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 15 of 76, ordered by
 

2089. Sanchis, M.R., O. Calvo, O. Fenollar, D. Garcia, and R. Balart, “Characterization of the surface changes and the aging effects of low-pressure nitrogen plasma treatment in a polyurethane film,” Polymer Testing, 27, 75-83, (Feb 2008).

In this work, low-pressure nitrogen plasma has been used to improve wettability in a polyurethane film. Evaluation of wettability changes has been carried out using contact angle measurements. Furthermore, plasma-treated films have been subjected to air aging to evaluate the extent of hydrophobic recovery. X-ray photoelectron spectroscopy (XPS) has been used to study surface functionalization; surface topography changes related with the etching mechanism have been followed by scanning electron microscopy (SEM), atomic force microscopy (AFM) and weight loss study. The results show a considerable improvement in surface wettability even for short exposure times, as observed by a remarkable decrease in contact angle values. The aging study shows a partial hydrophobic recovery due to the re-arrangement of polar species and migration of low molecular oxidized material (LMWOM). In addition to surface activation, SEM and AFM analyses show slight changes in surface topography as a consequence of the plasma-etching mechanism.

1690. Mount, E.M. III, “Substrate secrets: The best film optics for a particular application can be attained via rigid control of surface chemistry and internal and external light-scattering,” Converting, 26, 46-50, (Feb 2008).

1689. no author cited, “High definition corona treatment yields superior results,” Flexible Packaging, 10, 35, (Feb 2008).

2050. Morris, B.A., “Understanding why adhesion in extrusion coating decreases with diminishing coating thickness,” J. Plastic Film and Sheeting, 24, 53-88, (Jan 2008).

It is well known that in extrusion coating, the coating adhesion to the substrate decreases with decreasing thickness. The study on this phenomenon is divided into three parts. Part I explores the reduction in adhesion of LDPE to paper and other porous substrates. Several hypotheses are proposed for the origin of this decrease, including a reduction in oxidation time, faster cooling in the air gap, and more rapid quenching in the nip. A model of the molten polymer penetration into the substrate shows that the greatest effect is cooling in the nip; thinner coatings have less time to flow into the substrate interstices once the chill roll contact is made. The model results agree well with experimental adhesion data from the literature.

In Part II, adhesion to aluminum foil and other nonporous substrates is studied. Several hypotheses are proposed for why peel strength decreases in these structures, including a reduction in the air gap time, faster air gap cooling, more rapid nip quenching, and stress imposed during drawing. Modeling and experimental results show that cooling in the nip and imposed stress have the greatest impact.

In Part III, the peel test is analyzed to understand why the peel strength of better adhering adhesives are more sensitive to changes in coating thickness. The analysis shows that changes in the critical dimension of the deformation region at the peel front may be responsible.

1631. Mount, E.M. III, “Measuring treatment, part 3,” http://www.convertingmagazine.com/blog, Jan 2008.

1630. Mount, E.M. III, “Measuring treatment, part 2,” http://www.convertingmagazine.com/blog, Jan 2008.

1629. Mount, E.M. III, “Measuring treatment, part 1,” http://www.convertingmagazine.com/blog, Jan 2008.

1628. Mount, E.M., “Why is treatment such a poor indicator of surface quality?,” http://www.convertingmagazine.com/blog, Jan 2008.

1627. Arefi-Khonsari, F., and M. Tatoulian, “Plasma processing of polymers by a low-frequency discharge with asymmetrical configuration of electrodes,” in Advanced Plasma Technology, R. d'Agostino, P. Favia, Y. Kawai, H. Ikegami, N. Sato, F. Arefi-Khonsari, eds., 137-174, Wiley-VCH, Jan 2008.

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

2987. Balart, R., L. Sanchez, O. Fenollar, M. Pascual, and R. Lopez, “Hydrophobic recovery of low density polyethylene treated with corona discharge plasma,” Presented at International Federation of Associations of Textile Chemists and Colourists Congress 2008, 2008.

2841. Lee, S., J.-S. Park, and T.R. Lee, “The wettability of fluoropolymer surfaces: Influence of surface dipoles,” Langmuir, 24, 4817-4826, (2008).

The wettabilities of fluorinated polymers were evaluated using a series of contacting probe liquids ranging in nature from nonpolar aprotic to polar aprotic to polar protic. Fully fluorinated polymers were wet less than partially fluorinated polymers, highlighting the weak dispersive interactions of fluorocarbons. For partially fluorinated polymers, the interactions between the distributed dipoles along the polymer backbone and the dipoles of the contacting liquids were evaluated using both polar and nonpolar probe liquids. The results demonstrate that the surface dipoles of the fluoropolymers generated by substituting fluorine atoms with hydrogen or chlorine atoms can strongly interact with polar contacting liquids. The wettabilities of the partially fluorinated polymers were enhanced by increasing the density of dipoles across the surfaces and by introducing differentially distributed dipoles.

2785. Seppanen, R., M. Sundin, A. Swerin, and B. Brandner, “Relation between surface energy, topography, wettability and detailed surface chemistry by spectroscopy for coated printing papers,” in 2008 Advanced Coating Fundamentals Symposium, TAPPI Press, 2008.

2733. Gonzalez, E. II, M.D. Barankin, P.C. Guschl, and R.F. Hicks, “Remote atmospheric-pressure plasma activation of the surfaces of polyethylene terephthalate and polyethylene naphthalate,” Langmuir, 24, 12636-12643, (2008).

The surfaces of poly(ethylene terephthalate) (PET) and poly(ethylene naphthalate) (PEN) were treated with an atmospheric-pressure oxygen and helium plasma. Changes in the energy, adhesion, and chemical composition of the surfaces were determined by contact angle measurements, mechanical pull tests, and X-ray photoelectron spectroscopy (XPS). Surface-energy calculations revealed that after plasma treatment the polarity of PET and PEN increased 6 and 10 times, respectively. In addition, adhesive bond strengths were enhanced by up to 7 times. For PET and PEN, XPS revealed an 18-29% decrease in the area of the C 1s peak at 285 eV, which is attributable to the aromatic carbon atoms. The C 1s peak area due to ester carbon atoms increased by 11 and 24% for PET and PEN, respectively, while the C 1s peak area resulting from C-O species increased by about 5% for both polymers. These results indicate that oxygen atoms generated in the plasma rapidly oxidize the aromatic rings on the polymer chains. The Langmuir adsorption rate constants for oxidizing the polymer surfaces were 15.6 and 4.6 s(-1) for PET and PEN, respectively.

2656. no author cited, “Principle of surface modification by corona/gas plasma,” Kasuga Denki Inc., 2008.

2655. no author cited, “Configuration of a corona treater,” Kasuga Denki Inc., 2008.

2654. no author cited, “Wettability (wetting tension) and watt density,” Kasuga Denki Inc., 2008.

2653. no author cited, “How to measure the effect of treatment,” Kasuga Denki Inc., 2008.

2577. Pykonen, M., H. Sundqvist, M. Tuominen, J. Lahti, J. Preston, et al, “Influence of atmospheric plasma activation on sheet-fed offset print quality,” Nordic Pulp and Paper Research J., 23, 181-188, (2008).

The objective of this paper was to understand the effects of plasma activation, and thus influence of the surface energy and chemistry changes on offset print quality. Pigment coated and surface sized papers were treated with corona and atmospheric plasma in pilot and laboratory scales. The surface energy and surface chemistry changes were evaluated by contact angle and X-ray photoelectron spectroscopy (XPS). Offset printing was performed in laboratory scale with an IGT unit with predampening and in a pilot scale sheet-fed offset printing press. In addition, the ink setting rate was measured using an ink on paper tack tester. Plasma activation increased the surface energy of the papers. Furthermore, the polarity of the paper surface increased due to formed polar oxygen containing molecular groups. Due to differences in treatment times laboratory scale plasma treatment formed mainly carboxyl and ester groups, whereas pilot scale treatment induced mainly alcohol, ethers, aldehydes and/or ketones on paper surfaces. Printing evaluation showed that plasma activation influences both ink and water absorption properties. According to print tack results plasma activation led to faster ink-setting. With hydrophobic surface-sized paper plasma activation influenced the ink transfer, print gloss and density by changing dampening water absorption properties. The difference in surface chemistry with laboratory scale plasma treated samples was also reflected in the print quality properties. SEM imaging showed that too intense plasma activation can cause topography changes in addition to of the surface chemistry changes.

2514. Friedrich, J.F., R. Mix, and S. Wettmarshausen, “A new concept for adhesion promotion in metal-polymer systems by introduction of covalently bonded spacers at the interface,” J. Adhesion Science and Technology, 22, 1123-1143, (2008).

A new concept for molecular interface design in metal–polymer systems is presented. The main features of this concept are the replacement of weak physical interactions by strong covalent bonds, the flexibilization of the interface for compensating different thermal expansions of materials by using long-chain flexible and covalently bonded spacers between the metal and the polymer as well as its design as a moisture-repellent structure for hindering diffusion of water molecules into the interface and hydrolysis of chemical bonds. For this purpose, the main task was to develop plasmachemical and chemical techniques for equipping polymer surfaces with monotype functional groups of adjustable concentration. The establishing of monotype functional groups allows grafting the functional groups by spacer molecules by applying usual wet-chemical reactions. Four processes were favoured for production of monotype functional groups by highly selective reactions: the plasma bromination, the plasma deposition of plasma polymers, the post-plasma chemical reduction of O-functionalities to OH-groups, and the chemical replacement of bromine groups by NH2-groups. The grafting of flexible organic molecules as spacers between the metal layer and polymer improved the peel strength of the metal. To obtain maximal peel strength of aluminium coatings to polypropylene films and occurrence of cohesive failure in the polypropylene substrate, about 27 OH groups per 100 C-atoms or 6 COOH groups per 100 C-atoms were needed. Introducing C6–11-aliphatic spacers 1 OH or COOH group per 100 C-atoms contributed about 60% of the maximal peel strength of the Al–PP system, i.e. 2 or 3 spacer molecules per 100 C-atoms were sufficient for maximal peel strength.

2505. Borcia, C., G. Borcia, and N. Dumitrascu, “Relating surface modification to polymer characteristics,” Applied Physics A: Materials Science & Processing, 90, 507-515, (2008).

This paper aims to provide an analysis of the correlation between various plasma effects on polymers exposed to atmospheric pressure plasma. The relationship linking the surface polarity, the chemical structure and composition and the crystalline/amorphous phase contribution in the surface modification mechanisms of plasma-exposed polymers is explored. Different polymers were chosen comprising of various structures, functionality, degree of oxidation, crystallinity, and were treated under a particular experimental configuration, and dielectric barrier discharge-type. The plasma parameters and the treatment settings are observed, in relation to relevant surface properties, as surface energy components, surface topography, structural changes and chemical composition, under conditions where the gaseous environment chosen, He-N2, allows complex surface modification, by combined functionalisation and crosslinking.

2286. Gao, L., and T.J. McCarthy, “Teflon is hydrophilic: Comments on definitons of hydrophobic, shear versus tensile hydrophobicity, and wettability characterization,” Langmuir, 24, 9183-9188, (2008).

Comments are made concerning the recent use of adjectives to describe solid surfaces that exhibit anomalously high water contact angle values. We suggest that the meaning of the word hydrophobic be resolved before it is modified, for example, to superhydrophobic and further modified, for example, to sticky superhydrophobic and before the definitions of these new words become issues of contention. The case is made that the first statement in the title is appropriate with experiments that demonstrate significant attractive interaction between liquid water and the surface of solid Teflon. Four types of experiments are described: the interaction of a silicon-supported covalently attached perfluoroalkyl monolayer (a model Teflon surface) with a sessile water drop (1) and with a thin film of water on a clean silicon wafer surface (2), the interaction of 1 and 12 microm diameter solid Teflon particles with a water droplet surface (3), and the interaction of a thin (<5 microm) Teflon film with a water droplet (4). The concepts of shear and tensile hydrophobicity are introduced, and the recommendation that two numbers, advancing and receding contact angle values, should be considered necessary data to characterize the wettability of a surface. That the words hydrophobic, hydrophilic, and their derivatives can and should only be considered qualitative or relative terms is emphasized.

2178. Wolf, R.A., and A.C. Sparavigna, “Hidden problems in surface treatments I: Pinholing,” Converter: Flessibili, Carta, Cartone, 70, 96-104, (2008).

2177. Wolf, R.A., and A.C. Sparavigna, “Hidden problems in surface treatments II: Ground rolls,” Converter: Flessibili, Carta, Cartone, 71, 156-163, (2008).

2167. Kurihara, Y., H. Ohata, M. Kawaguchi, S. Yamazaki, and K. Kimura, “Improvement of adhesion between liquid crystalline polyester films by plasma treatment,” J. Adhesion Science and Technology, 22, 1985-2002, (2008).

Surface modification of thermotropic liquid crystalline aromatic polyester (LCP) films was carried out by low-pressure plasma treatment to improve the initial adhesion as well as the long-term adhesion reliability, a measure of durability between the LCP films used as substrates for printed circuit boards. Plasma irradiation was carried out in various plasma gases with different plasma modes such as reactive-ion-etching, and direct-plasma (DP) with pressures ranging from 6.7 Pa to 26.6 Pa. The introduction of polar groups on the film surface such as phenolic hydroxyl groups and carboxyl groups enhanced the initial adhesion by increased chemical interaction. However, if the concentration of polar groups became too high, the longterm adhesion reliability estimated by the pressure cooker test was degraded due to the acceleration of the penetration of water molecules into the interface. A large surface roughness was also effective in preventing the decrease in the long-term adhesion reliability. However, too much increase in surface roughness decreases the long-term adhesion reliability. The DP-treatment in the O2 atmosphere at a gas pressure of 6.7 Pa was found to be the best plasma condition for both the initial adhesion as well as the long-term adhesion reliability between the LCP films.

2165. Wolkenhauer, A., G. Avramidis, E. Hausweld, H. Militz, and W. Viol, “Plasma treatment of wood-plastic composites to enhance their adhesion properties,” J. Adhesion Science and Technology, 22, 2025-2037, (2008).

In this study, the adhesion properties of adhesives and paints on wood–plastic composites (WPCs) after plasma treatment at atmospheric pressure and ambient air were investigated. Surface energy determination by means of contact angle measurements according to the Owens–Wendt approach and atomic force microscopy to detect changes in surface topography were carried out. An increase in the polar component of surface energy and an increase in surface roughness after plasma treatment were detected, indicating enhanced bond strength. To confirm these results, bond strength tests were conducted. By tensile bond strength tests, increased adhesion of waterborne, solventborne and oil-based paints on plasma treated surfaces was found. Furthermore, by shear bond strength tests, an increase in bond strength of plasma treated WPCs bonded with poly(vinyl acetate) and polyurethane adhesives was ascertained.

2066. Tuominen, M., J. Lahti, and J. Kuusipalo, “Atmospheric plasma treatment equipment and its utilisation in paper converting,” in 2008 Advanced Coating Fundamentals Symposium Proceedings, TAPPI Press, 2008.

1845. Romero-Sanchez, M.D., and J.M. Martin-Martinez, “UV-ozone surface treatment of SBS rubbers containing fillers: Influence of the filler nature on the extent of surface modification and adhesion,” J. Adhesion Science and Technology, 22, 147-168, (2008).

SBS rubbers containing different loadings of calcium carbonate and/or silica fillers were surface treated with UV-ozone to improve their adhesion to polyurethane adhesive. The surface modifications produced on the treated filled SBS rubbers have been analyzed by contact angle measurements, ATR-IR spectroscopy, XPS and SEM. The adhesion properties have been evaluated by T-peel strength tests on treated filled SBS rubber/polyurethane adhesive/leather joints. The UV-ozone treatment improved the wettability of all rubber surfaces, and chemical (oxidation) and morphological modifications (roughness, ablation, surface melting) were produced. The increase in the time of UV-ozone treatment to 30 min led to surface cleaning (removal of silicon-based moieties) due to ablation and/or melting of rubber layers and also incorporation of more oxidized moieties was produced. Although chemical modifications were produced earlier in an unfilled rubber for short time of treatment with UV-ozone, they were more noticeable in filled rubbers for extended length of treatment, mainly for S6S and S6T rubbers containing silica filler. The oxidation process seemed to be inhibited for S6C and S6T rubbers (containing calcium carbonate filler). On the other hand, the S6S rubber containing silica filler and the lowest filler loading showed the higher extent of modification as a consequence of the UV-ozone treatment. The UV-ozone increased the joint strength in all joints, more noticeably in the rubbers containing silica filler, in agreement with the greater extents of chemical and morphological modifications produced by the treatment in these rubbers. Finally, the nature and content of fillers determined the extent of surface modification and adhesion of SBS rubber treated with UV-ozone.

1761. Wu, D., W. Ming, R.A.T.M. van Benthem, and G. de With, “Superhydrophobic fluorinated polyurethane films,” J. Adhesion Science and Technology, 22, 1869-1881, (2008).

A superhydrophobic polyurethane-based film is described, on which the water advancing and receding contact angles are 150° and 82°, respectively. The film was prepared from surface-fluorinated polyurethane (PU), obtained from a well-defined fluorinated isocyanate, with silica particles incorporated within the film. In the absence of the silica particles, smooth fluorinated PU films with about 2 wt% fluorine demonstrate water advancing and receding contact angles of 110° and 63°, respectively. A major cause for the large contact angle hysteresis, similar to the so-called 'sticky' superhydrophobic behavior, on the roughened PU films is believed to originate from the surface reorganization of the fluorinated PU upon contact with water, which is characteristic for the partially fluorinated PU film. When a similar poly(dimethylsiloxane) (PDMS)-based roughened film was made, the water contact angle hysteresis could be reduced significantly, since the long PDMS chain can effectively suppress the surface reorganization upon contact with water.

1760. Pascual, M., R. Sanchis, L. Sanchez, D. Garcia, and R. Balart, “Surface modification of low density polyethylene (LDPE) film using corona discharge plasma for technological applications,” J. Adhesion Science and Technology, 22, 1425-1442, (2008).

Surface modification by corona discharge plasma is one of the most interesting industrial applications for surface modification compared with other techniques which require vacuum conditions. In this work, we have used the corona discharge plasma technique to modify the wettability properties of low density polyethylene (LDPE) film. The effects of this treatment on the surface of LDPE film have been quantified by contact angle measurements, Fourier-transform Infrared Spectroscopy, X-ray photoelectron spectroscopy and atomic force microscopy. With these methods, we have determined how the treatment modifies, activates and functionalizes the surface of LDPE film, increasing its hydrophilic behavior, and how the process parameters influence the uniformity and homogeneity of the treated surface. The results obtained show good treatment homogeneity and an improvement of adhesion properties by the functionalization and etching of the film surface.

1759. Szymczyk, K., and B. Janczuk, “Wetting behavior of aqueous solutions of binary surfactant mixtures to poly(tetrafluoroethylene),” J. Adhesion Science and Technology, 22, 1145-1157, (2008).

Measurements of the surface tension (γLV) and advancing contact angle () on poly(tetrafluoroethylene) (PTFE) were carried out for aqueous solutions of cetyltrimethylammonium bromide (CTAB), cetylpyridynium bromide (CPyB), sodium decylsulfate (SDS), sodium dodecylsulfate (SDDS), p-(1,1,3,3-tetramethylbutyl) phenoxypoly(ethylene glycol)s, Triton X-100 (TX100) and Triton X-165 (TX165) and their mixtures. The results obtained indicate that the values of the surface tension and wettability of PTFE depend on the concentration and composition of the surfactants mixture. In contrast to Zisman finding, there was no linear dependence between cos and the surface tension of aqueous solutions of surfactants and their mixtures for all studied systems, but a linear dependence existed between the adhesional tension and solution surface tension for PTFE in the whole concentration range, the slope of which was –1, indicating that the surface excess concentration of surfactant at the PTFE–solution interface was the same as that at the solution–air interface for a given bulk concentration. It was also found that the work of adhesion of aqueous solutions of surfactants and their mixtures to PTFE surface did not depend on the type of surfactant, its concentration and composition of the mixture. This means that for the studied systems the interaction across PTFE–solution interface was constant, and it was largely of Lifshitz–van der Waals type. On the basis of the surface tension of PTFE and the Young equation and thermodynamic analysis of the work of adhesion of aqueous solutions of surfactants to the polymer surface it was found that in the case of PTFE the changes of the contact angle as a function of the total mixture concentration in the bulk phase resulted only from changes of the polar component of the solution surface tension.

1758. Dixon, D., R. Morrison, P. Lemoine, and B.J. Meenan, “Long term effects of air dielectric barrier discharge treatment of the surface properties of ethylene vinyl acetate (EVA),” J. Adhesion Science and Technology, 22, 717-731, (2008).

1757. Guo, C., S. Wang, H. Liu, L. Feng, Y. Song, and L. Jiang, “Wettability alteration of polymer surfaces produced by scraping,” J. Adhesion Science and Technology, 22, 395-402, (2008).

In this paper, we present a simple, yet novel, method, utilizing scraping to obtain continuous rough microstructures over large areas, leading to a tunable wettability conversion from hydrophilicity to superhydrophobicity on polymer surfaces. A series of polymers ranging from hydrophobic to hydrophilic were used, and we found that the wettability of these polymer surfaces could be modified by the scraping process, irrespective of their hydrophobicity or hydrophilicity. More importantly, those polymers with contact angle ranging from 65° to 90° on their smooth surfaces also exhibit enhanced hydrophobicity after scraping. Our results indicate that 65° is the critical value which is more suitable to define hydrophobicity and hydrophilicity for polymer materials.

3028. Durkee, J.B., and A. Kuhn, “Wettability measurements and cleanliness evaluation without substantial cost,” in Contact Angle, Wettability and Adhesion (Vol. 5), K.L. Mittal, ed., 115-138, VSP, 2008.

Cleanliness can be characterized in industrial applications via "simple" wettability measurements, and has been successfully done so for at least two centuries. A problem in much of general manufacturing and maintenance industries is not that more sophisticated measurement and evaluation technology is necessary to provide value, but that technology developed at least several generations ago has not been more widely and profitably used. This paper describes and references that technology, and identifies published case histories where it has been both successfully and unsuccessfully used.

3027. Della Volpe, C., and S. Siboni, “Principal component analysis and multicomponent surface free energy theories,” J. Mathematical Chemistry, 43, 1032-1051, (2008).

1732. Kumagai, H., T. Kusunoki, and T. Kobayashi, “Surface modification of polymers by thermal ozone treatments,” AZojomo J. Materials Online, 3, (Dec 2007).

Surface modification of polyethylene (PE), poly(vinylchloride) (PVC), and polystyrene (PS), was performed by thermal-ozone (O3) treatment to improve their properties. Polymer films were exposed to dried O3 gas with 3026 ppm at different temperatures. DRS-FT-IR and UV-Vis-NIR absorption methods were applied to observe the surface characteristics of polymers treated. Absorption band assigned to CDouble BondO stretching appeared near 1720 cm-1 in films treated with O3 at 65°C, whereas O3 treatment at 25°C showed no appearance of the CDouble BondO band on the surface. In PS, the O3 oxidation proceeded regardless of temperature. Comparison between PE-O3 and PS-O3 systems showed that different processes of the surface modification occured. Furthermore, contact angle measurements indicated that the surface wettability of PE and PS was improved by the thermal-O3 treatment.

1625. Cohen, E.D., “Corona treatment of metallized cast polypropylene,” AIMCAL News, 23, (Dec 2007).

1623. Bishop, C.A., “Adhesion problems with metallized CPP,” http://www.webcoatingblog.com, Dec 2007.

1622. Bishop, C.A., “Adding nitrogen as third gas in plasma treater,” http://www.webcoatingblog.com, Dec 2007.

1618. Morelock, C.R., Y. Htet, L.L. Wright, and E.C. Culbertson, “AFM studies of corona-treated, biaxially oriented PET film,” Converting, 25, 40-48, (Dec 2007).

 

<-- 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-->