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 61 of 76, ordered by
 

525. Markgraf, D.A., “Treatment required for printing with water-based inks,” in 1987 Polymers, Laminations and Coatings Conference Proceedings, 333-336, TAPPI Press, Aug 1987.

1983. Spelt, J.K., and A.W. Neumann, “Solid surface tension: The equation of state approach and the theory of surface tension components - theoretical and conceptual,” Langmuir, 3, 588-591, (Jul 1987).

611. no author cited, “Corona treatment improves ink adhesion,” Plastics Engineering, 43, 13, (Jul 1987).

596. Weiss, H., “Increasing the wettability of film and foil webs, II,” Paper Film & Foil Converter, 61, 74-78, (Jul 1987).

488. Ironman, R., “Corona treatment has key role for English flexible packager,” Paper Film & Foil Converter, 61, 74+, (Jun 1987).

1962. Hata, T., Y. Kitazaki, and T. Saito, “Estimation of the surface energy of polymer solids,” J. Adhesion, 21, 177-194, (Apr 1987).

The methods to estimate the surface tension of polymer solids using contact angles have been reviewed in the first part. They are classified into the following three groups depending on the theories or the equations applied: (1) the methods using the Young's equation alone, (2) the methods using the combined equation of Young and Good-Girifalco, and (3) the methods using the equations of work of adhesion. Some notes and comments are given for each method and results are compared with each other. The two-liquids method for rather high energy surface is also introduced.

Next, some new possibilities to evaluate the surface tension of polymer solids are presented by our new contact angle theory in consideration of the friction between a liquid drop and a solid surface. The advancing and receding angles of contact (θa and θr) are explained by the frictional tension γF and accordingly two kinds of the critical surface tension γC(γCa and γCr) are given.

This work has shown that one of the recommendable ways to evaluate γS is either the maximum γLV cos θa or the maximum γC using the advancing contact angle θa alone, and another way is the arithmetic or the harmonic mean of the γCa and γCr. A depiction to determine the γC such as ln(1 + cos θ0) vs. γLV with cos θ0 = (cos θ0 + cos θr)/2 has also been proposed.

913. Yasuda, H.K., ed., Plasma Polymerization and Plasma Treatment of Polymers: Applied Polymer Symposia 42, Wiley - Interscience, Apr 1987.

570. Sherman, P.B., “Corona treatment - label presses,” Converter, 24, 6-7, (Feb 1987).

610. no author cited, “Corona treatment tackles tough jobs,” Modern Plastics Intl., 17, 56-58, (Jan 1987).

366. Tirrell, M., “Polymer surface forces,” Physics Today, 40, 65-66, (Jan 1987).

198. Kogoma, M., H. Kasai, and K. Takahashi, “Wettability control of a plastic surface by CF4-O2 plasma and its etching effect,” J. Physics, 20, 147-149, (Jan 1987).

Any desired surface wettability of a plastic surface can be produced by changing the concentration of the plasma gas, which here is a mixture of oxygen and a compound which includes fluorine. In the plasma treatment, the use of a third electrode consisting of a metal mesh for ion trapping can significantly decrease the etching effect. The plastic surface wettability, given by the contact angle of a water drop, does not have any direct relationship with the surface roughness due to etching in this experiment.

2783. Aspler, J.S., S. Davis, and M.B. Lyne, “The surface chemistry of paper in relation to dynamic wetting and sorption of water and lithographic fountain soutions,” J. Pulp and Paper Science, 13, 355-360, (1987).

2777. Kinloch, A.J., “Interfacial contact,” in Adhesion and Adhesives: Science and Technology, 18-55, Springer, 1987.

1832. Strobel, M., S. Corn, C.S. Lyons, and G.A. Korba, “Plasma fluorination of polyolefins,” J. Polymer Science Part A: Polymer Chemistry, 25, 1295-1307, (1987).

ESCA and contact angle measurements were used to characterize the surfaces of Polyethylene and polypropylene films exposed to SF6, CF4, and C2F6 plasmas. None of these gases polymerized in the plasma. However, all plasma treatments grafted fluorinated functionalities directly to the polymer surfaces. SF6 plasmas graft fluorine atoms to a polyolefin surface. CF4 plasmas also react by a mechanism dominated by fluorine atoms, but with some contribution from CFx-radical reactions. Although C2F6 does not polymerize, the mechanism of grafting is still dominated by the reactions of CFx radicals. For all gases studied, the lack of polymerization is attributed to competitive ablation and polymerization reactions occurring under conditions of ion bombardment.

1827. van Oss, C.J., M.K. Chaudhury, and R.J. Good, “Monopolar surfaces,” Advances in Colloid and Interface Science, 28, 35-64, (1987).

Following the development of a methodology for determining the apolar components as well as the electron donor and the electron acceptor parameters of the surface tension of polar surfaces, surfaces of a number of quite common materials were found to manifest virtually only electron donor properties and no, or hardly, any electron acceptor properties. Such materials may be called monopolar; they can strongly interact with bipolar materials (e.g., with polar liquids such as water); but one single polar parameter of a monopolar material cannot contribute to its energy of cohesion. Monopolar materials manifesting only electron acceptor properties also may exist, but they do not appear to occur in as great an abundance. Among the electron donor monopolar materials are: polymethylmethacrylate, polyvinylalcohol, polyethyleneglycol, proteins, many polysaccharides, phospholipids, nonionic surfactants, cellulose esters, etc.

Strongly monopolar materials of the same sign repel each other when immersed or dissolved in water or other polar liquids. The interfacial tension between strongly monopolar surfaces and water has a negative value. This leads to a tendency for water to penetrate between facing surfaces of a monopolar substance and hence, to repulsion between the molecules or particles of such a monopolar material, when immersed in water, and thus to pronounced solubility or dispersibility. Monopolar repulsion energies can far outweigh Lifshitz-van der Waals attractions as well as electrostatic and “steric” repulsions. In aqueous systems the commonly observed stabilization effects, which usually are ascribed to “steric” stabilization, may in many instances be attributed to monopolar repulsion between nonionic stabilizing molecules. The repulsion between monopolar molecules of the same sign can also lead to phase separation in aqueous solutions (or suspensions), where not only two, but multiple phases are possible. Negative interfacial tensions between monopolar surfactants and the brine phase can be the driving force for the formation of microemulsions; such negative interfacial tensions ultimately decay and stabilize at a value very close to zero.

Strongly monopolar macromolecules or particles surrounded by oriented water molecules of hydration can still repel each other, albeit to an attenuated degree. This repulsion was earlier perceived as caused by “hydration pressure”.

A few of the relevant colloid and surface phenomena are reviewed and re-examined in the light of the influence of surface monopolarity on these phenomena.

1782. Dalal, E.N., “Calculation of solid surface tensions,” Langmuir, 3, 1009-1015, (1987).

1764. Yuk, S.H., and M.S. Jhon, “Temperature dependence of the contact angle at the polymer-water interface,” J. Colloid and Interface Science, 116, 25, (1987).

Contact angle measurements on polymer hydrogels were performed at various temperatures, and we obtained the dispersive (γsd) and nondispersive (γsp) components of the surface tension of polymer hydrogel at each temperature. Utilizing the temperature dependence values of γsd and γsp, we obtained the surface entropies of polymer hydrogels. The polymer hydrogels used were isotactic and syndiotactic poly(2-hydroxyethyl methacrylate) (HEMA), poly(2-hydroxyethyl methacrylate + aminoethyl methacrylate) (HEMA + AEMA), poly(2-hydroxyethyl methacrylate + N-vinyl pyrrolidone) (HEMA + VP), poly(2-hydroxyethyl methacrylate + methyl methacrylate) (HEMA + MMA), poly(2-hydroxyethyl methacrylate + methoxyethyl methacrylate) (HEMA + MEMA), and poly(2-hydroxyethyl methacrylate + methoxyethoxyethyl methacrylate) (HEMA + MEEMA), respectively. The contact angles were also measured by using droplets of water-immiscible liquids under conditions in which the polymer hydrogel was fully hydrated.

1744. Gerenser, L.J., “X-Ray photoemission study of plasma modified polyethylene surfaces,” J. Adhesion Science and Technology, 1, 303-318, (1987).

X-Ray photoelectron spectroscopy (XPS) was used to determine plasma induced chemical species on the surface of polyethylene (PE). Argon plasmas were found to have no detectable chemical effect on the PE surface, whereas oxygen and nitrogen plasmas created new chemical species which altered the chemical reactivity of the PE surface. Oxygen plasmas were found to react more rapidly with the PE surface than nitrogen plasmas. The degree of incorporation of new chemical species in the near surface region is approximately 20 at. % at the saturation level for both oxygen and nitrogen plasmas. Core level spectra for oxygen and nitrogen plasma treated PE suggest the formation of primarily C-O-C species in the former and C-N species in the latter. Angle-resolved XPS measurements indicate that the depth of incorporation of new chemical species is confined to the top 25 A.

1615. Shanahan, M.E.R., and P.G. deGennes, “Equilibrium of the triple line solid/liquid/fluid of a sessile drop,” in Adhesion 11, K.W. Allen, ed., 71-81, Elsevier, 1987.

1614. Carre, A., S. Moll, J. Schultz, and M.E.R. Shanahan, “A novel interpretation of contact angle hysteresis on polymer surfaces,” in Adhesion 11, K.W. Allen, ed., 82-96, Elsevier, 1987.

1597. Gaydos, J., and A.W. Neumann, “The dependence of contact angles on drop size and line tension,” J. Colloid and Interface Science, 76, 120+, (1987).

We report contact angle measurements of five n-alkanes, dodecane through hexadecane, on Teflon (FEP) as a function of drop size. In all cases the contact angles decreased by approximately 5° when the drop size was increased from approximately 1 to 4 mm contact radius. A complete solution to the problem of mechanical equilibrium of a sessile drop on a solid surface indicates that the dependence of the contact angle on drop size may be explained by including the effect of line tension in the Young equation. The observed drop size dependence of the contact angle yields a line tension of (2.5 ± 0.5) × 10−6 J/m. Over the range of n-alkanes studied it was not possible to discern any dependence of the line tension on liquid surface tension.

556. Sarabia, A., “Plasma surface treatment of poly(phenyl sulfide) and poly(etheretherketone) prior to adhesive bonding (MS thesis),” MIT, 1987.

526. Markgraf, D.A., “Corona treatment and water-borne technology: Implications for converting polyolefin substrates,” American Ink Maker, 65, 26-62, (1987).

512. Lee, H.Y., “Characterization of surface structure and properties in oriented polymers (MS thesis),” Univ. of Connecticut, 1987.

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.

450. Derjaguin, B.V., N.V. Churaev, and V.M. Muller, Surface Forces, Plenum Press, 1987.

439. Cherry, B.W., and P.B. Evely, “The interaction parameter and the strength of adhesive joints,” J. Adhesion, 22, 171-182, (1987).

A “blister test” technique has been used to determine the fracture surface energy of a range of adhesive joints formed using a polyurethane adhesive and a range of solid substrates. For each adhesive pair examined the work of adhesion was calculated from the contact angles formed by liquids for which the polar and dispersion force components of the surface tension are known. For each adhesive pair, the solubility parameter of adhesive and substrate were determined by swelling measurements in a range of liquids. Although cohesive failure of the joints was observed for some of the pairs for which the solubility parameters were matched, this was not true for all such pairs and an explanation of this behaviour has been sought in a new calculation of the volume interaction component of the molecular interaction parameters.

434. Chang, C.-A., “Enhanced Cu-teflon adhesion by presputtering treatment: effect of surfcae morphology changes,” Applied Physics Letters, 51, 1236-1238, (1987).

The recently observed enhancement in adhesion between Cu and Teflon due to a presputtering treatment of Teflon prior to the Cu deposition is analyzed. The sputtering treatment resulted in a morphology change of the Teflon, with the deposited Cu showing similar textures, and changes in chemical bondings between the two. A simple geometric model is used to analyze the contribution from the morphology changes to the observed peel strength. It is shown that, for a finite chemical bonding, an appreciable contribution to the peel strength is possible from the morphology changes observed.

398. Yasuda, H.K., D.L. Cho, and Y.-S. Yeh, “Plasma-surface interactions in the plasma modification of polymer surfaces,” in Polymer Surfaces and Interfaces, Feast, W.J., and H.S. Munro, eds., 149-162, John Wiley & Sons, 1987.

314. Schmidt, J.J., J.A. Gardella Jr., J.H. Magill, and R.L. Chin, “Surface spectroscopic studies of polymer surfaces and interfaces, II. Poly(tetramethyl-P-silphenylenesiloxane/poly(dimethylsiloxane) block copolymers,” Polymer, 28, 1462-1466, (1987).

The surface region of a series of poly(tetramethyl-p-silphenylenesiloxane)poly(dimethylsiloxane) block copolymers was investigated using X-ray photoelectron spectroscopy and attenuated total reflectance Fourier transform infra-red spectroscopy. Analysis of the results shows the surface region to be equivalent to the bulk composition for all but one sample. This indicates that for all but the most crystalline samples the surface region comprises a relatively thick layer of non-crystalline amorphous domains.

261. Nuzzo, R.G., and G. Smolinsky, “Preparation and characterization of functionalized polyethylene surfaces,” Macromolecules, 17, 1013-1019, (1987).

We describe a procedure to modify the surface of polyethylene (PE) film using a combination of gas discharge and wet chemical techniques. This method generates high densities (1014-1016 cm-2) of a specific functionality, largely unaccompanied by other groups, in a 50-100-Å surface layer. The topography of the polymer surface remains unchanged after treatment and functions as an effective starting material for subsequent derivatization by standard synthetic chemical reactions. A plasma of either oxygen, water, or hydrogen is generated under comparable experimental conditions. In all cases a 1-2-s, 5-W, 0.2-Torr treatment produces about the same degree of surface modification as does longer treatment. High-resolution X-ray photoelectron spectroscopy (XPS) shows that either an oxygen or a water plasma produces a variety of oxidation products ranging from alcohols to carboxylic acids. Chromic acid oxidizes the plasma-oxidized surface further to give high densities of carboxylic acid groups which can be readily converted to acid chlorides and derivatized. Borane/tetrahydrofuran reduces the plasma-oxidized surface to give alcohols which can be esterified readily. Contact-angle measurements show that the water-plasma-treated PE surface has a higher surface free energy (γs ∼ 62 dyn/cm) than the oxygen-plasma-treated surface (γs ∼ 50 dyn/cm). A 5-s, ambient-temperature, 0.2-Torr, 2-W hydrogen plasma generates a significant number of quenchable radical sites. XPS spectra of this treated surface, exposed to either nitric oxide or nitrosotrifluoromethane, show that both compounds bond to the surface.

163. Hook, Y.J., J.A. Gardella, Jr., and L. Salvati Jr., “Multitechnique surface spectroscopic studies of plasma-modified polymers, II. Water/argon plasma-modified polymethylmethacrylate/polymethylacrylic acid copolymers,” J. Materials Research, 2, 132-142, (1987).

Results from the x-ray photoelectron spectroscopy (XPS or ESCA), ion scattering spectroscopy (ISS or LEIS), and Fourier transform infrared spectrometry (FTIR) analyses are presented for unmodified and modified poly (methylmethacrylate)/poly (methacrylic acid) (PMMA/PMAA) copolymer films. Analyses of the unmodified PMMA/PMAA copolymer series, via ESCA, ISS, and FTIR, established the surface composition and functionality of the PMMA/PMAA copolymers before the H2O/Ar rf-plasma treatment was employed. The ESCA, ISS, and FTIR analysis of these modified PMMA/PMAA copolymers show that surface modification over a limited depth (50–200 Å) has occurred. The composition, bonding, and functionality changes of the surfaces are discussed. A two-step modification mechanism (surface reduction of the PMMA/PMAA copolymer followed by H2O adsorption) is proposed to interpret the spectroscopic results.

162. Hook, Y.J., J.A. Gardella, Jr., and L. Salvati Jr., “Multitechnique surface spectroscopic studies of plasma-modified polymers, I. Water/argon plasma-modified polymethylmethacrylates,” J. Materials Research, 2, 117-131, (1987).

Results from x-ray photoelectron spectroscopy (XPS or ESCA), low-energy ion scattering spectrometry (LEIS or ISS); and Fourier transform infrared spectroscopy (FTIR) analyses are presented for unmodified and modified poly (methylmethacrylate) (PMMA) polymer films. Analysis of the unmodified PMMA polymers (isotactic, syndiotactic, and atactic) via ESCA, ISS, and FTIR, established the surface composition, bonding, and functionality before the modification was employed. An rf-plasma glow discharge created from an Ar/H2gas mixture at different exposure times and power levels was used to treat the polymer surface. Subsequent ESCA, ISS, and FTIR analyses of these modified PMMA's show the effects of surface modification in terms of a model of structural differences, over a limited depth (50–100 Å). The composition and functionality changes of the resulting surfaces are discussed with respect to proposed mechanisms of the plasma reaction and differences in tacticity of the reactant. A two-step reaction mechanism involving reactive decarboxylation/reduction followed by H2O adsorption is proposed to understand the spectroscopic results.

132. Gilberg, G., “Polymer surface characterization: an overview,” J. Adhesion, 21, 129-154, (1987).

The properties of a polymer surface can be decisive for the function of the polymer. Both in the assessment of existing polymer systems and the development of new ones the possiblity of characterizing the chemical composition and structure of the polymer surface becomes important. Various instruments and chemical methods used to characterize polymer surfaces and interfaces are reviewed. The pros and cons of electron spectroscopy for chemical analysis and derivatization schemes to enhance the detectability of functional groups, Fourier transform infrared spectroscopic methods (ATR, RIFT, PAS, micro), Raman spectroscopy, static secondary ion mass spectrometry, high resolution solid state nuclear magnetic resonance, microscopy and contact angle measurements are presented. The importance of the fact that the polymer surface can undergo comparatively rapid reorientations leading to a changed surface chemistry is discussed and exemplified.

127. Gerenser, L.J., J.M. Pochan, J.F. Elman, and M.G. Mason, “Effect of corona discharge treatment of poly(ethylene terephthalate) on the adsorption characteristics of the fluorosurfactant Zonyl FSC as studied via ESCA and surface energy measurements,” Langmuir, 2, 765-770, (1987).

117. Garbassi, F., E. Occhiello, F. Polato, and A. Brown, “Surface effect of flame treatments on polypropylene (Part 2),” J. Materials Science, 22, 1450-1456, (1987).

Static secondary ion mass spectroscopy (fast atom bombardment mass spectroscopy), (SIMS (FABMS)) and Fourier transform infrared-photo-acoustic spectroscopy (FTIR-PAS) studies have been performed on samples of polypropylene subjected to different numbers of flame treatments. SIMS spectra allowed us to identify unambiguously the site of oxidation in the methyl pendant groups, because of the striking decrease in the intensity of the methyl fragment in positive-ion spectra. The behaviour of the surface concentrations of hydroxyl, formyl and carboxyl groups as a function of the number of flame treatments has also been observed, leading us to an hypothesis supporting the effectiveness of hydroxyl groups in promoting paint adhesion. FTIR-PAS spectra did not show evident changes on passing from untreated to flame-treated samples. This negative evidence is also important: it implies a limited depth of oxidation. In the light of previous XPS results and FTIR-PAS characteristics (thickness of the observed layer and sensitivity) we suggest a depth of oxidation of some 10 to 20 nm.

116. Garbassi, F., E. Occhiello, and F. Polato, “Surface effect of flame treatments on polypropylene (Part 1),” J. Materials Science, 22, 207-212, (1987).

A study of the effects of flame treatments on a high-impact polypropylene has been performed. Both physico-chemical and mechanical properties have been investigated. The surface chemical composition has been determined by XPS, while the surface tension and the polarity were obtained through contact angle measurements. A remarkable agreement in the behaviour of chemical composition and polarity has been found, emphasizing the role of carbonyl and carboxyl groups. The adhesion of treated and untreated samples to paint coatings hua been mechanically tested. The force of adhesion remains quite constant after the first flame treatment. This suggests the importance of chemical interactions of the coating with the first layers of the polymer.

93. Feast, W.J., and H.S. Munro, eds., Polymer Surfaces and Interfaces, John Wiley & Sons, 1987.

43. Briggs, D., “Analysis and chemical imaging of polymer surfaces by SIMS,” in Polymer Surfaces and Interfaces, Feast, W.J., and H.S. Munro, eds., 33-53, John Wiley & Sons, 1987.

2378. Lori, G., “Method of flame activation of substrates,” U.S. Patent 4622237, Nov 1986.

 

<-- 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-->