional alterations of the PPAAm layer occur during a storage time of one year in ambient air, (ii) whether they occur after γ-sterilization, and (iii) whether both processes influence the long-time cell adhesive capacity of this nanolayer. To achieve an understanding of surface chemistry effects on cell adhesion and spreading, the chemical alterations have to be investigated using cell experiments precisely at distinct, combined time points over a span of up to 360 days.

2. MATERIALS AND METHODS 2.1. Substrates: Titanium Alloy and Reference Specimens. Polished titanium alloy disks (Ti-6Al-4V_P, cp, grade 2; 1 or 3 cm in 13915

dx.doi.org/10.1021/la5019778 | Langmuir 2014, 30, 13914−13924

Langmuir

Article

a thickness of ∼50 nm deposited on a Ti-6Al-4V_P surface.14 The polished Ti-6Al-4V_P substrates plus PPAAm film were very smooth with an average roughness of Ra = 0.25 ± 0.1 μm. The XPS detected only the elements carbon, nitrogen, and oxygen in the first 10 nm of the surface. The film was homogeneously distributed, and no Ti was observable.14 The elemental XPS analysis showed an N/C ratio of 27.2 ± 0.3%, which is near the theoretical N/C value (33.3%) of the precursor allylamine. The O/C ratio was 5.6 ± 0.4%, most probably due to an oxygen uptake during initial air contact immediately after the plasma process. Within one year’s time, N/C decreased by ∼8% to 15%, some nitrogen content is lost in the form of gaseous ammonia NH3, and O/C increased by ∼11% to 17%. The altered PPAAm composition, in particular the evolution of the N/C and O/C element ratios as a function of storage time, is summarized in Figure 1a. The NH2/C ratio of 2.5 ± 0.5%, determined by derivatization with TFBA after plasma deposition, was much lower than the ratio in the allylamine monomer, an indication for a highly cross-linked polymer-like network. However, the density of primary amino groups NH2/C changed drastically in the course of one year (Figure 1b). The total loss of primary amino groups in this time

contact angles with different liquids. Water, ethylene glycol, and methylene iodide contact angles were determined with the help of the contact angle measuring system OCA 30 (Data Physics Instruments GmbH, Germany) by the sessile drop method. The drop size was 0.5 μL. The software SCA20 was employed for device control and data analysis. The surface free energy was calculated using the methods of Owens et al.39,40 The measurements were performed at the latest 30 min after preparation and repeated three times. 2.3.4. Zeta-Potential. Existing surface charges can be estimated by determining the so-called zeta-potential. Streaming potential measurements were carried out for various pressures using the Electrokinetic Analyzer SurPass (A. PAAR K.G., Germany) to determine the streaming potentials dependent on the pressure. The zeta-potential was calculated according to the method of Fairbrother and Mastin.41 The analysis was performed in a 0.001 M potassium chloride (KCl) solution at pH 6. Poly(diallyldimethylammonium chloride) (PolyDadMac) was used as a reference surface with a positive zeta potential. 2.4. Cell Culture. The human osteoblast-like cell line MG-63 (ATCC, CRL-1427, LGC Promochem, Germany) was used for all cell experiments. The cells were cultured in 75 cm2 flasks in Dulbecco’s modified Eagle’s medium (DMEM, Invitrogen, USA) with 10% fetal calf serum (FCS Gold, PAA Laboratories GmbH, Austria) and 1% gentamicin (Ratiopharm GmbH, Germany) at 37 °C in a humidified atmosphere with 5% CO2. During the experiments cells were cultured serum-free on all Ti-6Al-4V_P samples to avoid any masking of the charged groups.14 2.5. Cell Adhesion. Suspended MG-63 cells in serum-free DMEM were seeded onto all Ti-6Al-4V_P samples for 5 and 10 min at a density of 1 × 105/specimen, and nonadherent cells in the supernatant were counted and analyzed by flow cytometry (FACSCalibur, BD Biosciences, Germany). For data acquisition the software CellQuest Pro 4.0.1 was used. Cell adhesion was then calculated in percent of the cell number at 0 min. 2.6. Cell Spreading. Human MG-63 osteoblasts were trypsinated and washed in phosphate buffer solution (PBS), and the cell membrane was stained with the red fluorescent linker PKH26 (PKH26 General Cell Linker Kit, Sigma-Aldrich, USA) for 5 min in suspension. The cells were then seeded onto all Ti-6Al-4V_P discs and cultured for 30 min, 60 min, and 24 h. After fixation with 4% paraformaldehyde (PFA, Merck, Germany) the cells were embedded with a coverslip. Spreading (cell area in μm2) of 40 cells/specimen was measured using the software “area measurement” of the confocal microscope LSM 410 (Carl Zeiss, Germany), equipped with an argonion laser and a 63× water objective (1.25/0.17, Carl Zeiss, Germany). The image size was 512 pixels × 512 pixels. 2.7. Scanning Electron Microscopy (SEM). The sample surfaces were investigated using an SEM DSM 960A (Carl Zeiss, Germany). For cell analyses, cells were grown on the samples for 30 min, 60 min, and 24 h, fixed with 4% glutaraldehyde (1 h), dehydrated through a graded series of acetone and dried in a critical point dryer (K 850, EMITECH Germany). Gold sputtering was performed with a coater (SCD 004, BAL-TEC, Balzers, Liechtenstein). 2.8. Statistics. Statistical analyses were performed with the software SPSS Vers. 15.0 for Windows (SPSS Inc., USA) using the Mann-Whitney U-Test and ANOVA posthoc Bonferroni test. Data were presented as mean; a probability value of p < 0.05 was considered significant.

3. RESULTS The change of the surface composition was studied by several analyzing techniques such as XPS, FT-IR, surface free energy, water contact angle, and zeta potential. These have different analytical depths in the range of the interfacial forces, e.g., XPS, surface free energy, and FT-IR, with 2−10 nm, 2−10 nm, and 200 nm−2 μm, respectively. 3.1. Physicochemical Surface Properties of PPAAm during Storage in Ambient Air. The starting point was a freshly prepared PPAAm film, deposition rate 3.1 nm/min, with

Figure 1. a, b: XPS element ratios N/C, O/C (a) and the amino group density NH2/C (b) as a function of storage time of PPAAm until 600 d. Note the theoretical composition N/C of the precursor allylamine (AAm), solid line at 33.3% (a), and the rapid decrease of primary amino groups within the first 7 days in ambient air (b). Repeated investigations (open circles) confirmed the curve progression of NH2/ C (black triangle) (b). The plotted curves serve as guide to the eye. 13916

dx.doi.org/10.1021/la5019778 | Langmuir 2014, 30, 13914−13924

Langmuir

Article

CO, NCO (288.0 ± 0.3 eV), and COOH bonds (289.2 ± 0.2 eV) in the first 15 days was followed by a slower progress over the time of one year reflecting the proceeding oxidation process and the reaction with air moisture. Highly resolved XPS N 1s spectra confirmed these results (Figure 2b). Different NC bonds led to a peak broadening between 397 and 402 eV and could not be distinguished from each other. Hence, this broad peak was formally divided into three parts CN (1) , CN (2) , and CNOH: CN (1) and CN (2) for miscellaneous covalent CN bonds such as CNH2, CN, CN, CNC, OCNH2, and OCNHR. After 360 storage days the intensity of the total peak is clearly reduced. The CN(1), CN(2), and CNOH peaks decrease in the same relation. Bonds at binding energies >402 eV characteristic for an oxidation of nitrogen, for instance, nitroso CNO, nitro CNO2, nitrite CONO, or nitrate CNO3 bonds, were not found.43 An oxidation of nitrogen did not take place. In addition to the figures above, Figure 3 shows an overview of the starting and end conditions in N/C and O/C and the

period was at about 90%. A relatively quick aging process entailed a loss of 70% of primary amino groups within the first 5−7 days and in a slower, subsiding process a loss of a further 20% within the following half year. The same curve progression could be reconfirmed by repeating the investigation (open circles in Figure 1b). The plotted curves serve as guide to the eye. The equations of the curves in Figure 1a,b were obtained by fitting the data with exponential functions under the assumption of first order kinetics. From these, the time constants t were calculated. Those values are very different: for N/C ∼110 days, O/C ∼30 days, and NH2/C 2−5 days. This means that the decay and uptake processes seem to be very complex with few correlations. Not only oxidation reactions of primary amines but also the inclusions of free radicals, which occur in parallel, have to be kept in mind.42 Because of this, the single exponential fit is a rather crude approximation. High resolution XPS C 1s spectra of PPAAm (Figure 2a) verified the existence and changes of different CC-, CH-,

Figure 3. Changes of N/C- and O/C-ratios as well as CX bonds from C 1s high resolution spectra as deposited and stored for one year in ambient air.

fitted CX bonds of C 1s high resolution measurements of a freshly prepared and an aged PPAAm coating after 360 days. It is confirmed that primary amino groups (CF3 groups) have been lost as well as CN, CO, and CN bonds decreased, whereas amide NCO, carbonyl CO, and carboxyl OCO bonds were found to increase in this time period. FT-IRRAS studies of the PPAAm thin films immediately after deposition (0 h) and after storage in ambient air over 7, 30, 90, and 360 days, respectively, are shown in Figure 4a. In particular, the FT-IRRAS spectrum of PPAAm after deposition (0 h) confirmed a high retention of the structural properties of the monomer allylamine H2CCHCH2NH2 for the deposition method used here. Basic features of the monomer structure like the two stretching vibration bands of primary amines ν-NH at about 3500−3000 cm−1, the symmetric/ asymmetric stretching vibrations ν-CH2,3 at 2980−2880 cm−1, and the deformation vibrations of amines, δ-NH at 1650−1510 cm−1, as well as the deformation vibration δ-CH2,3 at 1465− 1375 cm−1 were still dominant in the spectrum.44−46 But typical plasma effects were also observable, e.g., significantly

Figure 2. a, b: High resolution XPS analysis of the C 1s (a) and N 1s (b) spectra of PPAAm at different time points up to 360 d in ambient air. Note (a) the decreasing primary amino groups (CF3 bonds near 297.7 eV) and the increasing acid amide bonds (CO, NCO at 288.0 ± 0.3 eV) and carboxylic acid bonds (COOH at 289.2 ± 0.2 eV). The N 1s peak (b) is formally fitted by different peaks which are marked by CN(1), CN(2), and CNOH. CN(1), CN(2) for miscellaneous covalent CN bonds: CNH 2 , CN, CN, CNC, OCNH 2 , OCNHR after deposition and storage in ambient air for 360 d.

nitrogen-, and also oxygen-containing bonds at a specific date; at the day of deposition and after 7, 30, and 360 days of storage in ambient air. The bond areas of CN (285.7 ± 0.1 eV) and CO, CN (286.6 ± 0.2 eV) ran through a maximum in the first 30 days (not shown here) while the primary amino groups (CF3 bonds near 292.7 eV) decreased promptly within even fewer days. The initial increase of 13917

dx.doi.org/10.1021/la5019778 | Langmuir 2014, 30, 13914−13924

Langmuir

Article

Figure 4. a−d: FT-IRRAS spectra for (a) PPAAm after deposition (0 h) and storage in air up to 360 days. (b) The difference between the FTIRRAS reflection spectra after deposition and after 360 days storage shows the formation and degradation of characteristic molecular groups. (c) Analysis of the IRRAS reflection band at 1750−1500 cm−1 reveals peak broadening and shifting to higher wave numbers as well as growing substructures during aging. (d) The substructures of the IRRAS reflection band (360 d) at 1750−1500 cm−1 can be assigned to ν-CO (amide I) at 1655 cm−1, ν-CN (amide II) for primary amides at 1602 cm−1, and a further ν-CN band (amide II) at 1563 cm−1 for secondary amides.

and are especially pronounced after 360 days of storage in ambient air (Figure 4c). The band minima as well as the band edge were shifted to higher wave numbers under band broadening with storage time. An analysis of the substructures was carried out. After one year of storage in ambient air, characteristic amide bands (Figure 4d) were found. For all amides, the reflection of amide I is at 1655 cm−1 (ν-CO stretching vibrations). The wavenumber of the amide II band depends on the type of amide group present. The band at 1602 cm−1 can be assigned to the amide II vibration of primary amides and the band at 1563 cm−1 to the amide II band of secondary amides. The amide III band is found at 1240 cm−1 νs-CN stretching/δ-NH bending vibrations (Figure 4a).49−51 Changes of the surface free energy and water contact angle with time were further indications for the aging process. Both measurements showed a characteristic curve progression (Figure 5a,b) which was confirmed by threefold determination on PPAAm samples prepared in different runs. The polar part and likewise the total surface free energy of the PPAAm thin film decreased about 7 mN/m in the first 7 days followed by an increase until saturation at about 37 or 57 mN/m, respectively, after half a year of storage. In parallel, the disperse part of the surface free energy went through a little initial maximum and into a diminished level at about 20 mN/m after approximately 180 days. The water contact angle rose from 47° after the preparation process to 57° in the first 7 days

broadened, disappearing, or arising bands. Amino groups of the precursor were partially transformed into amide, imine, or nitrile functional groups by the plasma deposition process. An indication for that is the new band between 2300−2100 cm−1 associated with stretching vibrations of nitrile and ethine groups, ν-CN and ν-CC,47 as well as the imine group CN and ethene group CC at 1690−1650 cm−1. The comparison of the IRRAS reflection spectrum measured directly after deposition (0 h) with that taken after storage in ambient air (360 d) exhibits significant molecular changes in the PPAAm film (Figure 4b). The difference spectrum gives important information about the formation and degradation of molecular groups and the broadening and shifting of IRRAS reflection bands. Characteristic molecular changes can be identified: the formation of OH groups due to water vapor absorption, ν-OH stretching vibrations at 3700−3000 cm−1, and carbonyl groups ν-CO stretching vibrations at 1700− 1650 cm−1, an indication for arising acid amides. The reflection at 1465−1300 cm−1 can be assigned to deformation vibrations of δ-CH and OH groups. Furthermore, CN and CC triple bond structures decreased, indicating the loss of unsaturated structures.48 The formation of acid amides over time is characterized by overlapping and peak broadening of δ-NH deformation vibration bands and ν-CO stretching vibration bands between 1750 and 1500 cm−1. In the FT-IRRAS spectra band substructures can be observed in this range already after 7 days 13918

dx.doi.org/10.1021/la5019778 | Langmuir 2014, 30, 13914−13924

Langmuir

Article

Figure 6. Effect of γ-sterilization on surface free energy of PPAAm (unsterile and γ-sterilized) at days 7 and 360. Note the relatively low influence of the γ-sterilization at day 7 and the somewhat higher at day 360.

3.2. Cell Adhesion Capacity during the Aging Process of PPAAm Thin Films. Initial cell adhesion experiments revealed a significant increase of adherent cells on the PPAAmmodified Ti-6Al-4V_P surface. A more than 2.3-fold increase compared to untreated Ti-6Al-4V_P was measured for all samples at every time point within the first week of storage. Although investigations were performed under standardized cell culture conditions, e.g., the use of cell passages between 5 and 25, variations in the control have occurred over the long period of one year. However, it is clearly visible that cell adhesion on PPAAm is consistently increased at all time points and the results show cell adhesion of 88.4−95.9% and 80.1− 96.4% after 5 and 10 min, respectively (Figure 7a). Furthermore, it was found that the surface modification with PPAAm boosts the spreading (30 min, 60 min, 24 h) of osteoblasts (Figure 7b). This significant increase was detected for all tests over the whole period of one year. It was even more pronounced in the cell spreading phase at the time point 30 min, where a nearly 2.2-fold increase was observed at day 0. After 24 h the PPAAm layer led to a 1.6-fold increased cell area compared to untreated Ti-6Al-4V_P. This increase can also be seen at day 180: 2.0-fold after 30 min and 1.6-fold after 24 h. Even after 360 days of storage the PPAAm layer can boost the 30 min as well as 24 h cell spreading by a factor of 1.8 and 1.6, respectively. The cell area is still visibly larger at all time points after 360 days of storage. These results underline the superior effect of PPAAm in the first hours of cell attachment and spreading and also demonstrate the positive effect of PPAAm modification that is stable throughout a one year period. Analyses of the cell morphology demonstrate that the cells on PPAAm-modified Ti-6Al-4V_P spread on the surface already after a contact time of only 30 min (Figure 8). Cells on the uncoated Ti-6Al-4V_P surface appear rather rounded and only loosely bound with few filopodia. Overall, on PPAAm the cell area is highly increased, resulting in an outstanding cellsurface contact even on a PPAAm surface stored for 360 d. The differences in short time (30, 60 min) and long term spreading (24 h) between unsterile and γ-sterilized PPAAm surfaces on days 7 and 360 are minimal (Figure 9a,b). SEM images (Figure 10) confirm this observation. Thus, this

Figure 5. a, b: (a) Surface free energy of PPAAm with its polar and disperse parts determined up to 360 days (three measuring points/ sample, n = 3 independent experiments until 200 days). Note the decrease of polar groups and total surface free energy in the first 7 days followed by a further increase. (b) Water contact angle measurements of PPAAm determined up to 360 days (one measuring point/sample, n = 3 independent experiments until 200 days). Note the increase of the water contact angle in the first 7 days followed by a subsequent decrease in the WCA.

of storage and decreased subsequently to 37° by a rising number of polar hydrophilic groups in the first half of the year. The influence of γ-sterilization on the PPAAm film is relatively low (Figure 6). The measurement of the surface free energy and water contact angle (not shown) of the unsterile PPAAm surface was carried out immediately after return of the γ-sterilized samples at about the seventh day. Therefore, both starting contact angles are comparatively high at about 60° (compare Figure 5b). The difference between the unsterile and sterile PPAAm surface at day 360 is higher. More polar groups and thus a lower water contact angle are found at the sterile surface. Investigations of the zeta potential confirmed a positive surface charge of the PPAAm surface during storage in ambient air for a half year (2 days: 13.9 ± 1.2 mV; 20 days: 20.9 ± 4.8 mV; 200 days: 26.3 ± 0.5 mV). This slightly increasing positive zeta potential was unexpected first since the density of primary amino groups decreased very quickly in the first 7 days. But other positively charged N-functional groups such as imines, nitriles, and acid amides contribute as well to the overall surface charge in a relatively long time span of more than half a year. Negatively charged functional groups seem to have only little influence in the first time. A small negative surface charge of −17 mV could be determined after storage of 2 years and longer. 13919

dx.doi.org/10.1021/la5019778 | Langmuir 2014, 30, 13914−13924

Langmuir

Article

sterilization method is well suited for PPAAm and does not further alter the chemical characteristics of PPAAm.

4. DISCUSSION The reason for the progressive aging processes of deposited plasma polymers is based on the physics and chemistry of the plasma process itself. In general, plasma polymerization processes at low pressure are due to radical reactions on the surface. Neutral transient reactive species from precursor fragmentation in the plasma volume are adsorbed on the surface and polymerize. Plasma−surface interactions and VUV/ UV photons contribute to the film growth.52 Accordingly, the plasma polymerization is a plasma-initiated and -supported process leading to a random hydrocarbon network of amorphous organic thin films with functional groups from the precursor molecule and amounts of free radicals remaining in the film and on the surface, bound to carbon or motile. Especially tertiary carbon radicals are long-lasting, living for many weeks and therefore stable for a long time. 53 Furthermore, the use of pulsed plasmas in turn enhances the pure chemical chain propagation, resulting in chemically betterdefined polymer structures. The radical chain reaction is terminated by disproportionation, a radical−radical recombination, or a reaction with oxygen molecules from the rest of the gas. Therefore, pulsed plasmas with long plasma-off periods and short plasma pulses offer a good compromise to produce chemically formed polymer structures with a minimum of irregularities and high chemical structure retention.44,54 After finishing the plasma process and sample transfer in ambient air, the functionalized polymer surface is promptly overlaid by the reaction of plasma-generated free radicals with oxygen from air, by the so-called post-plasma oxidation or autooxidation processes.55 The initial bound carbon radicals react with atmospheric oxygen to peroxy-radicals. Different intermediary and end products such as carbonyls, carboxyls, or even new carbon radicals can be formed in a series of reactions.55−57 Such processes are kinetically favored and, thus, very fast, because the oxygen molecule can be also considered as a biradical.53 Furthermore, it is assumed that motile radicals localize at the carbon atom anchor of the electron-rich primary amino group and promote their oxidation to acid amide groups and partially to imides.57 Figures 1b and 2a show the rapid conversion of primary amino groups into acid amides within 5− 7 days after deposition. XPS and FT-IR confirm both the formation of acid amides and carboxylic groups. An oxidation of nitrogen could not be found (Figure 2b). Stable oximes can also be detected. But also hydrolysis takes place during the storage of PPAAm in ambient air. Atmospheric moisture influences reactions of the existing and arising surface functional groups (possible reaction pathways in Figure 11a,b). Imines react to ketones or aldehydes plus ammonia (nitrogen loss) and nitriles react to acid amides. In contrast, acid amides are hydrolyzed to carboxylic acid groups and ammonia only under very harsh reaction conditions with concentrated acids or bases at temperatures over 100 °C. For this reason this reaction pathway seems to be irrelevant in our case. The nitrogen loss remains limited (Figure 1); only imines seem to contribute (Figures 3, 11a). CO, NCO, and COOH bonds increase slowly during the storage in ambient air in the course of one year (Figures 2a and 3). Acid amides are available in sufficient amounts during one year storage (Figures 2a, 3, and 4b−d), as physicochemical analysis indicates, displaying, e.g., a

Figure 7. (a) Initial cell adhesion after 5 and 10 min is significantly enhanced on Ti-6Al-4V_P surfaces biofunctionalized with PPAAm. Note the consistently good cell adhesive property of the plasma polymer film over the whole period of one year/360 days (n = 4 independent experiments, Mann-Whitney-U-Test, *p < 0.05). (b) Short time (30, 60 min) and long-term spreading (24 h) of MG-63 cells on PPAAm-modified surfaces. Note that the cell area is significantly enhanced over the whole period of 360 days (n = 40 cells, ANOVA posthoc Bonferroni ***p < 0.005; **p < 0.01). 13920

dx.doi.org/10.1021/la5019778 | Langmuir 2014, 30, 13914−13924

Langmuir

Article

Figure 8. Morphology (SEM) of MG-63 cells (after 30 min) on untreated Ti-6Al-4V_P (left) and a PPAAm surface (right), both stored for 360 d. Cells on the untreated Ti-6Al-4V_P are attached to the surface with only a few filopodes. In contrast, it is clearly visible that the cells spread far better on the PPAAm-samples, even though the samples had been stored for one year. The cell area is highly increased, resulting in outstanding cell− surface contact (magnification 2000×, 45°, DSM 960A, Carl Zeiss; bar = 10 μm).

oxidation productthe chemical amide groupobviously supports the attachment of osteoblasts at any time within the one-year period. The first seven days of storage in ambient air are very important for the PPAAm surface alteration. Not only were a decrease of polar surface groups and an increasing water contact angle and thus more hydrophobic surface properties observed in this time period, but also a rapid loss of primary amino groups and a change of the functionalities were observed. The reason for the increasing hydrophobicity during the initial phase of aging was explained by van Os58 and Griesser et al.;59 they cited the reorganization and restructuring of the PPAAm surface as the dominating process. The hydrophilic amino groups move away at a certain degree from the surface into the bulk, leading to a more hydrophobic surface with a higher density of hydrocarbon segments. Different processes, e.g., oxidation and hydrolysis, contribute in the first few days to the overall aging process (Figures 1 and 11). The formation of a higher fraction of polar groups by further oxidation leads to more hydrophilic surface properties with a lower water contact angle. The short time spreading for 30 and 60 min of MG-63 cells (Figure 7b) seems to be affected in the first 7 days by these PPAAm surface alterations. However, the values of the 24 h spreading leveled and the adhesion experiments also showed no influence of this effect. Nevertheless, the existing nitrogen functional groups can still counterbalance the negatively charged oxygen functional moieties of the surface. Zeta potential investigations confirmed a positive surface charge in the first half of the year and longer. Cell culture investigations confirmed the predominance of PPAAm thin film coated surfaces over the whole time compared to pure Ti-6Al-4V_P. Our experiments revealed that an additional sterilization process of PPAAm with γirradiation, necessary for medical applications, did not influence the improved spreading behavior of osteoblasts. It could be shown that MG-63 human osteoblastic cells adhered and spread out considerably also on aged and γ-sterilized PPAAm layers deposited on Ti-6Al-4V_P. The improvement of the cell adhesion capacity of PPAAm in comparison to Ti-6Al-4V_P could be preserved over the one-year period. Therefore, PPAAm thin films are very well-suitable for the finishing of cell adhesive implant surfaces.

Figure 9. Short time (30, 60 min) and long-term spreading (24 h) of MG-63 cells on untreated Ti-6Al-4V_P and PPAAm treated Ti-6Al-4V_P, unsterile and γ-sterilized, after about 7 days (a) and 360 days (b) of storage in ambient air. Note that the cell area is significantly enhanced over the whole period of 360 days and not influenced by the sterilization process (n = 40 cells, ANOVA posthoc Bonferroni ***p < 0.0001).

decrease and conversion of nitriles into acid amides in the FTIR spectra (Figures 4a,b and 11b). Chemically unstable functional groups are converted into stable amides and also oximes. These findings are consistent with conclusions reached earlier by Griesser et al.57 The zeta-potential is positive for half a year and longer. MG-63 human osteoblastic cells show considerably enhanced adhesion and spreading also on these “aged” PPAAm films, compared with untreated Ti-6Al-4V_P. These cell-relevant characteristics are highly significant over the whole period of one year and do not seem to be related to the existence of primary amino groups alone. Instead, their 13921

dx.doi.org/10.1021/la5019778 | Langmuir 2014, 30, 13914−13924

Langmuir

Article

Figure 10. Morphology (SEM) of MG-63 cells (24 h culture) on unsterile and γ-sterilized PPAAm surfaces after 360 days of storage. The cell area is still highly increased resulting in an outstanding cell−surface contact (magnification 1000×, DSM 960A, Carl Zeiss, bar = 20 μm).

PPAAm thin films it was important to demonstrate that the outstanding cell−surface contact of MG-63 osteoblasts (adhesion and spreading) was maintained even after storage of the PPAAm modified Ti-6Al-4V_P samples for up to one year and after γ-sterilization.



AUTHOR INFORMATION

Corresponding Author

*(B.F.) Tel.: +49 3834 554433. Fax: +49 3834 554301. E-mail: fi[email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the Federal Ministry of Education and Research (Grants 13N9779, 13N11183, and 13N11188, Campus PlasmaMed) as well as by the program TEAM of Mecklenburg-Vorpommern and the Helmholtz Association in Germany (UR 0402210, VH-MV1). We appreciate the technical support of R. Ihrke, Dr. A. Quade, U. Kellner, G. Friedrichs, U. Lindemann, D. Jasinski, and A. Schella from INP Greifswald, S. Staehlke from the University of Rostock, and H. Rothe from IBA Heilbad Heiligenstadt.



Figure 11. Possible reaction pathways: (a) hydrolysis of imines to carbonyl functional groups (ketones, aldehydes) and ammonia; (b) hydrolysis of nitriles to acid amides, see ref 60.

REFERENCES

(1) Puleo, D. A.; Nanci, A. Understanding and controlling the boneimplant interface. Biomaterials 1999, 20, 2311−2321. (2) Sul, Y. T.; Kwon, D. H.; Kang, B. S.; Oh, S. J.; Johansson, C. Experimental evidence for interfacial biochemical bonding in osseointegrated titanium implants. Clin. Oral Implants Res. 2013, Suppl. A 100, 8−19. (3) Coad, B. R.; Scholz, T.; Vasilev, K.; Hayball, J. D.; Short, R. D.; Griesser, H. J. Functionality of proteins bound to plasma polymer surfaces. ACS Appl. Mater. Interfaces 2012, 4, 2455−2463. (4) Sverzut, A. T.; Crippa, G. E.; Morra, M.; de Oliveira, P. T.; Beloti, M. M.; Rosa, A. L. Effects of type I collagen coating on titanium osseointegration: histomorphometric, cellular and molecular analyses. Biomed. Mater. 2012, 7, 035007. (5) Raynor, J. E.; Capadona, J. R.; Collard, D. M.; Petrie, T. A.; García, A. J. Polymer brushes and self-assembled monolayers: Versatile platforms to control cell adhesion to biomaterials. Biointerphases 2009, 4, FA3−16. (6) Schulz, M. C.; Korn, P.; Stadlinger, B.; Range, U.; Möller, S.; Becher, J.; Schnabelrauch, M.; Mai, R.; Scharnweber, D.; Eckelt, U.; Hintze, V. Coating with artificial matrices from collagen and sulfated

5. CONCLUSIONS Bioactive coatings on implant surfaces enhance the close connection between cells and material surface. Plasma polymer films in particular are very well suited for the improvement of cell-adhesive properties of titanium alloy implants, since they provide the required surface functionalities. In the present study nitrogen-rich films were deposited by plasma polymerization of allylamine in microwave plasma at low pressure and their properties studied over the time of one year and longer. It could be shown that such PPAAm coatings fulfill different requirements. Their long-term storage and the stability after γsterilization were analyzed by physicochemical methods and parallel cell physiological investigations within a period of one year. Surface alterations of PPAAm could be identified by postplasma oxidation and hydrolysis. A positive zeta-potential was verified for half a year and longer. For future applications of 13922

dx.doi.org/10.1021/la5019778 | Langmuir 2014, 30, 13914−13924

Langmuir

Article

hyaluronan influences the osseointegration of dental implants. J. Mater. Sci.: Mater. Med. 2014, 25, 247−258. (7) Morra, M.; Cassinelli, C.; Cascardo, G.; Calahan, P.; Calahan, L.; Fini, M.; Giardino, R. Surface engineering of titanium by collagen immobilization. Surface characterization and in vitro and in vivo studies. Biomaterials 2003, 24, 4639−4654. (8) Petrie, T. A.; Raynor, J. E.; Reyes, C. D.; Burns, K. L.; Collard, D. M.; García, A. J. The effect of integrin-specific bioactive coatings on tissue healing and implant osseointegration. Biomaterials 2008, 29, 2849−2857. (9) Schuler, M.; Owen, G. R.; Hamilton, D. W.; De Wilde, M.; Textor, M.; Brunette, D. M.; Tosatti, S. G. P. Biomimetic modification of titanium dental implant model surfaces using the RGDSP-peptide sequence: A cell morphology study. Biomaterials 2006, 27, 4003− 4015. (10) Petrie, T. A.; Capadona, J. R.; Reyes, C. D.; Garcia, A. J. Integrin specificity and enhanced cellular activities associated with surfaces presenting a recombinant fibronectin fragment compared to RGD supports. Biomaterials 2006, 27, 5459−5470. (11) Puleo, D. A. Activity of enzyme immobilized on silanized CoCr-Mo. J. Biomed. Mater. Res. 1995, 29, 951−957. (12) Puleo, D. A. Retention of enzymatic activity immobilized on silanized Co-Cr-Mo and Ti-6Al-4V. J. Biomed. Mater. Res. 1997, 37, 222−228. (13) Falconnet, D.; Csucs, G.; Grandin, H. M.; Textor, M. Surface engineering approaches to micropattern surfaces for cell-based assays. Biomaterials 2006, 27, 3044−3063. (14) Finke, B.; Luethen, F.; Schroeder, K.; Mueller, P. D.; Bergemann, C.; Frant, M.; Ohl, A.; Nebe, B. J. The effect of positively charged plasma polymerization on initial osteoblastic focal adhesion on titanium surfaces. Biomaterials 2007, 28, 4521−4534. (15) Hamerli, P.; Weigel, T.; Groth, T.; Paul, D. Surface properties of and cell adhesion onto allylamine-plasma-coated polyethylentherephthalate membranes. Biomaterials 2003, 24, 3989−3999. (16) Klee, D.; Böing, J.; Höcker, H. Surface modification of titanium to improve interface compatibility. Materialwiss. Werkstofftech. 2004, 35, 186−191. (17) Friedrich, J.; Kühn, G.; Mix, R.; Fritz, A.; Schönhals, A. Polymer surface modification with monofunctional groups of variable types and densities. J. Adhes. Sci. Technol. 2003, 17, 1591−1617. (18) Kylian, O.; Hanus, J.; Choukourov, A.; Kousal, J.; Slavinska, D.; Biederman, H. Deposition of amino-rich thin films by RF magnetron sputtering of nylon. J. Phys. D: Appl. Phys. 2009, 42, 142001. (19) Chu, L.; Knoll, W.; Förch, R. Biologically multifunctional surfaces using plasma polymerization methods. Plasma Process. Polym. 2006, 3, 498−505. (20) Zimmermann, E.; Geiger, B.; Addadi, L. Initial stages of cellmatrix adhesion can be mediated and modulated by cell-surface hyaluronan. Biophys. J. 2002, 82, 1848−1857. (21) Cohen, M.; Kam, Z.; Addadi, L.; Geiger, B. Dynamic study of the transition from hyaluronan- to integrin-mediated adhesion in chondrocytes. EMBO J. 2006, 25, 302−311. (22) Nebe, J. G. B.; Lüthen, F. Integrin- and Hyaluronan-mediated cell adhesion on titanium. In Metallic Biomaterial Interfaces; Breme, J., Kirkpatrick, C. J., Thull, R., Eds.; Wiley-VCH: Weinheim, 2008; pp 179−182. (23) Roessler, S.; Zimmermann, R.; Scharnweber, D.; Werner, C.; Worch, H. Characterization of oxide layers on TiAl6V4 and titanium by streaming potential and streaming current measurements. Colloids Surf., B 2002, 26, 387−395. (24) Faucheux, N.; Tzoneva, R.; Nagel, M. D.; Groth, T. The dependence of fibrillar adhesions in human fibroblasts on substratum chemistry. Biomaterials 2006, 27, 234−245. (25) Schweikl, H.; Müller, R.; Englert, C.; Hiller, K. A.; Kujat, R.; Nerlich, M.; Schmalz, G. Proliferation of osteoblasts and fibroblasts on model surfaces of varying roughness and surface chemistry. J. Mater. Sci. Mater. Med. 2007, 18, 1895−1905. (26) Nebe, B.; Finke, B.; Lüthen, F.; Bergemann, C.; Schröder, K.; Rychly, J.; Liefeith, K.; Ohl, A. Improved initial osteoblast’s functions

on aminofunctionalized titanium surfaces. Biomol. Eng. 2007, 24, 447− 454. (27) Rebl, H.; Finke, B.; Rychly, J.; Schroeder, K.; Nebe, J. B. Positively charged material surfaces generated by plasma polymerised allylamine enhance vinculin mobility in vital human osteoblasts. Adv. Eng. Mater. Adv. Biomater. 2010, 12, B356−B364. (28) Rebl, H.; Finke, B.; Schroeder, K.; Nebe, J. B. Time dependent metabolic activity and adhesion of human osteoblasts on sensor chips with an allylamine-plasma polymer nano coating. Int. J. Artif. Organs 2010, 33, 738−748. (29) Fritsche, A.; Luethen, F.; Lembke, U.; Finke, B.; Zietz, C.; Rychly, J.; Mittelmeier, W.; Bader, R. Measuring bone cell adhesion on implant surfaces using a spinning disc device. Materialwiss. Werkstofftech. 2010, 41, 83−88. (30) Hoene, A.; Walschus, U.; Patrzyk, M.; Finke, B.; Lucke, S.; Nebe, B. J.; Schroeder, K.; Ohl, A.; Schlosser, M. In vivo investigation of the inflammatory response against allylamine plasma polymer coated titanium implants in a rat animal model. Acta Biomater. 2010, 6, 676−683. (31) Finke, B.; Hempel, F.; Testrich, H.; Artemenko, A.; Rebl, H.; Kylián, O.; Meichsner, J.; Biederman, H.; Nebe, B.; Weltmann, K.-D.; Schröder, K. Plasma processes for cell-adhesive titanium surfaces based on nitrogen-containing coatings. Surf. Coat. Technol. 2011, 205, S520− S524. (32) Rebl, H.; Finke, B.; Lange, R.; Weltmann, K.-D.; Nebe, J. B. Impact of plasma chemistry versus titanium surface topography on osteoblast orientation. Acta Biomaterial. 2012, 8, 3840−3851. (33) Whittle, J. D.; Short, R. D.; Douglas, C. W. I.; Davies, J. Differences in the aging of allyl alcohol, acrylic acid, allylamine, and octa-1,7-diene plasma polymers as studied by X-ray photoelectron spectroscopy. Chem. Mater. 2000, 12, 2664−2671. (34) Girard-Lauriault, P.-L.; Dietrich, P. M.; Gross, T.; Wirth, T.; Unger, W. E. S. Chemical characterization of the long-term ageing of nitrogen-rich plasma polymer films under various ambient conditions. Plasma Process. Polym. 2013, 10, 388−395. (35) Tarasova, A.; Hamilton-Brown, P.; Gengenbach, T.; Griesser, H. J.; Meagher, L. Colloid probe AFM and XPS study of time-dependent aging of amine plasma polymer coatings in aqueous media. Plasma Process. Polym. 2008, 5, 175−185. (36) Beamson, G.; Briggs, D. High resolution XPS of organic polymers; The scienta ESCA 300 database; Wiley: Chichester, New York, 1992. (37) Everheart, D. E.; Reilly, C. N. Chemical derivatization in electron-spectroscopy for chemical analysis of surface functional groups introduced on low-density polyethylene film. Anal. Chem. 1981, 53, 665−676. (38) Greenler, R. G. Infrared study of adsorbed molecules on metal surfaces by reflection techniques. J. Chem. Phys. 1966, 44, 310−315. (39) Owens, D. K.; Wendt, R. C. Estimation of the surface free energy of polymers. J. Appl. Polym. Sci. 1969, 13, 1741−1747. (40) Rabel, W. Einige Aspekte der Benetzungstheorie und ihre Anwendungen auf die Untersuchung und Veränderung der Oberflächeneigenschaften von Polymeren. Farbe Lack 1971, 77, 997−1005. (41) Fairbrother, H.; Mastin, J. Studies in electro-endosmosis. J. Chem. Soc. 1924, 75, 2318. (42) St-Georges-Robillard, A.; Ruiz, J. C.; Petit, A.; Wang, H. T.; Mwale, F.; Elkin, B.; Oehr, C.; Lerouge, S.; Wertheimer, M. R. Adhesion of U-937 monocytes on different amine-functionalised polymer surfaces. Plasma Process. Polym. 2012, 9, 243−252. (43) Wilken, R.; Holländer, A.; Behnisch, J. Surface radical analysis on plasma-treated polymers. Surf. Coat. Technol. 1999, 116, 991−995. (44) Schröder, K.; Finke, B.; Ohl, A.; Lüthen, F.; Bergemann, C.; Nebe, J. B.; Rychly, J.; Walschus, U.; Schlosser, M.; Liefeith, K.; Neumann, H.-G.; Weltmann, K.-D. Capability of Differently Charged Plasma Polymer Coatings for Control of Tissue Interactions with Titanium Surfaces. J. Adhes. Sci. Technol. 2010, 24, 1191−1205. (45) Kim, J.; Jung, D.; Park, Y.; Kim, Y.; Moon, D. W.; Lee, T. G. Quantitative analysis of surface amine groups on plasma-polymerized ethylenediamine films using UV−visible spectroscopy compared to 13923

dx.doi.org/10.1021/la5019778 | Langmuir 2014, 30, 13914−13924

Langmuir

Article

chemical derivatization with FT-IR spectroscopy, XPS and TOF-SIMS. Appl. Surf. Sci. 2007, 253, 4112−4118. (46) Krishnamurthy, V.; Kamel, I. L. Analysis of plasma polymerization of allylamine by FT-IR. J. Polymer Sci. A: Polym. Chem. 1989, 27, 1211−1224. (47) Choukourov, A.; Biederman, H.; Kholodkov, I.; Slavinska, D.; Trchova, M.; Holländer, A. Properties of amine-containing coatings prepared by plasma polymerization. J. Appl. Polym. Sci. 2004, 92, 979− 990. (48) Testrich, H.; Rebl, H.; Finke, B.; Hempel, F.; Nebe, B.; Meichsner, J. Aging effects of plasma polymerized ethylenediamine (PPEDA) thin films on cell-adhesive implant coatings. Mater. Sci. Eng., C 2013, 33, 3875−3880. (49) Byler, D. M.; Susi, H. Examination of the secondary structure of proteins by deconvoluted FT-IR spectra. Biopolymers 1986, 25, 469− 487. (50) Gengenbach, T. R.; Griesser, H. J. Aging of 1,3-Diaminopropane plasma-deposited polymer films: mechanisms and reaction pathways. J. Polym. Sci., Part A: Polym. Chem. 1999, 37, 2191−2206. (51) Interpretation von C, H, O-Verbindungen: Amide, primäre Amide, sekundäre Amide, tertiäre Amide. p 21. Chemgapedia [Online]. http://www.chemgapedia.de/vsengine/vlu/vsc/de/ch/3/ anc/ir_spek/cho_verbindungen.vlu/Page/vsc/de/ch/3/anc/ir_spek/ schwspek/spektreninterpretation/ir_13_4_7_6/amid.vscml.html. (52) Yasuda, H.; Yasuda, T. The competitive ablation and polymerization (CAP) principle and the plasma sensitivity of elements in plasma polymerization and treatment. J. Polym. Sci., Part A: Polym. Chem. 2000, 38, 943−953. (53) Finke, B.; Schröder, K.; Ohl, A. Surface radical detection on NH3-plasma treated polymer surfaces using the radical scavenger NO. Plasma Process. Polym. 2008, 5, 386−396. (54) Förch, R.; Zhang, Zh.; Knoll, W. Soft plasma treated surfaces: tailoring of structure and properties for biomaterial applications. Plasma Process. Polym. 2005, 2, 351−372. (55) Meyer-Plath, A. Identification of surface radicals on polymers. Vak. Forsch. Prax. 2005, 17, 40−46. (56) Gengenbach, T. R.; Chatelier, R. C.; Griesser, H. J. Characterization of the ageing of plasma-deposited polymer films: Global analysis of x-ray photoelectron spectroscopy data. Surf. Interface Anal. 1996, 24, 271−281. (57) Griesser, H. J.; Chatelier, R. C.; Gengenbach, T. R.; Johnson, G.; Steele, J. G. Growth of human-cells on plasma polymers-putative role of amine and amide groups. J. Biomater. Sci., Polym. Ed. 1994, 5, 531− 554. (58) van Os, M. T.; Menges, B.; Foerch, R.; Vancso, G. J.; Knoll, W. Characterization of plasma-polymerized allylamine using waveguide mode spectroscopy. Chem. Mater. 1999, 11, 3252−3257. (59) Siow, K. S.; Britcher, L.; Kumar, S.; Griesser, H. J. Plasma methods for the generation of chemically reactive surfaces for biomolecule immobilization and cell colonization - A review. Plasma Process. Polym. 2006, 3, 392−418. (60) Gossauer, A. Struktur und Reaktivität der Biomoleküle. Eine Einführung in die Organische Chemie; Wiley-VCH: Weinheim, 2006; pp 299, 458.

13924

dx.doi.org/10.1021/la5019778 | Langmuir 2014, 30, 13914−13924