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 50 of 76, ordered by
 

961. Zenkiewicz, M., and J. Dzwonkowski, “Experimental evaluation of the process of decohesion of adhesive joints with polymer films,” Polimery, 45, 802-807, (2000).

Acrylic adhesive/PE film joints were prepared by using PE films preactivated by corona discharges. The tear-off work, a measure of strength of the joint, is the sum of several terms; the major terms include the work of adhesion and the work of deformation (here: stretching of film) involved in the tear-off process. The tear-off work (P-o, Fig.5) and the stretching work (P-r, Fig. 6) were determined for films that varied from 50 to 150 mum in thickness, in relation to specific treatment energies (E-j) ranging up to 7.5 kJ/m2. Within the interval 0 < E-j < 3 kJ/m2, P-r rises in an approximately linear way. As E-j is increased, evaluation of the unit adhesion work from tear-off tests becomes more and more burdened with error. Within the range of E-j adopted in technical preparation of upper layers of PE films (E-j = 2 kJ/m2), the error is kept within 9% and is only slightly related to film thickness.

967. Zenkiewicz, M., and J. Golebiewski, “Use of photoelectron spectroscopy in studies of the depth profile of polypropylene film,” Polimery, 44, 246-254, (1999).

1010. Zenkiewicz, M., “Flame modification of the surface layer of plastics products,” Polimery, 45, 81-88, (2000).

A review covering the general objectives of surface modification of plastics articles with particular reference to the flame modification technique. Hydrocarbon combustion conditions and flame modification parameters are described. Flame modification effects exercised on various, mostly polyolefinic, products are synthetically analyzed. The flame technique is shown to be advantageous and complementary to the corona-discharge technique. The effects of the polypropylene modification conditions on the oxygen content, the wetting angle for water, the free surface energy and the resistance of adhesion bonds are described.

3020. Zenkiewicz, M., “The analysis of principal conditions of van Oss-Chaudhury-Good's method in investigations of surface layers of polymeric materials,” Polimery, 51, 169-176, (2006).

The selected problems related to investigations of surface layers (WW) of solids were presented. The analysis of essential limits of van Oss - Chaunhury - Good's (vOCG) method, used for calculation of surface free energy (SEP) of polymeric materials, has been done. Some reasons of discrepancy between the results of calculations, obtained by various authors, were discussed in details. Namely, the need of use of algebraic analysis for selection of the set of three measured liquids, which are necessary in vOCG method, has been pointed. It makes possible to eliminate the sets of liquids being the reasons of bad conditioning of the sets of equations for SEP calculation. The effect of the proper selection of scale of components (acidic and basic ones) of SEP of water on the right evaluation of selected properties of the materials investigated was also presented (Table 1&2). General conclusions concerning the causes of controversy over van Oss - Chaunhury - Good's method were formulated.

2. Baszkin, A., and L. Ter-Minassian-Saraga, “Effect of temperature on the wettability of oxidized polyethylene films (letter),” Polymer, 15, 759-760, (1974).

37. Brennan, W.J., W.J. Feast, H.S. Munro, and S.A. Walker, “Investigation of the ageing of plasma oxidized PEEK,” Polymer, 32, 1527-1530, (1991).

Oxygen plasma treatment can be used for increasing the hydrophilicity of polymer surfaces, however, it is widely known that this effect decays significantly with time. This ageing phenomenon is thought to be caused by both migration of low molecular weight fragments and reorientation of modified polymer chains. It has recently been shown that the aged surface becomes transiently hydrophilic before attaining a final surface energy significantly lower than the initially treated surface. X-ray photoelectron spectroscopy (XPS) and contact angle measurements were used to monitor the changes in surface chemistry of plasma oxidized poly(ether ether ketone) (PEEK) during post treatment storage. The decay and transient increase in hydrophilicity were found to be dependent upon crystallinity and storage temperature.

39. Briggs, D., and C.R. Kendall, “Chemical basis of adhesion to electrical discharge treated polyethylene,” Polymer, 20, 1053-1055, (1979).

40. Briggs, D., D.G. Rance, C.R. Kendall, and A.R. Blythe, “Surface modification of poly(ethylene terephthalate) by electrical discharge treatment,” Polymer, 21, 895-900, (1980).

Poly(ethylene terephthalate) (PET) film has been discharge-treated under controlled conditions and the resulting surface modifications analysed via X.p.s. (ESCA), contact angle and surface energy measurements. Changes in surface properties have been followed as a function of ageing time. These measurements have been correlated with the adhesive properties of the treated surfaces using autoadhesion (treated-treated seals) as the probe. Discharge treatment introduces phenolic -OH and carboxylic acid -COOH groups into the surface resulting in increased wetting and much enhanced autoadhesion via hydrogen bonding of phenol groups to carbonyl groups. Much chain scission also occurs; the low molecular weight material is easily removed by washing and migrates into the film on ageing. The new functionalities in relatively immobile chains slowly reorientate and internally H-bond. The former process is largely responsible for the wettability change on aging, the latter for the loss of adhesive properties.

41. Briggs, D., D.R. Kendall, A.R. Blythe, and A.B. Wootton, “Electrical discharge treatment of polypropylene film,” Polymer, 24, 47-52, (1983).

The previously observed, but unexplained, deleterious effect of high relative humidity on the efficiency of electrical (‘corona’) discharge treatment for rendering polypropylene film printable has been re-examined. The effect of film temperature during treatment has also been studied. A consistent explanation of both effects based on the degree of surface coverage by physically adsorbed water is put forward, supported by X-ray photoelectron spectroscopy analysis of treated film surfaces.

42. Briggs, D., “New developments in polymer surface analysis,” Polymer, 25, 1379-1391, (1984).

Surface and interface characterization of polymeric materials has not enjoyed the multi-technique approach which typifies other types of materials. X-ray photoelectron spectroscopy (XPS) has dominated, despite several major disadvantages. New approaches are discussed which either improve XPS (particularly derivatization techniques) or utilize ‘static’ secondary ion mass spectrometry (SIMS) to overcome these limitations, with examples of their application in materials problem solving.

67. Corbin, G.A., R.E. Cohen, and R.F. Baddour, “Kinetics of polymer surface fluorination: elemental and plasma-enhanced reactions,” Polymer, 23, 1546-1548, (1982).

E.s.c.a. spectra of surface fluorinated polyethylene, poly(vinyl fluoride), and poly(vinylidene fluoride) are reported. Two reaction environments were used in this study: exposure to elemental fluorine and immersion in a glow discharge plasma. The systematic variation of fluorine composition in the polymer phase is shown to have a dramatic effect on the kinetics of the elemental reaction and little effect in the plasma reaction.

126. Gerenser, L.J., J.F. Elman, M.G. Mason, and J.M. Pochan, “ESCA studies of corona-discharge-treated polyethylene surfaces by use of gas-phase derivatization,” Polymer, 26, 1162-1166, (1985).

Chemically specific gas-phase reactions have been used to tag corona-discharge-induced chemical species on the surface of polyethylene. These tag reactions provide distinct moieties that can be detected via e.s.c.a. to provide a surface count of induced species. Hydroxyl, epoxy, hydroperoxy, carboxylic acid and carbonyl populations are discussed as a function of corona energy input, time after treatment and water washings.

137. Golub, M.A., T. Wydeven, and R.D. Cormia, “ESCA study of several fluorocarbon polymers exposed to atomic oxygen in low Eart h orbit or downstream from a radio-frequency oxygen plasma,” Polymer, 30, 1571-1575, (1989).

The ESCA (electron spectroscopy for chemical analysis) spectra of films of poly(vinyl fluoride) (Tedlar), tetrafluoroethylene-hexafluoropropylene copolymer (in the form of a Teflon FEP coating on Kapton H, i.e. Kapton F) and polytetrafluoroethylene (Teflon or Teflon TFE) exposed to atomic oxygen (O(3P)) - either in low Earth orbit (LEO) on the STS-8 Space Shuttle or within or downstream from a radio-frequency oxygen plasma - were compared. The major difference in surface chemistry of Tedlar induced by the various exposures to O(3P) was a much larger uptake of oxygen when etched either in or out of the glow of an O2 plasma than when etched in LEO. In contrast, Kapton F exhibited very little surface oxidation during any of the three different exposures to O(3P), while Teflon was scarcely oxidized.

138. Golub, M.A., and R.D. Cormia, “ESCA study of poly(vinylidene fluoride) tetrafluoroethylene-ethylene copolymer and polyethylene exposed to atomic oxygen,” Polymer, 30, 1576-1581, (1989).

The ESCA (electron spectroscopy for chemical analysis) spectra of films of poly(vinylidene fluoride) (PVDF), tetrafluoroethylene-ethylene copolymer (TFE/ET) and polyethylene (PE) exposed to atomic oxygen (O(3P)), in or out of the glow of a radio-frequency O2 plasma, were compared. ESCA spectra of PE films exposed to O(3P) in low Earth orbit (LEO) on the STS-8 Space Shuttle were also examined. Apart from O(3P)-induced surface recession (etching), the various polymer films exhibited surface oxidation, which proceeded towards equilibrium saturation oxygen levels. The maximum surface oxygen uptakes for in-glow or out-of-glow exposures were in the order: PE >TFEET >PVDF; for PE itself, the oxygen uptakes were in the order: in glow > out of glow > LEO. Given prior ESCA data on poly(vinyl fluoride) and polytetrafluoroethylene films exposed to O(3P), the extent of surface oxidation is seen to decrease regularly with increase in fluorine substitution in a family of ethylene-type polymers.

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

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

183. Kaelble, D.H., and J. Moacanin, “A surface energy analysis of bioadhesion,” Polymer, 18, 475-482, (1977).

This report applies recently developed surface energy and fracture mechanics relations to the analysis of bioadhesion and biocompatibility. The dispersion α and polar β components of 190 biological and implant surfaces are analysed. The surface energetics relations between bioadhesion and biocompatibility point out that a strongly adsorbed plasma protein film on the implant surface provides the best blood compatibility and low thrombogenic effects. The surface energy relations provide means of selecting optimum implant surface properties and mapping on surface energy diagrams the three phase interactions which define bioadhesion.

212. Leclercq, B., M. Sotton, A Baszkin, and L. Ter-Minassian-Saraga, “Surface modification of corona treated poly(ethylene terephthalate) film: adsorption and wettability studies,” Polymer, 18, 675-680, (1977).

Corona discharge treatment of poly(ethylene terephthalate) (PET) films produces chemical and physical modification of the surface leading to the formation of cavities and bumps. The roughness of the surface increases with the time of treatment and may be detected by scanning electron microscopy for the samples treated above 10 cycles, which corresponds to the duration of the exposure of the film under the electrodes. The degree of chemical modification, producing OH groups, is observed by adsorption of radioactive calcium ions and contact angle measurements. The results of these measurements are discussed and evidence presented shows that increase of the surface density of functional groups up to the value of 0.2 × 1014 sites/cm2 leads to a rapid increase in wettability of PET films.

225. Lub, J., F.C.B.M. van Vroohoven, E. Brunnix, and A. Benninghoven, “Interaction of nitrogen and ammonia plasmas with polystyrene and polycarbonate studied by X-ray photoelectron spectroscopy, neutron activation analysis and static secondary ion mass spectrometry,” Polymer, 30, 40-44, (1989).

The interactions of NH3 and N2 plasmas with the surfaces of polystyrene (PS) and bisphenol-A polycarbonate (PC) have been studied with X.p.s. and SSIMS. Primary amino groups could be detected at the surfaces of both polymers after treatment with the NH3 plasma but not with the N2 plasma, with the aid of derivatization reactions with salicylaldehyde and 5-bromosalicylaldehyde. PC differs in its reactivity from PS with respect to its ease of undergoing chain scission during the plasma treatments, which results in modified structures of low molecular weight at the surface. The surface coverage of primary amino groups on PS after treatment with the NH3 plasma was determined by means of neutron activation analysis after derivatization of these groups with 5-bromosalicylaldehyde and estimated to be approximately 0.5 amino groups per nm2.

237. Mercx, F.P.M., “Improved adhesive properties of high-modulus polyethylene structures, II. Corona grafting of acrylic acid,” Polymer, 34, 1981-1983, (1993).

High-modulus polyethylene (PE) tapes were grafted with acrylic acid using a two-step procedure. The tapes were first subjected to He/Ar corona discharge, immediately followed by exposure of the corona-treated tapes to acrylic-acid-saturated He gas. Evidence for the grafting was provided by X-ray photoelectron spectroscopy, which showed the surface of the treated tapes to consist of 64% acrylic acid and 36% PE. The grafting of acrylic acid is confined to the outermost surface layers, as indicated by reflection infra-red spectroscopy. Pull-out tests showed that the corona grafting of acrylic acid improves adhesion to epoxy resins by a factor of eight. Moreover, the increased adhesion is not achieved at the expense of a decrease in mechanical properties of the high-modulus PE tapes.

266. Occhiello, E., M. Morra, P. Cinquina, and F. Garbassi, “Hydrophobic recovery of oxygen-plasma-treated polystyrene,” Polymer, 33, 3007-3015, (1992).

Hydrophobic recovery of oxygen-plasma-treated polystyrene has been studied with regard to its dependence on temperature, molecular weight and plasma parameters, namely radio frequency power. Plasma chemistry has been studied by actinometry, while surface analysis has been performed by X-ray photoelectron spectroscopy (X.p.s.) and contact angle measurements. Molecular weight changes have been investigated by gel permeation chromatography. Plasma-treated surfaces become richer in oxygen and more crosslinked than untreated ones. Two basic mechanisms have been observed: one is based on short range motions within the plasma-modified layer, burying polar groups away from the surface without modifying its X.p.s. composition; the other involves long range motions, i.e. diffusion of non-modified macromolecules or segments from the bulk to the surface, altering its X.p.s. composition. The latter mechanism becomes important upon increasing ageing temperatures, or lowering molecular weight or extent of crosslinking of the surface layer.

288. Pochan, J.M., L.J. Gerenser, and J.F. Elman, “An ESCA study of the gas-phase derivatization of poly(ethylene terephthalate) treated by dry-air and dry-nitrogen corona discharge,” Polymer, 27, 1058-1062, (1986).

Gas-phase derivatization has been used along with e.s.c.a. to determine corona-discharge-induced chemical species on poly(ethylene terephthalate) (PET). Dry-air and dry-nitrogen coronas were studied. We showed that: (1) if the corona discharge treatment (CDT) power level is kept low enough, few water-soluble species are created; (2) 4% of oxygen is added to the surface with dry-air corona; (3) 75% of the oxidation products are identified as hydroperoxy, epoxy, hydroxyl, carboxylic acid and isolated carbonyl species (with hydroxyl and isolated carbonyl the prevalent species). Short-term time-dependent ageing studies show a one-to-one correspondence between the decrease in hydroperoxy species and the increase in hydroxyl and isolated carbonyl moieties. Reaction sequences are proposed to explain these data. At longer times these surface populations decrease. In general, the results from nitrogen coronas and dry-air coronas are similar.

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.

425. Blythe, A.R., D. Briggs, C.R. Kendall, D.G. Rance, and V.J. Zichy, “Surface modification of polyethylene by electrical discharge and the mechanism of autoadhesion,” Polymer, 19, 1273+, (Nov 1978).

A reexamination of previous studies concerning the electrical (‘corona’) discharge treatment of polyethylene and the resulting enhancement of autoadhesion has been carried out. X-ray photoelectron spectroscopic data provide new insight into the phenomenon by showing surface oxidation to result from treatment in ‘inert’ gases. Treatment in hydrogen is an exception and results in no autoadhesion enhancement even though energy input into the film is more efficient than in air. Autoadhesion theories based on electret formation are rejected; those based on hydrogen bonding are largely up-held and shown to be more generally applicable than at first imagined.

959. Bae, B., B.-H. Chun, and D. Kim, “Surface characterization of microporous polypropylene membranes modified by plasma treatment,” Polymer, 42, 7879-7885, (2001).

Scanning electron microscope and goniometer were used to investigate morphology and wetting property of polypropylene membrane surfaces modified by plasma treatment using different reagents. Surface morphology was significantly affected by the types of reagents. X-ray photoelectron spectroscopy and attenuated total refection-Fourier transform infrared spectroscopy were used to characterize the chemical structure of polypropylene membrane surfaces modified by Freon-116 gas plasma treatment. Many fluorine atoms were observed on the polypropylene surface, and its concentration increased to saturation with increasing plasma treatment time. The wetting behavior of plasma treated polypropylene membrane was well explained in relation with morphology and chemical structure.

981. Choi, D.M., C.K. Park, K. Cho, and C.E. Park, “Adhesion improvement of epoxy resin/PE joints by plasma treatment of PE,” Polymer, 38, 6243-6249, (1997).

Low density polyethylene (LDPE) and high density polyethylene (HDPE) were plasma-treated with N2 and O2 plasma. The wettability and polar component of surface free energy of plasma-treated polyethylene were investigated by contact angle measurement. The concentration of functional groups formed by plasma treatment such as hydroxyl and carbonyl groups was measured using attenuated total reflection Fourier transform infrared spectroscopy (ATR FTi.r.). The concentration of polar functional group increased rapidly with 5–10s of plasma treating time and then very slowly after that. The adhesion strength of epoxy resin/plasma-treated polyethylene joints was examined by a 90° peel test. The increase of the adhesion strength was similar to that of concentration of polar functional groups. The higher adhesion strength of epoxy resin/plasma-treated HDPE joints was observed than that of epoxy resin/plasma-treated LDPE joints since HDPE deformed more during the peel tests and had more polar functional groups on the surface.

982. Nihlstrand, A., T. Hjertberg, and K. Johansson, “Plasma treatment of polyolefins - influence of material composition, 2: Lacquer adhesion and locus of failure,” Polymer, 38, 3591-3599, (1997).

The adhesion properties achieved after oxygen plasma treatments of ten polypropylene (PP) and thermoplastic polyolefin (TPO) materials of different compositions were studied. It is shown that the adhesion between a polyurethane (PUR) lacquer and plasma-treated materials was strongly influenced by the plasma treatment conditions and the chemical composition of the materials. Generally, a low power-to-gas pressure (P/G) ratio during the plasma treatment and a high ethylene content, preferably in the form of blocks, and/or the presence of double bonds in the matrix, are favourable for adhesion properties. Moreover, the TPOs were less sensitive towards the plasma treatment conditions than the corresponding PPs. The properties and the type of rubber may also be important for the adhesion properties. Furthermore, it was shown by X-ray photoelectron spectroscopy (X.p.s.) and Fourier transform infrared (FTi.r.) spectroscopy (using the attenuated total reflectance (ATR) technique) that all failures—even the apparently interfacial failures—were located in the substrate, below the oxidized surface layer, the only difference being the depth of failure. The fracture surfaces of samples showing low peel forces generally had a more PP-like composition than fracture surfaces that were clearly cohesive in the substrate. This observation offers evidence that the lacquer adhesion is determined by the extent to which chain scission reactions occur in the near-surface region of the substrate during the plasma treatment. © 1997 Elsevier Science Ltd.

984. Nihlstrand, A., T. Hjertberg, and K. Johansson, “Adhesion properties of oxygen plasma-treated polypropylene-based copolymers,” Polymer, 38, 1557-1563, (1997).

Injection-moulded plates of four polypropylene-based copolymers with ethylene or an unconjugated diene as the comonomer were subjected to oxygen plasma treatments. The main objective was to investigate how the degree of wettability and the adhesion properties were influenced by the type and amount of comonomer and by selected plasma parameters. The change in wettability was monitored by static water contact angle measurements and the adhesion between plasma-treated polypropylene plates and a two-component polyurethane lacquer was evaluated by a 90° peel test. No significant difference in the degree of wettability depending on material composition or treatment conditions could be observed. However, the lacquer adhesion was shown to be a function of both material composition and discharge power, while the influence of gas pressure was less clear. For all procssing conditions used, the lacquer adhesion was distinctly improved as the diene content was increased. An increasing extent of crosslinking reactions combined with a reduction in the number of main chain scissions are proposed to account for the observed results.

1036. Nihlstrand, A., T. Hjertberg, and K. Johansson, “Plasma treatment of polyolefins - influence of material composition, 1: Bulk and surface characterization,” Polymer, 38, 3581-3589, (1997).

Injection-moulded plates of ten polypropylene (PP) and thermoplastic polyolefin (TPO) materials with varying material composition (different type of rubber, varying degree of ethylene etc.) were characterized before and after oxygen plasma treatments. Untreated materials were studied by means of differential scanning calorimetry (d.s.c.), size exclusion chromatography (s.e.c.), Fourier-transform infrared spectroscopy (FTi.r.), attenuated total reflectance (ATR) and transmission measurements, and the effect of plasma treatment conditions was followed by X-ray photoelectron spectroscopy (X.p.s.) and contact angle measurements. S.e.c. analysis revealed only minor variations among the materials, while the d.s.c. and FTi.r. experiments confirmed that the differences were to be expected as a result of the variation in material composition. The FTi.r.-ATR results showed that all samples had a gradient in material composition. The materials were generally more rich in PP in the topmost ∼ 200 nm than in the first ∼800 nm, and a lesser extent of ethylene modification and/or rubber was observed in the topmost ∼ 200 nm. It was also shown that the degree of surface crystallinity was normally greater at ∼ 800 nm than at ∼ 200 nm, and that a higher mould temperature led to a higher degree of surface crystallinity. The water contact angles and the atomic composition showed that the materials were more oxidized after plasma treatment at high power-to-gas pressure (P/G) ratios than at low ratios. Moreover, the dependence on material composition was weak for samples that were plasma-treated at low P/G ratios whereas the materials that were least ethylene-modified were less oxidized than the others at high P/G ratios. © 1997 Elsevier Science Ltd.

1196. Brewis, D.M., and D. Briggs, “Adhesion to polyethylene and polypropylene,” Polymer, 22, 7-16, (1981).

Adhesion to polyethylene and polypropylene is a complex subject requiring understanding of (a) the poor adhesive characteristics of these polymers; (b) the superior performance following certain pretreatments and (c) the nature of the changes brought about by these pretreatments and the mechanisms involved. This review discusses work on these topics and examines the impact of recent data resulting from the application of surface analytical techniques. The roles of ‘weak boundary layers’, surface energy and wettability and specific interactions are discussed in some detail.

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

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

1446. Badey, J.P., E. Urbaczewski-Espuche, Y. Jugnet, D. Sage, and T.M. Duc, “Surface modification of polytetrafluoroethylene by microwave downstream treatment,” Polymer, 35, 2472-2479, (Jun 1994).

The surface modification of polytetrafluoroethylene (PTFE) by microwave plasma treatment was investigated by means of contact angle measurement and e.s.c.a. studies. Various gases (e.g. O2, O2N2, NH3) were used. The influence of the various plasma parameters, such as power, gas flow, distance between the sample and the centre of the discharge, treatment time, etc., has been evaluated. No modification was induced by O2 and O2N2 treatment, whatever the treatment conditions. NH3 plasma irradiation, however, rendered the PTFE surfaces more hydrophilic, leading to an increase of the polar component of the surface energy from 4.5 to ∼ 57 mJ m−2 under optimized treatment conditions. NH3 treatment led to defluorination, crosslinking, hydrocarbon (CC,CH) bond formation, and incorporation of nitrogen-containing groups, as confirmed by e.s.c.a. Oxygen was also detected at the surface of treated PTFE. Correlations between the contact angle, defluorination rate, and surface nitrogen and oxygen contents, have been established. Optimization of operational NH3 plasma parameters, leading to the best wettability of the treated samples, is also reported.

1454. Youxian, D., H.J. Griesser, A.W.H. Mau, R. Schmidt, and J. Liesegang, “Surface modification of polytetrafluoroethylene by gas plasma treatment (to increase the surface energy),” Polymer, 32, 1126-1130, (1991).

Poly(tetrafluoroethylene) (PTFE) samples were surface modified in gas plasma atmospheres of air, oxygen, argon and water vapour in order to increase the surface energy. Its dispersive and polar components were determined by contact angle measurements after various treatment times. Plasma treatment times of only 15s were sufficient in all gases studied for substantial surface modification of PTFE. The chemical composition of the surfaces was studied by X-ray photoelectron spectroscopy (X.p.s.). The main results of all the plasma treatments were the abstraction of fluorine and the production of surface crosslinks, whereas only a low level of oxygen-containing groups was attached into the surface layer.

1800. Hwang, S.S., D.R. Iyengar, E.J. Kramer, and C.K. Ober, “Synthesis and characterization of fluorinated block copolymers for low surface energy applications,” Polymer, 36, 1321-1325, (Mar 1995).

In this paper we describe the preparation and preliminary characterization of diblock copolymers with a low surface energy block. These polymers were prepared by modifying the isoprene block in styreneisoprene-based block copolymers with either short perfluoroalkyl or dimethyl siloxy ‘fingers’. Specifically, the diene block of a styrene-isoprene block copolymer containing a large proportion of pendent vinyl groups (1,2- and 3,4-isoprene) was reacted with the appropriate hydrosilane in the presence of non-acidic Pt catalyst. The degree of attachment of hydrosilane was as high as 50% of the pendent unsaturations. Pendent vinyl groups were converted more efficiently than pendent methyl vinyl groups. These block copolymers, when mixed with the styrene homopolymer, exhibited surface segregation behaviour which depended on both polymer molecular weight and processing conditions. The surface segregation properties of the resulting block copolymers were studied by a variety of techniques which include contact angle measurements, and either X-ray photoelectron spectroscopy or Rutherford backscattering spectrometry. Contact angles as high as 110° were measured for both the siloxane- and perfluoroalkane-modified materials.

1915. Ferreira, L., B. Evangelista, M.C.L. Martins, P.L. Granja, et al, “Improving the adhesion of poly(ethylene terephthalate) fibers to poly(hydroxyethyl methacrylate) hydrogels by ozone treatment: Surface characterization and pull-out tests,” Polymer, 46, 9840-9850, (Nov 2005).

This work reports a methodology to improve the adhesion between poly(ethylene terephthalate) (PET) fibers and poly(hydroxyethyl methacrylate) (pHEMA) hydrogels by treating PET with ozone. The surface chemistry of PET was examined by water contact angle measurements, X-ray photoelectron spectroscopy (XPS), infrared reflection absorption spectroscopy (IRAS) and attenuated total reflectance infrared spectroscopy (ATR-IR) yielding information about the chemical functionalities at depths upon 0.6 μm. Ozone treatment introduces several polar groups in the surface of PET through oxidation and chain scission resulting in increased wettability. These groups include mostly carboxylic and anhydride groups and in small extent hydroxyl groups. Atomic force microscopy (AFM) analysis shows that the surface of ozone-treated PET films is fully covered with spherical particles that are removed after washing the film with water. During the washing step carboxylic functionalities were removed preferentially, as demonstrated by XPS and IR analysis. According to pull-out tests, PET monofilaments and bundles treated by ozone had a higher adhesion to pHEMA hydrogels than untreated ones. The apparent interfacial shear strength is 65% higher on pHEMA hydrogel containing an ozonated than an untreated PET monofilament. In addition, the force to pull-out an ozone-treated PET bundle from pHEMA hydrogel is ca. 81% higher than the one observed for the untreated bundle.

2553. Park, Y.W., and N. Inagaki, “Surface modification of poly(vinylidene fluoride) film by remote Ar, H2, and O2 plasmas,” Polymer, 44, 1569-1575, (Mar 2003).

The surface modification of poly(vinylidene fluoride) (PVDF) film induced by remote Ar, H2, and O2 plasmas have been investigated using contact angle measurement, X-ray photoelectron spectroscopy, and scanning probe microscope. The contact angle of water shows an improvement in the PVDF surface wettability during short plasma exposure time. Three remote plasmas treated PVDF sheet surfaces occurred dehydrofluorination and oxidation reactions simultaneously. Remote hydrogen plasma was the most effective in defluorination reactions and remote oxygen plasma was unfavorable to abstract fluorine atoms.

2775. Kano, Y., and S. Akiyama, “Critical surface tension of poly(vinylidene fluoride-co-hexafluoroacetone) by the contact angle method,” Polymer, 33, 1690-1695, (1992).

The contact angles θ of dispersion (D), polar (P) and hydrogen bonding (H) liquids on poly(vinylidene fluoride-co-hexafluoroacetone) (P(VDF-HFA); HFA content 6.5, 8.3 and 10.4 mol%) were measured. The critical surface tensions γc of P(VDF-HFA) were evaluated by the Zisman plot (cos θ versusγL), Young-Dupre-Good-Girifalco plot (1 + cos θ versus 1γ0.5L) and the log(1 + cos θ) versus log (γL) plot. The following results were obtained: the γc values of P(VDF-HFA) evaluated for the P liquids were larger than those for the D and H liquids; the γc values estimated by the Zisman plot were smaller than those obtained by the other plots; the surface tension γs values of P(VDF-HFA) revealed a minimum at the HFA content of 8.3 mol%. It was expected that P(VDF-HFA) with HFA = 8.3 mol% induced surface segregation most easily.

2828. Farris, S., S. Pozzoli, P. Biagioni, L. Duo, S. Mancinelli, and L. Piergiovanni, “The fundamentals of flame treatment for the surface activation of polyolefin polymers - A review,” Polymer, 51, 3591-3605, (Jul 2010).

This paper aims to provide an exhaustive and comprehensive overview on flame treatment as a valuable technique for improving the surface properties of polymers, especially polyolefins. It starts with a brief historical excursus on the origin of flame treatment, and the second section deals with the major fundamentals of flame chemistry, with a special focus on the combustion process and mechanism of surface activation. The most important parameters influencing the extent of the oxidation reaction along with relevant practical notes are discussed in the third section. The concluding section outlines how the most significant features of flame treatment can be profitably used to improve the wettability and adhesion properties of polyolefin surfaces, especially from the perspective of developing novel composite solutions such as polyolefins/bio-based coating pairs intended for many different applications.

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.

2083. Lee, J.H., and H.B. Lee, “Surface modification of polystyrene dishes for enhanced cell culture,” Polymer (Korea), 16, 680-686, (Nov 1992).

136. Golub, M.A., T. Wydeven, and R.D. Cormia, “ESCA study of Kapton exposed to atomic oxygen in low Earth orbit or downstream from a radio-frequency oxygen plasma,” Polymer Communications, 29, 285-288, (1988).

The ESCA spectra of Kapton polyimide film exposed to atomic oxygen O(3P), either in low earth orbit (LEO) on the STS-8 Space Shuttle or downstream from a radio-frequency oxygen plasma, were compared. The major difference in surface chemistry induced by the two types of exposure to O(3P), both of which caused surface recession (etching), was a much larger uptake of oxygen by Kapton etched in the O2 plasma than in LEO. This difference is attributed to the presence of molecular oxygen in the plasma reactor and its absence in LEO: in the former case, O2 can react with radicals generated in the Kapton molecule as it etches, become incorporated in the etched polymer, and thereby yield a higher steady-state ‘surface oxidation’ level than in LEO.

 

<-- 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-->