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 281 毫秒
1.
The aminolysis can effectively introduce primary amine (?NH2) groups onto polyester materials, enabling a variety of subsequent surface biofunctionalization reactions. However, less attention has been paid to the basic knowledge of aminolysis reaction in terms of reaction kinetics and its influences on materials properties. In this study, taking the widely used poly(?-caprolactone) (PCL) as a typical example, the influences of diamines and solvent property on the surface ?NH2 density are firstly assessed by using X-ray photoelectron spectroscopy (XPS) and colorimetric analysis. Results show that smaller diamine molecules and nonpolar alcohols could accelerate the reaction. The reaction kinetics with 1,6-hexanediamine is further investigated as a function of temperature, reaction time, and diamine concentration. During the initial stage, the reaction shows a 1st order kinetics with the diamine concentration and has an activation energy of 54.5 kJ/mol. Ionization state of the ?NH2 groups on the PCL surface is determined, revealing that the pKa of ?NH3 + (<5) is much lower than that of the corresponding diamine molecules in solution. After aminolysis, surface hydrophilicity of PCL membrane is significantly enhanced, while surface elastic modulus and average molecular weight are decreased to some extent, and others such as weight, surface morphology and bulk mechanical strength are not apparently changed. The introduced ?NH2 groups are found to be largely lost at 37 °C, but can be mostly maintained at low temperature.  相似文献   

2.
Poly(ethylene terephthalate) (PET) film surfaces were modified by argon (Ar), oxygen (O2), hydrogen (H2), nitrogen (N2), and ammonia (NH3) plasmas, and the plasma‐modified PET surfaces were investigated with scanning probe microscopy, contact‐angle measurements, and X‐ray photoelectron spectroscopy to characterize the surfaces. The exposure of the PET film surfaces to the plasmas led to the etching process on the surfaces and to changes in the topography of the surfaces. The etching rate and surface roughness were closely related to what kind of plasma was used and how high the radio frequency (RF) power was that was input into the plasmas. The etching rate was in the order of O2 plasma > H2 plasma > N2 plasma > Ar plasma > NH3 plasma, and the surface roughness was in the order of NH3 plasma > N2 plasma > H2 plasma > Ar plasma > O2 plasma. Heavy etching reactions did not always lead to large increases in the surface roughness. The plasmas also led to changes in the surface properties of the PET surfaces from hydrophobic to hydrophilic; and the contact angle of water on the surfaces decreased. Modification reactions occurring on the PET surfaces depended on what plasma had been used for the modification. The O2, Ar, H2, and N2 plasmas modified mainly CH2 or phenyl rings rather than ester groups in the PET polymer chains to form C? O groups. On the other hand, the NH3 plasma modified ester groups to form C? O groups. Aging effects of the plasma‐modified PET film surfaces continued as long as 15 days after the modification was finished. The aging effects were related to the movement of C?O groups in ester residues toward the topmost layer and to the movement of C? O groups away from the topmost layer. Such movement of the C?O groups could occur within at least 3 nm from the surface. © 2004 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 42: 3727–3740, 2004  相似文献   

3.
Summary The aminolysis of polyethylene sebacate with 1,6-hexanediamine was investigated and it was found that exchange of ethylene glycol and 1,6-hexanediamine residues occurs with formation of amide-ester copolymers or polyamides, depending on the relative amounts of reactants.  相似文献   

4.
以FeCl3·6H2O作为单一铁源,1,6-己二胺作为胺化试剂,利用无模板的溶剂热方法制备了胺基功能化的磁性Fe3O4纳米粒子,并利用其键合叶酸分子,制备出表面修饰了叶酸的磁性Fe3O4复合纳米粒子。利用傅里叶变换红外光谱仪、X-射线衍射仪、透射电镜、差热-热重分析仪和振动样品磁强计对所得纳米粒子的形貌、粒径、化学组成和磁性能进行了表征。结果表明,叶酸分子通过化学键牢固键合在磁性纳米Fe3O4粒子表面,叶酸修饰的复合纳米粒子仍然具有良好的磁性能。  相似文献   

5.
With the aim of introducing primary amino groups on the surface of poly(ethylene terephthalate) (PET), two methods were compared—the use of ammonia or a combination of nitrogen and hydrogen low-pressure microwave plasma. Several plasma parameters were optimized on the reactor to increase the –NH2 surface density, which was estimated by colorimetric titration and X-ray photoelectron spectroscopy (XPS). These techniques show that whatever the plasma treatment, almost 2 –NH2/nm2 are incorporated on PET films. Emission spectroscopy highlighted a correlation between the density of primary amino groups and the ratio between an NH peak intensity and an Ar peak intensity (INH/IAr). Variation in surface hydrophilicity with aging in air after plasma treatment was monitored with contact angle measurements and showed a hydrophobic recovery. This was confirmed by XPS, which suggests also that surfaces treated by NH3 plasma are more stable than surfaces treated by N2/H2.  相似文献   

6.
Poly(3-hydroxybutyric acid-co-3-hydroxyvaleric acid) (PHBV) was covered with an hydrophobic layer from plasma polymerization of tetradecafluorohexane, octadecafluorooctane, 3,3,4,4,5,5,6,6,6-nonafluoro-hex-1-ene and 3,3,4,4,5,5,6,6,7,7,8,8,8-tridecafluoro-oct-1-ene. The water contact angle increased from 74° for untreated films to 98° for the treated films and the surface energy decreased from 40.9 to 18.8 mJ m−2. XPS analysis showed the introduction of 54% fluorine and 3-7% oxygen, the binding energies were assigned to chemically differently bonded carbon atoms. CF2/CF3 molar ratio in plasma layers was lower than that in the monomers due to molecular fragmentation, however the extent of monomer structure retention in PFH, PFO saturated chains is higher than in NFH, TDFO unsaturated chains. Biodegradation tests under aerobic conditions showed that the fluorinated plasma layer inhibited the biodegradation of the PHBV film underneath.  相似文献   

7.
Crystallization behavior of blends of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV) and poly(1,2-propandiolcarbonate) (PR(CO2)) has been investigated by polarized light microscopy (PLM). The spherulite growth rates (SGR) of all blends were faster than that of pure PHBV, and the spherulite growth rates of PHBV in the PHBV/PR(CO2) blends reduced with increasing PR(CO2) weight fraction. There are two melting peaks in both the pure PHBV and the PHBV/PR(CO2) blends. The melting peak of PHBV/PR(CO2) blends was reduced by lower temperature about 20K as compared to PHBV and the higher temperature melting peak was increased by about 10K in the blends.  相似文献   

8.
Shaped binderless ZSM-11 zeolite catalysts were synthesized via a dry-gel conversion technique from 70ZSM-11/30 SiO_2mix extrudates. 1,6-hexanediamine combined with tetrabutylammonium bromide was proved to be the best structure directing agent for the synthesis of the binderless ZSM-11 catalyst, without adding other alkaline materials. The 70HZSM-11/30 SiO_2mix serials materials crystallized for different times were detected by X-ray diffraction(XRD), nuclear magnetic resonance(NMR), scanning electron microscopy(SEM), transmission electron microscopy(TEM), scanning transmission electron microscopy–energy dispersive spectroscopy(STEM–EDS) techniques, and so on. In order to investigate the possible crystallization mechanism, the textural and structural properties of 70HZSM-11/30 SiO_2mix serials samples were further characterized by N_2 adsorption–desorption. Acid properties were determined by temperature-programed desorption of NH_3(NH_3-TPD) and pyridine adsorption-infrared(Py-IR) measurements. In the alkylation of benzene with dimethyl ether, the serials catalysts exhibited different benzene conversions. 70HZSM-11/30 SiO_2mix showed the lowest benzene conversion while sample 70HZSM-11/30 SiO_2mix-6.5h synthesized only for 6.5h displayed a higher benzene conversion, even higher than the value over 70HZSM-11/30Al_2O_3mix. Extending the crystallization time, the obtained samples displayed the increased benzene conversion in general under the same reaction conditions. In the end, the relation of physicochemical properties with the reaction performance was investigated.  相似文献   

9.
The influence of the pulsed CO2 laser irradiation on the surface structure of the LDPE film was investigated. Significant changes were observed on the surface of laser treated films as it was verified by the attenuated total reflectance Fourier transform infrared (ATR-FTIR) spectroscopy, scanning electron microscopy and contact angle-measurement. Formation of polar functional groups onto the LDPE surfaces exhibited by the ATR-FTIR spectra was shown to be strongly dependent on the number of the CO2 laser pulses. The intensity of the polar groups increased with increasing the number of pulses up to two and then slightly decreased at three laser pulses. This was also confirmed with the contact angle measurements in which the sample subjected to two laser pulses showed the highest wettability i.e. the lowest water drop contact angle. The concentration of peroxide groups formed on the surface of the laser treated films was determined quantitatively by UV spectroscopic method using iodide procedure. The latter results showed a similar trend with the results obtained using FTIR spectroscopy.  相似文献   

10.
The density functional theory B3LYP/6-31G(d, p) method is employed to study the mechanism of aminolysis reaction of p-substituted phenyl acetates (CH3C(O)OC6H4X, X = H, NH2, and NO2) with ammonia in the gas phase. Two reaction pathways are considered: the concerted process and the stepwise pathway through neutral intermediates. The substituent effects of the leaving groups on the reactivity of phenyl acetates are discussed. The solvent effect of acetonitrile on the title reaction is also assessed by the polarizable continuum model (CPCM model) at B3LYP/6-31++G(d, p) level of theory. The calculated results show that the activation barriers of the concerted pathways are lower than those of the rate-controlling steps of the stepwise processes for all the three aminolysis reactions. This aminolysis of phenyl acetates is more favorable for X = NO2 than for X = H and NH2 in the gas phase and in acetonitrile.  相似文献   

11.
Due to a smooth hydrophobic surface of Kapton film, it is very difficult to coat an inorganic oxide coating on its surface. In this study, the surface of Kapton was treated by NH3·H2O and silicon coupling agent to improve the polymer surface wettability. Changes in surface hydrophilicity were studied by contact angle measurement. The samples were irradiated by atomic oxygen (AO) in a ground‐based simulation system. TiO2/SiO2 multi‐layer coatings were prepared on the surface of Kapton by sol–gel method to resist AO erosion. The optical transmittance and surface morphology of samples were investigated by UV–vis spectroscopy and scanning electronic microscope. The results indicated that the TiO2/SiO2 sol could easily form a uniform thin coating on the surface of pretreated Kapton. After AO exposure, the coatings became more compact, without peeling off. The AO erosion yield value of coated Kapton was sharply down, and the samples had good optical transparency. Copyright © 2017 John Wiley & Sons, Ltd.  相似文献   

12.
Kapton films were treated with seven plasmas: Ar-, N2-, O2-, CO-, CO2-, NO-, and NO2- plasmas. Surface properties and chemical composition of the plasma-treated Kapton films were investigated from the contact angle measurement, and the IR and XPS spectra. The plasmas, especially NO- and NO2-plasma, made the Kapton film surface hydrophilic. The XPS and IR spectra showed that the plasma led to the modification of the imide groups in the Kapton film to secondary amide and carboxylate groups.  相似文献   

13.
PBO fiber is one of the most promising reinforcements in resin matrix composite because of its excellent mechanical properties. However, the inert and smooth surfaces make it the poor interface adhesion with resin matrix, which seriously limits the application in composites. In this article, we report a method to modify the surface of PBO fibers with 2,2-Bis (3-amino-4-hydroxyphenyl) hexafluoropropane(BisAPAF)in supercritical CO2 to enhance interfacial properties. Chemical structures, surface elemental composition and functional groups, and surface morphology were characterized by FT-IR spectrometer, X-ray photoelectron spectroscopy (XPS), and scanning electron microscopy (SEM), respectively. The mechanical properties of the samples were tested by a tensile tester. Static contact angle and microdebonding tests were used to characterize the wetting ability and interfacial shear strength (IFSS) of the fiber and epoxy resin. The results showed that the BisAPAF could be solved in scCO2 and introduced more groups, –NH2, –OH, and –CF3 on the fiber surface, resulting in the mechanical properties and the wettability of PBO fiber slightly improved. Moreover, the fiber surface roughness was also increased obviously. The IFSS between the modified PBO fiber and epoxy resin increased from 8.18 MPa to 31.4 MPa when the treating pressure was 14 MPa. In general, the method to modify PBO fibers surface using BisAPAF in scCO2 can effectively improve their interfacial properties.  相似文献   

14.
1H-NMR spectroscopic analysis indicates that cucurbit[7]uril can form a stable inclusion complex with 1,6-hexanediamine, while cucurbit[5]uril cannot form pseudorotaxane with 1,6-hexanediamine under our experimental conditions. This was confirmed by the crystal structure of the complex. The cavity of cucurbit[8]uril seems to be large for binding 1,6-hexanediamine efficiently. And a simple, mild, high-yield (>80%) method has been described for the synthesis of rotaxanes through the self-assembly of pseudorotaxanes of cucurbit[n]uril (n=6, 7)/1, 6-hexanediamine and sodium tetraphenylborate. The obtained rotaxanes are held intact solely by noncovalent interactions, and are characterized by elemental analysis, 1H-NMR, ESI-MS and MALDI-TOF MS.  相似文献   

15.
The stirred interfacial polycondensation of phenylphosphonic dichloride and 1,6-hexanediamine has been studied as a function of several reaction variables. The reaction is rapid, being completed in less than 1 min. When organic solvent is varied and reactant molar ratio is varied with an excess of the acid chloride, yield is constant. When reactant molar ratio is varied polymer yield increases with increase in amine concentration. When reactant concentration is increased yield increases. With the addition of a soluble salt in the aqueous phase yield is increased. The above indicates that the diffusion of the amine to the reaction zone is of primary importance in determining polymerization rate and that the diffusion of the acid chloride is relatively unimportant. Polymer yield was found to be dependent on the pH of the amine in the aqueous phase. The observed trend is related to the apparent solubility of the amine in the aqueous phase such that the greater the apparent solubility of the amine, the less the polymer yield. Polymer molecular weight is found to be independent of reaction variables tested. Polymer was also formed from the condensation of phenylphosphonic dichloride with p-phenylenediamine, H2N-D-NH2 (where D is a 36-carbon hydrocarbon chain), 1,3-di-4-piperidylpropane, and 4-aminomethylpiperidine; phenyl phosphorodichloridate with 1,6-hexanediamine; chloromethyl phosphonic dichloride with 1,6-hexanediamine.  相似文献   

16.
Silicone rubber samples were treated by CF4 capacitively coupled plasma at radio frequency (RF) power of 60, 100 and 200 W for a treatment time up to 20 min under CF4 flow rate of 20 sccm, respectively. Static contact angle, ATR-FTIR and XPS, and AFM were employed to characterize the changes of surface on hydrophobicity, functional groups, and topography. The results indicate the static contact angle is improved from 100.7 to 150.2°, and the super-hydrophobic surface, which corresponds to a static contact angle of 150.2°, appears at RF power of 200 W for a 5 min treatment time. It is suggested that the formation of super-hydrophobic surface is ascribed to the co-action of the increase of surface roughness created by the ablation reaction of CF4 plasma and the formation of [–SiF x (CH3)2−x –O–] n (x = 1, 2) structure produced by the direct attachment of F atoms to Si.  相似文献   

17.
Incorporating amino groups is an efficient strategy for the tuning properties of energetic materials. However, there is no unanimous conclusion on the effect of the number of amino groups (−NH2) on performance. Therefore, in this study, different number of −NH2 of four energetic salts of triazolium based on oxadiazole and triazole were designed and synthesized. The structure features of energetic salts 4 – 6 were then investigated by single-crystal X-ray diffractions and Hirshfeld surfaces analyses. Afterward, the effects of −NH2 were evaluated by thermal stability, impact sensitivity and detonation performance. All these energetic salts were insensitive to mechanical stimulation (IS >40 J), but the thermal decomposition temperatures of energetic salt 5 – 7 with −NH2 are 24 °C to 54 °C higher than energetic salt 4 without −NH2. Moreover, energetic salt 5 with one −NH2 has the highest theoretical detonation properties compared to those without −NH2 ( 4 ) and with two −NH2 ( 6 , 7 ). These observations revealed that appropriate amount of −NH2 can lead to desirable increase in the energetic properties, and this work can offer guidance for the design and synthesis of further energetic salts.  相似文献   

18.
ESCA and contact angle measurements were used to characterize the surfaces of polypropylene and glass substrates exposed to CF4, CF3H, CF3Cl, and CF3Br plasmas. The use of both organic and inorganic substrates allowed clear distinction between treatments which led to plasma polymerization and treatments which caused grafting of functional groups directly to the substrate surfaces. CF4 plasmas were the only treatments studied which fluorinated polypropylene surfaces directly, without the deposition of a thin, plasma-polymerized film. CF3H polymerized in a plasma, while CF3Cl and CF3Br plasmas caused chlorination and bromination of polypropylene surfaces, respectively. Correlations were made between the active species present in the plasmas and the surface chemistry observed on the treated polypropylene substrates.  相似文献   

19.
The substitution reactions of the complexes [{trans‐Pt(NH3)2H2O}2(μ‐1,4‐diaminobutane)]4+ ( I ), [{trans‐Pt(NH3)2H2O}2(μ‐1,6‐diaminohexane)]4+ ( II ), and [{trans‐Pt(NH3)2H2O}2(μ‐1,8‐diaminooctane)]4+ ( III ), with nucleophiles L‐cysteine (L‐Cys), glutathione (GSH), guanosine‐5′‐monophosphate (5′‐GMP), L‐histidine (L‐His), and pyridine were studied in 0.1 M NaClO4 aqueous solutions at pH = 2.5. The substitutions were studied under pseudo‐first‐order conditions as a function of concentration and temperature using UV–vis spectrophotometry. At three different temperatures (288, 298, and 308 K) the reactions of the II and III complexes and 5′‐GMP were studied. The order of reactivity of study ligands is L‐Cys > GSH > 5′‐GMP > L‐His > pyridine and the order of reactivity of the complexes is I < II ≈ III . The obtained results indicate that the structure of the alkanediamine linker in the dinuclear Pt(II) complexes controls the substitution process. The negative values reported for entropy of activation confirmed the associative substitution mode. These results are discussed in order to find the connection between structure and reactivity of the dinuclear Pt(II) complexes.  相似文献   

20.
Because of the strong chemical inertness of high-modulus carbon fiber (HMCF), the surface treatment of HMCF becomes less effective. In this work, multistage anodization method was applied to modify HMCF. NH4HCO3 and (NH4)2SO4 were used as electrolytes to study the influence of electrolyte type on the oxidation effect in the multistage oxidation process. The HMCF was treated by multistage anodization in the following three ways: NH4HCO3 → NH4HCO3, (NH4)2SO4 → (NH4)2SO4, and NH4HCO3 → (NH4)2SO4. In order to compare with the multistage anodization methods, HMCF was also anodized by NH4HCO3 and (NH4)2SO4 using single anodization method, respectively. The treated HMCFs were characterized in detail using X-ray photoelectron spectroscopy (XPS), Raman spectroscopy, Scanning Electron Microscope (SEM), Atomic Force Microscope (AFM), and dynamic contact angle analysis, respectively. The results of XPS exhibited that HMCF modified by multistage anodization showed more functional groups on its surface. More sp2 hybridized bonds were converted into sp3 hybridized bonds after modification by multistage anodization, especially when (NH4)2SO4 was used as electrolyte. The results of SEM/AFM indicated that HMCF modified by multistage showed a greater surface roughness. Interlaminar shear strength (ILSS) was used to observe the interfacial properties of HMCF between epoxy resin. The results showed that ILSS of untreated carbon fiber composite was only 27.3 MPa; it was increased to 71.8 and 78.6 MPa after treated with single anodization by NH4HCO3 and (NH4)2SO4, respectively. After modification by multistage anodization, the ILSS increased in different degrees compared with single anodization method. Especially, the ILSS reached 93.1 MPa when the method of using NH4HCO3 followed by (NH4)2SO4 was used to treat HMCF.  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