首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The sessile drop technique was used to investigate the evolution of the physicochemical properties of cedar wood as a function of contact time with the Penicillium expansum spores. The most important finding showed that the impact of different contact periods (2, 4, 6, 8, 10, and 24 hr) on the wood surface were very indicative. In fact, after 2 hr of contact, the results have shown a significant impact of the bioadhesion of spores to the substrate on both the hydrophobic character (θW = 108.5°; ΔGiwi = ?28.25 mJ/m2), the electron donor (γ? = 13.63 mJ/m2), and the electron acceptor (γ+ = 4.35 mJ/m2) parameters that were significantly reduced compared to the initial wood (θW = 118.5°; ΔGiwi = ?6.29 mJ/m2; γ? = 32.1 mJ/m2; and γ+ = 9.1 mJ/m2). In addition, this decrease of parameters continued over time to stabilize after 10 hr of contact. Indeed, after 24 hr, the acid/base properties were almost zero and the contact angle with water decreased to 30°. Moreover, it was found that the coefficient of correlation (r2) was strong between the contact angle with water, the surface energy, and the electron acceptor character with the contact time parameter with values (r2 = 0.65), (r2 = 0.79), and (r2 = 0.68), respectively.  相似文献   

2.
The theory of the contact angle of pure liquids on solids, and of the determination of the surface free energy of solids, γs, is reviewed. The basis for the three components γLW s, γ⊕s, and γ?s is developed, and an algebraic expression for these properties in terms of measured contact angles is presented. The inadequacy of the 'two-liquid' methodology (which yields a parameter, 'γp') is demonstrated. Attention is given to contact angle hysteresis and to the film pressure, πe. Some recommendations are made with regard to contact angle measurements. A new treatment of hydrophilicity, and of the scale of hydrophobic/hydrophilic behavior, is proposed. It is shown that there are two kinds of hydrophilic behavior, one due to Lewis basicity (electron-donating or proton-accepting structures) and the other due to Lewis acidity (electron-accepting or proton-donating structures). The properties γ? and γ are the quantitative measures of these types of behavior and they are structurally independent of each other. A triangular diagram, with γLW at the hydrophobic corner, and γ and γ? at the two hydrophillic corners, is suggested.  相似文献   

3.
The adhesion of Bacillus subtilis and Bacillus sp. isolated from Fez cedar wood decay has been investigated. Furthermore, the physicochemical proprieties including hydrophobicity and electron donor/electron acceptor (Lewis acid–base) of both bacteria and substrata were evaluated using contact angle measurements. The results show that Bacillus subtilis has a hydrophobic character (ΔG iwi = –20 mJ/m2). In contrast, Bacillus sp. exhibits a hydrophilic (ΔG iwi = –20 mJ/m2), electron donating (γ) and weakly electron accepting (γ+) character. With respect to the substrata surface, we found that the cedar wood used in this work, was hydrophobic in character, having relatively more electron-donor that electron-acceptor properties (γ = 6 ± 4 mJ/m2; γ+ = 0 ± 3 mJ/m2). The phenomena of adhesion were observed by environmental scanning electron microscopy (ESEM) and cell adhesion was quantified using a Matlab program. The analysis of images obtained by ESEM show that the both cells was able to adhere to the wood substrata and the quantitative adhesion results showed that the surface coverage by Bacillus sp. (90%) was higher than that by the Bacillus subtilis strain (40%).  相似文献   

4.
The initial microorganism adhesion on substrate is an important step for biofilm formation. The surface properties of the silicone and Bacillus cereus were characterized by the sessile drop technique. Moreover, the physicochemical properties (hydrophobicity; electron donor/electron acceptor) of surface adhesion and the impact of bio adhesion on the silicone were determined at different time of contact (3, 7, and 24?h). The results showed that the strain was hydrophilic (Giwi?=?3.37?mJ/m2), whereas the silicone has hydrophobic character (Giwi?=??68.28?mJ/m2). Silicone surface presents a weak electron-donor character (γ ??=?2.2?mJ/m2) conversely to B. cereus that presents an important electron donor-parameter (γ ??=?31.6?mJ/m2). The adhesion of B. cereus to silicone was investigated using environmental scanning electron microscope and image analysis was assessed with the Matlab® program. After 3?h of contact, the data analysis, confirmed the bio adhesion with an amount of 9.6105?cfu/cm2 adhered cells. After 24?h, the percentage of silicone covered reached 93%. Furthermore, despite the difference in hydrophohbicity, the interaction between B. cereus and substrata was favoured by the thermodynamic model of adhesion (ΔG adhesion ?<?0). The real time investigation of the effect of B. cereus adhesion on the physicochemical properties of silicone has revealed that the substrata becomes hydrophilic (θ°?=?47.3, ΔGiwi?=?23.7?mJ/m2), after 7?h of contact. This bio adhesion had also favoured the increase of electron donor/acceptor character of silicone (γ ??=?53.1?mJ/m2 and γ +?=?5.3?mJ/m2).  相似文献   

5.
Biofilms are the most common mode of bacterial growth in nature and the formation will occur on organic or inorganic solid surfaces in contact with a liquid. The aims of this study were, by combining numeration and sessile drop technique, (i) to characterize the structural dynamics of dairy biofilm growth and the physico chemical properties on silicone and stainless steel and (ii) to evaluate the impact of bio-adhesion on chemistry of surfaces at different times of contact (2, 7, 9 and 24?h). Significantly, greater biofilm volumes were observed after 48?h on two materials. Gram-positive bacteria and fungal population exhibited a significantly higher biofilm organization than gram-negative (43–64%). Elsewhere, after 48?h, results showed a slight difference on gram-negative adhered cells on stainless steel than silicone (2.6?×?107?cfu/cm2 and 4.7?×?105?cfu/cm2, respectively). Moreover, the physico chemical properties of the surfaces showed that the silicone and stainless steel have a hydrophobic character (Giwi?=??68.28?mJ/m2 and ?57.6?mJ/m2, respectively). Also, both the surfaces present a weak electron donor character (γ ??=?2.2?mJ/m2 and 4.1?mJ/m2, respectively). The real-time investigation of the impact of dairy biofilm on the physico chemical properties of the materials has shown a decrease of hydrophobicity degree of the silicone surface that becomes hydrophilic (ΔGiwi?=?11.47?mJ/m2) after 7?h and the increase of electron donor character (γ ??=?75.8?mJ/m2). Elsewhere, bio-adhesion on stainless steel was accompanied with a decrease of hydrophobicity degree of the surface, which becomes hydrophilic after 7?h of contact (ΔGiwi?=?6.62?mJ/m2) and the increase of the electron donor character (γ ??=?44.8?mJ/m2). While, after 24?h of contact, results showed a decrease of the hydrophilicity degree and surface energy components of silicone and stainless steel that become hydrophobic (ΔGiwi?=??21.2?mJ/m2 and ΔGiwi?=??56.51?mJ/m2, respectively) and weak electron donor (γ ??=?14.0 and 2.3?mJ/m2, respectively).  相似文献   

6.
The solid surface tension γsv of hydrophobic polymer powders has been determined using the capillary penetration technique. By plotting Kγlv cos ζ, where K is a geometric factor, versus the liquid surface tension γlv, the following values of γsv were directly derived from the curves: poly(tetrafluoroethylene) γsv = 20.4 mJ/m2, polypropylene γsv = 30.2 mJ/m2, polyethylene γsv = 34.4 mJ/m2, and polystyrene γsv = 27.5 mJ/m2. These values are in good agreement with the γsv values obtained from contact angle measurements on flat and smooth solid surfaces of the same materials. If the contact angles were first calculated from the capillary penetration experiments, which is the usual procedure applied in the literature, distinctly higher contact angles were obtained. Obviously these angles are affected by the powder morphology and are therefore meaningless contact angles in terms of a surface energetic interpretation.  相似文献   

7.
The friction and adhesion between a fluorocarbon monolayer-coated surface against a hydrocarbon monolayer-coated surface has been directly measured. The friction was found to be lower than the friction between a hydrocarbon monolayer against a hydrocarbon monolayer and a fluorocarbon monolayer against a fluorocarbon monolayer. No stick-slip sliding was observed for speeds from 0.8?µm/s to 2.6?µm/s. The fluorocarbon–hydrocarbon interface was adhesive, with the energy of interaction measured to be 14.9?mJ/m2?±?1.0?mJ/m2. As predicted from theory, the magnitude of the adhesion of a fluorocarbon monolayer against a hydrocarbon monolayer is between that measured for a fluorocarbon monolayer against a fluorocarbon monolayer and a hydrocarbon monolayer against a hydrocarbon monolayer. One may note that the interfacial energy, γ, follows the general trend γFC/FC?<?γHC/FC?<?γHC/HC, whereas the shear stress, τ, varies according to τFC/HC?<?τHC/HC?<?τFC/FC.  相似文献   

8.
Surface energies of amorphous cellulose “beads” were measured by IGC at different temperatures (50 to 100°C) using n-alkane probes (pentane to undecane). The equation of Schultz and Lavielle was applied which relates the specific retention volume of the gas probe to the dispersive component of the surface energy of the solid and liquid, γd s and γd l, respectively, and a parameter (“a”) which represents the surface area of the gas probe in contact with the solids. At 50°C, γd s was determined to be 71.5 mJ/m2, and its temperature dependence was 0.36 mJ m?2 K?1. Compared with measurements obtained by contact angle, IGC results were found to yield higher values, and especially a higher temperature dependence, d(γd s)/dT. Various potential explanations for these elevated values were examined. The surface energy, as determined by the Schultz and Lavielle equation, was found to depend mostly on the parameter “a”. Two experimental conditions are known to affect the values of “a”: the solid surface and the temperature. While the surface effect of the parameter “a” was ignored in this study, the dependence of the surface energy upon temperature and probe phase was demonstrated to be significant. Several optional treatments of the parameter “a” were modeled. It was observed that both experimental imprecision, but mostly the fundamental difference between the liquid-solid vs the gas-solid system (and the associated theoretical weakness of the model used), could explain the differences between γd s and d(γd s)/dT measured by contact angle and IGC. It was concluded that the exaggerated temperature dependence of the IGC results is a consequence of limitations inherent in the definition of parameter “a”.  相似文献   

9.
The thin-layer wicking technique was used to determine the surface free energy components and the surface character of three celluloses (Sigmaccll 101, Sigmacell 20, and Avicel 101), using an appropriate form of the Washburn equation. For this purpose, the penetration rates of probe liquids into thin porous layers of the celluloses deposited onto horizontal glass plates were measured. It was found that the wicking was a reproducible process and that the thin-layer wicking technique could be used for the determination of the celluloses' surface free energy components. The size of the cellulose particles was characterized with the Galai CIS-100 system and their crystallinity was measured by X-ray diffraction. The three celluloses have high apolar (yLWS = 50-56 mJ/m2) and electron donor (γs = 42-45 mJ/m2) components, while the electron acceptor component (γS+ ) is practically zero. The free energy interactions of cellulose/water/cellulose calculated from the components are positive, regardless of the cellulose crystallinity. This would mean that the cellulose surfaces have a hydrophilic character. However, the work of spreading of water has a small negative value (3-9 mJ/m2), indicating that the surfaces are slightly hydrophobic. It is believed that the work of spreading characterizes better the hydrophobicity of the surface than the free energy of particle/water/particle interaction, because in the latter case, no electrostatic repulsion is taken into account in the calculations.  相似文献   

10.
A method has been developed to calculate the interfacial tension of sessile drops and captive bubbles of arbitrary contact angle by measuring the drop diameter and vertical distance to the apex at arbitrary horizontal planes within the drop. The procedure works in theory for any contact angle with an accuracy on the order of 0.1%. However, practical limitations reduce the range of angles to roughly 50°–180° but do not restrict the range of interfacial tensions (at least 0.01 mJ/m2 to 72.0 mJ/m2). The optimal strategy is to use the method at several points on a single drop and to calculate the mean and standard deviation of the resulting interfacial tensions.  相似文献   

11.
The surface tensions of fluorinated polysiloxanes prepared by hydrosilylation of unsaturated perfluoroalkyl esters derived from undecylenic acid [CH2?CH? (CH2)8? COO? CH2? CH2? RF, with RF = C6F13, C8F17, and C8F17? (CH2)10COO? CH2? CH2? CH?CH2] by methylhydrodimethylsiloxane copolymers of various Si? H contents have been measured. The critical surface tensions, γc, and the solid surface tensions, γDs, were deduced from n-alkane and water contact angle data. They decrease as the perfluoroalkyl graft content of the copolymers increases. Some of them, which are in the range of the lowest surface tension fluoro polymers known, are observed when the fluorinated segments are self-organized at the interface, i.e. when the polymers are mesomorphous or crystalline at room temperature.  相似文献   

12.
Physicochemical characterization of microorganism is very important in a wide range of scientific and technological fields. In this study, we reported the isolation and the molecular identification of actinomycetes recovered from cedar wood decay. The isolates named H5 and H8 were identified by 16S rDNA sequencing and were shown to belong to the genus Nocardia and Streptomyces, respectively. Furthermore, physicochemical proprieties including hydrophobicity, electron donor/acceptor, and the Lifshitz–van der Waals (γLW) of these strains were evaluated using contact angle measurements. The results showed that Nocardia sp. (H5) had a hydrophobic (ΔGiwi?=??78.56?mJ/m2) and a weak electron donor/acceptor character. In contrast, results from contact angle measurements showed that the surface free energy of Streptomyces strains (H2, H3, and H8) were ΔGiwi?=?20.71?mJ/m2, ΔGiwi?=?30.63?mJ/m2, and ΔGiwi?=?15.35?mJ/m2, respectively, classifying these microorganisms as hydrophilic bacterium. Moreover, the three strains were predominantly electron donating (γ–?) and exhibit a weak electron-accepting (γ+) character.  相似文献   

13.
A cationic Gemini surfactant with a benzene ring (abbreviated as C14‐CGB) was synthesized in 2 steps with aniline, epichlorohydrin, and N,N‐dimethyltetradecylamine as starting materials. This product was characterized using mass spectroscopy and nuclear magnetic resonance (1H NMR). The critical micelle concentration (CMC) and surface tension (γcmc) of C14‐CGB were measured from 298 to 313 K and thermodynamic parameters of micellization were calculated. The results showed that the CMC and γcmc were 1.269 × 10?3 mol L?1 and 38.33 mN m?1 at 298 K, respectively. Moreover, upon increasing the temperature, the CMC increases, γcmc decreases, the maximum surface adsorption capacity (Γmax) decreases, and the minimum molecular area (Amin) increases. The emulsified asphalt test showed that C14‐CGB is a slow‐breaking asphalt emulsifier exhibiting excellent emulsifying ability.  相似文献   

14.
An average value for the surface energy (γ¯S) of palygorskite was determined from experimental data of spreading pressure, the surface tension of water and its contact angle using a formula based on the combination of the Young equation with a general equation of pair interaction. The value found is 226.6 mJ m− 2. Some factors that affect the determination of surface energy via solid-liquid interaction are exposed. The backgrounds of previous formulas for the calculation of the surface energy of palygorskite are also critically examined.  相似文献   

15.
The design, synthesis and interfacial behaviors of six asymmetric carboxyl betaine surfactants (BCm?n, m, n = 8, 10, 12, or 14, mn) derived from s‐triazine, which were prepared from cyanuric chloride, aliphatic amines, N,N‐dimethylpropane‐1,3‐diamine, followed by the reaction with sodium chloroacetate, are reported. The structures were confirmed by MS, 1H NMR and FT‐IR. Compared with symmetric surfactants (BCn?n, n = 8, 10, 12, or 14) we previously synthesized, the asymmetric series show superior surface activity. The γCMC of surfactants BC10?8, BC12?8, BC14?8 and BC12?10 is all below 30 mN/m. The minimum alkane carbon number of these ten surfactants is determined to be between 10 and 14. The interfacial behaviors between the alkanes and the solutions of triazine carboxyl betaine surfactants show that surfactants with a total carbon number in hydrophobic chains between 16 and 22 exhibit the ability to reduce the interfacial tension to an ultra‐low value (10?3 mN/m). The surfactants with longer hydrocarbon chains display strong affinity to the alkanes with longer chains.  相似文献   

16.
Biodegradation of poy(L ‐lactide) (PLLA) films and filaments recovered with hydrophilic layer (contact angle = 14°, surface energy γs = 70.9 mJ m?2) from allylamine plasma polymerization was investigated under aerobic conditions in sludge. XPS and FTIR‐ATR analysis of the plasma layer showed 14.4% N and 16.6% O mainly as amide group. Optical microscopy showed much bacteria colonies on treated PLLA surface than on untreated one. Weight loss and oxygen consumption after 65 days were 4–5% and 4 mg h?1 per gram polymer respectively. The fact that biodegradation lag‐phase for treated PLLA was released quicker (7 days) than untreated one (14 days), could be related to the presence of hydrophilic plasma layer that improved swelling‐dissolution of hydrolyzed molecular fragments. © 2007 Wiley Periodicals, Inc. J Appl Polym Sci, 2007  相似文献   

17.
The direct potentiometric determination of the normal potentials of Cu/Cu+, Cu+/Cu2+, Ag/Ag+, Hg/Hg2+2 couples in nitromethane and the comparison of these values with the normal potentials in aqueous solution makes possible the calculation of solvation coefficients γ(t)(Mn+) of the corresponding cations; the Strehlow hypothesis (γ(Fc) = γ(Fc+)) has been used. An indirect method using the determination of the solubility constants of alkaline chlorides and alkaline earth chlorides leads to the normal potentials of alkaline and Mg2+ cations and thus to the γ(t) values of these ions.The measures of the AgX-2 stability constants and of the AgX solubility constants lead to the γ(t) values of X? and AgX?2 anions (X? = Cl?, Br?, I?, Scn?, CH3CO?2, CN?, BrO?3, (C6H5)4B?). A comparison of γ(t)(H+) and γ(t)(Ag+) for 20 solvents is given.  相似文献   

18.
Low-rate dynamic contact angles of 13 liquids on a polystyrene polymer are measured by an automated axisymmetric drop shape analysis – profile (ADSA-P). It is found that 7 liquids yielded non-constant contact angles, and/or dissolved the polymer on contact. From the experimental contact angles of the other 6 liquids, it is found that the liquid-vapor surface tension times cosine of the contact angle changes smoothly with the liquid-vapor surface tension, i.e. γlvcosθ depends only on γlv for a given solid surface (or solid surface tension). This contact angle pattern is in harmony with those from other inert and non-inert (polar and non-polar) surfaces (7–13, 24–26). The solid-vapor surface tension calculated from the equation-of-state approach for solid-liquid interfacial tensions (33) is found to be 29.8 mJ/m2, with a 95% confidence limit of ±0.5 mJ/m2 from the experimental contact angles of 6 liquids.  相似文献   

19.
In 1987, van Oss, Chaudhury and Good introduced the Lewis acid (or hydrogen-bond acidic) component, γ+, and Lewis base (or hydrogen-bond basic) component, γ?, and assumed the ratio of γ+ and γ? for water at 20°C to be 1.0. With that ratio, the base components, γ?, for other liquids and polymers appeared to be overestimated. Recently, we unexpectedly found a correlation between γ+ and γ? and the linear solvation energy relationship (LSER) parameters α (hydrogen-bond-donating ability, HBD) and β (hydrogen-bond-accepting ability, HBA), introduced by Taft and Kamlet in 1976. Interestingly, we found the ratio for the normalized α and β for water at ambient temperature to be 1.8 instead of 1.0. Based on this new ratio for the corresponding γ+ and γ?, the calculated total surface tensions for other liquids and polymers at 20°C are generally unchanged, as expected, despite the favorable changes in the γ+ and γ? ratio to make them less basic. In addition, the implications of other LSER parameters, e.g. ∏? and δ2 H on surface properties will be briefly mentioned.  相似文献   

20.
A novel ultraviolet (UV)‐curable monomer α,ω‐dichloropolysiloxane was synthesized by the telomerization of dichlorodimethylsilane and octamethylcyclotetrasiloxane (D4). The products with very low peel strength (<0.332 N/cm) could be used as release agents in pressure‐sensitive adhesives. Moreover, the values of the dispersion component of surface energy (γ) from the films of UV‐curable prepolymers (26.40–33.75 mJ/m2) were determined and the effects of γ on the reduction of adhesion were investigated. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 86: 2135–2139, 2002  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号