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 26 of 76, ordered by
 

784. Helt, S., J. Evieux, Y. Baziard, V. Nassiet, and J.-A. Petit, “Characterization of acid-base properties of surfaces by contact angle titration:application to the adhesive bonding of silicon carbide,” in Acid-Base Interactions: Relevance to Adhesion Science and Technology, Vol. 2, K.L. Mittal, ed., 399-418, VSP, Dec 2000.

2626. Henry, E.B., “Determination of the surface energy for UV-curable, easy-release coatings,” Presented at RadTech 2016, May 2016.

1580. Herbert, P.A.F., and E. Bourdin, “New generation atmospheric pressure plasma technology for industrial on-line processing,” J. Coated Fabrics, 28, (1999).

1571. Herbert, T., “Atmospheric-pressure cold plasma processing technology,” in Plasma Technologies for Textiles, R. Shishoo, ed., 79-128, Woodhead Publishing, Mar 2007.

Although the power of plasma surface engineering across vast areas of industrial manufacturing, from microelectronics to medical and from optics to packaging, is demonstrated daily, plasma in the textile industry has been cynically described as the technology where anything can happen... but never does. Research into the application of plasmas to textiles goes back to the 1960s but, despite the reporting of novel and potentially commercial effects, it is only in recent years that plasma processing systems have begun to emerge into textile manufacturing in the production of specialty/high value fabrics. It is instructive to look at major criteria for the introduction of new technology into the textile market and to assess plasma processing against such criteria. They can be separated into qualifiers (must be satisfied by the new technology as a minimum) and winners (motivate take-up of the new technology by the industry). Here are ‘qualifier’criteria for new textile technologies:

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

158. Heusch, C., “Understanding surface tension,” Flexo, 18, 42-43, (Jul 1993).

893. Hibbard, D., “Sticky science: new polymer technology makes it easier to paint and glue plastic parts,” Modern Paints & Coatings, 91, 31, (Dec 2001).

2801. Hild, F., “Surface energy of plastics,” https://www.tstar.com/blog/bid/33845/Surface-Energy-of-Plastics, Dec 2009.

1896. Hill, J.M., E. Karbashewski, A. Lin, M. Strobel, and M.J. Walzak, “Effects of aging and washing on UV- and ozone-treated poly(ethylene terephthalate) and polypropylene,” J. Adhesion Science and Technology, 9, 1575-1591, (1995) (also in Polymer Surface Modification: Relevance to Adhesion, K.L. Mittal, ed., p. 273-290, VSP, May 1996).

In this study we investigated the stability of poly(ethylene terephthalate) (PET) and polypropylene (PP) surfaces modified using three combinations of UV light and ozone: ozone only, UV light in air (producing ozone), and UV light in air supplemented by additional ozone in the incoming air. Analysis was done using X-ray photoelectron spectroscopy and dynamic contact angle measurements. Our results showed that PET film is oxidized using these treatment conditions and it changes significantly within the first week of aging and after washing with water. These changes are reflected in the decrease in the Δ(O : C) ratio and the increase in the contact angle. Conversely, PP changes very little on aging or washing. Low-molecular-weight oxidized material (LMWOM), produced on the polymer surfaces treated with UV/air or UV/air + ozone, is easily removed with water washing. On aging PET, a number of the oxidized groups at the surface disappear, seeming to migrate into the bulk. The PP, however, does not favour migration as a path to reduce the overall free energy of the system, so the oxidized groups remain at the surface. Treatment with ozone only, in the absence of UV light, is a much different modification process in terms of the mechanism and the functional groups formed on the surface. This is reflected in the aging and washing behaviour of both the PET and the PP treated with ozone only.

695. Hill, M., “Flame treatment meets quality management,” in 2002 PLACE Conference Proceedings, TAPPI Press, Sep 2002.

2989. Hillborg, H., and U.W. Gedde, “Hydrophobicity recovery of polydimethylsiloxane after exposure to corona discharges,” Polymer, 39, 1991-1998, (May 1998).

A high-temperature-vulcanized polydimethylsiloxane (PDMS) elastomer has been subjected to corona discharges for different periods of time in dry air. The loss and recovery of hydrophobicity of the surface have been characterized by contact angle measurements. Immediately after exposure to corona discharges, samples showed a low surface hydrophobicity and, on storage in dry air, a continuous increase in hydrophobicity finally approaching the hydrophobicity of the unexposed material. The activation energy of the hydrophobicity recovery was two to four times greater than the activation energy of the diffusivity of low molar mass PDMS in PDMS elastomers, indicating that the diffusivity properties of the oxidized surface layer were different from that of the bulk. PDMS elastomers quenched in liquid nitrogen or subjected to small mechanical deformation ( < 1% strain) after exposure to corona discharges for 1 h or more recovered their hydrophobicity faster than untouched specimens kept under identical conditions. X-ray photoelectron spectroscopy confirmed the early formation of a silica-like surface layer, with a thickness of at least 10–12 nm. The atomic composition of the oxidized surface layer remained essentially unchanged after the first hour of corona discharges. It is suggested that the silica-like surface layer delayed the recovery of hydrophobicity by inhibiting the transport of low molar mass PDMS to the surface. It is also suggested that thermally or purely mechanically induced stresses lead to a cracking of the brittle silica-rich layer and that this in turn facilitates the transport of low molar mass PDMS to the surface and to a more rapid recovery of the hydrophobicity. Data obtained by reflection infrared spectroscopy assessing the outermost micrometer, confirmed the oxidation and the formation of hydroxyl groups at a progressively higher concentration with increasing exposure time of corona discharges.

2196. Hine, C., “Corona collaboration,” Paper Film & Foil Converter, 77, (Nov 2003).

1971. Hirotsu, T., and S. Ohnishi, “Surface modification of some fluorine polymer films by glow discharges,” J. Adhesion, 11, 57-67, (1980).

Extensive study has been made of the effects of various types of glow discharge plasmas on the changes of the surface properties of some fluorine polymers. The properties were investigated as a function of such factors as the exposing period, aging after exposure, type of plasma, and so on.

It was found that the wettability and the critical surface tensions were changed considerably with plasma exposure and that periods of several tens of seconds are long enough to cause changes. The extents of change were not so prominent for fluorine polymers as for polyethylene, and this fact may show the important role of the fluorine atom in the surface properties even after the plasma treatments.

1762. Hitchcock, S.J., N.T. Carroll, and M.G. Nicholas, “Some effects of substrate roughness on wettability,” J. Materials Science, 16, 714, (1981).

The influence of substrate roughness on wettability has been investigated at room and high temperatures using sixteen material combinations, mostly liquid metals and solid ceramics but also water, glycerol and solid nickel. The contact angles assumed by both wetting and non-wetting drops of all but two material combinations increased linearly with the relative steepness of the surface features, the effect being less for experiments conducted at high temperatures. In contrast, the contact angles of good wetting drops of glycerol and exceptionally good wetting drops of Easy-flo decreased when their silica and nickel substrates were roughened. Similarly, contact angles of both wetting and non-wetting drops were decreased by ultrasonic vibration. The experimental data can best be interpreted in terms of the metastable equilibrium configuration models in which an advancing liquid front has to overcome energy barriers associated with surface features. This occurs more readily if these barriers are small relative to the energy of the liquid which our data suggest can be equated with the enthalpy of the liquid. This interpretation enables the effects of substrate roughness at one temperature or with one liquid to be used to predict behaviour at other temperatures and with other liquids.

159. Hjertberg, Y., B.A. Sultan, and E.M. Soervik, “The effect of corona discharge treatment of ethylene copolymers on their adhesion to aluminum,” J. Applied Polymer Science, 37, 1183-1195, (1989).

The efficiency of different techniques of obtain improved adhesion in polyethylene-aluminum laminates have been studied. Both surface treatments, such as thermal oxidation and corona discharge, and the use of copolymers with polar comonomers, i.e., vinyl acetate (EVA) and butyl acrylate (EBA), have been included. Thermal oxidation performed by high temperature extrusion including an ozone shower seems to be more effective than corona discharge. In a model experiment thermal oxidation was studied in more detail. The adhesion, as measured by a T-peel test, increased with the content of carbonyl measured by reflexion IR, except for relatively long thermal treatments. In the latter case molecular scission gave a large fraction of low molecular weight material with low cohesive strength. For EBA and EVA the peel strength increased linearly with the bulk concentration of comonomer from about 100 N/m for untreated polyethylene to 450 and 300 N/m, respectively, at 5 mol % comonomer. Corona discharge treatment of these copolymers had, however, a most remarkable effect on the adhesion properties. The increases, relative to untreated EBA and EVA, were much more dramatic compared to polyethylene, e.g., three to four and less than two times, respectively. The higher values obtained with EBA are suggested to be due to the conversion of acrylate groups into carboxylic acid. In the case of EVA, loss of acetic acid might instead decrease the content of polar groups.

160. Ho, C.-P., and H. Yasuda, “Coatings and surface modification by methane plasma polymerization,” J. Applied Polymer Science, 39, 1541-1542, (1990).

Polymers formed from plasma-polymerized methane were employed to modify the surface properties of silicone rubber membrane. Polymers were evaluated based on the energy input parameter W/FM, where W is the discharge power, F is the monomer flow rate, and M is the molecular weight of the monomer. Dealing with the characteristics of plasma polymerization and the deposited polymer film, the effect of pumping rate on deposition rate and the coating thickness, surface energy, and gas permeabilities of methane-plasma-polymer-coated silicone rubber membrane were investigated in three plasma regions. Because more reactive species are expelled at high pumping rates, the monomer-deficient region is reached at lower W/FM in the high pumping rate system than that in the low pumping rate system. The composite parameter W/FM had a strong influence on coating thickness, gas permeability, surface energy, and the polar component of the surface energy but little effect on its dispersion component. Examination of gas permeabilities indicated that coating thickness was another important controlling factor on the properties of plasma polymer.

479. Hobbs, J.P., C.S.P. Sung, K. Krishnann, and S. Hill, “Characterization of surface structure and orientation in polypropylene and poly(ethylene terephthalate) films by modified attenuated total reflection IR dichromism studies,” Macromolecules, 16, 193-199, (1983).

480. Hobin, T.P., “Surface tension in relation to cohesive energy with particular reference to hydrocarbon polymers,” J. Adhesion, 3, 327+, (1972).

A known relationship between heat of vaporisation, surface tension and molar volume applicable to spherical non-polar molecules is modified to apply also to linear molecules; the treatment involves calculation of molar surface areas corresponding to the appropriate “fully-packing” molecular shapes.

A linear relationship between the ratio cohesive energy density/surface tension and the reciprocal of molar volume is predicted for members of homologous series and demonstrated, with data for the n-paraffins.

1373. Hochart, F., J. Levalois-Mitjaville, R. De Jaeger, L. Gengembre, J. Grimblot, “Plasma surface treatment of poly (acrylonitrile) films by fluorocarbon compounds,” Applied Surface Science, 142, 574-578, (Apr 1999).

This paper concerns the use of the low-pressure plasma process to confer an hydrophobic character to poly(acrylonitrile) films, without altering their bulk properties. Plasma based fluorination processes using saturated fluorine compounds such as CF4, C6F14 and mixtures CF4/H2 were used. Such treated polyacrylonitrile films were characterized by XPS analysis, infrared (ATR) spectroscopy and contact angle measurements. The use of CF4 led to the fluorination of the film surfaces. Indeed, XPS spectra showed the presence of -CHF-, -CF2- and -CF3 groups. Moreover, the contact angle θ which was 63° for untreated PAN films increased to 115°. Fluorine incorporation in the PAN disappeared when hydrogen is added to CF4. In the case of treatment by C6F14, fluorination was accompanied by a polymerization, as evidenced by XPS with the disappearance of the N 1s photopeak corresponding to the C≡N functions in PAN. By considering the deposition of F- containing layer thickness, it was possible in this latter case to follow the surface modifications by ATR-i.r. spectroscopy

1159. Hockley, P., and M. Thwaites, “A remote plasma sputter process for high rate web coating of low temperature plastic film with high quality thin film metals and insulators,” AIMCAL News, 28-29, (Dec 2005).

161. Hoebergen, A., Y. Uyama, T. Okada, and Y. Idada, “Graft polymerization of fluorinated monomer onto corona-treated PVA cellulose films,” J. Applied Polymer Science, 48, 1825-1829, (1993).

Surface graft polymerization of a hydrophobic monomer, 2,2,3,3,3-pentafluoropropyl methacrylate (5FMA), onto hydrophilic poly(vinyl alcohol) (PVA) and cellulose films was studied after corona discharge of the films. It was found that grafting strongly depended on the reaction medium; especially, addition of alcohol to the monomer greatly accelerated graft polymerization. For instance, when an ethanol/ water /5FMA mixture (65/25/10, by volume) was used as the polymerization medium. the PVA and cellulose films corona-discharged for a few minutes exhibited a high contact angle up to 100° after 30 min polymerization, the graft density being approximately 170 μg/cm2 for cellulose and 80 μg/cm2 for PVA. © 1993 John Wiley & Sons, Inc.
https://onlinelibrary.wiley.com/doi/abs/10.1002/app.1993.070481016

639. Hoffman, A.S., “Biomedical applications of plasma gas discharge processes,” in Plasma Polymerization and Plasma Treatment of Polymers, Yasuda, H.K., ed., 251-267, John Wiley & Sons, 1988.

481. Hollahan, J.R., and G.L. Carlson, “Hydroxylation of polymethylsiloxane surfaces by oxidizing plasmas,” J. Applied Polymer Science, 14, 2499-2508, (1970).

The reaction of a radiofrequency-excited oxygen plasma with the surfaces of cured and uncured polymethylsiloxane produces intense hydroxylation of the surface region as followed by FMIR spectroscopy. Characteristic infrared features indicative of intraor intermolecular hydrogen bonding are evident. Plasma oxidation differs markedly from thermal oxidation processes. Reactions of polymethylsiloxane with nitrogen and air plasmas were also investigated and compared to corona reactions of oxygen, nitrogen, and air. In air corona, nitrogen moieties appear to be introduced. The behavior of polymethylsiloxane surfaces in oxidizing acids is also described.

2376. Holland, G.J., “Subjecting film to corona discharge prior to compression rolling,” U.S. Patent 4548770, Oct 1985.

1517. Holland, L., “Glow discharge excitation and surface treatment in low-pressure plasmas,” in Low Energy Ion Beams (Conference Series No. 54), L.H. Wilson and K.G. Stephens, eds., 220-228, Institute of Physics, Apr 1980.

Advances in glow discharge excitation for sputtering and surface treatment are reviewed. Conditions for the production of the glow discharges either by using a DC supply with a cold cathode or using an RF supply capacitively coupled to an electrode are discussed. Examples are given of surface treatment processes at present under study including RF magnetron sputtering of silica, a-C film growth in hydrocarbon plasmas and plasma nitriding.

868. Hollander, A., J. Behnisch, and M.R. Wertheimer, “Plasma vacuum UV effects on polymers,” in Plasma Processing of Polymers (NATO Science Series E: Applied Sciences, Vol. 346), d'Agostino, R., P. Favia, F. Fracassi, eds., 411-422, Kluwer Academic, Nov 1997.

A low pressure plasma comprises a complex mixture of electrons, charged and neutral molecules and fragments in the ground state and excited states, and a broad spectrum of radiation ranging from the infrared to the far ultraviolet. The specific role of each of these components in a plasma treatment of polymers is still not understood completely. The experimental data reported in the literature seem to be contradictory.

895. Holman, S., “What's your problem?,” Australian Flexo, (Apr 2001).

2369. Hood, J.L., “Method and apparatus for the corona discharge treatment of webs, and webs treated therewith,” U.S. Patent 4298440, Nov 1981.

482. Hook, T.H., R.L. Schmitt, and J.A. Gardella Jr., “Analysis of polymer surface structure by low-energy ion scattering spectroscopy,” Analytical Chemistry, 58, 1285-1290, (1986).

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.

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.

2263. Horakova, M., P. Spatenka, J. Hladik, J. Hornik, J. Steidl, and A. Polachova, “Investigation of adhesion between metal and plasma-modified polyethylene,” Plasma Processes and Polymers, 8, 983-988, (Oct 2011).

The polyethylene (PE) coatings could be very promising for various branches of industry due to their chemical stability and impact resistance. Plasma modification of powder has recently attracted much interest because of new prospects to control the interfacial properties. Plasma modification also significantly enhanced the adhesion of the polymer to the substrate. Powders find wide application in various branches of industry like paintings, biotechnology, filling for composite materials etc., but the plasma modification of powder surface has not found such application as plasma modification of flat solid materials. This is due to problems connected with the three dimensional geometry, necessity of solid mixing (due to the aggregation phenomenon) and the large surface area of powders which should be treated. We investigated plasma modification of PE powder, its adhesion properties on steel surface and mechanism influencing this adhesion. PE powder was modified using various working gases and chemicals. It was found that adhesion properties were strongly influenced by concentration of oxygen containing groups and also by PE crosslinking after modification. The value of crosslinking depends on used working gas and chemicals. The ternary mixture of O2/H2O/methanol was found to be an appropriate working gas for plasma treatment of PE for adhesion purposes. The treated PE had good wettability, low crosslinking and very high adhesion to the steel substrate.

1585. Hossain, M.M., D. Hegemann, A.S. Herrmann, and P. Chabrecek, “Contact angle determination on plasma-treated poly(ethylene terephthalate) fabrics and foils,” J. Applied Polymer Science, 102, 1452-1458, (2006).

The surfaces of polyester (PET) fabrics and foils were modified by low-pressure RF plasmas with air, CO2, water vapor as well as Ar/O2 and He/O2 mixtures. To increase the wettability of the fabrics, the plasma processing parameters were optimized by means of a suction test with water. It was found that low pressure (10–16 Pa) and medium power (10–16 W) yielded a good penetration of plasma species in the textile structure for all oxygen-containing gases and gaseous mixtures used. While the wettability of the PET fabric was increased in all cases, the Ar/O2 plasma revealed the best hydrophilization effect with respect to water suction and aging. The hydrophilization of PET fabrics was closely related to the surface oxidation and was characterized by XPS analysis. Static and advancing contact angles were determined from the capillary rise with water. Both wetting and aging demonstrated a good comparability between plasma-treated PET fabrics and foils, thus indicating a uniform treatment. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 102: 1452–1458, 2006
https://onlinelibrary.wiley.com/doi/abs/10.1002/app.24308

2246. Hou, W., L. Zhang, and Y. Long, “Study on the wettability of polyethylene film fabricated at lower temperature,” J. Colloid and Interface Science, 362, 629-632, (Oct 2011).

Polyethylene films were prepared with phase separation at lower temperatures. The wettability of such films varied from hydrophobicity to superhydrophobicity as the processing temperature decreased owing to the increase of surface roughness. Storing the as-prepared films at subzero temperature (−15 °C), it was found that the water contact angle of the film decreased obviously, and the decrease depended on the corresponding roughness. Further keeping the as-prepared films at room temperature for 30 min, the water contact angle would return to the normal value, which indicated that the reversible switching of surface wettability can be controlled by the environmental temperature.

483. Hoy, K.L., “New values of the solubility parameters from vapor pressure data,” J. Paint Technology, 42, 76+, (1970).

The solubility parameters of a broad spectrum of solvents and chemicals are calculated from vapor pressure data using an expression derived from the relationship of Haggenmacher. In the case of high boiling liquids, the available vapor pressure data are found to be unreliable when extrapolated to room temperature and an alternate method of calculation is proposed. A structure correlation is made using the method of Small and new values of the molar cohesion constants are developed. The problem of associations of certain molecular species is discussed and the concept of chameleonic character introduced as a qualitative explanation.

484. Hoy, K.L., “Tables of Solubility Parameters,” Union Carbide Corp., Chemicals and Plastics Research and Development Dept., 1985.

2144. Hozbor, M., “Plasma processes boost bondability of rubber and metal,” Adhesives Age, (Dec 1993).

2071. Hozumi, A., H. Inagaki, and T. Kameyama, “The hydrophilization of polystyrene substrates by 172-nm vacuum ultraviolet light,” J. Colloid and Interface Science, 278, 383-392, (Oct 2004).

This paper describes the photochemical surface modification of polystyrene (PS) substrates using vacuum ultraviolet (VUV) light 172 nm in wavelength. We have particularly focused on the effects of atmospheric pressure during VUV irradiation on the obtained surface's wettability and the stability of the wettability, in addition to its chemical structure, morphology, and photooxidation rate. Samples were photoirradiated with VUV light under pressures of 10, 10(3), or 10(5) Pa. Although, in each case, the originally hydrophobic PS surface became highly hydrophilic, the final water-contact angle and photooxidation rate depended on the atmospheric pressure. The samples treated at 10 Pa were less wettable than those prepared at 10(3) and 10(5) Pa due to the shortage of oxygen molecules in the atmosphere. The minimum water-contact angles of the samples treated at 10, 10(3), and 10(5) Pa were about 8 degrees, 0 degrees, and 0 degrees, respectively. With the samples prepared at 10 and 10(3) Pa, photooxidation reactions proceeded in the topmost region closest to the surface, while at 10(5) Pa photooxidation was found to be greatly enhanced in the deeper regions, as evidenced by angle-resolved X-ray photoelectron spectroscopy. Photoetching rates were determined through atomic force microscope observation of microstructured PS samples prepared by a simple mesh-contact method. As estimated from AFM images of the latticed microstructures obtained, the rates of samples prepared at 10(3) and 10(5) Pa were about 1.5 and 1.3 nm/min, respectively. However, no photoetched features were observable on the sample surface prepared at 10 Pa. Hydrophilic stability also varied greatly depending on atmospheric pressure. The hydrophilicity of samples treated at 10 and 10(3) Pa gradually decreased as they were exposed to air. On the other hand, the sample surface prepared at 10(5) Pa showed excellent hydrophilicity even after being left in air for 30 days.

2070. Hozumi, A., N. Shirahata, Y. Nakanishi, S. Asakura, and A. Fuwa, “Wettability control of a polymer surface through 126 nm vacuum ultraviolet light irradiation,” J. Vacuum Science and Technology, A22, 1309-1314, (Jul 2004).

The control of the surface wettability of poly (methyl methacrylate) (PMMA) substrates has been successfully demonstrated using an Ar2* excimer lamp radiating 126 nm vacuum ultraviolet (VUV) light. Each of the samples was exposed to 126 nm VUV light in air over the pressure range of 2×10−4-105 Pa. Although at the process pressures of 10, 103, and 105 Pa, the PMMA surfaces became relatively hydrophilic, the degree of hydrophilicity depended markedly on the pressure. The minimum water contact angles of the samples treated at 10, 103, and 105 Pa were about 50°, 33°, and 64°, respectively. These values were larger than those of PMMA substrates hydrophilized through 172 nm VUV irradiation conducted under the same conditions. On the other hand, after 126 nm VUV irradiation conducted under the high vacuum condition of 2×10−4 Pa, the PMMA substrate surface became carbon-rich, probably due to preferential cross-linking reactions, as evidenced by x-ray photoelectron spectroscopy. This surface was hydrophobic, showing a water contact angle of about 101°. Although the 126 nm VUV-irradiated surfaces appeared relatively smooth when observed by atomic force microscope, very small particles with diameters of 30-60 nm, which probably originated from the readhesion of photodecomposed products, existed on all of the sample surfaces.

2797. Hrinya, G., “Corona treaters: This valuable converting process helps avoid delivery delays and costly reprints,” Label & Narrow Web, 24, 76-79, (Oct 2019).

 

<-- 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-->