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 57 of 76, ordered by
 

1105. Schubert, G., and O. Plassmann, “Shedding a new light on corona-treated alu-foil,” in 2004 PLACE Conference Proceedings, TAPPI Press, Sep 2004.

560. Schuelke, G.W., “Corona treatment: troubleshooting your system,” in 1987 Polymers, Laminations and Coatings Conference Proceedings, 217-219, TAPPI Press, Aug 1987.

1407. Schuelke, G.W., “Modern trends in corona treating,” in 1984 Polymers, Laminations and Coatings Conference Proceedings, 249+, TAPPI Press, Aug 1984.

561. Schultz, J., K. Tsutsumi, and J.B. Donnet, “Surface properties of high-energy solids, I. Determination of the dispersive component of the surface free energy of mica and its energy of adhesion to water and n-alkanes,” J. Colloid and Interface Science, 59, 272-276, (1977).

A method for measuring the dispersive part of the surface free energy γsD of a high-energy solid, and its interaction energy with water and n-alkanes, WSL, has been developed. It is based on the measurement of the contact angle of water on the solid under n-alkanes. Muscovite mica was chosen as a model high surface energy solid. The results obtained for γsD and WSL of mica are in good agreement with the results obtained by other techniques. The present method can be considered to be applicable for other solids.

562. Schultz, J., K. Tsutsumi, and J.B. Donnet, “Surface properties of high-energy solids, II. Determination of the nondispersive component of the surface free energy of mica and its energy of adhesion to polar liquids,” J. Colloid and Interface Science, 59, 277-282, (1977).

A method of determining the polar term of the adhesion energy of several liquids to a high-energy solid, ISLP, has been developed, based on the measurement of the contact angle of water on a solid in a liquid medium. The ISLP values for mica are found to be a linear function of the square root of the polar term of the surface free energy of liquids. This finding agrees with the suggestion that the polar term of the energy of adhesion may be represented by the geometric mean of the polar term of the surface free energy of a solid and a liquid. The slope of the straight line provides the value of γSP = 90 ergs/cm2 for the polar term of the surface free energy of mica. The results were compared with those obtained by a cleavage method and also discussed in terms of each component of the surface free energy of mica. The present method is useful for the determination of the polar part of the energy of adhesion of a high-energy solid to liquids, and its surface free energy.

326. Schultz, J., and M. Nardin, “Determination of the surface energy of solids by the two-liquid-phase method,” in Modern Approaches to Wettability: Theory and Applications, Schrader, M.E., and G.I. Loeb, eds., 73-100, Plenum Press, Oct 1992.

The surface free energy of solids is a characteristic parameter that determines most of the surface properties such as adsorption, wetting, adhesion, etc. The surface energetics of solids may be characterized by measurement of contact angles of different liquids. Nevertheless, the calculation of surface free energy from contact angle measurements has been the subject of much controversy. Indeed, this characteristic of a solid cannot be measured directly because of elastic and viscous restraints of the bulk phase, which necessitate indirect methods.

751. Schultz, J., and M. Nardin, “Theories and mechanisms of adhesion,” in Adhesion Promotion Techniques, K.L. Mittal and A. Pizzi, eds., 19-34, Marcel Dekker, May 1994 (also in Handbook of Adhesive Technology, 2nd Ed., A. Pizzi and K.L. Mittal, eds., p. 53-68, Marcel Dekker, Aug 2003).

Adhesion phenomena are relevant to many scientific and technological areas and in recent years have become a very important field of study. The main application of adhesion is bonding by adhesives, which is replacing, at least partially, more classical mechanical attachment techniques such as bolting or riveting. It is considered to be competitive primarily because it saves weight, ensures better stress distribution, and offers better aesthetics because the glue line is practically invisible. Applications of bonding by adhesives can be found in many industries, particularly in advanced technological domains such as the aeronautical and space industries, automobile manufacture, and electronics. Adhesives have also been introduced in areas such as dentistry and surgery. However, adhesive joints are not the only application of adhesion. Adhesion is concerned whenever solids are brought into contact, for instance, in coatings, paints, and varnishes; multilayered sandwiches; polymer blends; filled polymers; and composite materials. Because the final performance or use properties of these multicomponent materials depend significantly on the quality of the interface that is formed between the solids, it is understandable that a better knowledge of adhesion phenomena is required for practical applications. The field of adhesion began to create real interest in scientific circles only about 50 years ago. Thus, adhesion became a scientific subject in its own right, but it is still a subject in which empiricism and technology are slightly ahead

2059. Schuman, T., B. Adolfsson, M. Wikstrom, and M. Rigdahl, “Surface treatment and printing properties of dispersion-coated paperboard,” Progress in Organic Coatings, 54, 188-197, (Nov 2005).

Paperboard was coated on a pilot scale using aqueous dispersions of styrene–butadiene (SB) copolymers in order to improve its surface characteristics (including printability) and barrier properties with regard to the transmission of water vapour. Coating the paperboard with the dispersion in two steps gave a smoother surface with a remarkable increase in gloss. The printing properties of the smoother double-coated surface were slightly better than those of the single-coated surface. Paraffin wax added to the latex dispersion reduced the water vapour transmission rate (WVTR) but had a negative effect on the printability of the board.

The effect of two commonly used surface treatment techniques (corona and plasma at atmospheric pressure) on the printing and barrier properties of dispersion-coated (containing wax) paperboard was evaluated. A fairly intense corona treatment led to an undesirable increase in the WVTR-value. A less intense corona treatment preserved the WVTR-value to a great extent, but the printability remained at an unsatisfactory level. With plasma treatment, the water vapour barrier was not impaired, and the printability of the plasma-treated dispersion-coated (wax-containing) substrate was good. It is suggested that a better result using corona treatment may be obtained by optimising the power and controlling the time between the treatment and the printing, although this was not investigated here.

1663. Schussler, J., “Ensuring that folding box seams do not burst,” VR Verpackungs-Rundschau, 56-57, (Jun 2006).

1348. Schut, J.H., “Plasma treatment: The better bond,” Plastics Technology, 38, 64-69, (Oct 1992).

1396. Schwab, F.C., et al, “Effect of resin additives on corona treatment of polyethylene,” in 1985 Polymers, Laminations and Coatings Conference Proceedings, 95, TAPPI Press, Aug 1985 (also in J. Plastic Film and Sheeting, V. 2, p. 119+. 1986).

A systematic study was made of seven common polyolefin resin stabilizers. Surface analysis techniques were used to characterize the surfaces of films containing these additives. Films were evaluated before and after corona treatment. Results of this study showed that a surprising number of additives are surface active. In some cases these additives have a dramatic effect on the surface chemistry produced by corona treatment, yet they do not affect subsequent ink adhesion. Conversely, an additive may not significantly affect the corona treatment chemistry but yet still reduce the adhesion performance of the film product.

1992. Schwartz, A.M., “Contact angle hysteresis: A molecular interpretation,” J. Colloid and Interface Science, 75, 404-408, (Jun 1980).

“Intrinsic contact angle hysteresis” is defined as hysteresis that cannot be ascribed to roughness, heterogeneity, or penetrability of the solid surface. It can be explained if we postulate that the layer of liquid immediately adjacent to the solid surface has an ordered structure similar to that of a liquid monocrystal. This structure is fluid (zero yield point in shear) in the plane of the solid surface and presents no obstacle to an increase of the solid—liquid interfacial area (advancing of the three-phase boundary line). In planes normal to the solid surface the structure has a positive yield point in shear, which prevents decrease of the solid—liquid interfacial area (receding of the three-phase line) until the yield point is exceeded by the surface pressure πSL. Mechanical stability of the system at all values of the contact angle between the “advancing” and “receding” angles θA and θR is ascribed to a continuously changing value of πSL and of the corresponding specific interfacial free energy γSL in this interval. This change reflects the elastic response in shear of the solid—liquid interfacial film in planes normal to the solid surface in this interval.

327. Schwartz, A.M., and S.B. Tejada, “Studies of dynamic contact angles on solids,” J. Colloid and Interface Science, 38, 359-375, (1972).

In forced spreading systems, three different modes of θd-V behavior have been found, each of which predominates in a different velocity range. In the lowest velocity range, with systems involving the low viscosity, low boiling, nonpolar liquid hexane, θd was found equal to θeq (the Elliott-Riddiford or Hansen-Miotto mode). In the next higher velocity range, which extends to very low velocities for all other systems studied, the behavior described by Eq. 6 (the Blake-Haynes mode) predominates. At still higher velocities, the behavior described by Eq. [9] (the Friz mode) becomes superposed on the Blake-Haynes mode, and eventually predominates up to the range where θd approaches 90° and Eq. [9] becomes inapplicable. In the Blake-Haynes mode the major force opposing advance of the liquid front is the solid-liquid interfacial viscosity. In the Friz mode it is the bulk viscosity of the liquid.

The roughness of solid surfaces has no appreciable effect on the θd-V relationship, provided the physicochemical character of the surfaces is the same and the roughness is random. If the process of roughening alters the physicochemical character the θd-V behavior of the roughened surface may differ from that of the smooth one.

There is no qualitative difference between the θd-V behavior of systems in which θeq is zero and systems in which θeq is positive.

564. Schwartz, J., “The importance of low dynamic surface tension in waterborne coatings,” J. Coatings Technology, 64, 65-73, (Sep 1992).

Low dynamic surface tension is an important factor in achieving superior film formation in water-borne coatings. Dynamic coating application methods require surfactants with low dynamic surface tensions in order to prevent defects such as retraction, crawling and cratering. Comparative basic and empirical data are presented that will demonstrate the ability of acetylenic diols to lower the dynamic surface tension of water-borne coatings and hence improve the quality of the cured film.

1121. Sciarratta, V., D. Hegemann, M. Muller, U. Vohrer, and C. Oehr, “Upscaling of plasma processes for carboxyl functionalization,” in Plasma Processes and Polymers, d'Agostino, R., P. Favia, C. Oehr, and M.R. Wertheimer, eds., 39-64, Wiley-VCH, 2005.

565. Seaman, R., “Surface preparation by corona discharge: clean, green, and cost-effective,” in Decorating Div. ANTEC 1995, Society of Plastics Engineers, 1995.

1926. Sedev, R., M. Fabretto, and J. Ralston, “Wettability and surface energetics of rough fluoropolymer surfaces,” J. Adhesion, 80, 497-520, (Jun 2004).

Hydrophobic solid surfaces with controlled roughness were prepared by coating glass slides with an amorphous fluoropolymer (Teflon® AF1600, DuPont) containing varying amounts of silica spheres (diameter 48 μm). Quasi-static advancing, θA, and receding, θR, contact angles were measured with the Wilhelmy technique. The contact angle hysteresis was significant but could be eliminated by subjecting the system to acoustic vibrations. Surface roughness affects all contact angles, but only the vibrated ones, θV, agree with the Wenzel equation. The contact angle obtained by averaging the cosines of θA and θR is a good approximation for θV, provided that roughness is not too large or the angles too small. Zisman's approach was employed to obtain the critical surface tension of wetting (CST) of the solid surfaces. The CST increases with roughness in accordance with Wenzel equation. Advancing, receding, and vibrated angles yield different results. The θA is known to be characteristic of the main hydrophobic component (the fluoropolymer). The θV is a better representation of the average wettability of the surface (including the presence of defects).

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.

1381. Seebock, R., H. Esrom, M. Charbonnier, and M. Romand, “Modification of polyimide in barrier discharge air-plasmas: Chemical and morphological effects,” Plasmas and Polymers, 5, 103-118, (Jun 2000).

Chemical and physical modifications of polyimide (PI) surfaces caused by an air plasma have been studied. The plasma-induced surface changes of PI were investigated by using a local dielectric barrier discharge (DBD) in air at atmospheric pressure and room temperature as a function of the plasma exposure time and plasma power, while the excitation frequency was kept constant at about 130 kHz. The first results obtained in this work suggest that DBDs operating in air at atmospheric pressure can be an efficient alternative plasma source for surface treatment of polymers: a short time air plasma treatment of few seconds leads to chemical and physical changes including the rise of wettability, surface oxidation, and enhancement of surface roughness. Therefore, this simple kind of dry surface treatment seems to be an effective, low cost method for production of well-adhering subsequent layers such as metal films, paints, glues, etc. on DBD pretreated polymers.

566. Seffins, W., “A model treatment of solid/liquid interfacial energies for non-zero contact angle systems (MS thesis),” Univ. of Texas, El Paso, 1981.

1674. Seidel, C., C. Damm, and H. Muenstedt, “Surface modification of films of various high temperature resistant thermoplastics,” J. Adhesion Science and Technology, 21, 423-439, (2007).

The influence of different surface treatments on the physical and chemical surface properties of poly(etheretherketone) (PEEK), poly(phenylenesulfide) (PPS) and a liquid crystal polymer (LCP) was studied. For all the three polymers, the adhesion strength of an adhesively-bonded copper foil could be increased significantly by a chemical etching process using chromic sulphuric acid or a low pressure air-plasma treatment. However, for LCP the enhancement of adhesion by the surface treatments was lower than for the other polymers. Peel tests were employed for determining the adhesion strength of the copper foil. The physical surface properties were investigated by laser scanning microscopy (LSM). Contact-angle measurements and X-ray photoelectron spectroscopy (XPS) provided detailed information on the chemical surface properties. The detailed XPS analyses revealed different chemical mechanisms of the surface treatments depending on the polymer investigated. In all cases an incorporation of oxygen containing groups by the surface treatments was found to be responsible for a better adhesion of the copper foil on the treated polymer films compared to the untreated.

1734. Seidel, C., H. Kopf, B. Gotsmann, T. Vieth, H. Fuchs, and K. Reihs, “Ar plasma treated and Al metallised polycarbonate: an XPS, mass spectroscopy and SFM study,” Applied Surface Science, 150, 19-33, (1999).

Ar plasma etched and Al metallised bisphenol A carbonate was analysed by mass spectroscopy, photoelectron spectroscopy (XPS), and scanning force microscopy (SFM). We mainly used a technical polymer (Makrolon 2808, Bayer) made by injection-moulding, as well as spin coated bisphenol A carbonate (n=1) and polycarbonate (PC) (n=115). The mass spectroscopy during the etching process shows the degradation of the PC in the form of carbon monoxide, carbon dioxide and methyl groups. The photoelectron spectroscopy shows in detail the surface modification after Ar plasma treatment and metallisation. The plasma induces a reduction of the carboxylic carbon (C 1s), a strong reduction of singly bonded oxygen (O 1s) and also a slight reduction of doubly bonded oxygen. After Al metallisation, a reaction of Al with the oxygen groups and an interaction with the aromatic system is documented. Ar plasma etching increases the chemical interaction of Al mainly with the aromatic carbon. The X-ray photoelectron spectroscopy of metallised PC under different initial conditions shows a strong influence of incorporated water in the PC bulk that cannot be seen by XPS on uncoated PC. The O 1s signal increases during metallisation and results in an oxidation of Al probably caused by the fact that the hydrophobic surfaces becomes hydrophillic. Temperature-dependent XPS was done on technical PC samples and on spin coated samples (n=1, n=115) and supports the influence of the bulk state for the Al–PC interface. For n=1 carbonate, a diffusion of Al into the PC volume was observed. The SFM measurements showed a roughening effect on the nanometer scale even after short treatment times. Al can be seen as a weakly bound cluster on the virgin PC, and if a pre-etching is done, Al seems to grow as a good wetting film. The adhesion force of Al films on PC without any influence of the volume can be explained by the chemical bonding of Al to the carboxylic and aromatic systems. The adhesion can be increased by plasma pre-treatment. A breakdown of the adhesion on technical PC is probably induced by a reaction of Al with mobile intercalated gas, that is enriched near the surface after Al coating.

764. Seifert, A.M., “The spinning drop tensiometry,” in Drops and Bubbles in Interfacial Research, Mobius, D., and R. Miller, eds, 187-238, Elsevier, Jun 1998.

The spinning drop technique (SDT) has been developed to measure extremely low interfacial tensions (from 10 -6 mNm -1 to 10 mNm -1). It uses profile analysis of deformed droplets similar to the pendent drop method. Unlike in pendent drop experiments, where the droplets are deformed by the gravitational force, SDT is based on the balance of centrifugal and interracial forces in rapidly rotating systems. Apart from purely tensiometric applications SDT has been found to be a versatile tool for surface and interface science. It allows the study of adsorption phenomena and even permits the “simulation” of spontaneous structure formation processes, e.g., the break-up of liquid threads and the coalescence of droplets. This chapter reviews both standard and non-standard SDT applications. After a brief description of basic principles and properties, the equilibrium properties of a rotating drop, i.e., its shape and its stability, are considered in detail. Experimental aspects of SDT: Both commercial and laboratory SDT set-ups are introduced. Problems arising from sample preparation (particularly in the case of highly viscous polymers) and the determination of the droplet dimensions are discussed.

2834. Seitz, V., K. Azrt, S. Mahnel, C. Rapp, S. Schwaminger, M. Hoffstetter, E. Wintermantel, “Improvement of adhesion strength of self-adhesive silicone rubber on thermoplastic substrates - Comparison of atmospheric pressure plasma jet (APPJ) and a Pyrosil flame,” Intl. J. Adhesion and Adhesives, 66, 65-72, (Apr 2016).

Polymeric hard/soft combinations consisting of a rigid, thermoplastic substrate and an elastomeric component offer many advantages for plastic parts in industry. Manufactured in one step by multi-component injection moulding, the strength of the thermoplastics can be combined with sealing, damping or haptic properties of an elastomer. Bonds of self-adhesive liquid silicone rubber (LSR) on high performance thermoplastics such as polyetheretherketone (PEEK) or polyphenylene sulphide (PPS) are especially interesting e.g. for medical applications due to their outstanding resistance properties. To ensure good adhesion between the two components, surface treatments from an atmospheric pressure plasma jet (APPJ) and a Pyrosil® flame are applied. Chemical changes on the thermoplastic surfaces are verified by water contact angle measurement (CA) and X-ray photoelectron spectroscopy (XPS). Plasma treatment causes a decline in water contact angle, indicating the formation of functional groups, especially –OH, on the surface. XPS measurements confirm the increase of oxygen on the surface. Thus, the number of functional groups on the thermoplastic surface is enlarged by plasma treatment, leading to stronger bonding to the organofunctional silanes of the self-adhesive silicone rubber. A thin layer of silanol groups is created by the Pyrosil® flame on the thermoplastic substrates, which could be verified by XPS. A hydrophilic behaviour of the coated surface is noticed. Both surface modification methods lead to enhanced adhesion properties of self-adhesive LSR on thermoplastic surfaces. This is confirmed by 90°- peel tests of the injection-moulded composites leading to an increase in peel force by the applied surface modification techniques.

1332. Sell, P.J., and A.W. Neumann, “Estimation of surface and interfacial tensions of solids,” Z. Physik. Chem. Neue Folge, 41, 191-196, (1964).

2408. Selwyn, G., I. Henins, S.E. Babayan, and R.F. Hicks, “Large area atmospheric-pressure plasma jet,” U.S. Patent 6262523, Jul 2001.

1489. Semal, S., T.D. Blake, V. Geskin, M.J. de Ruijter, G. Castelein, J. de Coninck, “Influence of surface roughness on wetting dynamics,” Langmuir, 15, 8765-8770, (1999).

Using the molecular-kinetic theory of wetting, we analyze the dynamic contact angle of a sessile drop of squalane spreading spontaneously on Langmuir−Blodgett multilayer substrates (behenic acid on glass). This allows the effect of microscale roughness on the parameters appearing in the theory to be determined. In particular, it is shown that the jump frequency of liquid molecules at the wetting line decreases with microroughness, supporting the idea that surface defects induce additional pining potentials. The increase in pinning potential can be explained in terms of a linear increase in the activation free energy of wetting with increasing RMS microroughness.

2874. Sengupta, A., and H.P. Schreiber, “Surface characteristics of polyurethane adhesive formulations,” J. Adhesion Science and Technology, 5, 947-957, (1991).

The surface characteristics of a two-part polyurethane adhesive formulation, based on controlled amounts of polyol, isocyanate, and catalyst, have been studied by methods including contact angle analysis, 1R spectroscopy, and inverse gas chromatography (IGC). The response of surface properties to various cure regimes and to exposure to water has been established. IGC analyses show that the adhesive surface is mildly basic, and as first evaluated by contact angle methods, has a surface energy close to 40 mJ/m2. This is largely accounted for by dispersion forces. Following immersion in water at 60°C, however, the surface energies change, the most important effect being an increase in the non-dispersive component. FTIR spectra show that immersion in water also produces chemical changes in the surface region, likely related to enolization effects. On subsequent immersion of the adhesive surface in non-polar n-heptane, the non-dispersive component of the surface energy is again reduced, showing that surface restructuring of polyurethane chains contributes significantly to the observed surface dynamics. The magnitude of the restructuring effects was shown to vary with, but to persist for, all cure regimes applied to the formulation. The documented surface dynamics of the polymer are fully analogous to earlier results obtained for a series of two-part (soft-segment) polyurethanes. As expected, the surface dynamics in this family of polymers affect the bond strength of joints using the polyurethanes as adhesives.

567. Sengupta, K.S., and H.K. Birnbaum, “Structural and chemical effects of low-energy ion bombardment of PMMA-ODA surfaces,” J. Vacuum Science and Technology, A9, 2928-2935, (1991).

The effects of ion irradiation on polyimide surfaces have been studied using x‐ray photoemission techniques. Ion bombardment with energies in the range 0.5–2.0 keV and doses between 8×1013 and 1×1015 ions/cm2 were carried out in situ in the x‐ray photoelectron spectrometer and the chemistry of the modified surface was monitored using core level spectral changes. At low doses and energies, carbonyl groups were preferentially sputtered keeping the rest of the monomer intact. Loss of nitrogen was insignificant compared to losses of carbon and oxygen. At higher energies and doses, the polymer undergoes extensive bond scission, restructuring of various functional groups and species, together with radical and anion formation. High resolution spectra indicated a binding energy scale shift to a lower value, which increased with ion energy and dose, and which was related to the creation of a surface negative charge. The effects of exposure to moisture in the ambient on the surface charge, on the surface structure, and on the surface chemistry was studied.

1038. Seok-Keun, K., P. Sung-Chul, K. Sung-Ryong, et al, “Surface modification of polytetrafluoroethylene by Ar+ irradiation for improved adhesion to other materials,” J. Applied Polymer Science, 64, 1913-1921, (Jun 1997).

Ion Irradiation on polytetrafluoroethylene(PTFE) has been carried out to improve adhesion to metal and to adhesive cement. Argon ion was irradiated on the polymer, and amount of Ar+ was changed from 1014 ions/cm2 to l×1017 ions/cm2 at 1 keV, and 4 ml/min of oxygen gas was flowed near the polymer surface during the ion irradiation. Wetting angle was changed from 100 degree to 70 - 150 degree depending on the ion beam condition. The changes of wetting angle and effects of Ar+ irradiation in oxygen environment were explained in a view of surface morphology due to the ion beam irradiation onto PTFE and formation of hydrophilic group due to a reaction between irradiated polymer chain and the blown oxygen. Strongly enhanced adhesions were explained by interlock mechanism, formation of electron acceptor groups on the modified PTFE, and interfacial chemical reaction between the irradiated surface and the deposited materials.

2785. Seppanen, R., M. Sundin, A. Swerin, and B. Brandner, “Relation between surface energy, topography, wettability and detailed surface chemistry by spectroscopy for coated printing papers,” in 2008 Advanced Coating Fundamentals Symposium, TAPPI Press, 2008.

1554. Sesetyan, T., “Testing equipment,” Label & Narrow Web, 6, 36-43, (Jan 2001).

689. Sessler, G.M., J.E. West, F.W. Ryan, and H. Schonhorn, “Increase of gold-teflon FEP joint strength by electron bombardment,” J. Applied Polymer Science, 17, 3199-3209, (1973).

The strength of joints between Teflon FEP (Type A) and 500- to 1000-Å gold layers deposited by evaporation can be greatly increased if the Teflon surface is subjected to electron-beam bombardment prior to the evaporation process. Typically, joint strengths of about 60 kg/cm2, approaching the bulk strength of Teflon, are obtained for treatments with electron-beam energies in the range of 5 to 20 keV and intercepted charge densities of about 5 X 10−6 C/cm2. This compares with gold–Teflon joint strengths of about 10 kg/cm2 for untreated material. The increase in joint strength is believed to be primarily due to crosslinking caused by the electron bombardment. Compared to the other known treatments to improve gold–Teflon joints, the present method has the advantage that the charge-storage properties of the Teflon are not irreversibly degraded. It is possible, for example, to store charge densities up to 3 X 10−8 C/cm2, on 25-μm films treated with this method, with the same favorable charge-retention properties and thermally stimulated current characteristics as obtained for untreated Teflon.

1277. Seto, F., Y. Muraoka, N. Sakamoto, A. Kishida, and M. Akashi, “Surface modification of synthetic fiber nonwoven fabrics with poly(acrylic acid) chains prepared by corona discharge induced grafting,” Angewandte Makromolekulare Chemie, 266, 56-62, (May 1999).

The surface modification of synthetic fiber fabrics via corona discharge treatment and subsequent graft polymerization was investigated. Polyethylene (PE) nonwoven fabric and polyamide-6 (PA-6) nonwoven fabric were used as base fabrics. Acrylic acid (AAc) was graft polymerized onto the fabrics via corona discharge pre-treatment. The grafted amounts of PAAc were dependent on the grafting time, that of PA-6 being higher than that of PE. It was confirmed that the surface of the fibers constructing the fabric was fully covered with PAAc after the 20 min reaction. The surface of the PAAc grafted fabrics was characterized by X-ray photoelectron spectroscopy. The leakage of electrostatic charge from the fabric was determined and the dyeability was studied with methylene blue. The period of time in which the charge potential attenuated to 1/2 of the initial potential decreased drastically by grafting with PAAc. The grafted amount was enough for dyeing the entire fabrics.

1028. Seto, F., Y. Muraoka, T. Akagi, A. Kishida, and M. Akashi, “Surface grafting of poly(vinylamine) onto poly(ethylene) film by corona discharge-induced grafting,” J. Applied Polymer Science, 72, 1583-1587, (Jun 1999).

Poly(vinylamine) (PVAm) was grafted on a poly(ethylene) (PE) film surface via the surface graft polymerization of N-vinylformamide (NVF) and N-vinylacetamide (NVA) and the subsequent hydrolysis of those grafted polymers. The surface was characterized by X-ray photoelectron spectroscopy (XPS), contact angle, moisture absorption, and the leakage of electrostatic charge from the films. PNVF and PNVA were introduced onto the surface of the PE film successfully, in spite of the fact that the initiator for polymerization was a peroxide group. The grafted amounts of PNVF and PNVA were dependent on the grafting time. A PVAm-grafted surface was obtained via the hydrolysis of the grafted PNVF. The grafted-PNVA was not hydrolyzed under mild hydrolysis. The obtained PVAm-grafted surface appeared to be useful for various applications, such as protein immobilization or chemical modification. © 1999 John Wiley & Sons, Inc. J Appl Polym Sci 72: 1583–1587, 1999
https://onlinelibrary.wiley.com/doi/abs/10.1002/%28SICI%291097-4628%2819990620%2972%3A12%3C1583%3A%3AAID-APP11%3E3.0.CO%3B2-0

2900. Seveno, D., A. Vaillant, R. Rioboo, H. Adao, J. Conti, and J. DeConinck, “Dynamics of wetting revisited,” Langmuir, 25, 13034-13044, (Oct 2009).

We present new spreading-drop data obtained over four orders of time and apply our new analysis tool G-Dyna to demonstrate the specific range over which the various models of dynamic wetting would seem to apply for our experimental system. We follow the contact angle and radius dynamics of four liquids on the smooth silica surface of silicon wafers or PET from the first milliseconds to several seconds. Analysis of the images allows us to make several hundred contact angle and droplet radius measurements with great accuracy. The G-Dyna software is then used to fit the data to the relevant theory (hydrodynamic, molecular-kinetic theory, Petrov and De Ruijter combined models, and Shikhmurzaev’s formula). The distributions, correlations, and average values of the free parameters are analyzed and it is shown that for the systems studied even with very good data and a robust fitting procedure, it may be difficult to make reliable claims as to the model which best describes results for a given system. This conclusions also suggests that claims based on smaller data sets and less stringent fitting procedures should be treated with caution.

2100. Severini, F., L. Di Landro, L. Galfetti, L. Meda, G. Ricca, and G. Zenere, “Flame surface modification of polyethylene sheets,” Macromolecular Symposia, 181, 225-244, (May 2002).

High density polyethylene sheets 2 mm thick were flame treated to modify the surface properties. Sheets treated using a flame with air to gas (methane) ratio ∼ 10:1 at different distances between the inner cone tip of the flame and the polymer surface were investigated. Grafting of selected monomers as maleic anhydride, acrylamide and glycidyl methacrylate was attempted by flame treatment of sheets covered with a monomer layer. Good grafting results were obtained with acrylamide and maleic anhydride. The surface temperature-time dependence during the flame treatment was measured with a high resolution thermocouple. Scanning Electron Microscopy (SEM) allowed evidencing a modified thickness of about 120 μ. The chemical surface modification was studied by X ray Photoelectron Spectroscopy (XPS) and Diffuse Reflectance Infrared Fourier Transform Spectroscopy (DRIFT). The hydroxyl, carbonyl and carboxyl content was measured after derivatization with reagents containing an elemental tag to facilitate XPS analysis of surface functional groups. In comparison to the untreated polyethylene, wetting tension and contact angle of the flamed materials showed a strong variation. This variation was almost independent of the distance between the flame and the polymer surface. Adhesion between treated polyethylene and a polyurethane adhesive was determined using T-peel test measurements. High adhesion levels were found with flame treated polyethylene at 5 mm distance. XPS results indicate that when adhesion is high, the hydroxyl is in excess compared to the other measured functions, i.e. carbonyl and carboxyl species.

2333. Severn, I.D., and S.L. Burring, “The wetting properties of lithographic printing surfaces,” in Wetting, Spreading and Adhesion, J.F. Padday, ed., 403-421, Academic Press, 1978.

328. Sewell, J.H., “Polymer critical surface tensions,” Modern Plastics, 48, 66-72, (Jun 1971).

329. Shafrin, E.G., and W.A. Zisman, “Constitutive relations in the wetting of low energy surfaces and the theory of the retraction method of preparing monolayers,” J. Physical Chemistry, 64, 519-524, (1960).

 

<-- 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-->