Accudynetest logo

Products available online direct from the manufacturer

ACCU DYNE TEST ™ Bibliography

Provided as an information service by Diversified Enterprises.

3040 results returned
showing result page 28 of 76, ordered by
 

682. Polischuk, T., “Corona treaters: up to the challenge,” Package Printing, 50, 36-40, (Jan 2003).

2759. Arlt, G., “Treatment electrode topology - some secrets for success,” in 9th TAPPI European PLACE Conference Proceedings, TAPPI Press, 2003.

2114. Smith, R.E., “Substrate surface energy testing,” http://www.accudynetest.com/qctest.html, 2003.

2113. Smith, R.E., “Using ACCU DYNE TEST Marker Pens to measure substrate surface energy,” http://www.accudynetest.com/pentest.html, 2003.

2096. Zekonyte, J., “Mechanisms of argon ion-beam surface modification of polystyrene,” Surface Science, 532-535, 1040-1044, (2003).

The surface characteristics of polymers are important factors determining their interfacial properties and their technological performance. Changes in physical and chemical properties of a polymer film may be induced by subjecting the material to a variety of surface modification techniques, one of which is ion-beam modification. In order to understand the underlying mechanisms X-ray photoelectron spectroscopy (XPS) was used to study the alterations of the polystyrene (PS) surface after Ar-ion treatment under well controlled conditions with low ion doses from 1012 to 1016 cm-2. The ion bombardment leads to surface functionalization, loss of aromaticity, and free radical formation. Induced surface cross-linking and the formation of polar groups raised the surface glass transition temperature of PS film.

2093. Wang, M.-J., Y.-I. Chang, and F. Poncin-Epaillard, “Effects of the addition of hydrogen in the nitrogen cold plasma: The surface modification of polystyrene,” Langmuir, 19, 8325-8330, (2003).

Polystyrene, a polymer extensively used in the biomedical field, causes a problem for some applications because of its surface hydrophobicity. Nitrogen plasma could transform this shortage through polar group attachment. To understand the role of hydrogen during surface functionalization in the nitrogen cold plasma, the effects of the nitrogen and the mixture of N2/H2 plasma are investigated by both the examinations of the densities of attached amine groups and the in-situ diagnostic analyses such as optical emission spectroscopy and mass spectrometry. An increase of functionalization has been proved to be controlled by the gaseous NH radical formation when H2 is added.

2065. Mesic. B., M. Lestelius, G. Engstrom, and B. Edholm, “Printability of PE-coated paper-board using water-based flexographic ink,” Presented at Surf-Treat Karlstad 2003, 2003.

1856. Romero-Sanchez, M.D., M.M. Pastor-Blas, J.M. Martin-Martinez, and M.J. Walzak, “UV treatment of synthetic styrene-butadiene-styrene rubber,” J. Adhesion Science and Technology, 17, 25-45, (2003).

The effectiveness of the treatment with ultraviolet light (UV) on several polymeric surfaces has previously been established. In this study, a low pressure mercury vapour lamp was used as a source of UV radiation for the surface treatment of a difficult-to-bond block styrenebutadiene-styrene rubber (S6), the treatment time ranging from 10 s to 30 min. The UV-treated S6 rubber surfaces were characterized by contact angle measurements (ethylene glycol, 25°C), ATR-IR spectroscopy, XPS, Scanning Electron Microscopy (SEM), and Atomic Force Microscopy (AFM). T-peel tests on UV-treated S6 rubber/polyurethane (PU) adhesive/ leather joints (before and after ageing) were carried out to quantify adhesion strengths. The UV treatment of S6 rubber produced improved wettability, the formation of CSingle BondO, CDouble BondO and COOSingle Bond moieties, and ablation (removal of a thin rubber layer from the surface). The extent of these modifications increased with increasing treatment time. The extended UV treatment produced greater surface modifications, as well as the incorporation of nitrogen moieties at the surface. Furthermore, noticeable ablation of S6 rubber surface occurred. Peel strength values increased with increased treatment time of UV treatment of S6 rubber. Also, with increasing treatment time, the adhesive joints showed different loci of failure: adhesional failure for the as-received and 2 min-UV treated S6 rubber/polyurethane adhesive/leather joints changed to mixed failure (cohesive in the treated S6 rubber + adhesional failure) for the 30 min-UV treated S6 rubber/polyurethane adhesive/leather joint.

1855. Sohn, S., S. Chang, I. Hwang, “The effects of NaOH and corona treatments on triacetyl cellulose and liquid crystal films used in LCD devices,” J. Adhesion Science and Technology, 17, 453-469, (2003).

One of technologically imminent problems related to the use of pressure sensitive adhesives (PSAs) in the LCD industry is how to properly control the surface properties of various polymeric films used in devices to obtain sufficient bond strength with PSAs. To provide practical solutions to this issue, we used two types of surface treatments, NaOH and corona, to control the surface properties of polymeric films that are widely used in the LCD industry. Here we report a significant increase in surface tension in triacetyl cellulose (TAC) and discotic liquid crystal (D-LC) films along with a remarkable enhancement of bond strength in TAC/PSA and D-LC/PSA systems. The major portion of surface tension increase, in both types of films, was found to be due to an increase of polar component. The continuous increase of OH functionality in TAC with NaOH treatment time supported this observation. Furthermore, we established a map of surface treatment by studying the sequential effects of the two treatments, and based on this, we clearly demonstrated that each treatment had its own limiting value that could not be altered regardless of the sequence of surface treatment.

1854. Inagaki, N., K. Narushim, S. Ejima, Y. Ikeda, S.K. Lim, Y.W. Park, K. Miyazaki, “Hydrophobic recovery of plasma modified film surfaces of ethylene-co-tetrafluoroethylene co-polymer,” J. Adhesion Science and Technology, 17, 1457-1475, (2003).

Ethylene-co-tetrafluoroethylene copolymer (ETFE) films were modified by four plasmas: direct and remote H2 plasmas and direct and remote O2 plasmas; and the hydrophobic recovery process of these plasma-modified surfaces was investigated using water contact angle measurements and angular XPS. The water contact angle measurements showed important aspects for the hydrophobic recovery process. (1) All plasma-modified ETFE surfaces, regardless of the kind and mode of plasmas, showed increases in the contact angle with increasing aging time. The increase continued for 5 days after finishing the plasma modification, and stopped after 5 days. (2) The plasmamodified surfaces after the aging process never reverted back to the same level of the contact angle as for the unmodified (original ETFE) surfaces. (3) The contact angle after the aging process was strongly dependent on to what plasma the ETFE surfaces were exposed in the modification. (4) The aging temperature influenced the contact angle value after the aging process. The angular XPS measurements also provided a detailed description of the chemical composition of the topmost layer. (1) The chemical composition at the topmost layer of the surfaces altered during the aging process. (2) CH2-CH2-CHF, and CH2-CHF-CH2 and CH2-CH(OH)-CF2 groups disappeared from the topmost layer during the aging process; and CH2-CH2-CH2, and CH2-CH2-CF2 and CH2-CH(OH)-CHF groups appeared at the topmost layer. (3) Such disappearance and appearance occurred on all plasma-modified surfaces regardless of the kind (H2 or O2 plasma) or mode (direct or remote plasma) of plasmas used for the modification. This may be due to segmental mobility of CH2-CH2-CH2 sequences rather than of CF2-CF2-CF2 sequences.

1853. Della Volpe, C., D. Maniglio, S. Siboni, and M. Morra, “Recent theoretical and experimental advancements in the application of the van Oss-Good acid-base theory to the analysis of polymer surfaces I: General aspects,” J. Adhesion Science and Technology, 17, 1477-1505, (2003).

The acid-base theory as developed by van Oss, Chaudury and Good is a powerful tool to analyze the surface free energy of polymeric materials; however, some problems are encountered in its application and some authors have shown that these problems can be theoretically solved considering this theory as an example of the so-called LFER theories. From this point of view, the definition of a well-defined scale of acid-base strength and the use of a wide and well-equilibrated, appropriate set of liquids is very important. In the present paper some recent results are presented which are based on the mathematical approach discussed by Della Volpe and Siboni in previous papers. The treatment is developed as a list of questions, Frequently Asked Questions (FAQs), whose theoretical implications are discussed using numerical examples chosen from the literature. Some literature data, collected by the opponents of the acid-base theory and recently published, are re-analysed using these methods, showing that they constitute a well-defined set to calculate, with a good precision, the acidbase components of the considered materials and the interfacial energies of liquids used. The present paper is the premise of a second one, in which a set of contact angles data collected by the authors and by other researchers will be analysed following the principles discussed here.

1765. Chen, J.S., Z. Sun, P.S. Guo, Z.B. Zhang, D.Z. Zhu, and H.J. Xu, “Effect of ion implantation on surface energy of ultrahigh molecular weight polyethylene,” J. Applied Physics, 93, 5103-5108, (2003).

The effect of ion implantation including ion species (N2 and C3H8+) and the fluences (1x1014-5x1015 ions/cm2) on the surface energy of ultrahigh molecular weight polyethylene (UHMWPE) were investigated. The total surface energy increases significantly after implanting with the fluence of 1x1014 ions/cm2 regardless of ion species, then, the total surface energy slightly increases for N2+ implanted UHMWPE and decreases slightly for C3H8+ implanted UHMWPE with a further increase of fluence. The structural changes of UHMWPE with different fluence for different ion species are very similar. The linear chains of UHMWPE are damaged and cross linking is generated after implantation. As the fluence increases, the polymer surface becomes more disordered, and the surface becomes hydrogenated amorphous carbon when the fluence exceeds 1x1015 ions/cm2. The surface roughness increases with the increase of the fluence regardless of ion implantation species.

1746. Grace, J.M., and L.J. Gerenser, “Plasma treatment of polymers,” J. Dispersion Science and Technology, 24, 305-341, (2003).

Plasma treatment of polymers encompasses a variety of plasma technologies and polymeric materials for a wide range of applications and dates back to at least the 1960s. In this article we provide a brief review of the United States patent literature on plasma surface modification technologies and a brief review of the scientific literature on investigations of the effects of plasma treatment, the nature of the plasma environment, and the mechanisms that drive the plasma–surface interaction. We then discuss low‐radio‐frequency capacitively coupled nitrogen plasmas and their characteristics, suggesting that they provide significant plasma densities and populations of reactive species for effective plasma treatments on a variety of materials, particularly when placing the sample surface in the cathode sheath region. We further discuss surface chemical characterization of treated polymers, including some results on polyesters treated in capacitively coupled nitrogen plasmas driven at 40 kHz. Finally, we connect plasma characterization with surface chemical analysis by applying a surface sites model to nitrogen uptake of poly(ethylene terephthalate) (PET) and poly(ethylene naphthalate) (PEN) treated in a 40 kHz nitrogen plasma. This example serves to suggest an interesting practical approach to comparisons of plasma treatments. In addition, it suggests an approach to defining the investigations required to conclusively identify the underlying treatment mechanisms.

1735. Abdrashitov, E.F., and A.N. Ponomarev, “Plasma modification of elastomers,” High Energy Chemistry, 37, 279-285, (2003).

The treatment of elastomer articles in a low-temperature glow-discharge plasma in fluorinated organics is an effective method for the enhancement of their wear resistance without changing the formulation of rubbers. As a result of plasma-assisted deposition on an elastomer of a fluorinated antifriction film chemically bound to the substrate, the elastomer friction coefficient is considerably decreased, sticking to a counterface is prevented, and the wear resistance of the elastomer is enhanced, retaining their bulk properties. Based on the study of the structure of the antifriction film at different modification stages and its transformation during friction, a conclusion on the mechanism of elastomer surface failure under dynamic friction conditions was made.

1667. Eckert, W., “Improvement of adhesion on polymer film, foil and paperboard by flame treatment,” in 2003 European PLACE Conference Proceedings, TAPPI Press, 2003.

1513. Barankova, H., and L. Bardos, “Cold atmospheric plasma sources for surface treatment,” in 46th Annual Technical Conference Proceedings, 427-430, Society of Vacuum Coaters, 2003.

1471. Packham, D.E., “Surface energy, surface topography and adhesion,” Intl. J. Adhesion and Adhesives, 23, 437-448, (2003).

In this paper are discussed some of the fundamental principles which are relevant to an understanding of the influence that interfacial roughness may have on adhesion. The surface energies of the adhesive, substrate and of the interface between them determine the extent of wetting or spreading at equilibrium. Numerical values for surface energies may be obtained either from contact angle measurements or from analysing force–displacement curves obtained from the surface forces apparatus. The extent to which the relationships, appropriate for plane surfaces, may be modified to take into account interfacial roughness are discussed. For modest extents of roughness, the application of a simple roughness factor may be satisfactory, but this is unrealistic for many of the practical surfaces of relevance to adhesive technology which are very rough, and is ultimately meaningless, if the surface is fractal in nature. Some examples are discussed of published work involving polymer–metal and polymer–polymer adhesion, where the roughness of the interface exerts a significant influence on the adhesion obtained. Roughness over a range of scales from microns to nanometres may strengthen an interface, increasing fracture energy by allowing bulk energy dissipating processes to be activated when the bond is stressed.

1432. Tanaka, K., and M. Kogoma, “Investigation of a new reactant for fluorinated polymer surface treatments with atmospheric pressure glow plasma to improve the adhesive strength,” Intl. J. Adhesion and Adhesives, 23, 515-519, (2003).

Poly(tetrafluoroethylene-co-perfluoro [alkyl vinyl ether]) (PFA) and polytetrafluoroethylene (PTFE) films were treated by three kinds of atmospheric pressure glow plasmas: an untreated sample was treated by He plasma or trimethoxyborane (TMB)/H2/He plasma, and a TMB-absorbed sample was treated by H2/He plasma. TMB was a new reactant for the treatment, to increase the films’ adhesive strength with an epoxy glue. These films were also treated by a wet method using a sodium solution (Tetra-Etch compound) and such films were used as the control samples. The peel strength values of the controls of PFA and PTFE were 3.5 and 9.5 N cm−1, respectively. The adhesive strengths of all plasma-treated PFAs were stronger than those of untreated one. Especially, the peel strength of the TMB/H2/He plasma-treated PFA showed the maximum value of 4.5 N cm−1, which was bigger than that of the control one. The adhesive strength of the TMB/H2/He plasma-treated PTFE films also showed the maximum peel strength, 7.9 N cm−1, but this value did not exceed that of the control PTFE. Such results suggested that the TMB/H2/He plasma had the advantage of providing better adhesive improvement of those polymers, especially PFA than the wet method could provide. The results of XPS and SEM indicated that TMB actively removed fluorine atoms from the polymer surface. Therefore, boron compounds are effective for the improvement of the adhesive strength between the fluorinated polymer and the epoxy glue.

1379. Romero-Sanchez, M.D., M.M. Pastor-Blas, and J.M. Martn-Martinez, “Treatment of a styrene-butadiene-styrene rubber with corona discharge to improve the adhesion to polyurethane adhesive,” Intl. J. Adhesion and Adhesives, 23, 49-57, (2003).

A styrene-butadiene-styrene (S6) rubber was treated with corona discharge to increase its surface energy and adhesion to polyurethane (PU) adhesive. The influence of the length of treatment (the speed of the upper plate was varied from 80 to 900cm/min) during corona discharge was analyzed. The corona energy applied to S6 rubber surface ranged from 0.4 to 4.6J/cm 2 . The surface modifications produced as a consequence of the corona discharge were characterized immediately after treatment was carried out and were monitored by means of different surface analysis techniques, mainly contact angle measurements (ethylene glycol, 25 o C), ATR-IR spectroscopy, XPS and Scanning Electron Microscopy (SEM). T-peel tests of corona-discharge-treated S6-rubber/polyurethane (PU) adhesive/leather joints (72h after joint formation) were carried out to evaluate the influence of the surface modifications produced by the corona discharge on the adhesion properties of the treated S6 rubber.The corona discharge improved the wettability of the S6 rubber due to the formation of polar moieties, mainly C-O, C=O and COO - groups. These chemical modifications were not detected by ATR-IR spectroscopy (depth of analysis about 5μm), indicating that a nanometer-range oxidized layer was created on the S6 rubber surface by treatment with corona discharge. Besides, surface cleaning and removal of rubber contaminants (mainly silicon moieties) were produced but roughness was not created as a consequence of the treatment. These modifications were enhanced when a low speed treatment (long treatment and high corona energy) was carried out. Peel strength values of corona-discharge-treated S6 rubber/PU adhesive/leather joints only moderately increased (mainly for long length of the corona discharge). Although corona treatment chemically modified the surface of the S6 rubber, the absence of surface roughness might likely be responsible for the slight improvement in its adhesion properties.

1374. Hwang, Y.J., Y. Qiu, C. Zhang, B. Jarrard, R. Stedeford, J. Tsai, et al, “Effects of atmospheric pressure helium/air plasma treatment on adhesion and mechanical properties of aramid fibers,” J. Adhesion Science and Technology, 17, 847-860, (2003).

In order to investigate the effect of atmospheric pressure plasmas on adhesion between aramid fibers and epoxy, aramid fibers were treated with atmospheric pressure helium/air for 15, 30 and 60 s on a capacitively-coupled device at a frequency of 5.0 kHz and He outlet pressure of 3.43 kPa. SEM analysis at 10 000× magnification showed no significant surface morphological change resulted from the plasma treatments. XPS analysis showed a decrease in carbon content and an increase in oxygen content. Deconvolution analysis of C1s, N1s and O1s peaks showed an increase in surface hydroxyl groups that can interact with epoxy resin. The microbond test showed that the plasma treatment for 60 s increased interfacial shear strength by 109% over that of the control (untreated). The atmospheric pressure plasma increased single fiber tensile strength by 16-26%.

1362. Borcia, G., C.A. Anderson, and N.M.D. Brown, “Dielectric barrier discharge for surface treatment: Application to selected polymers in film and fibre form,” Plasma Sources Science and Technology, 12, 335-344, (2003).

In this paper, we report and discuss a surface treatment method, using a dielectric barrier discharge (DBD) of random filamentary type. This offers a convenient, reliable and economic alternative for the controlled modification (so far, largely dependent on surface oxidation) of various categories of material surfaces. Remarkably uniform treatment and markedly stable modified surface properties result over the entire area of the test surfaces exposed to the discharge even at transit speeds simulating those associated with continuous on-line processing. The effects of air-DBD treatment on the surfaces of various polymer films and polymer-based fabrics were studied. The dielectric barrier concerned has been characterized in terms of the energy deposited by the discharge at the processing electrodes and the resultant modifications of the surface properties of the treated samples were investigated using x-ray photoelectron spectroscopy, contact angle/wickability measurement and scanning electron microscopy. The influence of the surface treatment parameters, such as the energy deposited by the discharge, the inter-electrode gap and the treatment time were examined and related to the post-treatment surface characteristics of the materials processed. Relationships between the processing parameters and the properties of the DBD treated samples were thus established. Of the three process variables investigated, the duration of the treatment was found to have a more significant effect on the surface modifications found than did the discharge energy or the inter-electrode gap. Very short air-DBD treatments (fractions of a second in duration) markedly and uniformly modified the surface characteristics for all the materials treated, to the effect that wettability, wickability and the level of oxidation of the surface appear to be increased strongly within the first 0.1–0.2 s of treatment. Any subsequent surface modification following longer treatment (>1.0 s) was less important. The modification of the surface properties also appears to be stable with time, as minimal recovery of the surface properties is shown on ageing post-treatment. The behaviour of the woven textile polymers examined was found to be very similar, under DBD treatment, to that of thin-film variants based on the same polymers. For the porous textile fabrics examined, rapid and efficient treatment (fractions of a second) on both sides of the treated samples was found to be ensured. Thereby the system regime used offers the attractive prospect of controlling the modification of non-compact materials of various texture, porosity, etc. The DBD described system thus provides a chemically mild and mechanically non-destructive means of altering surface properties targeting improved surface characteristics and potentially better application performance.

1253. Strobel, M., and C.S. Lyons, “The role of low-molecular-weight oxidized materials in the adhesion properties of corona-treated polypropylene film,” J. Adhesion Science and Technology, 17, 15-23, (2003).

The effects of low-molecular-weight oxidized materials generated by corona treatment on the adhesion properties of polypropylene (PP) film were investigated by adhering four different materials to the modified PP: a polyamide printing ink, vapor-coated aluminum, a synthetic-rubber pressure-sensitive adhesive, and an acrylate-based pressure-sensitive adhesive. The low-molecularweight materials enhanced the adhesion of the ink and acrylate-based material, but hindered the adhesion of the metal and the rubber-based adhesive. This seemingly contradictory adhesion behavior can be readily explained using the principles outlined by Brewis and Briggs in the 1980s.

1244. Park, J., C.S. Lyons, M. Strobel, M. Ulsh, M.I. Kissinger, M.J. Prokosch, “Characterization of non-uniform wettability on flame-treated polypropylene-film surfaces,” J. Adhesion Science and Technology, 17, 643-653, (2003).

The flame treatment of polypropylene (PP) film involves the use of impinging, conical flames to oxidize the surface of the PP. Depending on treatment conditions, the PP film can be exposed to an inhomogeneous environment because of the conical shape of the flames. This environment can lead to cross-web variations, or 'lanes', in the wettability of the film. We have developed a simple method to quantify these non-uniformities using the information provided by the Wilhelmy plate technique of dynamic contact angle measurement. Both surface-averaged and spatially resolved surface-energy data can be obtained by this technique. In the case of our PP film, the spatial nonuniformities were found to be caused by variations in surface chemistry, not topography. These nonuniformities are not observed on untreated PP. Use of this method enables a quantitative evaluation of the effects of flame-treatment process variables on treatment uniformity.

1235. Mikula, M., Z. Jakubikova, and A. Zahoranova, “Surface and adhesion changes of atmospheric barrier discharge-treated polypropylene in air and nitrogen,” J. Adhesion Science and Technology, 17, 2097-2110, (2003).

Oriented polypropylene treated by atmospheric barrier discharges in air and nitrogen was investigated using several techniques: contact angle measurements, ATR-FT-IR spectroscopy and two adhesion tests based on the stripping of an applied ink layer. The activation in an air discharge was found to be much weaker compared to the activation in industrial grade nitrogen, particularly with respect to adhesion. The adhesion was found to be much better in nitrogen in spite of the common use of air in industrial 'corona discharges'. A new 'abrasive shear-stripping' (AS) test for ink coating adhesion was designed and performed. It was shown that the AS test was much more sensitive than the classical adhesive tape test and was sensitive enough to monitor ageing and overtreatment. The contact angle measurements did not correlate completely with the adhesion properties and could not monitor the overtreatment, while the ATR-FT-IR technique indicated changes just for overtreated foils.

1231. Martinez-Garcia, A., A. Sanchez-Reche, S. Gisbert-Soler, et al, “Treatment of EVA with corona discharge to improve its adhesion to polychloroprene adhesive,” J. Adhesion Science and Technology, 17, 47-65, (2003).

Ethylene vinyl acetate (EVA) material containing 20 wt% vinyl acetate (EVA20) was treated with corona discharge to improve its adhesion to polychloroprene adhesive. Several experimental variables in the corona discharge treatment of EVA20 were considered: corona energy, type of electrode, and number of consecutive treatments. Advancing contact angle measurements (water, 25°C) showed an increase in the wettability of EVA20 after treatment with corona discharge, which corresponds to an increase in the O/C ratio on the treated surface. The higher the corona energy (i.e. the higher discharge power and longer treatment times), the greater the degree of surface oxidation. Peel strength values of the joints produced with EVA20 using a polychloroprene adhesive containing 5 wt% isocyanate increased from 1.5 kN/m (as-received EVA20) to 4.3 kN/m (corona-treated EVA20). A mixed (adhesional + cohesive in EVA20) locus of failure was obtained in all adhesive joints produced with corona discharge-treated EVA20. Finally, the number of consecutive corona discharge treatments and the surface area of the electrode (spherical versus hook-shaped electrode) did not greatly influence the adhesion of EVA20 to polychloroprene adhesive.

1220. Jaehnichen, K., J. Frank, D. Pleul, and F. Simon, “A study of paint adhesion to polymeric substrates,” J. Adhesion Science and Technology, 17, 1635-1654, (2003).

In order to explore the fundamental mechanism of paint adhesion to polymer substrates the surface of polypropylene- ethylene propylene rubber (PP-EPR) blends was modified by flame or plasma treatments. The changes in surface composition and properties were investigated and discussed in light of the results of simple adhesion tests. The topography and surface properties of the PP-EPR samples were studied by employing various surface sensitive techniques. Additionally, the surface properties of the pre-treated PP-EPR were compared with the model polymers poly(methyl methacrylate) (PMMA) and polycarbonate (PC) displaying a poor and an excellent paint adhesion, respectively. Differential scanning calorimetry (DSC) measurements showed that the miscibility of the polymer substrate with paint components was an essential factor for the understanding of the adhesion mechanism. A general model of paint adhesion to polymer surfaces is proposed, where the degree of interdiffusion of the polymer chains of the substrate and paint in the interphase determines the adhesion strength.

1210. Dorai, R., and M.J. Kushner, “A model for plasma modification of polypropylene using atmospheric pressure discharges,” J. Physics D: Applied Physics, 36, 666-685, (2003).

Atmospheric pressure plasmas are commonly used to improve the wetting and adhesion properties of polymers. In spite of their use, the mechanisms for achieving these properties are unclear. In this regard, we report on a computational investigation of the gas phase and surface kinetics during humid-air corona treatment of polypropylene (PP) and the resulting modification of its surface properties while varying energy deposition, relative humidity (RH), web speed, and gas temperature. Using results from a global plasma chemistry model validated against experiments, we found that increasing energy deposition increased the densities of alcohol, carbonyl, acid, and peroxy radicals on the PP surface. In doing so, significant amounts of gas phase O3 and NxOy are produced. Increasing the RH increased the production of peroxy and acid groups, while decreasing those of alcohol and carbonyl groups. Production of O3 decreased while that of HNO3 increased. Increasing the temperature decreased the concentrations of alcohol, carbonyl, and acid groups on PP while those of the peroxy radicals increased. For a given energy deposition, higher web speeds resulted in decreased concentrations of alcohols, peroxy radicals, carbonyl, and acid groups on PP.

1207. Della Volpe, C., S. Siboni, D. Maniglio, M. Morra, C. Cassinelli, et al, “Recent theoretical and experimental advancements in the applications of the van Oss-Chaudhury-Good acid-base theory to the analysis of polymer surfaces, II: Some peculiar cases,” J. Adhesion Science and Technology, 17, 1425-1456, (2003).

According to the general guidelines presented in the accompanying paper, some relevant examples of common polymer surfaces are analysed and discussed; a number of polymers commercially available or laboratory synthesized have been analysed. In particular, the case of poly(vinyl chloride) (PVC), often considered as peculiar in the literature, is fully analysed on the basis of a new set of well-prepared samples, whose compositions were checked by high-vacuum spectroscopies. 'Equilibrium' contact angles, obtained by a new experimental technique, are presented. The results are, however, preliminary, because the final set of liquids used is not so 'well equilibrated' as possible, from the point of view of acid–base properties. The contact angle data obtained are analysed in a non-linear way to calculate the acid–base components of all the liquids and solids. The results are discussed and compared with those obtained from liquid–liquid interfaces presented in the accompanying paper. The physico-chemical features of these samples have also been compared with the adhesion properties of some bacterial cells, commonly found as infective agents on biomaterials surfaces of medical devices, in order to rationalize these results within the theoretical framework of acid–base theory.

1200. Charbonnier, M., and M. Romand, “Polymer pretreatments for enhanced adhesion of metals deposited by the electroless process,” Intl. J. Adhesion and Adhesives, 23, 277-285, (2003).

Metallization techniques based on electroless plating are widely used to coat polymer materials in a large variety of technological applications. Traditionally, dilute tin chloride (SnCl2) and palladium chloride (PdCl2) solutions in HCl are used to render the surface of such non-conductive substrates catalytically active towards metal deposition in the electroless plating solution. In the present work, it is shown how X-ray photoelectron spectroscopy has allowed to monitor the chemical and compositional surface modifications of polymer substrates (polypropylene, polycarbonate) subjected to plasma and UV or VUV irradiation (use of ArF* excimer laser and Xe2* incoherent excimer lamp, respectively) in oxygenated (O2, air) or nitrogenated (N2, NH3) atmospheres, as well as to understand the mechanisms of the catalyst (palladium species) chemisorption on the so-grafted surfaces through the use of a simple dilute palladium chloride solution in HCl. In addition, this work has allowed to bring into light the precise role that the reducer plays (sodium hypophosphite) present in the electroless nickel bath. In short, this research has been successful in allowing the development of new approaches for the electroless metallization of polymer surfaces.

1199. Cepeda-Jiminez, C.M., R. Torregrosa-Macia, and J.M. Martin-Martinez, “Surface modifications of EVA copolymers induced by low pressure RF plasmas from different gases and their relation to adhesion properties,” J. Adhesion Science and Technology, 17, 1145-1159, (2003).

Two ethylene vinyl acetate (EVA) copolymers (12 and 20 wt% of vinyl acetate,VA, content) have been treated with low pressure RF plasmas from non-oxidizing gases (Ar, N2) and oxidizing gases (air, a mixture of 4N2: 6O2 (v/v), O2 and CO2). The formation of polar moieties on both EVAs was more noticeable by treatment with plasmas from non-oxidizing gases than from oxidizing ones (the higher the reactivity, the lower the difference with respect to untreated EVA surfaces). The surface etching with the non-oxidizing plasmas, giving rise to a high roughness, depends on the wt% of VA in the composition of the copolymer because of the different resistances of VA (low) and PE (high) to the non-oxidizing plasma particles bombardment. The adhesion properties obtained using a polyurethane adhesive (PU) showed high T-peel strength values and adhesion failure in EVAs treated with plasmas from oxidizing gases, due to roughness produced causing mechanical interlocking of the adhesive. Lower T-peel strength values were obtained with non-oxidizing plasmas: the values for EVA12 being, in general, lower than those obtained for EVA20. The durability of the treated EVAs/PU adhesive joints after ageing in humidity and temperature was quite good.

954. Owen, M.J., “Surface energy,” in Comprehensive Desk Reference of Polymer Characterization and Analysis, Brady, R.F. Jr., ed., 361-374, Oxford University Press, 2003.

942. Markgraf, D.A., “Atmospheric plasma - the new functional treatment for extrusion coating and lamination processes,” in 2003 European PLACE Conference Proceedings, TAPPI Press, 2003.

817. Kim, J.K., H.S. Kim, and D.G. Lee, “Adhesion characteristics of carbon/epoxy composites treated with low- and atmospheric-pressure plasmas,” J. Adhesion Science and Technology, 17, 1751-1771, (2003).

Although an adhesive joint can distribute load over a larger area than a mechanical joint, requires no holes, adds very little weight to structures and has superior fatigue resistance, it requires careful surface preparation of adherends for reliable joining and low susceptibility to service environments. The load transmission capability of adhesive joints can be improved by increasing the surface free energy of the adherends with suitable surface treatments. In this study, two types of surface treatment, namely the low pressure and the atmospheric pressure plasma treatment, were performed to enhance the mechanical load transmission capabilities of carbon/epoxy composite adhesive joints. The suitable surface treatment conditions for carbon/epoxy composite adhesive joints for both low and atmospheric pressure plasma systems were experimentally investigated with respect to chamber pressure, power intensity and surface treatment time by measuring the surface free energies of the specimens. The change in surface topography of carbon/epoxy composites was measured with AFM (Atomic Force Microscopy) and quantitative surface atomic concentrations were determined with XPS (X-ray Photoelectron Spectroscopy) to investigate the failure modes of composite adhesive joints with respect to surface treatment time. From the XPS investigation of carbon/epoxy composites, it was found that the ratio of oxygen concentration to carbon concentration for both low and atmospheric pressure plasma-treated carbon/epoxy composite surfaces was maximum after about 30 s treatment time, which corresponded with the maximum load transmission capability of the composite adhesive joint.

3026. Sagit-Levanon, S., and M. Marmur, “Validity and accuracy in evaluating surface tension of solids by additive approaches,” J. Colloid and Interface Science, 262, 489-499, (2003).

1851. Dadbin, S., “Surface modification of LDPE film by CO2 pulsed laser irradiation,” European Polymer J., 38, 2489-2495, (Dec 2002).

The influence of the pulsed CO2 laser irradiation on the surface structure of the LDPE film was investigated. Significant changes were observed on the surface of laser treated films as it was verified by the attenuated total reflectance Fourier transform infrared (ATR–FTIR) spectroscopy, scanning electron microscopy and contact angle-measurement. Formation of polar functional groups onto the LDPE surfaces exhibited by the ATR–FTIR spectra was shown to be strongly dependent on the number of the CO2 laser pulses. The intensity of the polar groups increased with increasing the number of pulses up to two and then slightly decreased at three laser pulses. This was also confirmed with the contact angle measurements in which the sample subjected to two laser pulses showed the highest wettability i.e. the lowest water drop contact angle. The concentration of peroxide groups formed on the surface of the laser treated films was determined quantitatively by UV spectroscopic method using iodide procedure. The latter results showed a similar trend with the results obtained using FTIR spectroscopy.

2412. Zhang, D., P.D. Spence, Q. Sun, and L.C. Wadsworth, “Plasma treatment of polymer materials for increased dyeability,” U.S. Patent 6479595, Nov 2002.

2411. Geibart, D., “Method of surface preparation using plasma in air,” U.S. Patent 6476342, Nov 2002.

1549. Oller, S., “Printing on plastic,” American Printer, (Nov 2002).

1524. Chen, Q., “PTFE electret negative charge stability after RF plasma treatment,” J. Physics D: Applied Physics, 35, 2939-2944, (Nov 2002).

An 18 μm nonmetallized polytetrafluoroethylene (PTFE) film is treated in radio frequency (RF) plasma before a point-to-grid corona charged. The isothermal (170°C) surface potential measurement shows that the surface charge stability is significantly dependent on the plasma sources and treatment conditions. Oxygen (O2), oxygen/helium (O2/He) mixture gases and helium (He) plasma treatment enhance the film negative charge stability significantly but not hydrogen (H2) plasma. Electron spectroscopy chemical analysis confirms that this superior negative charge retention for O2 plasma treatment is a result of the high concentration of oxide groups on the subsurface during the plasma treatment.

1171. Brynolf, R., “Method and apparatus, with redundancies, for treating substrate plastic parts to accept paint without using adhesion promoters,” U.S. Patent #6716484, Nov 2002.

 

<-- Previous | 1 | 2 | 3 | 4 | 5 | 6 | 7 | 8 | 9 | 10 | 11 | 12 | 13 | 14 | 15 | 16 | 17 | 18 | 19 | 20 | 21 | 22 | 23 | 24 | 25 | 26 | 27 | 28 | 29 | 30 | 31 | 32 | 33 | 34 | 35 | 36 | 37 | 38 | 39 | 40 | 41 | 42 | 43 | 44 | 45 | 46 | 47 | 48 | 49 | 50 | 51 | 52 | 53 | 54 | 55 | 56 | 57 | 58 | 59 | 60 | 61 | 62 | 63 | 64 | 65 | 66 | 67 | 68 | 69 | 70 | 71 | 72 | 73 | 74 | 75 | 76 | Next-->