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 19 of 76, ordered by
 

1015. Bhowmik, S., P.K. Ghosh, S. Ray, and S.K. Barthwal, “Surface modification of high density polyethylene and polypropylene by DC glow discharge and adhesive bonding to steel,” J. Adhesion Science & Technology, 12, 1181-1204, (1998).

The surface modification of high density polyethylene (HDPE) and polypropylene (PP) has been carried out by exposure to a DC glow discharge in air at different power levels of 5.28, 11, and 13 W. The surface energies of polymers exposed to glow discharge were estimated by measuring the contact angles of two test liquids: de-ionized water and formamide, whose surface energy components are known. Both the polar and the dispersion components of the surface energy increased rapidly at short exposure times but the increase of the polar component was relatively more than that of the dispersion component. At low power levels of 5.28 and 11 W, the polar component of the surface energy reached a maximum plateau depending on the exposure time, but at a 13 W power level the polar component of the surface energy decreased from a maximum value to a saturation level. For PP, this saturation level could not be attained in this study. The maximum total surface energy measured in this study corresponds to the maximum polar component at 13 W for an exposure time of 120 s. The contact angle of the adhesive, Araldite AY 105 mixed with hardener HY 840 in a weight ratio of 2 : 1, was minimum at this maximum surface energy attained with HDPE and PP by exposure to a glow discharge in air. The measured lap shear strengths of HDPE or PP-Araldite-mild steel joints show a maximum corresponding to the maximum surface energy measured on the above-mentioned polymers.

1021. Nihlstrand, A., T. Hjertberg, H.P. Schreiber, and J.E. Klemberg-Sapieha, “Plasma treatment and adhesion properties of a rubber-modified polypropylene,” J. Adhesion Science & Technology, 10, 651-675, (1996).

34. Bonnerup, C., and P. Gatenholm, “The effect of surface energetics and molecular interdiffusion on adhesion in multicomponent polymer systems,” J. Adhesion Science and Technology, 7, 247-262, (1993) (also in Contact Angle, Wettability and Adhesion: Festschrift in Honor of Professor Robert J. Good, K.L. Mittal, ed., p. 753-768, VSP, Nov 1993).

The interfacial region of coated plastics is an example of a multicomponent polymer system. Practical adhesion, as determined by the peel test, has been found to be strongly dependent on the composition of the system and the degree of interaction between its components. Several interactions are possible during the coating process of polypropylene (PP)/ethylenepropylene-diene-monomer (EPDM) blends with chlorinated polyolefin (primer) and polyurethane (PUR) paint. Wettability, a necessary but not sufficient condition alone for molecular interdiffusion, was found to be good in all cases. The lack of interfacial adhesion between PP and PUR and between EPDM and PUR was explained by high interfacial tensions calculated from surface energetics, which, in turn, were determined by contact angle and inverse gas chromatography (IGC) measurements. The improvement of interfacial adhesion between PUR and PP by chlorinated polyolefin was explained by acid-base interactions detected by IGC. The creation of surface topography by extraction of low molecular weight fractions during the coating process does not influence the adhesion. Molecular interdiffusion was shown to be facilitated by solvents.

101. Foerch, R., J. Izawa, and G. Spears, “Comparative study of the effects of remote nitrogen plasma, remote oxygen plasma, and corona discharge treatments on the surface properties of polyethylene,” J. Adhesion Science and Technology, 5, 549-564, (1991).

The effects of remote nitrogen plasma, remote oxygen plasma, and corona discharge treatments on linear low-density polyethylene were studied with regard to the chemical and physical surface modification, depth of modification, and surface stability. An attempt was made to correlate the type and the extent of modification with the printing and adhesion properties of the modified surfaces. Surface topography was studied using scanning electron microscopy. The relative percentages of nitrogen and oxygen on the surfaces were determined by X-ray photoelectron spectroscopy. Printing and adhesion tests were performed using standard, commercially available inks and adhesives.

102. Foerch, R., G. Kill, and M.J. Walzak, “Plasma surface modification of polyethylene: short-term vs. long-term plasma treatment,” J. Adhesion Science and Technology, 7, 1077-1089, (1993).

A remote plasma reactor, with air as the plasma gas, has been used for in-line surface modification of linear low-density polyethylene tape (LLDPE) passing 10 cm below the main plasma zone. Line speeds of up to 0.70 m/s were tested, allowing the study of 0.014 s exposure times to the plasma. Oxygen to carbon (O/C) ratios averaging 0.11 were observed on a reproducible basis. The reactor was also used for static plasma treatment under similar experiment onditions. This allowed a comparative study of short-term (milliseconds) vs. long-term (several seconds) plasma treatment. High-resolution X-ray photoelectron spectroscopy (XPS) analysis of the treated polymer surface suggested the formation of hydroxyl (CSingle BondOH), carbonyl (CDouble BondO) and carboxyl (OSingle BondCDouble BondO) groups, even after short plasma treatment. The intensities of these components were seen to increase in approximately equal quantities with increasing O/C ratio. Water washing of polyethylene surfaces with high O/C ratios showed a loss of oxygen, apparent as a decrease in OSingle BondCDouble BondO groups in the C 1s spectra. A smaller loss in oxygen was observed when washing samples that had been plasma-treated for milliseconds. A surface ageing study revealed that polyethylene surfaces that had been plasma-treated for short time periods showed only a negligible loss of oxygen on prolonged exposure to air. Surfaces treated for longer time periods showed a loss of up to 50% of the total oxygen on the surface within a few days of treatment. Static secondary ion mass spectrometry has provided some supporting evidence for surface damage of the treated films.

195. Kinbara, A., A. Kikuchi, S. Baba, and T. Abe, “Effect of plasma treatment of PTFE substrates on the adhesion characteristics of vacuum-deposited Au films,” J. Adhesion Science and Technology, 7, 457-466, (1993).

PTFE foils were plasma-treated in order to enhance their adhesion to thin films. The effect of plasma treatment using argon and oxygen discharge gases on the surface energy of PTFE foils was examined by measuring the contact angles of water droplets placed on the foil surface. Exposure to the plasma for only about 10-20 s was very effective in enhancing the surface energy. By depositing gold films onto the PTFE substrates, it was found that this enhancement in surface energy was directly related to an increase in the film adhesion. It was also found that Ar plasma treatment of a few tens of seconds followed by O2 plasma treatment for 10 s was even more effective for adhesion enhancement.

218. Lee, L.-H., “Roles of molecular interactions in adhesion, adsorption, contact angle, and wettability,” J. Adhesion Science and Technology, 7, 583-634, (1993) (also in Contact Angle, Wettability and Adhesion: Festschrift in Honor of Professor Robert J. Good, K.L. Mittal, ed., p. 45-96, VSP, Nov 1993).

This study is aimed at understanding the controversy between the surface tension component (STC) theory and the equation of state (EQS) approach for interfacial tensions. We attempt to relate molecular interactions to various components of surface tension. Molecular interactions consist of electrostatic (ES), charge transfer (CT), polarization (PL), exchange-repulsion (EX), dispersion (DIS), and coupling (MIX) components. These interactions can be the basis for the STC theory involving Lifshitz-van der Waals (LW) and the short range acid-base (AB) or donor-acceptor interaction. Each of these components is shown to contain two major parameters. New equations for the interaction energy and surface tension for polar molecules are proposed to include the ES and EX parameters, which happen in some cases to balance each other or nearly cancel out without being detected. The roles of molecular interactions on adhesion, adsorption, contact angle, and wettability are illustrated through the spreading coefficient S, the Hamaker coefficient A, and Derjaguin's disjoining pressure . We have found that the STC theory is applicable to the systems involving either physisorption or chemisorption, whlie the EQS applies to those involving ony physisorption.

246. Morra, M., E. Occhiello, and F. Garbassi, “Chemical reactions on plasma-treated polyethylene surfaces,” J. Adhesion Science and Technology, 7, 1051-1063, (1993).

Oxygen plasma treatment as a surface functionalization technique is discussed. Oxygen-containing functionalities were introduced on the surface of high- (HDPE) and low-density polyethylene (LDPE) by glow discharge. The number of surface hydroxyl groups was increased by a post-discharge wet treatment in a reducing solution. The effects of the substrate nature, the discharge parameters, and the post-discharge wet treatment on the surface functional groups are discussed, and the effectiveness of functionalized surfaces on the yield of coupling reactions is shown.

283. Papirer, E., D.Y. Wu, and J. Schultz, “Adhesion of flame-treated polyolefins to styrene butadiene rubber,” J. Adhesion Science and Technology, 7, 343-362, (1993).

Samples of polyethylene and polypropylene have been submitted to repeated short duration (75 ms) flame treatments, at optimum flaming conditions. Surface energies of untreated and flamed specimens were determined by liquid contact angle measurements. It appears that the surface energy of polyethylene increases much more than that of polypropylene after flame treatment. The flamed polymer surfaces were further examined by electron spectroscopy, Fourier Transform IR spectroscopy and secondary ions mass spectrometry. The adhesion properties of modified polymer surfaces were studied by testing in peel the bonded Styrene Butadiene Rubber/polyolefins assemblies. Scanning electron microscopy (SEM) and water contact angle measurements have been used to observe the locus of failure. Good correlations were obtained between surface energy and adhesion strength, the increase in adhesion strength being particularly important for flamed PE/SBR assemblies. In addition, the peeling in a liquid medium allowed the determination of the respective contribution to adhesion of chemical and physical interactions. It is shown that a major part of the adhesion strength increase is of chemical origin, particularly for the bonded flamed PE/SBR assemblies.

438. Chen, K.S., Y. Uyama, and Y. Ikada, “Adhesive-free adhesion of grafted surfaces with different wettabilities,” J. Adhesion Science and Technology, 6, 1023-1035, (1992).

A polyester film surface was graft-polymerized with various water-soluble monomers by a combination of plasma pretreatment and photoirradiation. The grafted surfaces showed strong adhesion to another non-grafted or grafted film when brought into direct contact in the presence of water and subsequently dried. The adhesion force depended on the hydrophilicity of the adhering surfaces and the graft density. The film having a larger graft density generally showed stronger adhesion in the final stage of drying, but it took longer to achieve high adhesion because of the larger amount of water existing in the interfacial region between the two surfaces. On the other hand, substantial adhesion was obtained almost instantaneously upon contact when one was grafted with an anionic polymer and the other grafted with a cationic polymer. Adhesion between similarly charged surfaces was very weak at the beginning of drying, probably because of the electrostatic repulsion between the charged groups.

458. Fowkes, F.M., “Role of acid-base interfacial bonding in adhesion,” J. Adhesion Science and Technology, 1, 7-27, (1987).

The strength of macroscopic adhesive bonds of polymers is known to be directly proportional to the microscopic exothermic interfacial energy changes of bond formation, as measured by Dupre's 'work of adhesion'. Since the work of adhesion can be very appreciably increased by interfacial acid-base bonding with concomitant increases in adhesive bond strength, it is important to understand the acid-base character of polymers and of the surface sites of substrates or of the reinforcing fillers of polymer composites. The best known acid-base bonds are the hydrogen bonds; these are typical of acid-base bonds, with interaction energies dependent on the acidity of the hydrogen donor and on the basicity of the hydrogen acceptor. The strengths of the acidic or basic sites of polymers and of inorganic substrates can be easily determined by spectroscopic or calorimetric methods, and from this information one can start to predict the strengths of adhesive bonds. An important application of the new knowledge of interfacial acid-base bonding is the predictable enhancement of interfacial bonding accomplished by surface modification of inorganic surfaces to enhance the interfacial acid-base interactions.

465. Golander, C.-G., and B.-A. Sultan, “Surface modification of polyethylene to improve its adhesion to aluminum,” J. Adhesion Science and Technology, 2, 125-135, (1988).

The effects of surface modification of polyethylene (PE) and aluminum on the adhesion strength have been investigated. PE was modified by KMnO4/H2SO4 treatment followed by adsorption of different cations, Ca2+, Ba2+, and Zn2+. The aluminum surface was treated with alkali and was also modified by adsorption of titanates. The surfaces were characterized by means of multiple internal reflection (MIR) IR and ESCA. The adhesion strength was measured by the T-peel test. Both ESCA and MIR analyses show the presence of hydroxyl, carbonyl, ester, and carboxyl groups on the KMnO4/H2SO4-treated PE surface. In addition, sulfate and sulfonate groups are present. The sulfonate groups are apparently localized in crevices extending beneath the ESCA sampling depth of 50 A. Vinylidene groups are also present on the surface. Cation adsorption on the oxidized PE surface seems to be determined by the solubility constant of the corresponding sulfate salts and is in the order Ba2+ > Ca2+ > Zn2+. Adsorption of Ca2+ and Ba2+ increases the relative concentration of oxygen-containing groups on the KMnO4/H2SO4-treated surface. A further increase is seen after annealing. KMnO4/H2SO4 treatment almost doubled the adhesion strength of PE to aluminum. Ca2+ adsorption on the surface prior to lamination increased the adhesion strength nearly three times and caused cohesive failure in the T-peel test. However, when Zn2+ was adsorbed, the adhesion strength decreased drastically. Alkaline treatment of the aluminum surface had only a minor effect on adhesion. The chemical structure of the adsorbed titanates has a great influence on the adhesion strength.

486. Inagaki, N., S. Tasaka, H. Kawai, and Y. Kimura, “Hydrophilic surface modification of polyethylene by NO-plasma treatment,” J. Adhesion Science and Technology, 4, 99-107, (1990).

The surface modification of polyethylene surfaces by NO-plasma irradiation was investigated from the point of view of the hydrophilicity and chemical composition. The hydrophilicity was evaluated from the advancing contact angle of water and the surface energy. The chemical composition of the modified surfaces was determined by diffuse reflectance Fourier transform infrared spectroscopy and XPS. NO-plasma irradiation for 5 min made the polyethylene surfaces hydrophilic. The advancing contact angle of water on the modified polyethylene surfaces reached 28 deg, and the surface energy was 57.6 mJ/m2. The incorporation of oxygen and nitrogen moieties on the polyethylene surfaces occurred during the NO-plasma irradiation. The main oxygen moieties were carbonyl groups, hydroxyl groups, and ether linkages; the nitrogen moieties were amino groups. NO-plasma irradiation was more effective in improving the hydrophilicity than the O2 plasma, N2 plasma, or corona discharge treatment.

569. McCafferty, E., and J.P. Wightman, “Determination of the acid-base properties of metal oxide films and of polymers by contact angle measurements,” J. Adhesion Science and Technology, 13, 1415-1436, (1999) (also in Apparent and Microscopic Contact Angles, J. Drelich, J.S. Laskowski, and K.L. Mittal, eds., p. 149-170, VSP, Jun 2000).

580. Strobel, J.M., M. Strobel, C.S. Lyons, C. Dunatov, and S.J. Perron, “Aging of air-corona-treated polypropylene film,” J. Adhesion Science and Technology, 5, 119-130, (1991).

X-ray photoelectron spectroscopy (ESCA), wettability measurements, and an ink adhesion test were used to characterize changes in the surface properties of air-corona-treated polypropylene (PP) films upon aging under a variety of storage conditions. No changes in ESCA O/C atomic ratios as a function of aging were observed for corona-treated PP films. The wettability data indicated a slight decrease in wettability upon aging. Aging did not affect ink adhesion for the particular PP and ink studied. The responses obtained were independent of the various film storage conditions employed. The slight decrease in wettability observed upon aging was attributed to reorientation of oxidized functionalities within the surface region.

581. Strobel, M., C.S. Lyons, J.M. Strobel, and R.S. Kapaun, “Analysis of air-corona-treated polypropylene and polyethylene terephthalate films by contact angle measurement and X-ray photoelectron spectroscopy,” J. Adhesion Science and Technology, 6, 429-443, (1992) (also in Contact Angle, Wettability and Adhesion: Festschrift in Honor of Professor Robert J. Good, K.L. Mittal, ed., p. 493-507, VSP, Nov 1993).

Contact-angle measurements in air and water environments and X-ray photoelectron spectroscopy (XPS) were used to characterize the surface properties of air-corona-treated polypropylene (PP) and poly(ethylene terephthalate) (PET) films. Surface properties were examined as a function of the storage time at various temperatures. Corona treatment forms water-soluble, low-molecular-weight oxidized materials on both polymer films. Corona-treated PP and corona-treated PET films have markedly different responses to aging. For corona-treated PP stored at ambient temperatures, only a slight decrease in wettability was observed. This decrease was attributed to the reorientation of oxidized functionalities within the surface region. At elevated storage temperatures, migration of oxidized species out of the surface region occurred under some conditions. For corona-treated PET, extensive migration and reorientation of oxidized groups occurred even at ambient temperatures, leading to significant decreases in wettability and a loss of surface oxidation. The contrasts in the responses of PP and PET to corona treatment are primarily due to differences in the properties of the base polymer resins.

587. Varughese, K.T., P.P. De, and S.K. Sanyal, “Contact angle behavior of poly(vinyl chloride)/epoxidized natural rubber miscible blends,” J. Adhesion Science and Technology, 3, 541-550, (1989).

Contact angle studies of miscible poly(vinyl chloride)/epoxidized natural rubber (PVC/ ENR) blends were carried out in air using water and methylene iodide. The solid surface free energy was calculated from harmonic mean equations. Blending of PVC and ENR decreased their contact angle or increased their solid surface free energy due to the improved chain mobility, and the accumulation of excess polar sites at the surface through conformational alterations resulting from the specific interaction of PVC and ENR. The work of adhesion, interfacial free energy, spreading coefficient, and Girifalco-Good's interaction parameter changed markedly with the blend composition. In blends, PVC and ENR improved hydrophilicity, and wettability with polar and non-polar liquids. The presence of a plasticizer in PVC, in general, further improved the wettability and hydrophilicity in blends.

770. Neimark, A.V., “Thermodynamic equilibrium and stability of liquid films and droplets on fibers,” J. Adhesion Science and Technology, 13, 1137-1154, (1999) (also in Apparent and Microscopic Contact Angles, J. Drelich, J.S. Laskowski, and K.L. Mittal, eds., p. 301-318, VSP, Jun 2000).

772. Shi, S.Q., and D.J. Gardner, “A new model to determine contact angles on swelling polymer particles by the column wicking method,” J. Adhesion Science and Technology, 301-314, (2000) (also in Apparent and Microscopic Contact Angles, J. Drelich, J.S. Laskowski, and K.L. Mittal, eds., p.431-444, VSP, Jun 2000).

808. Ringenbach, A., Y. Jugnet, and T.M. Duc, “Interfacial chemistry in Al and Cu metallization of untreated and plasma treated polyethylene and polyethylene terephthalate,” J. Adhesion Science and Technology, 9, 1209-1228, (1995) (also in Polymer Surface Modification: Relevance to Adhesion, K.L. Mittal, ed., p. 101-120, VSP, May 1996).

811. Zhang, J.-Y., H. Esrom, U. Kogelschatz, and G. Emig, “Modifications of polymers with UV excimer radiation from lasers,” J. Adhesion Science and Technology, 9, 1179-1218, (1995) (also in Polymer Surface Modification: Relevance to Adhesion, K.L. Mittal, ed., p.153-184, VSP, May 1996).

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.

818. Lunkwitz, K., W. Burger, U. Lappan, H.-J. Brink, and A. Ferse, “Surface modification of fluoropolymers,” J. Adhesion Science and Technology, 9, 297-310, (1995) (also in Polymer Surface Modification: Relevance to Adhesion, K.L. Mittal, ed., p. 349-362, VSP, May 1996).

819. Lee, K.-W., “Modification of polyimide morphology: relationship between modification depth and adhesion strength,” J. Adhesion Science and Technology, 8, 1077-1092, (1994) (also in Polymer Surface Modification: Relevance to Adhesion, K.L. Mittal, ed., p. 363-378, VSP, May 1996).

825. Liston, E.M., L. Martinu, and M.R. Wertheimer, “Plasma surface modification of polymers for improved adhesion: a critical review,” J. Adhesion Science and Technology, 7, 1091-1127, (1993) (also in Plasma Surface Modification of Polymers: Relevance to Adhesion. M. Strobel, C.S. Lyons, and K.L. Mittal, eds., p. 3-42, VSP, Oct 1994).

826. Gerenser, L.J., “XPS studies of in situ plasma-modified polymer surfaces,” J. Adhesion Science and Technology, 7, 1019-1040, (1993) (also in Plasma Surface Modification of Polymers: Relevance to Adhesion, M. Strobel, C.S. Lyons, and K.L. Mittal, eds., p. 43-64, VSP, Oct 1994).

827. Clouet, F., M.K. Shi, R. Prat, Y. Holl, P. Marie, et al, “Multitechnique study of hexatriacontane surfaces modified by argon and oxygen RF plasmas: effect of treatment time and funtionalization, and comparison with HDPE,” J. Adhesion Science and Technology, 8, 329-361, (1994) (also in Plasma Surface Modification of Polymers: Relevance to Adhesion, M. Strobel, C.S. Lyons, and K.L. Mittal, eds., p. 65-98, VSP, Oct 1994).

828. Foerch, R., G. Kill, and M.J. Walzak, “Plasma surface modification of polyethylene: short-term vs. long-term plasma treatment,” J. Adhesion Science and Technology, 7, 1077-1089, (1993) (also in Plasma Surface Modification of Polymers: Relevance to Adhesion, M. Strobel, C.S. Lyons, and K.L. Mittal, eds., p. 99-112, VSP, Oct 1994).

833. Morra, M., E. Occhiello, and F. Garbassi, “Chemical reactions on plasma-treated polyethylene surfaces,” J. Adhesion Science and Technology, 7, 1051-1063, (1993) (also in Plasma Surface Modification of Polymers: Relevance to Adhesion, M. Strobel, C.S. Lyons, and K.L. Mittal, eds., p. 183-196, VSP, Oct 1994).

835. Collaud, M., S. Nowak, O.M. Kuttel, and L. Schlapbach, “Enhancement of the sticking coefficient of Mg on polypropylene by in situ ECR-RF Ar and N2 plasma treatments,” J. Adhesion Science and Technology, 8, 435-453, (1994) (also in Plasma Surface Modification of Polymers: Relevance to Adhesion, M. Strobel, C.S. Lyons, and K.L. Mittal, eds., p. 255-274, VSP, Oct 1994).

839. Berger, E.J., “A method of determining the surface acidity of polymeric and metallic materials and its application to lap shear adhesion,” J. Adhesion Science and Technology, 4, 373-391, (1990) (also in Acid-Base Interactions: Relevance to Adhesion Science and Technology, K.L. Mittal and H.R. Anderson Jr., eds., p. 207-228, VSP, Nov 1991).

A method has been developed to measure the surface acidity of solids using the contact angles of seven probe liquids. The geometric mean model was used to calculate the surface free energy. Then the value of the solid polarity, from the geometric mean, was compared with the polarity values calculated using the geometric mean dispersive component and the contact angles of two Lewis acids (liquefied phenol and glycerol) and two Lewis bases (formamide and aniline.) The deviation between these values was used to determine a value for the acidity, referred to as D. D was measured for a series of polymeric and metallic materials. Lap shear joints were fabricated and tested using these substrates and two adhesives, a urethane and an epoxy. The acidity of the substrate surfaces was found to affect the lap shear joint strength.

841. Dwight, D.W., F.M. Fowkes. D.A. Cole, M.J. Kulp, et al, “Acid-base interfaces in fiber-reinforced polymer composites,” J. Adhesion Science and Technology, 4, 619-632, (1990) (also in Acid-Base Interactions: Relevance to Adhesion Science and Technology, K.L. Mittal and H.R. Anderson Jr., eds., p. 243-256, VSP, Nov 1991).

The role of Lewis acid-base interactions at the fiber-matrix interface in composites is studied with both glass and Teflon fibers. In the glass fiber case, surface chemistry is modified with amino-, methacryloxy- and glycidoxy-silane coupling agents (A-1100, A-174 and A-187, respectively). Silane adsorption mechanisms as well as the properties of filament-wound, unidirectional epoxy and polyester composites are explained by a combination of X-ray photoelectron spectroscopy (XPS), scanning electron microscopy (SEM), and flow microcalorimetry. The heats of adsorption of pyridine and phenol prove that the coupling agents add acidic sites to the glass fiber surface as well as stronger basic sites. The subsequent adhesion of the matrix polymers and the short beam shear strengths of composites are explained on this basis. The Teflon fibers are first etched with sodium naphthalene solutions, and then sequentially hydroborated and acetylated, producing approximately monofunctional hydroxyl (acidic) and ester (basic) groups on the surfaces, as determined by XPS, FTIR, and electrophoretic mobility analyses. Composites prepared with the acetylated fibers and a chlorinated polyvinyl chloride (acidic) matrix are superior in tensile properties, and SEM fractography shows PTFE fibrillation, indicative of good fiber-matrix adhesion and stress transfer, in this case only.

842. Borch, J., “Thermodynamics of polymer-paper adhesion: A review,” J. Adhesion Science and Technology, 5, 523-541, (1991).

A review of studies of polymer-paper adhesion illustrates the thermodynamic nature of the bondability of polymers to plain, uncoated paper surfaces. The bond strength depends strongly on the chemical nature of the polymer surface and on that of the fibrous paper surface. Adhesion to paper may be characterized indirectly through thermodynamic analysis of the paper substrate, or directly through paper laminate or adhesion tape peel testing. The need for adequate paper adhesion is emphasized, particularly for some of the newer printing processes (electrophotographic and thermal imaging). It is concluded that some of the indirect methods of adhesion characterization (surface energetics analysis via contact angle measurements or the inverse gas chromatography technique) may serve to characterize paper adhesion in these processes.

843. Finlayson, M.F., and B.A. Shah, “The influence of surface acidity and basicity on adhesion of poly(ethylene-co-acrylic acid) to aluminum,” J. Adhesion Science and Technology, 4, 431-439, (1990) (also in Acid-Base Interactions: Relevance to Adhesion Science and Technology, K.L. Mittal and H.R. Anderson Jr., eds., p. 303-312, VSP, Nov 1991).

This work demonstrates the usefulness of flow microcalorimetry for surface characterization of metal foils (aluminum) and polymer [poly (ethylene-co-acrylic acid)] fibers. It shows that the polymer to aluminum adhesion is dominated by Lewis acid/Lewis base type interactions. These interactions are predictable from the measured heats of surface adsorption and desorption of probe molecules from dilute solution. The heats of interaction are a measure of the strengths of these sites. Adhesion between basic aluminum foil and acidic polymer resin increases with increasing numbers of either acidic sites on the polymer or basic sites on the foil. The calorimetry and adhesion results are in good agreement. This study supports recent observations vide infra that wettability of the aluminum is much less important for polymer/aluminum adhesion than chemical bonding.

844. Tiburcio, A.C., and J.A. Manson, “The effects of filler/polymer acid-base interactions in model coating systems,” J. Adhesion Science and Technology, 4, 653-668, (1990) (also in Acid-Base Interactions: Relevance to Adhesion Science and Technology, K.L. Mittal and H.R. Anderson Jr., eds., p. 313-328, VSP, Nov 1991).

The water vapor permeability of glass-bead-filled phenoxy films was shown to depend strongly on the degree of interfacial interaction between the filler and matrix; the greater the adhesion, the lower the permeability. Scanning electron microscopy (SEM) was used to characterize the glass surface and the corresponding degree of adhesion between the filler and polymer matrix. Maximum interaction between the acidic phenoxy and glass filler was obtained when the glass had been treated with an aminopropyltriethoxysilane, which yielded a basic surface overall. Retention of the cellosolve acetate solvent was also reduced by the glass filler, especially by the more basic glass. The dynamic mechanical properties were affected primarily by the presence of residual solvent.

958. Cho, D.L., K.H. Shin, W.-J. Lee, and D.-H. Kim, “Improvement of paint adhesion to a polypropylene bumper by plasma treatment,” J. Adhesion Science and Technology, 15, 653-664, (2001).

Improvement of the paint adhesion to a polypropylene (PP) bumper has been investigated without using a primer by treating the bumper surface with O2, H2O, and acetylene plasmas. All the plasma treatments resulted in an increase of the adhesion strength in dry conditions. The adhesion strength could be increased up to a value comparable to that obtained by applying a primer. The treated surfaces were quite stable for 7 days in air. After exposure to wet conditions, however, the adhesion strengths for both O2 and H2O plasma-treated samples decreased significantly, while the adhesion strength for the acetylene plasma-treated sample did not change much.

964. Cho, C.K., B.K. Kim, and C.E. Park, “The aging effects of repeated oxygen plasma treatment on the surface rearrangement and adhesion of LDPE to aluminum,” J. Adhesion Science and Technology, 14, 1071-1083, (2000).

The effects of aging temperature and time on the adhesion properties of oxygen plasmatreated low-density polyethylene (LDPE) were investigated. As the aging temperature and time increased, surface rearrangement and the migration of molecules containing polar functional groups into the bulk were accelerated to the surface to form a hydrophobic surface. The adhesion strength of oxygen plasma-treated LDPE/aluminum joints was measured using a 90° peel test by varying the plasma treatment time and aging temperature. The adhesion strength was constant, regardless of the plasma treatment time. As the aging temperature increased, the adhesion strength of the LDPE/aluminum joints decreased and the locus of failure changed from cohesive to interfacial failure. It was also found that the polar functional groups buried in the bulk could be reoriented to the surface in a polar environment. This study also investigated whether repeated oxygen plasma treatment would increase the concentration of polar functional groups at the surface and reduce the surface rearrangement and the migration of molecules containing polar functional groups during aging. Contact angle measurements and X-ray photoelectron spectroscopy (XPS) showed that repeated oxygen plasma treatments increased the concentration of polar functional groups at the surface. However, the aging time between plasma treatments had a negligible effect on the concentration of polar functional groups at the surface.

969. Nakamatsu, J., L.F. Delgado-Aparicio, R. Da Silva, and F. Soberon, “Ageing of plasma-treated poly(tetrafluoroethylene) surfaces,” J. Adhesion Science and Technology, 13, 753-761, (1999).

970. Kawabe, M., S. Tasaka, and N. Inagaki, “Effects of nitrogen plasma treatment of pressure-sensitive adhesive layer surfaces on their peel adhesion behaviour,” J. Adhesion Science and Technology, 13, 573-592, (1999).

The influence of the surface modification of pressure-sensitive adhesive tapes on their adhesion behavior has been investigated. PBA [poly(butyl acrylate)] and PIB [poly(isobutylene)] adhesives were chosen as pressure-sensitive adhesives and nitrogen plasma was used for the surface modification of the adhesives. The peel force of PBA or PIB adhesive/stainless steel joints was evaluated. The nitrogen plasma treatment showed large effects on the adhesion behavior of both the PBA and the PIB adhesives. The peel force for the PBA adhesive/stainless steel joint decreased by 57 times as a result of the nitrogen plasma treatment and that for the PIB adhesive/stainless steel joint increased by 2.2 times. There are essential differences in the modification reactions caused by the nitrogen plasma between the PBA and PIB adhesives. For the PBA adhesive, cross-linking reactions occurred among the PBA polymer chains and the surface was hardened. For the PIB adhesive, degradation reactions occurred and products with a low molecular weight were formed on the surface. These differences are due to the different responses of the PBA and PIB adhesives towards the nitrogen plasma. The mechanism of the changes in adhesion behavior caused by the nitrogen plasma is discussed.

977. Della Volpe, C., A. Deimichei, and T. Ricco, “Multiliquid approach to the surface free energy determination of flame-treated surfaces of rubber-toughened polypropylene,” J. Adhesion Science and Technology, 12, 1141-1180, (1998).

In the present work, rubber-toughened polypropylenes (TPOs) with different properties of the rubbery phase, consisting in different degrees of rubber dispersion and different levels of rubber crystallinity, were considered. The flame-treated surfaces of these materials and their interfaces with a commonly used primer were studied by dynamic contact angle (DCA) analysis using seven liquids and by scanning probe microscopy (SPM). The contact angle data were analysed using the concepts and the equations of the surface free energy acid-base component theory. A new approach consisting in the use of a large number of probe liquids and of a proper mathematical method is proposed; it allows higher precision in the determination of the surface energy components and the work of adhesion, the reduction of possible artefacts, and the calculation of standard deviations of obtained quantities. It was found that: (i) the characteristics of the flame-treated surfaces were largely independent of the composition and morphology of the rubbery dispersed phase; (ii) the flaming effect was better shown by receding angles and the observed hysteresis allowed a quantitative evaluation of the surface heterogeneity induced by the flame process; and (iii) the flame treatment induced fragmentation of the macromolecules with the production of fragments, soluble in all the test liquids, depending on their surface tension and their acid-base character, as shown by repeated DCA immersions. A comparison has been made with the ASTM 2578-67 method ('swab'). The SPM analysis, both in contact and in 'force spectroscopy' modes, confirmed the surface model obtained by the DCA data.

 

<-- 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-->