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 0 of 76, ordered by
 

2630. Abbott, S., “Adhesion Apps: What is NOT important in adhesion?,” Converting Quarterly, 6, 10, (Jan 2016).

2648. Abbott, S., “Adhesion Apps: What IS important in adhesion?,” Converting Quarterly, 6, 12-13, (May 2016).

2666. Abbott, S., “Adhesion Apps: Why does a higher level of cross-linking actually make adhesion weaker?,” Converting Quarterly, 6, 16-17, (Aug 2016).

2670. Abbott, S., “Adhesion Apps: Why is 'real' adhesion 'unknowable'?,” Converting Quarterly, 6, 12-13, (Nov 2016).

2689. Abbott, S., “Adhesion Apps: How does entanglement result in strong adhesion?,” Converting Quarterly, 7, 14-15, (May 2017).

2697. Abbott, S., “Adhesion Apps: How do we achieve strong adhesion with polymers that have to be weak?,” Converting Quarterly, 7, 14-15, (Jul 2017).

2702. Abbott, S., “What is the real science behind PSAs and release coatings?,” Converting Quarterly, 7, 12-13, (Nov 2017).

2996. Abdel-Fateh, E., and M. Alshaer, “Polyimide surface modification using He-H2O atmospheric pressure plasma jet-discharge power effect,” Coatings, 10, (Jul 2020).

The atmospheric pressure He- H 2 O plasma jet has been analyzed and its effects on the Kapton polyimide surface have been investigated in terms of discharge power effect. The polyimide surfaces before and after plasma treatment were characterized using atomic force microscopy (AFM), X-ray photoelectrons spectroscopy (XPS) and contact angle. The results showed that, increasing the discharge power induces remarkable changes on the emission intensity, rotational and vibrational temperatures of He- H 2 O plasma jet. At the low discharge power ≤5.2 W, the contact angle analysis of the polyimide surface remarkably decrease owing to the abundant hydrophilic polar C=O and N–C=O groups as well as increase of surface roughness. Yet, plasma treatment at high discharge power ≥5.2 W results in a slight decrease of the surface wettability together with a reduction in the surface roughness and polar groups concentrations.

1356. Abdel-Salam, M., H. Singer, and A. Ahmed, “Effect of the dielectric barrier on discharges in non-uniform electric fields,” J. Physics D: Applied Physics, 34, 1219-1234, (2001).

This paper is aimed at calculating the electric field in the point-to-plane electrode system with the plate covered with a dielectric layer. With charge accumulation on the dielectric surface by corona discharge, the field in the dielectric is increased at the expense of a decrease in the gas gap. The charge accumulation on the dielectric surface proceeds to the maximum possible value when the normal component of the surface field vanishes. With the dielectric layer fully-charged, the percentage decrease of the field in the gas gap is maximum at the dielectric surface and declines along the gap axis to vanish at the point tip. The percentage decrease of the field becomes more pronounced with the increase of the diameter of the dielectric layer. The effect of inter-electrode spacing and the dielectric layer thickness on the field distribution is investigated. An accurate method of charge simulation was used for field calculation irrespective of the thickness of the dielectric layer and the gap geometry. With ion flow along the flux lines from the stressed point to the ground plane, the field enhancement factor increases and the volume charge density decreases along the flux lines. The voltages of the ion flow threshold and corona quenching are calculated and compared with previous measurements. The method of calculation is extended to calculate how high the surface potential of the charged dielectric needs to be to trigger a micro-spark in the electrostatic discharges from grounded point electrodes.

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

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

836. Abenojar, J., R. Torregrosa-Coque, M.A. Martinez, and J.M. Martin-Martinez, “Surface modifications of polycarbonate (PC) and acrylonitrile butadiene styrene (ABS) copolymer by treatment with atmospheric plasma,” Surface and Coatings Technology, 203, 2173-2180, (May 2009).

Two engineering thermoplastic polymers (polycarbonate, PC, and acrylonitrile butadiene styrene copolymer, ABS) were treated with atmospheric plasma torch using different treatment rates (1, 5 and 10 m/min). The modifications produced by the treatment were analysed by contact angle measurements, XPS, SEM and ATR-IR spectroscopy. Particular emphasiswas placed on the ageing (up to 30 days) after atmospheric plasma treatment on both polymers. The slower the atmospheric plasma treatment, the greater the wettability of the treated polymers. The decrease in water contact angle was mainly ascribed to a significant increase in oxygen content due to the formation of carboxylic and hydroxyl groups and a decrease in the carbon content on the polymer surfaces. After natural ageing, there was an increase in the water contact angle, although the values of the untreated polymer surface were never reached.

1783. Ada, E.T., O. Kornienko, and L. Hanley, “Chemical modification of polystyrene surfaces by low-energy polyatomic ion beams,” J. Physical Chemistry B, 102, 3959-3966, (Apr 1998).

The chemical modification of polystyrene surfaces by low-energy (10−100 eV) SF5+, C3F5+, and SO3+ ions was studied by X-ray photoelectron spectroscopy and two-laser ion trap mass spectrometry. The mechanism of fluorination was found to be dissimilar for SF5+ and C3F5+ ions in this energy range at fluences of 1014−1016 ions/cm2. SF5+ was found to induce fluorination of the polymer surface by grafting reactive F atoms upon dissociation at impact. SFn fragments were not found to be grafted or implanted into the polymer. Sulfur was detected on the polymer surface only at incident energies above 50 eV and was found to be sulfidic in nature. In contrast, C3F5+ ions induced grafting of both reactive F atoms and molecular CmFn fragments from the dissociation of the incident projectile. Larger proportions of highly fluorinated sites and thicker fluorocarbon layers were found for C3F5+ at all energies and fluences. A variety of aliphatic and aromatic fluorine bonding environments were detected on both SF5+ and C3F5+ modified polystyrene surfaces.

15. Adelsky, J., “Effects of corona pre-treatment on surface characteristics of oriented polypropylene film,” TAPPI J., 72, 181-184, (Sep 1989).

2864. Agbezuge, L., and F. Wieloch, “Estimation of interfacial tension components for liquid-solid system from contact angle measurements,” J. Applied Polymer Science, 27, 271-275, (Jan 1982).

A technique has been developed for estimating the hydrogen bonding and London dispersion force components of liquid surface tension and solid surface free energy levels. The technique relies on (a) measuring contact angles generated by sessile drops of liquids on solids and (b) performing calculations based on theories of thermodynamic wetting of solids by liquids. The technique is used to estimate interfacial force components of certain liquids and papers typical of those used in xerographic processing.

7. Agler, S., “Are your bottles print ready?Understanding treatments for surface tension,” ScreenPrinting, 84, 100, (Jan 1994).

2319. Ahlbrandt, A., “Corona treater for plastic film,” U.S. Patent 4533523, Aug 1985.

2416. Ahmed, Q.U., M.D. Christy, and P.A. Wallis, “Treatment of plastics containers,” U.S. Patent 6866810B2, Mar 2005.

1777. Ahn, D., and K.R. Shull, “Effects of substrate modification on the interfacial adhesion of acrylic elastomers,” Langmuir, 14, 3646-3654, (1998).

Axisymmetric adhesion tests are used to probe the adhesion between a carboxylated, poly(n-butyl acrylate) (PNBA) elastomer and a variety of substrates. The elastomer is in the form of a hemispherical cap, which is pressed against a flat substrate to give a circular contact area. A standard fracture mechanics approach is used to relate the applied load and the radius of the contact area to the energy release rate, 𝒢, which can be viewed as an adhesion energy. For a given substrate, 𝒢 is a function of the crack velocity, v, defined as the time derivative of the contact radius. The contact pressure does not appear to play a role in the development of the adhesive forces, and the specific relationship between 𝒢 and v depends on the substrate. In all cases this relationship can be adequately described by the empirical expression, 𝒢(v) = 𝒢0(1 + (v/v*)0.6). Values of 𝒢0 are within a factor of about 2 of the expected thermodynamic work of adhesion, but values of v* vary from 5.5 to 215 nm/s, depending on the detailed nature of the substrate. Decreases in the adhesion energy, as quantified by a decrease in 𝒢0 and/or an increase in v*, are determined primarily by the segmental mobility of the substrate molecules. Comparison of results for free and grafted PNBA chains indicates that translational diffusion of molecules at the substrate is not required in order to substantially reduce the adhesion energy.

1193. Akishev, Y.S., M.E. Grushin, A. Napartovich, and N.I. Trushkin, “Novel AC and DC non-thermal plasma sources for cold surface treatment of polymer films and fabrics at atmospheric pressure,” Plasmas and Polymers, 7, 261-289, (Sep 2002).

Novel types of non-thermal plasma sources at atmospheric pressure based on multi-pin DC (direct current) diffusive glow discharge and AC (alternative current) streamer barrier corona have been elaborated and tested successfully for cold surface treatment of polymer films [polyethylene (PE), polypropylene (PP), polyethylene terephthalate (PET),] and polyester fabric. Results on physical properties ofdischarges mentioned and output energy characteristics of new plasma sources as well as data on after-treatment changes in wettability of films and fabrics are presented. The main goal of this study was to find out the experimental conditions for gas discharge and surface processing to achieve a remarkable wettability change for a short treatment time.

1192. Akishev, Y.S., M.E. Grushin, A.E. Monich, A.P. Napartovich, and N.I. Trushkin, “One-atmosphere argon dielectric-barrier corona discharge as an effective source of cold plasma for the treatment of polymer films and fabrics,” High Energy Chemistry, 37, 286-291, (Sep 2003).

The properties of an ac dielectric-barrier corona discharge in argon under atmospheric pressure were studied and the results of testing of this type of gas discharge in the low-temperature treatment of polymer films and fabrics for the purpose of enhancement of their wettability were reported.

1691. Al-Turaif, H., “Relationship between surface chemistry and surface energy of different shape pigment blend coatings,” J. Coatings Technology and Research, 5, 85-91, (Mar 2008).

The influence of pigment shapes and pigment blends on the surface energy was investigated and compared with the surface chemistry of pigmented latex coatings. The coatings were made of different volume ratios of two pigments: plate-like kaolin clay pigment and prismatic precipitated calcium carbonate (PCC) pigment. These were mixed together with carboxylated styrene–butadiene–acrylonitrile latex (SBA), and applied over nonabsorbent substrates as well as absorbent substrates. The composition of the surface of the coatings was investigated by X-ray photoelectron spectroscopy (XPS). Two approaches were used to estimate the total surface energy and the components of the coatings: a conventional approach—“the Kaelble approach”—and a more modern approach—“the van Oss approach.” Pigment blends with different shapes and increments caused a change in the surface chemistry and the surface energy of the latex coatings. As the prismatic PCC pigment particles increased in the kaolin/SBA coating system, the SBA latex content at the coating surface increased and the total surface energy of the coating decreased. This is valid for both nonabsorbent as well as absorbent substrates. It was found that there was a strong correlation between the surface energy and the surface composition. The surface energy of the coatings estimated by the Van Oss approach was always lower than that estimated by the Kaelble approach. Colloidal interactions between pigment–pigment and/or pigment–binder were thought to play an essential role in determining the final coating surface energy and its components. Changes in the surface latex content and the surface energy due to the different pigment blends investigated were found to fit straight-line equations.

1696. Al-Turaif, H., D.W. Bousfield, and P. LePoutre, “The influence of substrate absorbency on coating surface energy,” Progress in Organic Coatings, 49, 62-68, (2004).

The surface energy of coating layers influences their final properties such as their ability to repel or absorb fluids. Recent work has shown that the substrate, due to absorption, can alter the surface chemistry of the top coating layer. However, the influence of substrate properties on coating surface energies is not reported in the literature.Three coatings, based on a pigment and a latex binder, are applied on three different substrates that differ in terms of absorption properties. The three coatings were also modified with a soluble polymer. Contact angle measurements of three different probe fluids were measured. These contact angles were used to estimate the polar, dispersive, and total surface energy of the coating layers. Surface energies were also determined for the latex and pigments.The contact angles and surface energies of the latex films and pigments agree with the expected results. Most of the results for the coating layers agree with the reported surface chemistry of these coatings. Large pigment systems on absorbent substrates have a high contact angle and low surface energy. These results agree with the expected results based on the surface chemistry reported in past work. The results for the fine pigment system had low contact angles and high surface energies and did not agree with the expected results. The contact angles may be influenced by the surface roughness of the coatings or the expected surface energy of a heterogeneous surface may not be a simple function of the surface composition.

1695. Al-Turaif, H., W.N. Unertl, and P. LePoutre, “Effect of pigmentation on the surface energy and surface chemistry of paper coating binders,” J. Adhesion Science and Technology, 9, 801-811, (1995).

The effect of the addition of clay and TiO2 pigments on the surface energy and surface chemistry of films made from polymers used in paper coating formulations was evaluated. The polymers were carboxymethyl cellulose, polyvinyl alcohol and a protein-based polymer - all water-soluble - and two styrene-butadiene latexes of different carboxylation levels. The morphology of the surfaces was characterized by SEM examination, gloss measurement and stylus profilometry. Chemical composition was determined by EDS and XPS techniques. Surface energy and its Lifshitz-van der Waals and acid-base components were obtained from contact angle measurements using the van Oss et al. approach. Even though the addition of pigment increasingly upset the planar surface of the films, their surface chemistry and surface energy were only slightly affected over the pigmentation range studied (up to 40% by volume) and were dominated by the characteristics of the binder polymer.

2496. Ala-Kuha, A., “The influence of surface treatment on the polyolefin coating (Master's thesis),” Tampere University of Technology, Nov 2011.

1507. Alam, P., M. Toivakka, K. Backfolk, and P. Sirvio, “Dynamic spreading and absorption of impacting droplets on topographically irregular porous substrates,” Presented at ISCST 13th International Coating Science and Technology Symposium, Sep 2006.

1357. Alemskaya, O., V. Lelevkin, A. Tokarev, and V. Yudanov, “Synthesis of ozone in a surface barrier discharge with a plasma electrode,” High Energy Chemistry, 39, 263-267, (Jul 2005).

The synthesis of ozone from oxygen in a cylindrical ozonizer operating under surface discharge conditions with a plasma electrode was studied. The conditions of ozone synthesis were optimized. The dependence of ozone concentration and specific energy consumption on gas pressure in the plasma electrode and on distance between the coils of a corona electrode was determined. The results were compared with data obtained with the use of classical surface barrier discharge.

2829. Alexander, C.S., M.C. Branch, M. Strobel, M. Ulsh, N. Sullivan, and T.Vian, “Application of ribbon burners to the flame treatment of polypropylene films,” Progress in Energy and Combustion Science, 34, 696-713, (Dec 2008).

This article focuses on recent advances in the understanding of industrial gas burners. Ribbon burners have been chosen as the focus of the review because of the advantages presented by the burner arrangement and burner performance. The ribbon burner configuration, because of its ability to provide large flame surface and flame stabilization, has a large range of stability as flow rate, equivalence ratio and reactant gas composition are varied. Discussed in detail is the application of ribbon burners in the surface modification, or flame treatment, of polymer films to increase the wettability of a polymer surface. Optimum treatment requires a spatially homogeneous post-flame reaction zone even with burners up to 3 m in length. For methane/air flames, the optimum equivalence ratio is near 0.93 where the active oxidizing-species concentration near the surface is a maximum. Chemical kinetic models of the impinging flame and surface oxidation chemistry of a polymer film are also discussed. The model predictions are in good qualitative agreement with the available understanding of the flame variables affecting surface treatment and the expected oxidized species on the polymer surface.

1490. Allain, C., D. Ausserre, and F. Rondelez, “A new method for contact angle measurement of sessile drops,” J. Colloid and Interface Science, 107, 5-13, (1985).

A new method to measure the contact angle θ of sessile drops deposited onto solid substrates is presented. The crux of the method is to use the drop as a convex mirror for large, collimated, incident laser beams. The reflected light is intercepted on a screen in the far field. It appears as a bright circular spot, the diameter of which is simply related to the beam angular divergence 4θ. Overall measuring accuracies can be made better than 0.1° for θ angles less than 45θ. Reproducibility is also excellent because of the natural averaging over the entire three-phase line boundary of the drop. It is estimated to be ±0.25°. An extension of the method, valid for all angles less than 90°, is to use the drop as a plano-convex lens. This is of course applicable only if the solid substrate is optically clear. The refracted light is then intercepted on a screen and analyzed to yield the contact angle through simple geometrical optics considerations. These two methods are demonstrated on a series of short-chain alkanes (hexane to hexadecane) in contact with hydrophobic glass slides and/or polytetrafluoroethylene plates.

1920. Allan, A.J.G., “Surface properties of polyethylene: Effect of an amphiphatic additive,” J. Colloid Science, 14, 206-221, (Apr 1959).

1141. Allen, K.W., “Dispersion forces,” in Handbook of Adhesion, 2nd Ed., D.E. Packham, ed., 111-113, John Wiley & Sons, Jul 2005.

1965. Allen, K.W., L. Greenwood, and T.C. Siwela, “Surface treatment of metal surfaces by corona discharge,” J. Adhesion, 16, 127-131, (Nov 1983).

Aluminium and titanium surfaces have been treated by corona discharge in air and gave bonds of strength similar to those obtained by conventional chemical treatment.

1358. Allen, N.L., and A.A.R. Hashem, “The role of negative ions in the propagation of discharges across insulating surfaces,” J. Physics D: Applied Physics, 35, 2551-2557, (2002).

The relative importance of atmospheric negative ions in corona formation and breakdown in a radial electric field has been investigated, with special reference to effects occurring at an insulating surface of a cycloaliphatic polyurethane with dolomite filler. The inception times, corona charge, light emission (by UV photography) and sparkover voltages under lightning impulses have been recorded under natural atmospheric conditions and under the influence of negative ions introduced from an auxiliary corona. The presence of excess negative ions on the surface is shown to increase both the charge injected and the mean radius of the corona, attributable to an augmentation of streamer discharges. In similar experiments in air, the excess ions cause a transition to a `glow' discharge. Effects of negative ions on sparkover are not significant on the insulator surface, but they cause a small increase in air. Comparisons between the two cases lead to the conclusion that the most important effect of the ions on the surface is to provide `seed' electrons for streamer propagation, following ion detachment in the field of advancing streamer tips.

2747. Allen, R., “How to obtain good adhesion of extruded polypropylene to film and foil substrates by using ozone and primers,” in 2006 PLACE Conference Proceedings, 1354-1359, TAPPI Press, Sep 2006.

787. Allred, R.A., and S.P. Wesson, “Effects of acid base interactions on carbon/polycarbonate composite interfacial adhesion,” in Acid-Base Interactions: Relevance to Adhesion Science and Technology, Vol. 2, K.L. Mittal, ed., 551-580, VSP, Dec 2000.

2724. Alm, H.K., G. Strom, J. Schoelkopf, and P. Gane, “Ink-lift-off during offset printing: a novel mechanism behind ink-paper coating adhesion failure,” J. Adhesion Science and Technology, 29, 370-391, (2015).

This paper reports on a special pilot coating and industrial printing trial designed to gain fundamental knowledge on ink adhesion failure on coated papers. We found that ink adhesion failure resulted in white spots without ink on the paper, referred to as uncovered areas and these spots gave print mottle problems. The white spots were due to two fundamentally different types of ink adhesion failure. One is the well-known ink rejection, which simply means that ink is not transferred to the surface. The other is a new type of ink adhesion failure, confirming a previous hypothesis suggested from laboratory observations. We refer to this as ink-lift-off adhesion failure, meaning that ink initially deposited on the paper surface becomes lifted off from the surface in a subsequent print unit. Adhesion failure by this mechanism was seen to occur more frequently than failure due to the well-known ink rejection.

2650. Altay, B.N., “Smart ink for flexo,” Flexo, 41, 70-75, (Jun 2016).

2965. Altay, B.N., R. Ma, P.D. Fleming, M.J. Joyce, A. Anand, et al, “Surface free energy estimation: A new methodology for solid surfaces,” Advanced Materials Interfaces, 7, (Mar 2020).

An interpretation of solid surfaces is generated based on physical considerations and the laws of thermodynamics. Like the widely used Owens–Wendt (OW) method, the proposed method uses liquids for characterization. Each liquid provides an absolute lower bound on the surface energy with some uncertainty from measurement variations. If multiple liquids are employed, the largest lower bound is taken as the most accurate, with uncertainty due to measurement errors. The more liquids used, the more accurate is the greatest lower bound. This method links generalizations of the Good–Girifalco equation with a general thermodynamic inequality relating the three-interfacial tensions in a three-phase equilibrium system. The method always satisfies this inequality with better than a 65% certainty. However, the OW seldom, if ever, conforms to this inequality and even then, the degree of satisfaction is insignificant. A reconciliation of the two methods is proposed based on rescaling the OW surface energies to conform to the inequality. This enables interpretations of dispersion and polar components of the surface energy, which are thermodynamically self-consistent. The proposed method is also capable of dealing with material exchange between liquid and solid phases, when the surface tension and contact angle of the saturated liquids can be measured.

1626. Amanatides, E., and D. Mataras, “Modeling and diagnostics of He discharges for treatment of polymers,” in Advanced Plasma Technology, R. d'Agostino, P. Favia, Y. Kawai, H. Ikegami, N. Sato, F. Arefi-Khonsari, eds., 55-74, Wiley-VCH, Jan 2008.

2374. Andrade, J.D., P.M. Triolo, L.M. Smith, and F.J. Miller, “Process for treating polymer surfaces to reduce their friction resistance characteristics when in contact with non-polar liquid, and resulting products,” U.S. Patent 4508606, Apr 1985.

8. Andrade, J.D., S.M. Ma, R.N. King, and D.E. Gregonis, “Contact angles at the solid-liquid interface,” J. Colloid and Interface Science, 72, 488-494, (1979).

The study of polymer—water interfaces by contact angle methods can be accomplished directly at the polymer—water interface. Using two water-immiscible liquids or a liquid and a vapor, one can deduce the dispersion and polar components of the hydrated solid surface free energy and the solid—water interfacial free energy. The theory is presented and a numerical analysis procedure is developed to solve the equations in the general case. The special case of n-octane and air is also presented. Data and results are given for poly(hydroxyethyl methacrylate-methoxyethyl methacrylate) copolymers of varying composition and equilibrium water contents. The results show that the hydrophilic component dominates the polymer—water interfacial properties, even at relatively low hydrophilic component compositions. The method presented should be useful for the study of polymer—water interfaces, particularly for hydratable or mobile polymers which can reorient to equilibrate differently with a water environment than with the air or vapor environment commonly used in contact angle studies.

 

<-- 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-->