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 29 of 76, ordered by
 

174. Janczuk, B., T. Bialopiotrowicz, and W. Wojcik, “The components of surface tension of liquids and their usefulness in determinations of surface free energy of solids,” J. Colloid and Interface Science, 127, 59-66, (1989).

Measurements of the interfacial tension of glycerol-dodecane, formamide-dodecane, ethylene glycol-dodecane, and aqueous ethylene glycol solution-dodecane and the surface tension of ethylene glycol-water solutions were carried out. On this basis the surface tension components of these liquids were calculated and they were compared with values from the literature. It was found that they are close to J. Panzer's (J. Colloid Interface Sci.44, 142, 1973) results obtained by using solubility parameters. In order to verify whether the determined components of the surface tension of polar liquids are valid, measurements of equilibrium contact angles for these liquids were made on the surface of paraffin, polytetrafluoroethylene, polyethylene, polyethylene terephthalate, and polymethyl methacrylate. The measured values of contact angles were compared with those calculated, using the well-known components of the surface free energy. Good agreement was achieved among measured and calculated contact angle values and those obtained by other researchers. It was found that the calculated components of the surface tension of polar liquids worked well in the studied systems, and the geometric mean used for dispersion and nondispersion interfacial interactions gives good results despite existing intermolecular forces due to hydrogen bonding.

175. Janczuk, B., and T. Bialopiotrowicz, “Surface free energy components of liquids and low energy solids and contact angles,” J. Colloid and Interface Science, 127, 189-204, (1989).

Employing the values of organic liquid surface tension and interfacial surface tension of water-organic liquid, values of dispersion and nondispersion components of these liquids were calculated and compared with those obtained in another way. For these organic liquids and water, the values of the contact angle on paraffin wax, polytetrafluoroethylene, polyethylene, polyethylene terephthalate, and polymethyl methacrylate were measured. The values of dispersion and nondispersion components of surface free energy of these polymers and paraffin wax were calculated using the measured values of the contact angle for diiodomethane and water and the calculated values of the components of their surface tension. These calculated data were in agreement with the literature data. Taking our values of free energy components of liquids and solids, the values of the contact angle for these solids were calculated and compared with those measured, obtaining good agreement. On the basis of the measurements and calculations it was found that dispersion and nondispersion components of surface free energy of liquids and solids “work well” in the systems studied.

176. Janczuk, B., and T. Bialopiotrowicz, “The total surface free energy and the contact angle in the case of low energetic solids,” J. Colloid and Interface Science, 140, 362-372, (1990).

Using the literature data of the refractive index, the structural unit molar volume of polymers and their dipole moment, as well as the literature data of the polarizability, ionization potential, and dipole moment of many liquids, values of the Φ parameter for paraffin—liquid and polymer—liquid interfaces were calculated. Next, introducing these values of Φ and the earlier measured values of the contact angle for many liquids to the Young equation, values of the surface free energy (γS) of paraffin, polytetrafluoroethylene (PTFE), polyethylene (PE), polyethylene terephthalate (PET), and polymethacrylate (PMMA), were determined. It was found that the average values of γS for these solids were in agreement with those calculated on the basis of geometric, harmonic, or harmonic—geometric mean approaches. The values of the surface free energy of paraffin, PTFE, PE, PET, and PMMA were also calculated from the Young equation modified by Neumann et al. and, using the earlier measured values of the contact angle for many liquids, they were compared with the values obtained by other methods. Next, employing the mean value of the surface free energy, values of the contact angles for many liquids were calculated and compared with those measured earlier for the same liquids. It was found that for paraffin, PTFE, and PE there were big differences among the values of their surface free energies calculated from the contact angles for some liquids; however, the average values were in agreement with those obtained by other methods. The average values of the surface free energies of PET and PMMA were also in the range of the results obtained by other authors. It was also found that the average deviations of the contact angles calculated from the Young equation modified by Neumann et al. from the measured ones were slightly larger than those of the contact angles calculated from equations employing the geometric and harmonic means of the surface free energy components; the method of Neumann et al. may also be used to predict the wettability in some systems.

2917. Janule, V.P., “On-site surface and wetting tension measurements of water-based coatings and substrates,” Pigment & Resin Technology, 24, 7-12, (1995).

2786. Jarnstrom, J., B. Grandqvist, M. Jarn, C.-M. Tag, and J.B. Rosenholm, “Alternative methods to evaluate the surface energy components of ink-jet paper,” Colloids and Surfaces A: Physicochemical and Engineering Aspects, 294, 46-55, (2007).

The surface free energy is an essential paper property affecting liquid/ink interaction with the ink-jet paper surface. Different ways of calculating surface energy components for ink-jet papers is introduced. The results given by the very useful van Oss–Chaudhury–Good (vOCG) bi-bidentate model are compared with simpler mono-bidentate and mono-monodentate models. The unbalance in the acid–base (AB) values of the vOCG-model is compensated for, and occasional negative roots obtained are removed when applying the simpler mono-bidentate- and mono-monodentate models. The simple and elegant mono-monodentate model produces comparable values with the other models, and is thus recommended. The calculated percent work of adhesion between the probe liquids and substrates correlates well with surface energy component values. Also the percent work of adhesion between the inks and substrates correlates with surface energy values.

697. Jarvis, S.P., “Adhesion on the nanoscale,” in Nano-Surface Chemistry, Rosoff, M., ed., 17-58, Marcel Dekker, Oct 2001.

1718. Jaycock, M.J., and G.D. Parfitt, “The study of liquid interfaces,” in Chemistry of Interfaces, John Wiley & Sons, 1981.

944. Jensen, W.B., “Lewis acid-base interactions and adhesion theory,” Rubber Chemistry and Technology, 55, 881-901, (1982).

The above results are intended to be suggestive rather than definitive. Nevertheless, they strongly support the premise that a cross-fertilization of both concepts and experimental data from the apparently unrelated fields of Lewis acid-base chemistry and adhesion theory can lead to potentially valuable results for both fields, emphasizing again the value of using a generalized acid-base vocabulary in describing the phenomena of chemistry, whatever one's area of specialization.

1750. Jensen, W.B., “The Lewis acid-base concepts: recent results and prospects for the future,” in Acid-Base Interactions: Relevance to Adhesion Science and Technology, K.L. Mittal and H.R. Anderson Jr. eds., 3-24, VSP, Nov 1991.

641. Jhon, M.S., and S.H. Yuk, “Contact angles at polymer - water interface; temperature dependence and induced deformation,” in Polymer Surface Dynamics, Andrade, J.D., ed., 25-44, Plenum Press, 1988.

2919. Jin, M., F. Thomsen, T. Skrivanek, and T. Willers, “Why test inks cannot tell the whole truth about surface free energy of solids,” in Advances in Contact Angle, Wettability and Adhesion (Vol. 2), K.L. Mittal, ed., 419-438, Wiley, Sep 2015.

960. Jingxin, L., H. Guangjian, L. Qiman, and L. Xiaohong, “Surface structure and adhesive properties of biaxially oriented polypropylene film grafted with poly(acrylic amide) using corona discharge,” Polymer Engineering and Science, 41, 782-785, (May 2001).

Corona discharge was explored as a means of forming chemically active sites on the surface of biaxially oriented polypropylene (BOPP) film. The active species formed in air was used to induce graft copolymerization of acrylic amide (AAM) in aqueous solution. The surface structure, hydrophilicity and adhesion of the grafted BOPP film were characterized by the extent of grafting, electron spectroscopy for chemical analysis (ESCA), scanning electron microscopy (SEM), peel strength and contact angle measurements. Surface graft-copolymerization of AAM onto BOPP film by corona discharge in air can be carried out with high efficiency. With increasing copolymerization time, the degree of grafting of AAM onto BOPP increases. The degree of grafting achieved a relatively high value of 2.13 wt% for the conditions of 1 min corona discharge and a copolymerization reaction time of 2.5 hr in 20% AAM aqueous solution at 70°C. After corona discharge grafting, the contact angle of water on the BOPP film decreased and the peel strength increased compared with those for ungrafted BOPP film. The hydrophilicity and adhesion of BOPP were improved by surface graft copolymerization with AAM induced by corona discharge.

1485. Joanny, J.F., and P.G. de Gennes, “A model for contact-angle hysteresis,” J. Chemical Physics, 81, 552-562, (1984).

We discuss the behavior of a liquid partially wetting a solid surface, when the contact angle at equilibrium θ0 is small, but finite. The solid is assumed to be either flat, but chemically heterogeneous (this in turn modulating the interfacial tensions), or rough. For weak heterogeneities, we expect no hysteresis, but the contact line becomes wiggly. For stronger heterogeneities, we first discuss the behavior of the contact line in the presence of a single, localized defect, and show that there may exist two stable positions for the line, obtained by a simple graphic construction. Hysteresis shows up when the strength of the defect is above a certain threshold. Extending this to a dilute system of defects, we obtain formulas for the “advancing” and “receding” contact angles θa, θr, in terms of the distribution of defect strength and defect sharpness. These formulas might be tested by controlled contamination of a solid surface.

1165. Johans, C., I. Palonen, P. Suomalainen, and P.K.J. Kinnunen, “Making surface tension measurement a practical utility for modern industrial R & D,” American Laboratory (News Edition), 37, 14-16, (Dec 2005).

1573. Johansson, K., “Plasma modification of natural cellulosic fibres,” in Plasma Technologies for Textiles, R. Shishoo, ed., 247-281, Woodhead Publishing, Mar 2007.

This chapter provides a general summary of the current state of knowledge of plasma modification of various natural cellulosic fibres. Much of the information reported here is taken from the references cited at the end of the chapter, which should be consulted for a more in-depth treatment. Several aspects of plasma modification of various natural cellulose fibres are thoroughly treated in a number of excellent works. [1, 2]

1426. Johansson, K.S., “Ammonia plasma-simulating treatments and their impact on wettability of PET fabrics,” in Contact Angle, Wettability and Adhesion, Vol. 4, K.L. Mittal, ed., 335-350, VSP, Jul 2006.

Ammonia plasma treatments were performed on both thermoplastic plates and fabrics made of poly (ethylene terephthalate)(PET). The plates became more hydrophilic with improved adhesion properties as expected, whereas the fabrics became more hydrophobic, yet positively charged. Plasma treatments of PET fabrics using gas mixtures such as NH3/N2 and HZ/NZ were performed in order to simulate pure ammonia plasma treatments since such treatments are not always applicable in industrial applications, due to environmental and safety reasons. It was shown that the recommended and allowed gas compositions, 15% NH; in N2 and 5% H2 in N2, did not show any similarity with pure ammonia plasma treatments with respect to surface charge, wettability and chemical surface composition. At least 80% NH; in N; or 80% H2 in N2 is needed to simulate an ammonia-like plasma treatment of PET fabrics.

492. Johnson, B.A., “Studies of advancing and receding contact angles (MS thesis),” Univ. of Wisconsin, Madison, 1982.

177. Johnson, R.E. Jr., and R.H. Dettre, “An evaluation of Neumann's 'Surface equation of state' (comments),” Langmuir, 5, 293-295, (1989).

178. Johnson, R.E. Jr., and R.H. Dettre, “Wetting of low energy surfaces,” in Wettability, Berg, J.C., 1-74, Marcel Dekker, Apr 1993.

Wetting involves the interaction of a liquid with a solid. It can be the spreading of a liquid over a surface, the penetration of a liquid into a porous medium, or the displacement of one liquid by another. It can help to characterize surfaces and to determine solid/liquid interactions. Wettability is most often described by a sessile or resting drop. A schematic diagram is shown in Fig. 1. The contact angle (6) is a measure of wettability. A low contact angle means high wettability and a high contact angle means poor wettability. Zero contact angles are possible but they are always less than 180.(The highest commonly observed angle, mercury on glass, has been reported to be as high as 148 [1].) Systems having more than one stable contact angle are said to show contact-angle hysteresis.

1605. Johnson, R.E. Jr., and R.H. Dettre, “Contact angle hysteresis, 1: Study of an idealized rough surface,” in Contact Angle, Wettability and Adhesion: The Kendall Award Symposium Honoring William A. Zisman (Advances in Chemistry Series 43), F.M. Fowkes and R.F. Gould, eds., 112-135, American Chemical Society, 1964.

The effect of roughness on the wettability of an idealized sinusoidal surface has been studied with a digital computer. The equations of Wenzel and of Cassie and Baxter are discussed in relation to the model. The heights of the energy barriers between metastable states of a drop are seen to be of utmost importance in determining the magnitude of contact angle hysteresis.

2297. Johnson, R.E. Jr., and R.H. Dettre, “Wettability and contact angles,” in Surface and Colloid Science, Vol. 2, E. Matijevic, ed., 85-153, Wiley - Interscience, 1969.

2301. Johnson, R.E. Jr., and R.H. Dettre, “Contact angle hysteresis III: Study of an idealized heterogeneous surface,” J. Physical Chemistry, 68, 1744-1750, (Jul 1964).

1998. Johnson, R.E., Jr., R.H. Dettre, andD.A. Brandreth, “Dynamic contact angles and contact angle hysteresis,” J. Colloid and Interface Science, 62, 205-212, (Nov 1977).

Contact angles have been measured as a function of the three-phase-boundary velocity. Large velocity effects observed with other techniques were not seen using the plate method. It is possible to relate the dependence of contact angles on velocity to surface heterogeneity.

728. Jones, R.A.L., and R.W. Richards, Polymers at Surfaces and Interfaces, Cambridge University Press, Jun 1999.

2279. Jones, V., “Development of poly(propylene) surface topography during corona treatment,” Plasma Processes and Polymers, 2, 547-553, (Aug 2005).

Atomic force microscopy (AFM), contact-angle measurements, and X-ray photoelectron spectroscopy (XPS or ESCA) were used to characterize biaxially oriented poly(propylene) (PP) films modified by exposure to a corona discharge. Surface analysis was performed on PP films modified at various corona energies to explore the changes in surface topography, wettability, and oxidation state resulting from the corona treatment. Even at low corona energies, water-soluble low-molecular-weight oxidized materials (LMWOM) are formed. These LMWOM products agglomerate into small topographical mounds that are visible in the AFM images. For the detection of LMWOM on corona-treated surfaces, AFM appears to be at least as sensitive as contact-angle measurements or ESCA. A major advantage of AFM relative to the other surface analytical techniques used to confirm the presence of the LMWOM is that no washing of the surface with water is required in conjunction with the AFM analysis.

179. Jones, W.C., “Testing surfaces for cleanliness,” Metal Finishing, 83, 13-15, (Oct 1985).

2780. Jones, W.C., and M.C. Porter, “A method for measuring contact angles on fibres,” J. Colloid and Interface Science, 24, 1+, (1967).

A technique has been developed for rapid and extremely accurate measurements of contract angles formed by liquids on the surface of small-diameter filaments. The light beam reflection technique first de- scribed by Langmuir and Sehaeffer (1) and recently by Fort and Patterson (2) for liquid drops on flat plates has been refined for use with filaments and microscope equipment.

493. Jones, W.R., “Contact angle and surface tension measurements of a five-ring polyphenyl ether,” ASLE Translations, 29, 276-282, (Apr 1986).

Contact angle measurements were performed for a five-ring polyphenyl ether isomeric mixture on M-50 steel in a dry nitrogen atmosphere. Two different techniques were used: (1) a tilting-plate apparatus, and (2) a sessile drop apparatus. Measurements were made for the temperature range 25 to 190°C. Surface tension was measured by a differential maximum bubble pressure technique over the range 23 to 220°C in room air. The critical surface energy of spreading (γc) was determined for the polyphenyl ether by plotting the cosine of the contact angle (θ) versus the surface tension (γLV). The straight line intercept at cos θ = 1 is defined as γc. γc was found to be 30.1 dyn/cm for the tilting-plate technique and 31.3 dyn/cm for the sessile drop technique. These results indicate that the polyphenyl ether is inherently autophobic (i.e., it will not spread on its own surface film until its surface tension is less than γc). This phenomenon is discussed in light of the wettability and wear problems encountered with this fluid.

Presented at the 40th Annual Meeting in Las Vegas, Nevada May 6–9, 1985

670. Joos, P., Dynamic Surface Phenomena, VSP, Sep 1999.

494. Joos, P., and E. Rillaerts, “Theory on the determination of the dynamic surface tension with the drop volume and maximum bubble pressure methods,” J. Colloid and Interface Science, 79, 96-100, (1981).

The paper presents a theory on determining the dynamic surface tension using two methods: the drop volume method and the maximum bubble pressure method.

1733. Jorda-Vilaplana, A., V. Fombuena, D. Garcia-Garcia, M.D. Samper, and L. Sanchez-Nacher, “Surface modification of polylactic acid (PLA) by air atmospheric plasma treatment,” European Polymer J., 58, 23-33, (Jun 2014).

The main objective of this experimental study is the validation of the technique of atmospheric plasma with the aim of improving the surface energy of the polylactic acid (PLA) for further adhesion uses. The wettability of PLA has been improved with the application of an atmospheric plasma surface treatment. This method provides good adhesion properties with the optimizing the process parameters in terms of the nozzle–substrate distance and sample advance rate. In order to achieve that goal, a new and environmentally friendly technology has been used which is based on the use of air atmospheric plasma. The effects of the surface treatment on this type of substrates have been analyzed. The macroscopic effects of the process parameters have been determined using contact angle measurements and subsequent surface free energy (SFE) calculation. In addition, the chemical changes at the topmost layers have been studied using X-ray photoelectron spectroscopy (XPS) and Fourier transformed infrared spectroscopy (FTIR). Surface topography changes due to the plasma-acting mechanisms have been evaluated with scanning electron microscopy (SEM) and atomic force microscopy (AFM). The obtained results show a remarkable increase in surface free energy from 37.1 mJm−2 up to values of 60 mJm−2 thus indicating the effectiveness of the air plasma treatment. The main advantage of this technology is that the industrial process is continuous, it is easy to establish in current production systems and it does not generate wastes.

2420. Jordan, J.F., A. Yahiaoui, and P.R.R. Wallajapet, “Durable hydrophilic treatment for a biodegradable polymeric substrate,” U.S. Patent 7700500, Apr 2010.

2406. Jorgensen, M., “Electric discharge surface treating electrode and system,” U.S. Patent 6007784, Dec 1999.

2273. Joshi, R., R.-D. Schulze, A. Meyer-Plath, and J.F. Friedrich, “Selective surface modification of poly(propylene) with OH and COOH groups using liquid-plasma systems,” Plasma Processes and Polymers, 5, 695-707, (Sep 2008).

Underwater plasma and glow discharge electrolysis are interesting new methods for polymer surface functionalization. The achievable content of O-containing functional groups exceeds that of oxygen glow discharge gas plasmas by a factor of two (up to ca. 56 O/100 C). The percentage of OH groups among all O-containing groups can reach 25 to 40%, whereas it is about 10% in the gas plasmas. Addition of hydrogen peroxide increases the fraction of OH groups to at most 70% (27 OH/100 C). The liquid plasma systems are also able to polymerize acrylic acid and deposit the polymer as very thin film on substrate surfaces or membranes, thereby retaining about 80% of all COOH functional groups (27 COOH/100 C).

2421. Jung, J., and T. Gottfreund, “Biaxially oriented polyolefin film having improved surface properties,” U.S. Patent 7824600, Nov 2010.

495. Junnila, J., A. Savolainen, and D. Forsberg, “Adhesion improvements between paper and polyethylene by pretreatment of substrates,” in 1989 Polymers, Laminations and Coatings Conference Proceedings, TAPPI Press, Aug 1989.

2074. K. Kato, V.N. Vasilets, M.N. Fursa, M. Meguro, Y. Ikada, and K. Nakamae, “Surface oxidation of cellulose fibers by vacuum ultraviolet irradiation,” J. Polymer Science Part A: Polymer Chemistry, 37, 357-361, (1999).

The efficacy of vacuum ultraviolet irradiation for oxidizing the surface of cellulose fibers was compared to that of the conventional wet and dry processes. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 357–361, 1999
https://onlinelibrary.wiley.com/doi/abs/10.1002/%28SICI%291099-0518%2819990201%2937%3A3%3C357%3A%3AAID-POLA13%3E3.0.CO%3B2-2

180. Kaainoa, S., “What you should know about bare-roll corona treaters,” Plastics Technology, 32, 85-88, (Feb 1986).

1514. Kabza, K., J.E. Gestwicki, and J.L. McGrath, “Contact angle goniometry as a tool for surface tension measurements of solids, using Zisman plot method: A physical chemistry experiment,” J. Chemical Education, 77, 63-65, (Jan 2000).

The paper describes a physical chemistry experiment for measurement of surface tension of solids. It includes description of the Zisman-plot method for obtaining the surface tensions of solids as well as a detailed experimental procedure for surface tension measurements using contact angle goniometry. Four surfaces were analyzed and the critical surface tension (gc) for each of them was obtained. Excellent reproducibility was achieved despite measurement of less accurate Young's contact angles. The experiment includes an extensive data analysis section combined with graphical interpretation of the results. It can be included in an undergraduate physical chemistry laboratory sequence on its own or combined with existing experiments pertaining to surface phenomena.

1450. Kaczinski, M.B., and D.W. Dwight, “Enhancement of polymer film adhesion using acid-base interactions determined by contact angle measurements,” J. Adhesion Science and Technology, 7, 165-177, (1993) (also in Contact Angle, Wettability and Adhesion: Festschrift in Honor of Professor Robert J. Good, K.L. Mittal, ed., p. 739-751, VSP, Nov 1993).

Quantitative correlations among surface chemical composition, acid-base thermodynamics, adhesion strength, and locus-of-failure are demonstrated. Four types of functional Teflon surfaces were prepared: two acidic (containing hydroxyl and carboxyl groups), and two basic (containing acetyl and dinitrobenzoate groups). X-Ray photoelectron spectroscopy (XPS) and attenuated total reflection infrared (ATR-IR) spectroscopy were used to characterize the molecular structure of the surface region. Contact angle adsorption isotherms were determined using phenol as an acidic probe and tetrahydrofuran (THF) as a basic probe. The carboxylated surface had a higher molar ▵Hab with basic THF than the hydroxylated surface, and neither surface had any interaction with the acidic phenol probe. The acetylated surface behaved as a base, interacting with phenol but not with THF, while the dinitrobenzoyl surface had both acidic and basic character. Adhesion tests were carried out in the 180° peel mode using post-chlorinated poly(vinyl chloride) as a model acidic adhesive between pairs of each type of film. The two surfaces with basic character had significant peel strengths, while the two acidic surfaces had very low peel strengths. Scanning electron microscopy (SEM) of the basic failure surfaces showed significant plastic deformation of the Teflon polymer, while the acidic failure surfaces showed no deformation. XPS analysis of the failure surfaces confirmed interfacial failure for the acid-acid pairs, and bulk FEP failure for the acid-base pairs. These results demonstrate directly and quantitatively the enhancement of adhesive bond strength through acid-base interactions.

 

<-- 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-->