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 44 of 76, ordered by
 

717. Neumann, A.W., and J.K. Spelt, eds., Applied Surface Thermodynamics, Marcel Dekker, Jun 1996.

2049. Drummond, C.J., G. Georgaklis, and D.Y.C. Chen, “Fluorocarbons: Surface free energies and van der Waals interaction (letter),” Langmuir, 12, 2617-2621, (May 1996).

Surface free energies have been calculated for solid fluorocarbon materials by employing a method that utilizes dielectric data and theoretical predictions of van der Waals (dispersion) interactions. Excellent agreement between the results of direct force measurements and those of the theory for retarded van der Waals interactions supports the methodology. Two relatively new fluorocarbon polymers have been identified as having the lowest known surface free energies of all bulk homogeneous polymeric solids. This study provides confirmation that estimates of solid surface free energies based on contact angle measurements with dispersive organic liquids depend on the dielectric properties of both the liquids and the solid.

1467. Chan, C.-M., T.-M. Ko, and H. Hiraoka, “Polymer surface modification by plasmas and photons,” Surface Science Reports, 24, 1-54, (May 1996).

Polymers have been applied successfully in fields such as adhesion, biomaterials, protective coatings, friction and wear, composites, microelectronic devices, and thin-film technology. In general, special surface properties with regard to chemical composition, hydrophilicity, roughness, crystallinity, conductivity, lubricity, and cross-linking density are required for the success of these applications. Polymers very often do not possess the surface properties needed for these applications. However, they have excellent bulk physical and chemical properties, are inexpensive, and are easy to process. For these reasons, surface modification techniques which can transform these inexpensive materials into highly valuable finished products have become an important part of the plastics and many other industries. In recent years, many advances have been made in developing surface treatments to alter the chemical and physical properties of polymer surfaces without affecting bulk properties. Common surface modification techniques include treatments by flame, corona, plasmas, photons, electron beams, ion beams, X-rays, and γ-rays.

Plasma treatment is probably the most versatile surface treatment technique. Different types of gases such as argon, oxygen, nitrogen, fluorine, carbon dioxide, and water can produce the unique surface properties required by various applications. For example, oxygen-plasma treatment can increase the surface energy of polymers, whereas fluorine-plasma treatment can decrease the surface energy and improve the chemical inertness. Cross-linking at a polymer surface can be introduced by an inert-gas plasma. Modification by plasma treatment is usually confined to the top several hundred ångströms and does not affect the bulk properties. The main disadvantage of this technique is that it requires a vacuum system, which increases the cost of operation.

Thin polymer films with unique chemical and physical properties are produced by plasma polymerization. This technology is still in its infancy, and the plasma chemical process is not fully understood. The films are prepared by vapor phase deposition and can be formed on practically any substrate with good adhesion between the film and the substrate. These films, which are usually highly cross-linked and pinhole-free, have very good barrier properties. Such films find great potential in biomaterial applications and in the microelectronics industry.

Very high-power microwave-driven mercury lamps are available, and they are used in UV-hardening of photoresist patterns for image stabilization at high temperatures. Other applications of UV irradiation include surface photo-oxidation, increase of hydrophilicity, and photocuring of paintings.

Pulsed UV-lasers are used in surface modification in many areas. Pulsed UV-laser irradiation can produce submicron periodic linear and dot patterns on polymer surfaces without photomask. These interference patterns can be used to increase surface roughness of inert polymers for improved adhesion. These images can also be transferred to silicon surfaces by reactive ion etching. Pulsed laser beams can be applied to inert polymer surfaces for increased hydrophilicity and wettability. Polymer surfaces treated by pulsed UV-laser irradiation can be positively or negatively charged to enhance chemical reactivity and processability. Pulsed UV-laser exposures with high fluence give rise to photoablation with a clean wall profile. There are many other practical applications of laser photoablation, including via-hole fabrication, and diamond-film deposition. The present review discusses all these current applications, especially in the biomedical and microelectronics areas.

1366. Comyn, J., L. Mascia, G. Xiao, and B.M. Parker, “Plasma-treatment of polyetheretherketone (PEEK) for adhesive bonding,” Intl. J. Adhesion and Adhesives, 16, 97-104, (May 1996).

Polyetheretherketone (PEEK) has been treated with oxygen-, air-, argon- and ammonia-plasmas, which greatly improve adhesion to an epoxide film adhesive. Treated surfaces can be stored under laboratory conditions for up to 90 days without significant loss of the improved adhesion properties. Contact angle measurements show that the surface energy of PEEK is much increased by plasma-treatment. X-ray photoelectron spectroscopy shows that the plasmas increase the amounts of oxygen and in some cases the amounts of nitrogen, and that new surface groups include -OH and -CO-. Wiping treated surfaces with acetone can reverse the effects of plasma-treatment.

1074. Tavakoli, S.M., and S.T. Riches, “Laser surface modification of polymers to enhance adhesion, I: Polyolefins,” in Antec '96 Vol. 1, 1219-1224, Society of Plastics Engineers, May 1996.

806. Shi, M.K., A. Selmani, L. Martinu, E. Sacher, M.R. Wertheimer, and A. Yelon, “Fluoropolymer surface modification for enhanced evaporating,” in Polymer Surface Modification: Relevance to Adhesion, K.L. Mittal, ed., 73-86, VSP, May 1996.

350. Stobbe, B.D., “Corona treatment 101: Understanding the basics from a narrow web perspective,” Label & Narrow Web Industry, 1, 33-36, (May 1996).

2898. Drelich, J., J.D. Miller, and R.J. Good, “The effect of drop (bubble) size on advancing and receding contact angles for heterogeneous and rough solid surfaces as observed with sessile-drop and captive-bubble techniques,” J. Colloid and Interface Science, 179, 37-50, (Apr 1996).

Sessile-drop and captive-bubble techniques were used for contact angle measurements. The advancing and receding contact angles were measured for water and ethylene glycol at self-assembled monolayer surfaces of dodecanethiol, for water at methylated quartz surfaces, and for water at roughened polyethylene and polytetrafluoroethylene surfaces. It was found that for each technique used, sessile-drop and captive-bubble, different advancing contact angles and different receding contact angles were frequently obtained for nonideal systems with rough and heterogeneous solid surfaces. The disagreement between contact angles, as measured with the two different techniques, increased with increasing imperfection of the solid surface. Also, it was observed that solid surface roughness and heterogeneity affected a variation of the advancing and receding contact angles with drop (bubble) size. No contact angle change with respect to drop (bubble) size (in the range 1–7 mm base diameter) was observed when smooth and homogeneous solid surfaces were well prepared. It is possible that metastable states, which are responsible for the contact angle hysteresis, also affect the contact angle/drop (bubble) size relationship. These three-phase systems with sessile drop and captive bubble at heterogeneous and/or rough solid surfaces are complex because solid surface heterogeneity and roughness cause contortions in the shape of the three-phase contact line and the drop (bubble) surface in the vicinity of the three-phase contact line. These contortions may affect a variation of the internal free energy of the liquid drop (gas bubble). It is shown that a slight variation in the advancing contact angle value over a few millimeters change in drop (bubble) diameter does not guarantee a high-quality surface state. Measurements of the receding contact angles provide more information on the quality of the solid surface and they should always be included with the measurements of advancing contact angles.

1944. Feinerman, A.E., Y.S. Lipatov, and V.I. Minkov, “On the hysteresis of polymer wetting,” J. Adhesion, 56, 97-105, (Apr 1996).

The reasons for the appearance of the hysteresis of wetting are considered. The model is proposed according to which the hysteresis is the result of the orientations of molecules of wetting liquids which is preserved due to the action of surface forces even after the flow ceases.

145. Gorzynski, M.R., “Goniometer provides accurate measurement of bottle coatings,” Packaging Technology & Engineering, 5, 48-51, (Apr 1996).

169. Inagaki, N., Plasma Surface Modification and Plasma Polymerization, Technomic, Mar 1996.

985. Herranz, M., “Coextrusion and printing problems,” Plast' 21, 49, 43-45, (Feb 1996).

927. Salmaggi, H.L, “Flexo finds the answer: How does the treater roll get affected by dirt, dust, or ink, and how should it be cleaned?,” Flexo, 21, 96, (Feb 1996).

447. Dan, N., “The effect of polymer additives on the spreading of partially wetting films,” Langmuir, 40, 1101-1104, (Feb 1996).

410. no author cited, “Ceramic rollers boost corona-treating uptime,” Plastics Technology, 42, 92, (Feb 1996).

2937. no author cited, “Standard T565: Contact angle of water droplets on corona-treated polymer film surfaces,” TAPPI, 1996.

2763. Markgraf, D.A., “Corona treatment: An adhesion promoter for water-based & UV-cured printing,” in 1996 New Printing Technologies Symposium Proceedings, TAPPI Press, 1996.

1874. Niem, P.I.F., T.L. Lau, and K.M. Kwan, “The effect of surface characteristics of polymeric materials on the strength of bonded joints,” J. Adhesion Science and Technology, 10, 361-372, (1996).

The degree of roughness and the linear direction of the abrasion process operated over the adherend surface are two important design factors for the adhesive joint. Thus, in the first part of this study, the surface roughness was varied by means of different grades of abrasive paper and its effect on the joint strength was studied. An investigation involving changing the linear direction with respect to the loading direction was also carried out. These experiments were done to determine the effectiveness of the abrasion process for the pretreatment of the adherend. A significant increase in joint strength was found for the abrasion treatment. However, it was shown that different linear directions did not have any significant effect on the joint strength. In the second part of this study, thermodynamic analysis of the bonding of dissimilar polymeric materials using different adhesives in terms of their surface tension, critical surface tension, and joint strength was carried out. The aim of the study was to determine the thermodynamic criteria for maximum joint strength in bonding dissimilar materials. The results showed that the joint strength was dictated by the adherend with the lower critical surface tension. Maximum joint strength for bonding dissimilar materials is attained when the surface tension of the adhesive used is close to that of the adherend with the lower critical surface tension.

1873. Chen, H.H., and M.D. Ries, “Surface energy modification and characterization of a plasma-polymerized fluoropolymer,” J. Adhesion Science and Technology, 10, 495-513, (1996).

This paper describes two methods of modifying the surface energy of a plasma-polymerized film. One method is to use diphenylamine (DPA) to stabilize the surface energy increase of the polymer caused on exposure to air or a polar liquid. Another method is to use heptafluorobutyric anhydride (HFBA) to reduce the surface energy of aged (oxidized) film. The HFBA-treated film displays the same surface energy (20 mJ/m2) as the freshly prepared film. It is, however, much more stable than the as-polymerized film in propylene glycol. Other silylation and fluorinated esterification reagents were found to be much less effective. The changes in surface energy were caused by changes in the chemical structure. The chemical changes were analyzed by infrared (IR) spectroscopy and electron spectroscopy for chemical analysis (ESCA). These changes were either caused by oxidation of the film in air, water, and propylene glycol or were induced by fluorination of the oxidized film. The polymer used in this study is a copolymer of perfluoropropane (PFP) and 3,3,3-trifluoropropylmethyldimethoxysilane (TFPS). Other physical properties, such as solubility, thickness, coefficient of friction, adhesion, and thermal transitions of the polymer, have also been studied.

1872. Leonard, D., P. Bertrand, A. Scheuer, et al, “Time-of-flight SIMS and in-situ XPS study of O2 and O2-N2 post-discharge microwave plasma-modified high-density polyethylene and hexatriacontane surfaces,” J. Adhesion Science and Technology, 10, 1165-1197, (1996).

The O2 and O2-N2 ([N2] < 15%) post-discharge microwave plasma modifications of high-density polyethylene (HDPE) and hexatriacontane (HTC) surfaces have been studied as a function of the distance from the discharge and the gas composition. They are compared in terms of the in-situ XPS O/C ratios and C 1s components, and the ex-situ ToF-SIMS O-/CH- ratios and relative intensities of series of peaks. The results on the effect of the distance from the discharge showed a clear correlation between the in-situ XPS results and the O2 post-discharge modeling, exhibiting the double role of oxygen atoms: functionalization initialization by creating radicals (which react with molecular oxygen) and degradation due to the energy released by the oxygen atom recombination process. Specific in-situ XPS and ex-situ ToF-SIMS signatures of this in-situ degradation related to the oxygen atom recombination process were exhibited. When N2 was introduced in the plasma gas, the in-situ XPS results and the ex-situ ToF-SIMS results were very different. The in-situ functionalization decreased as a function of the N2 addition and the ex-situ functionalization exhibited a high maximum for the 5% N2-95% O2 post-discharge plasma and then decreased. Despite the absence of a complete O2-N2 post-discharge modeling, it can be assumed that there is also a maximum of the oxygen atom content for the 5% N2-95% O2 post-discharge. Thanks to the in-situ XPS and ex-situ ToF-SIMS specific signatures, it was also shown that this maximum corresponds to a low in-situ degradation effect. Nitrogen introduction could influence the role of oxygen atoms in such a way that there is a decrease in oxygen atom recombination processes (thus in degradation) for small N2 addition and even a decrease in oxygen functionalization initialization for further N2 incorporation in the plasma gas. No nitrogen was observed in the in-situ XPS results, whereas some ex-situ ToF-SIMS nitrogen-containing ions were observed for the O2 and O2-N2 post-discharge. However, their relative intensities followed the variation in oxidation and not the variation in N2 concentration in the plasma gas. They could be related to a post-treatment functionalization effect. Differences observed between HDPE and HTC are explained in terms of their structural differences (desorption of low molecular weight oxygen-containing fragments for HTC).

1871. Flitsch, R., and D.-Y. Shih, “An XPS study of argon ion beam and oxygen RIE modified BPDA-PDA polyimide as related to adhesion,” J. Adhesion Science and Technology, 10, 1241-1253, (1996).

Modification of polymer surfaces by changing the chemical structure, surface energy, and bonding characteristics has considerable technological importance in the area of adhesion. Reactive ion etching (RIE) and ion beam (IB) bombardment were employed to modify the surfaces of fully imidized 3,3',4,4'-biphenyl tetracarboxylic acid dianhydride-p-diaminophenyl (BPDA-PDA)-based polyimide (PI) films. These modification techniques affect only a shallow surface region, approximately 10-20 nm, and the bulk properties of the polymer are unaffected. The angle-resolved X-ray photoelectron spectroscopy (XPS) technique was used to characterize the PI surfaces modified by argon IB bombardment or oxygen RIE treatment. On the argon ion-bombarded surfaces, the XPS spectra indicate that the carbonyl and imide groups are decreased. Oxygen RIE treatment resulted in an increase in the atomic concentration of oxygen. To understand the surface aging effect, the freshly modified PI surfaces were exposed to laboratory air for 1 and 2 days. The changes in composition as a function of the depth of the modified surface region right after treatment and after aging were determined by the angle-resolved XPS technique (ARXPS). Contact angle measurements were used to determine the polar and dispersion components, the sum of which is the surface free energy. The polar component of the surface free energy shows the greatest change, with an increase of 8.0-9.4 times for both the oxygen RIE and ion beam treatments as compared with the as-cured PI surface. Aging of these modified surfaces resulted in a decrease of surface free energy as compared with the just-modified surfaces. In the case of oxygen RIE treatment, the dispersion component of the surface free energy showed little or no change from the as-cured sample. Adhesion of chromium/copper/chromium (Cr/Cu/Cr) films on PI was determined by peel strength measurements. Significant increases in peel strength, by a factor of 10-80, were shown for the modified surfaces. A good correlation between the peel strength and the experimentally determined polar component of surface energy was shown.

1805. Iyengar, D.R., S.M. Perutz, C.-A. Dai, C.K. Ober, and E.J. Kramer, “Surface segregation studies of fluorine-containing diblock copolymers,” Macromolecules, 29, 1229-1234, (1996).

A diblock copolymer of deuterated styrene and isoprene (dPS−PI) with a small volume fraction of isoprene was chemically modified to incorporate pendant fluorinated side chains (“fingers”). The composition distribution of the diblock copolymers within a high molecular weight polystyrene (PS) homopolymer was determined by forward recoil spectrometry. Surface segregation and interfacial segregation of the modified block copolymers from a polystyrene matrix are observed in as-spun films. Equilibrium segregation was achieved on annealing at 160 °C for several days. The segregation isotherms at the air−polymer interface are shown to be quantitatively described by a self-consistent mean field theory (SCMF), and these permit us to estimate an effective Flory parameter which describes the attraction of the fluorinated segments to the surface and their repulsion from the bulk. The change in the surface tension as a result of the adsorption of the block copolymers at the air−homopolymer interface was evaluated from the predictions of SCMF theory and compared with the changes in the water contact angle observed. Advancing water contact angle data are consistent with the presence of a nonuniform layer of PS, CF2, and CF3 segments on the surface of the segregated samples.

1742. Coates, D.M., and S.L. Kaplan, “Modification of polymeric surfaces with plasma,” MRS Bulletin, 21, 43-45, (1996).

As adaptable as polymeric materials are in their many applications to our daily lives, the need exists to tailor the polymer surfaces to provide even more flexibility in regard to their uses. Plasma treatments offer an unprecedented spectrum of possible surface modifications to enhance polymers, ranging from simple topographical changes to creation of surface chemistries and coatings that are radically different from the bulk polymer. Furthermore plasma treatments are environmentally friendly and economical in regard to their use of materials.

Plasma processing can be classified into at least four categories that often overlap. These are the following: (1) surface preparation by breakdown of surface oils and loose contaminates, (2) etching of new topographies, (3) surface activation by creation or grafting of new functional groups or chemically reactive, excited metastable species on the surface, and (4) deposition of monolithic, adherent surface coatings by polymerization of monomeric species on the surface. Key features of these processes will be briefly discussed, with a rudimentary introduction to the chemistries involved, as well as examples. Focus is placed on capacitively coupled radio-frequency (rf) plasmas (see Figure 1 in the article by Lieberman et al. in this issue of MRS Bulletin) since they are most commonly used in polymer treatment.

1714. Markgraf, D.A., Surface Treatment of Plastics: Technology and Applications, Technomic, 1996.

1442. Badey, J.P., E. Espuche, D. Sage, B. Chabert, Y. Jugnet, C. Batier, T.M. Duc, “Comparative study of the effects of ammonia and hydrogen plasma downstream surface treatment on the surface modification of polytetrafluoroethylene,” Polymer, 37, 1377-1386, (1996).

Polytetrafluoroethylene (PTFE) was treated with hydrogen and ammonia microwave plasmas and the effects of treatment were evaluated by means of advancing and receding contact angle measurements, X-ray photoelectron spectroscopy, secondary-ion mass spectroscopy and atomic force microscopy analysis. Hydrogen plasma downstream treatment principally leads to defluorination and creation of CC and CH groups. This surface modification results in a slight decrease of the water contact angle and a large decrease of the methylene iodide contact angle. No evolution of the surface properties occurs over a period of at least two months following treatment. Ammonia plasma downstream treatment leads to defluorination and creation of CC and CH groups, as already observed with the H2 plasma, but also to the introduction of nitrogen-containing groups. The modification produces a decrease of both water and methylene iodide contact angles. A large hysteresis is found with water contact angles due to the reorientation of the polar groups when the surface is in contact with a polar liquid. The surface modifications that result after a NH3 plasma treatment are less stable than after a H2 treatment. Nevertheless, after two days of ageing the water contact angle reaches a constant value, which is largely inferior to that of the untreated PTFE.

1330. Grundke, K., T. Bogumil, T. Gietzelt, H.-J. Jacobasch, D.Y. Kwok, A.W. Neumann, “Wetting measurements on smooth, rough and porous solid surfaces,” Progress in Colloid and Polymer Science, 101, 58-68, (1996).

The solid-vapour surface tension has been determined by contact angle measurements with polar and non-polar liquids on flat solid surfaces using Axisymmetric Drop Shape Analysis (ADSA) and by capillary penetration experiments on rough and porous solids. For smooth and inert, well prepared solid surfaces (PTFE, FC 721 on mica, FEP, PET) the plot of γlvcosΘ versus γlv yields smooth curves which are consistent with the equation of state approach to calculate solid-vapour and solid-liquid interfacial tensions. Other experimental patterns of contact angle data are caused by surface roughness and non-inert solids which may result in contact angles incompatible to Young’s equation. An alternative way to obtain the solidvapour surface tension of rough and porous solids are capillary penetration experiments. The determination of the penetration velocity of liquids into rough and porous solids yields lv coΘ versus γlv plots, which provide γsv values for these systems; K is an unknown parameter of the constant geometry of the porous solid. The application of this concept was demonstrated for a hydrophobic PTFE powder and for hydrophilic Cellulose membranes.

1311. Kwok, D.Y., and A.W. Neumann, “A simple experimental test of the Lifshitz-van der Waals/acid-bsae approach to determine interfacial tensions,” Canadian J. Chemical Engineering, 74, 551-553, (1996).

A study was conducted which showed that the Lifshitz-van der Waals / acid-base approach yielded liquid-liquid interfacial tensions which were incompatible with experimental results. The approach allowed correct prediction of interfacial tensions for only four completely miscible liquid/liquid pairs. This result calls into question traditionally held interfacial tension theories, and points up the need for caution in the application to solid/liquid contact angle systems if an approach should fail the liquid/liquid test.

1310. Sedev, R.V., J.G. Petrov, and A.W. Neumann, “Effect of swelling of a polymer surface on advancing and receding contact angles,” J. Colloid and Interface Science, 180, 36-42, (1996).

The kinetics of modification of a fluoropolymer coating (FC 722, 3M Company) during its contact with octane, dodecane, and hexadecane is studied via measurement of quasi-static (velocity independent) advancing and receding dynamic contact angles. A decrease in both angles with the time of contact between solid and liquid is observed and it is interpreted as the result of swelling of the polymer. By means of a theoretical extrapolation of the θR(t) data tot= 0, based on an equation relating θR(t) to swelling kinetics, the experimentally inaccessible receding contact angle on dry coating, θ0R, is determined. The contact angle hysteresis on such a surface, θ0A− θ0R, is found to be less than the hysteresis, θA− θR, obtained on samples that were soaked in the alkanes long enough to reach saturation. This increase is thought to be due to loosening of the polymer chains during the swelling, leading to an exposure of higher-energy segments to the nonpolar liquid and to an enlargement of the solid surface pores filled with liquid. The contact angle data are also interpreted in terms of interfacial free energies.

1228. Lee, L.-H., “Correlation between Lewis acid-base surface interaction components and linear solvation energy relationship solvatochromic alpha and beta parameters,” Langmuir, 12, 1681-1687, (1996).

In this paper, we report our unexpected finding about the correlation between Lewis acid−base surface interaction components and linear solvation energy relationship (LSER) solvatochromic parameters α and β. In 1987, van Oss, Chaudhury, and Good proposed to split the asymmetric acid−base parts of a bipolar system into separate surface tension components:  Lewis acid (electron acceptor) γ+ and Lewis base (electron donor) γ-. It was assumed that the ratio of γ+ and γ- for water at 20 °C was to be 1.0. With that ratio as a reference, the base components, γ- for other liquids, biopolymers, polymers, and solids appeared to be overestimated. Recently, we unexpectedly found a correlation for liquids between γ+ and γ-, and α (solvent hydrogen-bond-donating ability) and β (solvent hydrogen-bond-accepting ability) introduced since 1976 by Taft and Kamlet. From that correlation, we obtained a more realistic ratio for the normalized α and β values for water at ambient temperature to be 1.8 instead of 1.0. Based on this new ratio, we calculated total surface tensions for related materials at 20 °C. They are generally unchanged as expected, despite the considerable, favorable change in the γ+ and γ- values in the direction of lowering the Lewis basicity. The predicability of solubility with interfacial tension is also unaffected. For example, the sign of those negative interfacial tensions that favor solubility remains the same. In addition, the implications of other LSER parameters, e.g. Π* and δH2, on surface properties will be briefly mentioned.

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).

990. Mathieson, I., and R.H. Bradley, “Improved adhesion to polymers by UV/ozone surface oxidation,” Intl. J. Adhesion and Adhesives, 16, 29-31, (1996).

An ultraviolet-ozone oxidation process is shown to be an effective adhesion pretreatment for polyethylene (PE) and polyetheretherketone (PEEK). The data obtained indicate that the treatment gives considerable oxidation and improved wettability for PE and PEEK surface types. Treated surfaces were analysed using X-ray photoelectron spectroscopy (XPS) and water contact angles. XPS was also used to follow the chemistry and mechanism of the oxidation. Adhesion with a two-part epoxy was measured for PE and PEEK and was observed to improve significantly after pretreatment.

987. Good, R.J., S. Li Kuang, C. Hung-Chang, and C.K. Yeung, “Hydrogen bonding and the interfacial component of adhesion: Acid/base interactions of corona treated polypropylene,” J. Adhesion, 59, 25-37, (1996).

The effect of activation of the surface of polypropylene sheet, by a corona discharge, upon the contact angles of liquids and on the surface free energy parameters γLW, γ and γ, was determined. Both advancing and retreating contact angles were measured. The “acid/base” theory of the components of surface free energy was employed.

The contact angles of water and glycerol were initially lower by as much as 30°, after treatment, and that of diiodomethane was lower by about 5°. With time, the advancing angles rose, and the γ and γ parameters fell, towards the values on the untreated solids, and attained more or less steady values after 5 to 10 days. The basic component, γ, was the most strongly affected by the corona treatment; it rose, typically, from 2.2 to as high as 25 mJ/m2. The acidic component, γ, rose from zero to as high as 1.9 mJ/m2. Its decay with time was only qualitatively the same as that of γ. The retreating angles, and the corresponding energy components, were changed in the same direction, and somewhat more strongly, than were the “advancing” data.

The well-known improvement in the property of forming strong joints or adherent coatings, after corona treatment, is no doubt due to the formation of sites or areas on the polymers where hydrogen bonds can be formed. The decay of the strength of adhesion with time is, no doubt, due to the decay of these sites or areas.

975. Matousek, P., G. Kreuger, and O.-D. Hennemann, “Adhesion tests with corona-pretreated plastics,” Gummi Fasern Kunststoffe, 49, 630-631, (1996).

955. Markgraf, D.A., “Corona treater station design & construction: Meeting the blown film challenge,” in 1996 Polymers, Laminations and Coatings Conference Proceedings, TAPPI Press, 1996.

574. Sherman, P.B., “The benefits of ozone in extrusion coating,” in 1996 Polymers, Laminations and Coatings Conference Proceedings, TAPPI Press, 1996.

181. Kaczmarek, H., “Changes to polymer morphology caused by UV irradiation, I. Surface damage,” Polymer, 37, 189-194, (1996).

The influence of u.v. irradiation on the surface morphology of some commercial polymers was investigated using scanning electron microscopy (SEM). The photodegraded samples exhibited a high degree of surface damage. The formation of cracks and holes resulting from the degradation and evolution of volatile products was observed. In polymers undergoing photocrosslinking, the agglomeration of particles was clearly seen. Polymer photodegradation in the presence of hydrogen peroxide as an acceleration agent revealed its etching action.

1946. Fritz, J.L., and M.J. Owen, “Hydrophobic recovery of plasma-treated polydimethylsiloxane,” J. Adhesion, 54, 33-45, (Dec 1995).

Plasma treatment of silicone surfaces is a useful, environmentally-sound method of increasing wettability to improve adhesion. A thin, wettable silica-like layer is produced with various plasma gases such as argon, helium, oxygen and nitrogen. However, in each case the surfaces gradually recover their hydrophobicity. The silica-like layer is brittle and microcracking is evident at more severe levels of plasma treatment. The onset of cracking is a function of plasma gas, RF power, pressure and treatment time. Scanning electron microscopy has been used to characterize the cracks.

The hydrophobic recovery has been monitored by water contact angle changes. It occurs with both cracked and uncracked treated surfaces. There is an initial jump in hydrophobicity at the onset of cracking. Thereafter, the recovery of both cracked and uncracked surfaces broadly parallels each other with virtually complete recovery of original hydrophobicity within one week. These effects can be accounted for by rapid surface diffusion of low molecular weight material out of fresh cracks followed by slower bulk diffusion through the polymer matrix. Significant differences in recovery rates are also evident between different plasma gases.

1945. Brewis, D.M., and G.W. Critchlow, “Adhesion and surface analysis,” J. Adhesion, 54, 175-199, (Dec 1995).

In the last 25 years, surface sensitive analytical techniques have made a major contribution to our understanding of adhesion phenomena and problems. There are several areas where these techniques have provided important information including the identification of failure modes, the chemistry of a substrate before and after pretreatments, the stability of surfaces and interfaces, the identification of surface contaminants, the interaction across an interface and the nature of interphases. X-ray photoelectron spectroscopy (XPS or ESCA), Auger electron spectroscopy (AES) and static secondary ion mass spectrometry (SSIMS) have proved to be especially useful. Many examples of the usefulness of these techniques are given.

2395. Kusano, Y., T. Inagaki, M. Yoshikawa, S. Akiyama, and K. Naitoh, “Corona discharge surface treating method,” U.S. Patent 5466424, Nov 1995.

896. Tomasino, C., J.J. Cuomo, and C.B. Smith, “Plasma treatments of textiles,” in The Fifth Annual International Conference on Textile Coating and Laminating, W.C. Smith, ed., Technomic, Nov 1995.

 

<-- 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-->