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 18 of 76, ordered by
 

1942. Tingey, K., K. Sibrell, K. Dobaj, K. Caldwell, M. Fafard, and H.P. Schreiber, “Surface restructuring of polyurethanes and its control by plasma treatment,” J. Adhesion, 60, 27-38, (Jan 1997).

It was shown that when polyurethanes designed for use in biopolymer applications were immersed in orienting fluids, significant increases in their non-dispersive surface energies took place. The kinetics of the surface energy response were found to be a function of the immersion medium's acid-base interaction potential. Restructuring from the as-cast state, similar to that reported for two-component polyurethane adhesives, occurs in response to thermodynamic demands and is attributable to a preferential concentration of low energy segments in the surface region. Since shifting surface energies in polyurethanes may pose problems in biological applications, an attempt was made to crosslink the surface of the polymers by the use of cold, microwave plasma discharges with Argon as the treatment gas. Plasma treatments proved to be successful, in that polyurethane surfaces so modified responded much more weakly to changes in the polarity of contact media.

1943. Nakamura, Y., and K. Nakamae, “Adhesion between plasma-treated polypropylene films and thin aluminum films,” J. Adhesion, 59, 75-86, (Aug 1996).

Polypropylene (PP) film was treated with radio-frequency-induced oxygen plasma, followed by the vacuum deposition of aluminum (Al) thin film, and the peel strength of the Al deposited PP film (Al/PP) was examined. The peel strength of plasma-treated PP film varied widely in the range of 6.7 to 157 N/m depending upon the plasma treatment conditions, whereas that of the untreated PP was 5.2 N/m. The peel strength was minimized at oxygen pressure near 13.3 Pa (0.1 Torr), and decreased with increasing discharge power. The peel strength rapidly increased at the initial stage of plasma treatment (∼ several seconds), decreased at the second stage, and slightly increased again at the third stage. A good agreement was found between the peel strength of Al/PP and the amounts of oxygen introduced onto the PP surface at the initial stage. A short-time treatment was very effective to improve the adhesion of Al/PP. At the end of the second stage, a large amount of carbon was detected by XPS on the Al layer of the peeled interface of Al/PP, which gave a minimum peel strength. Cohesive failure of PP film might have occurred. SEM photograph showed that PP surface was etched by oxygen plasma at the thrid stage. These peel behaviors of Al/PP were explained by the chemical and physical changes of the PP surface caused by oxygen plasma treatment: (1) introduction of O-functional groups onto the PP surface at the initial stage, (2) formation of weak booundary layers resulting from the partial scission of PP molecules at the second stage, and (3) plasma etching of the PP surface at the third stage.

1944. Feinerman, A.E., Y.S. Lipatov, and V.I. Minkov, “On the hysteresis of polymer wetting,” J. Adhesion, 56, 97-105, (Apr 1996).

The reasons for the appearance of the hysteresis of wetting are considered. The model is proposed according to which the hysteresis is the result of the orientations of molecules of wetting liquids which is preserved due to the action of surface forces even after the flow ceases.

1945. Brewis, D.M., and G.W. Critchlow, “Adhesion and surface analysis,” J. Adhesion, 54, 175-199, (Dec 1995).

In the last 25 years, surface sensitive analytical techniques have made a major contribution to our understanding of adhesion phenomena and problems. There are several areas where these techniques have provided important information including the identification of failure modes, the chemistry of a substrate before and after pretreatments, the stability of surfaces and interfaces, the identification of surface contaminants, the interaction across an interface and the nature of interphases. X-ray photoelectron spectroscopy (XPS or ESCA), Auger electron spectroscopy (AES) and static secondary ion mass spectrometry (SSIMS) have proved to be especially useful. Many examples of the usefulness of these techniques are given.

1946. Fritz, J.L., and M.J. Owen, “Hydrophobic recovery of plasma-treated polydimethylsiloxane,” J. Adhesion, 54, 33-45, (Dec 1995).

Plasma treatment of silicone surfaces is a useful, environmentally-sound method of increasing wettability to improve adhesion. A thin, wettable silica-like layer is produced with various plasma gases such as argon, helium, oxygen and nitrogen. However, in each case the surfaces gradually recover their hydrophobicity. The silica-like layer is brittle and microcracking is evident at more severe levels of plasma treatment. The onset of cracking is a function of plasma gas, RF power, pressure and treatment time. Scanning electron microscopy has been used to characterize the cracks.

The hydrophobic recovery has been monitored by water contact angle changes. It occurs with both cracked and uncracked treated surfaces. There is an initial jump in hydrophobicity at the onset of cracking. Thereafter, the recovery of both cracked and uncracked surfaces broadly parallels each other with virtually complete recovery of original hydrophobicity within one week. These effects can be accounted for by rapid surface diffusion of low molecular weight material out of fresh cracks followed by slower bulk diffusion through the polymer matrix. Significant differences in recovery rates are also evident between different plasma gases.

1947. Collaud Coen, M., S. Nowak, L. Schlapbach, M. Pisinger, and F. Stucki, “Plasma treatment of polyacetal-copolymer, polycarbonate, polybutylene terephthalate, and nylon 6,6 surfaces to improve the adhesion of ink,” J. Adhesion, 53, 201-216, (Oct 1995).

Polyacetal-copolymer (POMB), polycarbonate (PC), polybutylene terephthalate (PBT), and nylon 6, 6 (PA6, 6) have been treated in an electron cyclotron resonance (ECR) plasma chamber to improve their adhesion properties towards ink. The chemical composition, the surface free energy, and the macroscopic adhesion have been studied by X-ray photoelectron spectroscopy (XPS), contact angle measurements, cross-cut tests, and the Scotch Tape test. Their dependence on the neutral gas, the treatment time, the pressure, and the ageing in air have been investigated. The XPS results reveal that the plasma treatment allows one to clean the surface and, if reactive gases are used, to incorporate new chemical species. The static and dynamic contact angles decrease with the plasma treatment and continue to decrease after contact with air. Very slow hydrophobic recovery is visible in the advancing contact angle, whereas the receding contact angle remains non-measurable even after more than a week of air exposure. Lower pressures and longer treatment times (120 s) lead to better macroscopic adhesion and reproducibility. For optimal treatment conditions (0.5 Pa, 120s N2 plasma treatment time), the improvement of the adhesion remains excellent after seven days exposure of the sample in air.

1948. Jacobasch, H.J., K. Grundke, S. Schneider, and F. Simon, “Surface characterization of polymers by physico-chemical measurements,” J. Adhesion, 48, 57-73, (Jan 1995).

The possibility of characterizing dispersion forces and acid-base interactions by means of physico-chemical measurements is demonstrated by the examples of contact angle and zeta potential measurements, with special attention being given to the latter. This method has been applied, to characterize the effect of plasma and flame treatment on the adhesion behaviour of injection moulded poly(propylene) specimens. The results with respect to acidic or basic functional surface sites, as obtained by zeta potential measurements, correlate with the elemental surface compositions determined by XPS. There is no general interrelation between acidic and basic parameters determined by contact angle measurements and the results of zeta potential and XPS measurements.

1950. Woods, D.W., P.J. Hine, R.A. Duckett, and I.M. Ward, “Effect of high modulus polyethylene fibre surface treatment on epoxy resin composite impact properties,” J. Adhesion, 45, 173-189, (Sep 1994).

1951. Sutherland, I., E. Sheng, D.M. Brewis, and R.J. Heath, “Flame treatment and surface characterisation of rubber-modified polypropylene,” J. Adhesion, 44, 17-27, (Oct 1994).

1952. Carre, A., and J. Vial, “Simple methods for the prediction of surface free energy and its components: Application to polymers,” J. Adhesion, 42, 265-276, (Oct 1993).

The surface free energy of a polymer can be easily calculated by the Group Contribution Method developed by the authors. After having briefly recalled the method and illustrated it with new examples, the latest developments including the Weighted Group Contribution Method and the study of the molecular weight dependence of surface free energy are also expounded.

Finally, very simple means to determine the dispersive contribution to the surface energy are described. The dispersive component values calculated from the Lifshitz theory, and from the solubility parameters, are in good agreement with those obtained from wettability measurements.

1953. Cueff, R., G. Baud, J.P. Besse, M. Jacquet, and M. Benmalek, “Surface free energy modification of PET by plasma treatment - influence on adhesion,” J. Adhesion, 42, 249-254, (Oct 1993).

Different cold plasmas have been used to treat the surface of polyethylene terephtalate (PET) in order to improve the adhesion of alumina thin films deposited by RF sputtering. The influence of these treatments on the surface free energy of the polymer is shown by a study of wettability. ESCA analysis of the PET surface suggests that chemical changes occur as the polymer is plasma treated.

The adhesion of alumina films on PET is studied by using tensile testing. The results show that the surface treatment of the PET by a slightly oxidizing plasma, such as carbon dioxide, increases by a factor of 1.7 the adhesion of alumina coatings.

1954. Kloubek, J., and H.P. Schreiber, “Futher comments on contact angle measurements on polymer solids,” J. Adhesion, 42, 87-90, (Aug 1993).

In recent years much emphasis has been placed on the importance of specific interactions in determining a wide range of polymer properties. Following the work of Fowkes and coworkers,1,2 all non-dispersion force interactions falling into the “specific” category may be considered to arise from acid/base interactions. This places heavy emphasis on methods for determining suitable acid/base parameters for polymers, with inverse gas chromatography (IGC) as a convenient choice. The IGC approach employed by one of us3 uses Gutmann's theory4 of (Lewis) acids and bases, and determines polymer interaction properties by placing these in contact with vapors of selected fluids for which acidlbase indexes, AN and DN, are available. Retention volume data for interacting vapor/polymer combinations are compared with the retention volumes of dispersion-force probes (e.g., n-alkanes), leading to the identification of AN and DN parameters for the polymer.3

1955. Brewis, D.M., “Pretreatments of hydrocarbon and fluorocarbon polymers,” J. Adhesion, 37, 97-107, (Feb 1992).

Pretreatments of polyolefins and fluoropolymers are usually necessary to achieve satisfactory adhesion for bonding and related technologies. In this paper results for various pretreatments of these polymers are presented. These are the treatment of polyolefins with aqueous reagents, dilute fluorine and a natural gas flame, the treatment of PTFE with sodium naphthalenide and the treatment of ECTFE with sodium naphthalenide and a flame. X-ray photoelectron spectroscopy was used to investigate the chemical changes caused by the treatment and the adhesion levels were discussed in relation to wetting, interactions across interfaces and weak boundary layers.

1956. Schreiber, H.P., “Specific interactions and contact angle measurements on polymer solids,” J. Adhesion, 37, 51-61, (Feb 1992).

The present work examined the susceptibility of contact angle data to specific interactions taking place between solids and contacting liquids. The polymers involved were polystyrene, polyvinyl chloride and polyethylene, representing respectively basic, acidic and neutral substrates. Contacting fluids also were chosen to represent acid and base interaction categories.

Significant time-dependent changes in contact angles were observed when acid/base pairs were involved in the experimental sequence. In specific cases it was possible to identify initial (zero contact time) contact angles, as well as equilibrium values, attained after prolongued contact times. Local solvation, or plasticization, of the polymer by the wetting fluid was postulated as the operative mechanism. The differences between initial and final values of the contact angles were correlated with parameters of specific interaction, calculated from the acceptor/donor numbers for the pertinent materials as measured by inverse gas chromatography. In contrast, when acid/acid or base/base combinations of polymer and wetting fluid were studied, equilibrium values of the contact angle were established rapidly. Since accurate information on acid/base properties of polymers and wetting fluids is not always available, it seems prudent to record contact angles as a function of contact time, and by extrapolation to determine the initial (true) value for further use in surface characterizations of polymers.

1957. Chin, J.W., and J.P. Wightman, “Adhesion to plasma-modified LaRC-TPI, I: Surface characterization,” J. Adhesion, 36, 25-37, (Nov 1991).

LaRC-TPI, an aromatic thermoplastic polyimide, was exposed to oxygen, argon and ammonia plasmas as pretreatments for adhesive bonding. Chemical changes which occurred in the surface as a result of the plasma treatments were investigated using x-ray photoelectron spectroscopy (XPS) and infrared reflection-absorption spectroscopy (IR-RAS). Water contact angle analysis was utilized to characterize the changes in surface wettability, and the ablative effects of the plasmas were monitored using ellipsometry. Both XPS and IR-RAS results indicated the formation of polar functional groups at the surface. Contact angle analysis showed enhanced water wettability of the plasma-treated surface. Oxygen and argon plasmas were highly ablative, whereas ammonia plasma was only moderately so. Oxygen and argon plasmas appear to react with the LaRC-TPI via a fragmentation/oxidation mechanism; the effect of ammonia plasma is postulated to be imide ring-opening resulting in the formation of amide functional groups.

1958. Bascom, W.D., and W.-J. Chen, “Effect of plasma treatment on the adhesion of carbon fibers to thermoplastic plastics,” J. Adhesion, 34, 99-119, (Jun 1991).

A study has been made of the effect of RF plasmas on the adhesion of carbon fibers to polycarbonate and polysulfone. Treatment in oxygen plasma significantly increased the adhesion to both polymers. The effect is lost if the treated fiber is stored in air for a week. Surface analysis using XPS indicated an increase in atom percent oxygen but the spectra were unchanged for the stored fibers even though there had been a significant loss in adhesion. It is suggested that oxygen surface functionality is responsible for the improved adhesion but that this surface activation is lost on storage. Due to a sampling depth of 5-10 nm, XPS would not be expected to detect this small change in surface functionality.

1959. Kinloch, A.J., and G.K.A. Kodokian, “On the calculation of dispersion and polar force components of the surface free energy,” J. Adhesion, 34, 41-44, (Jun 1991).

Contact angle measurements have been widely used1–6 to calculate the values of the dispersion force, γ d s , and polar force, γ p s , components to the total surface free energy of a material using a derivation originally proposed by Kaelble.2 In this analysis a pair of simultaneous equations is derived which for two liquids, i and j, on a common solid surface may be written as:

where α is the contact angle of the liquid on the solid surface. Thus, if the values of α, γ d l , γ p l and γ l (where γ l = γ d l + γ p l ) for the two liquids are known, these equations may be solved to yield the dispersion, γ d d ;, and the polar, γ d s , force components to the surface free energy of the solid surface. The total surface free energy, γ s , is then simply the sum of these components.

1960. Dillard, J.G., T.F. Cromer, C.E. Burtoff, A.J. Cosentino, R. Cline, G. Maciver, “Surface properties and adhesion of flame treated sheet molded composite (SMC),” J. Adhesion, 26, 181-198, (Oct 1988).

The surface chemistry of sheet molded composite (SMC) following interaction with a natural gas/air flame operated under reducing, stoichiometric, and oxidizing conditions has been investigated. The SMC surface chemistry is altered to contain in addition to hydrocarbon, ether, and ester functional groups, carbonyl and a greater carboxyl concentration. The extent of surface oxidation varies with the flame condition in the manner oxidizing ∼ stoichiometric > reducing. Lap shear tests carried out at 82°C (180°F) for coupons bonded with a urethane adhesive did not fail by fiber tear. Surface analysis results indicate failure at an oxidized SMC-adhesive/non-oxidized SMC interface and within the non-oxidized SMC surface.

1961. Cho, K., and A.N. Gent, “Adhesion between polystyrene and polymethylmethacrylate,” J. Adhesion, 25, 109-120, (Apr 1988).

Measurements have been made of the energy required to break through unit area of polystyrene (PS), polymethylmethacrylate (PMMA), and joints prepared by molding the two polymers in contact. The results were: 1.23 ± 0.5 kJ/m2 (PS), 0.46 ± 0.10 kJ/m2 (PMMA), and 0.22 ± 0.04 kJ/m2 for the bonded joint. Thus, the interface was significantly weaker than either adherend, but surprisingly strong for two incompatible materials. Microscopy and selective dyeing revealed that fracture took place at the interface itself, with no appreciable transfer of material from one side to the other. It is concluded that Van der Waals interactions are sufficient to create relatively strong bonds.

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

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

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

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

1964. Yetka-Fard, M., and A.B. Ponter, “Surface treatment and its influence on contact angles of water drops residing on teflon and copper,” J. Adhesion, 18, 197-205, (1985).

The variation of contact angle of liquid sessile drops on solids has been attributed to roughness (Wenzel2), the static charge effect (Holly,3 Ponter and Yekta-Fard3) and contamination at the solid surface or in the liquid and gaseous phases.

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.

1966. Sharma, A.K., and H. Yasuda, “Effect of surface energetics of substrates on adhesion characteristics of poly(p-xylylenes),” J. Adhesion, 13, 201-214, (Apr 1982).

In investigating the effect of the surface energetics of substrate materials on the adhesion characteristics of poly(p-xylylene) and poly(chloro-p-xylylene) by the “Scotch Tape” method, it was found that if the substrates had not been preconditioned (treated with argon or a methane plasma), the adhesion was poor. The characteristics of water resistant adhesion that were observed when coated substrates were boiled in 0.9% sodium chloride solution were found to vary from excellent (when the polymer did not peel from the substrate after three cycles of 8 hours of boiling and 16 hours at room temperature) to poor (when the polymer peeled off almost immediately). It was noticed that water resistant adhesion depends on the hydrophobicity of the substrate material (the greater the hydrophobicity, the greater the adhesion) and is not related to the dry adhesive strength of poly(p-xylylene). The oxygen glow discharge treatment of the substrates decreased both the dry and wet adhesive strength of the polymer. The effect of the argon glow discharge treatment depended on the surface energetics of the substrate, and the methane glow discharge treatment increased both the dry and wet adhesive strength of the polymer. These preconditioning processes are discussed in terms of the sputtering of the material from the wall of the reactor in contact with the plasma and the deposition of the plasma polymer of the sputtered material on the substrate surface.

1967. Carter, A.R., “Adhesion to polyolefins with flexible adhesives,” J. Adhesion, 12, 37-49, (May 1981).

Compounds based on polyolefins may find further use in the footwear industry as solings. However, a significant problem is the poor adhesion obtained with the urethane adhesives currently used. SATRA has recently attempted to develop practical bonding systems for commercial olefinic compounds. The use of flame treatments for polyethylene appeared to be a possible method of improving compatibility between the adhesive and substrate if an isocyanate is present at the interface. Polypropylene does not respond to the flame treatment but reasonable bonds have been obtained after surface oxidation or by using a sensitiser in conjunction with UV irradiation. The use of dual compound moulding is described as a possible alternative means of obtaining adequate adhesion to difficult surfaces.

1968. Smith, T., “Surface energetics and adhesion,” J. Adhesion, 11, 243-256, (1980).

The relationships between surface energetics and adhesion are critically reviewed. New data that confirm such relationships, for peel tests as well as lap shear tests, are presented. The effect of hydrothermal aging of aluminum surfaces on surface energetics can be used to predict degradation in bond strength. The mechanism of failure for elastic adhesives (such as Scotch® tape) in peel tests may be essentially the same as for more brittle adhesives (such as epoxies) in lap shear tests. This mechanism may involve brittle fracture that forms a critical flaw at the adherend-adhesive interface (on a microscopic level), followed by crack propagation which then may include considerable elastic and plastic deformation. The locus of propagation (fractography) is generally not (but may be) relevant to the problem of how to remedy mechanical weakness in an adhesive joint, since the local region of critical flaw formation rather than the general surface area determines the joint strength.

1969. Kruger, R., and H. Potente, “Corona-discharge treatment of polypropylene films: Effects of process parameters,” J. Adhesion, 11, 113-124, (1980).

Corona treatment of films, mainly polypropylene (PP)-copolymers, was studied at commercial levels in a 2.7 kVA treater. The films were produced on a flat-film extruder with chill rolls. Degree of treatment was characterized by power of the generator divided by web speed and width of film (m Ws/cm2).

The effectiveness of the treatment was measured in terms of the polar and dispersion components of surface-energy, the peel adhesion of pressure sensitive tape (similar to ASTM Adhesion Ratio) and the peel adhesion of polyurethane adhesives.

The polar component of surface energy is a measure of the effectiveness of corona pretreatment. For a given degree of treatment, the polar surface energy component becomes greater as the film cooling rate increases (and the degree of crystallization falls).

A comparison of homopolymers and copolymers does, however, reveal that even where these have the same density or the same degree of crystallization one cannot count on them having equally-sized polar components.

Peel strengths of pressure-sensitive tapes and polyurethane-bonded patches confirm the influence of cooling conditions on wetting properties.

1970. Schreiber, H.P., M.D. Croucher, and C. Prairie, “On multi-valued surface properties of PMMA films,” J. Adhesion, 11, 107-112, (1980).

An apparent link between the surface properties of polar group-containing polymers, such as PMMA and Styrene/Acrylic copolymers, and the thermodynamic quality of solvents used in solutions from which the polymers were cast, was described in earlier papers.1,2 In these polymers, significant variations have been observed in critical surface tensions(γc), and in the thermodynamic interaction parameters for selected vapor-polymer pairs, when the configuration of the polymer in solution was varied through the suitable selection of solvents of differing thermodynamic quality. The “solvent history” effect on surface properties of solid film was not detected however for non-polar polymers such as polystyrene (PS).1,2 Apparently the distinct chain configurations adopted in solution by PMMA are carried over into the solid and result in different proportions of non-polar (backbone) and polar (side chain) moieties being located in the surface layer of the solid. Since only one surface state can correspond to a thermodynamic equilibrium, it may be expected that the film surface properties will change with time, as the thermodynamically preferred state is attained. As a consequence, use properties of these films should also display (initially) the “solvent history” effect, and should vary similarly with time. The present communication is concerned with these points.

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.

1972. Evans, J.R.G., and D.E. Packham, “Adhesion of polyethylene to metals: The role of surface topography,” J. Adhesion, 10, 177-191, (1979).

Previous work established the importance of the fibrous substrate topography in obtaining good adhesion of polyethylene to matt black oxide films formed on copper in alkaline solution. In this paper the effect of the very rough surface topography is shown to be general. Anodising treatments for copper and zinc and a high temperature oxidation for steel are described which give a very rough surface consisting (respectively) of fibrous, dendritic and blade-like growths. The peel strength of polyethylene to these substrates is high even under circumstances, for example when the polymer is stabilised with anti-oxidant, where adhesion to a chemically similar smooth surface is low. The high peel strength is associated with large amounts of energy being dissipated during peeling in plastic deformation of the polymer near the interface. It is suggested that this is caused by the development of high shear stress concentration at the fibre ends causing yielding in a large volume of polymer.

1973. Huntsberger, J.R., “Reply to A.W. Neumann,” J. Adhesion, 9, 93-94, (1977).

1974. Neumann, A.W., and A.V. Rapacchietta, “Comments to J.R. Huntsberger: Surface chemistry and adhesion - a review of some fundamentals,” J. Adhesion, 9, 87-91, (1977).

1975. Stradal, M., and D.A.I. Goring, “The corona-induced autohesion of polyethylene: The effect of sample density,” J. Adhesion, 8, 57-64, (1976).

With increase in sample density, corona treatment was found to be decreasingly effective in enhancing the autohesion of polyethylene sheets. The effect of higher density could be offset in part by an increase in temperature of lamination. This parallel behaviour suggests that similar molecular mechanisms govern the phenomena of thermally-induced and corona-induced autohesion.

1976. Huntsberger, J.R., “Surface chemistry and adhesion: A review of some fundamentals,” J. Adhesion, 7, 289-299, (1976).

A critical review of some fundamentals of surface chemistry revealed several areas in which current interpretations of data or interrelationships are erroneous or misleading.

Correct forms of fundamental equations interrelating surface energies, equilibrium contact angles and adhesion are given and plotted in a convenient, illuminating, dimension-less form. These curves provide a basis for comparing some recently published empirical equations with the fundamental ones showing that discrepancies result from changing values of the interaction parameter φ.

1977. Sherriff, M., “Polar and dispersion contributions to solid surface tension: A reconsideration of their mathematical evaluation,” J. Adhesion, 7, 257-259, (1976).

One technique for the experimental determination of the dispersion and polar contributions to solid tension, γs d and γs p , is to measure the contact angle θ of a set of m liquids of known dispersion and polar contributions to surface tension on the solid and then to calculate γs d and γs p . There are two common techniques for this calculation, graphically1 or analytically.2,3 The graphical technique is limited in that it only considers dispersion forces (i.e., nonpolar systems) and so only isolates γs d . For this reason the analytical procedures which isolate both γs d and γs p are more commonly used, and they can be expressed in matrix notation as:

>where A is a 2 × 2 matrix containing information about the characterizing liquids and their contact angles, and the vector ◯ is related to γs d and γs p . Equation (1) is solved for all mC2 different liquid pairs to give a set of values for γs d and γs p which can then be subjected to statistical analysis.

1978. Evans, J.M., “The influence of oxygen on the nitrogen corona treatment of polyolefins,” J. Adhesion, 5, 9-16, (Jan 1973).

The resultant surface activation of polymers by corona discharges has been found to be markedly influenced by the type and purity of gases used in the corona. In this work it is shown that for the nitrogen gas corona treatment (15 KV, 15 mins) of polyethylene and polypropylene, traces of oxygen, >0.5% and <0.15% respectively, are sufficient to produce chemical changes in the polymer surface.

1979. Evans, J.M., “Nitrogen corona activation of polyethylene,” J. Adhesion, 5, 1-7, (Jan 1973).

Experiment has shown that the nitrogen corona-induced autohesion of polyethylene and the nitrogen-corona induced sorption of iodine by polyethylene both follow similar mechanisms. The controlling factor is postulated to be the formation of short-lived electrets within the polymer surface.

1981. Kiyozumi, K., T. Kitakoji, K. Uchiyama, and J. Goto, “Surface treatment of plastics by plasmajet,” J. Adhesion, 3, 77-81, (Sep 1971).

By applying a simple device comprising a power supply for arc welding and a plasmajet torch, a new method for plastic surface treatment to improve adhesion of the plastic was developed. The method enables such surface treatment instantly in the air atmosphere by applying the plasmajet to test pieces and is effective to various k-inds of plastics, especially to crystalline plastics as polyethylene.

When several pieces of polyethylene were treated under the following conditions in our experiment, the contact angle of water on the surface was improved from 80" to 20" and the adhesive strength by the tensile test was also remarkably improved from a few kg/cm2 to 120kg/cm2.

2329. Wu, S., “Polar and nonpolar interactions in adhesion,” J. Adhesion, 5, 39-55, (1973) (also in Recent Advances in Adhesion, L.-H. Lee, ed., Gordon and Breach, p. 45-63, 1973).

Equations for polar and nonpolar interactions across the interface are developed by using energy additivity concept in a semi-continuum model. Interfacial and surface tensions of molten polymers are measured directly and used to test the resulting equations: The first expression may be called the harmonic-mean equation preferred for low energy systems such as organic liquids, water, polymers, and organic pigments. The second may be called the geometric-harmonic-mean equation preferred for high energy systems such as mercury, glass, metal oxides and graphite. The third may be called the geometric mean equation which is found unsatisfactory. The harmonic-mean equation is used to obtain the “optimum” wettability condition for adhesion. The importance of polar interactions and matching of the polarity are analyzed and emphasized.

2331. Gent, A.N., and J. Schultz, “Effect of wetting liquids on the strength of adhesion of visoelastic materials,” J. Adhesion, 3, 281-294, (1973) (also in Recent Advances in Adhesion, L.-H. Lee, ed., Gordon and Breach, p. 253-268, 1973).

The effect of a variety of wetting liquids on the resistance to peeling separation for a lightly crosslinked rubbery adhesive in contact with a Mylar substrate has been studied over a wide range of peeling rates and at two temperatures. Although the magnitude of the peel strength is much greater than the thermodynamic work of detachment, it is reduced by alcohols and alcohol/water mixtures in good agreement with calculated reduction factors. It is concluded that the measured strength is a product of two terms: the thermodynamic work, and a numerical factor, generally large, denoting inefficiency. The latter term is strongly dependent on peel rate and temperature for viscoelastic adhesives. Two anomalies are pointed out: particularly low adhesion is observed at low rates of peel for certain liquids, attributed to swelling of the adhesive, and smaller effects are found for some other liquids than predicted.

2816. Dynes, P.J., and D.H. Kaelble, “Surface energy analysis of carbon fibers and films,” J. Adhesion, 6, 195-206, (1974).

Amorphous and graphitic carbon fibers and film surfaces are characterized by wettability measurements and surface energy analysis which isolate the (London-d) dispersion γd svand (Keesom-p) polar γp sv contribution to solid-vapor surface tension γsvd sv + γp sv Graphitized carbon fibers which are surface treated to provide strong bonding to polar matrix resins show consistent strong polar contributions to total surface tension with γd svsv ≃ γp svsv ≃ 0.50. Amorphous carbon films prepared for biological implant applications display dominant dispersion character in surface energy with γd svsv ≃ 0-74 to 0.95 and γp svsv ≃ 0.05 to 0.24.

 

<-- 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-->