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 31 毫秒
1.
Electroless plating reactions are classified according to four overall reaction schemes in which each partial reaction is either under diffusion control or electrochemical control. The theory of a technique based on the observation of the mixed potential as a function of agitation, concentration of the reducing agent and concentration of metal ions is presented. Using this technique it is shown that in electroless copper plating the copper deposition reaction is diffusion-controlled while the formaldehyde decomposition reaction is activation-controlled. Values of the kinetic and mechanistic parameters for the partial reactions obtained by this method and by other electrochemical methods indicate that the two partial reactions are not independent of each other.Nomenclature a Tafel slope intercept - A electrode area - b M Tafel slope for cathodic partial reaction - b R Tafel slope for anodic partial reaction - B M diffusion parameter for CuEDTA2– complex - diffusion parameter for dissolved oxygen - B R diffusion parameter for HCHO - C M bulk concentration of copper ions - bulk concentration of dissolved oxygen - C R a surface concentration of HCHO - C R bulk concentration of HCHO - D R diffusion coefficient of HCHO - E electrode potential - E M thermodynamic reversible potential for the metal deposition reaction - E M 0 standard electrode potential for copper deposition - E MP mixed potential - E R thermodynamic reversible potential for reducing agent reaction - E R 0 standard electrode potential for HCHO - F Faraday constant - i M current density for metal deposition - i M total cathodic current density - i M k kinetic controlled current density for metal deposition - i M 0 exchange current density for metal deposition - i M D diffusion-limited current density for metal deposition - i M D diffusion-limited current density for total cathodic reactions - current density for oxygen reduction - i plat plating current density - i R current density for HCHO oxidation - i R 0 exchange current density for HCHO oxidation - i R D diffusion-limited current density for HCHO oxidation - n M number of electrons transferred in metal deposition reaction - n R number of electrons transferred in the HCHO oxidation reaction - R gas constant - T absolute temperature - stoichiometric number - M transfer coefficient for metal deposition - R transfer coefficient for HCHO oxidation - M symmetry factor - number of steps prior to rate determining step - M overpotential for metal deposition - R overpotential for HCHO oxidation - v kinematic viscosity - rotation rate of electrode  相似文献   

2.
Summary The weight-average MW and number-average Mn molecular weights of gum arabic are identical after a proteolysis treatment with pronase. The value (1.8×105) is closed from Mn early reported in the literature whereas MW before treatment are dispersed for a large lot of samples up to more 106. This can be interpreted by the wattle blossom model for which some homogeneous chains of molecular weight c.a. 2.105 are still linked to a protein core, the crude gum being a mixture of this complex and free chains.  相似文献   

3.
Summary Quasiliving cationic polymerization of styrene was obtained in the system 2-phenyl-2-propano-/AlCl3·OBu2/Bu2O in a mixture of 1,2-dichloroethane and n-hexane (55:45 viv) at -15 °C. The molecular weights of the polymers (Mn) increased in direct proportion to the monomer conversion. However, the experimental Mns are essentially higher than theoretical ones, indicating that slow initiation relative to propagation takes place. The molecular weight distributions were broad (Mw/Mn2.5), probably due to the slow initiation and slow exchange between reversibly terminated and propagating species.  相似文献   

4.
Summary Living polymerizations of 2-vinyloxyethyl methacrylate and 2-vinyloxyethyl cinnamate were successfully performed with a mixture of hydrogen iodide and iodine (HI/I2) as an initiating system in toluene at –15 to –40 °C, Although the two monomers have an unsaturated ester pendant group, their living polymerizations proceeded exclusively via the vinyloxyl group without undesirable side reactions of the pendant group. The product polymers had a very narrow molecular weight distribution (Mw/Mn 1.1), and {Mn} directly proportional to monomer conversion. For both vinyloxyethyl monomers, the addition of a new feed of monomer to a polymerization mixture led to an increase in polymer molecular weight which was again proportional to the conversion.  相似文献   

5.
Acid polymers, –[N=P(OC6H5)2–x (OC6H4SO3H) x ] n (II), having an entirely inorganic chain of fifth-group elements, with acid equivalent values between 2.90 and 5.19 mEq/g and molecular weights (M w) of 105–106 (205n3582), have been obtained from –[N=P(OC6H5)2]– n (I), 4325n 20,300, in very strong acid medium (SO3/–P=N –=1.15–3.10 mol/mol). Sulfonation of the pendant substituents occurs first in the meta position and successively at the para carbons, presumably due to reduced conformational mobility as the degree of substitution (x) in II increases.  相似文献   

6.
Summary The intrinsic viscosity [] and the molecular weight MW of 8 pure, unbranched polyacrylamide (PAAm) samples — prepared in our laboratory — were determined and lead to the following formula [] = 0,0194 · MW 0,70(cm3/g) Solvent: aqeous 0.1 M Na2SO4 solution; T=298K; MW/Mn-2,5 . In addition a comparison to the []-M-relationships of PAAm in other solvents is given.  相似文献   

7.
Summary Polymerization of isoprene was investigated by using a novel ternary catalyst system composed of neodymium(III) isopropoxide (Nd(OiPr)3), dimethylphenylammonium tetrakis(pentafluorophenyl)borate ([HNMe2Ph]+[B(C6F5)4]-; borate), and triisobutylaluminum (i-Bu3Al). The mole ratios of borate and aluminum compounds to Nd catalyst significantly affected the polymerization behavior. Both yield and cis-1,4 content of polyisoprene decreased in the case of [borate]/[Nd] < 1.0, while at [borate]/[Nd] > 1.0 the formation of multiple active species resulted in the polymer showing bimodal peaks in GPC. When the [Al]/[Nd] ratio was lower than 30, the polymer yield sharply decreased, whereas the cis-1,4 content became relatively low with use of a large excess of Al ([Al]/[Nd] > 50). Thus, the optimal catalyst composition was [Nd]/[borate]/[Al] = 1/1/30, which gave in > 97% yield polyisoprene with high molecular weight (Mn2×105) and relatively narrow molecular weight distribution (Mw/Mn2.0) and mainly cis-1,4 structure (90%).  相似文献   

8.
The non-Newtonian behavior of commercial linear polyethylene samples and their fractions were studied at 190°C. The viscosity η versus shear rate \documentclass{article}\pagestyle{empty}\begin{document}$ \dot \gamma $\end{document} curves of whole polymers could be superimposed onto a single master curve despite the variations of their molecular weights and molecular weight distributions. For fractions, however, the same master curve was inapplicable, and the sensitivity of the viscosity to shear rate was found to be greater than those of the whole polymers. The zero-shear viscosities η0 of fractions were related to the 3.42 power of the weight-average molecular weight Mu as follows: For whole polymers, the zero-shear viscosities were found to be considerably higher at the same Mw and markedly lower at the same z-average molecular weight Mz than those of the fractions. Thus, it was concluded that η0 corresponds to an average of molecular weight between Mw and Mz. It was found that the molecular relaxation time τ is proportional to Mz5.3 for whole polymers and to η0Mw for fractions. Using these relations it was possible to relate the flow ratio, the ratio of flow rates at two different shear stresses, with the molecular weight distribution.  相似文献   

9.
The kinetics and molecular weight averages of the hyperbranched polymers formed by the alternating copolymerization of equimolar allyl methyl maleate (AMM) and N‐n‐propyl maleimide (PMI) were investigated. The yields, molecular weight averages, and polydispersity indices as well as the branching degrees of the produced copolymers increased with increasing initiator concentrations and prolonged polymerization time. The trends of the experimental molecular weights as determined by size exclusion chromatography were in good agreement with the theoretical predictions. The molecular weight distribution indices fit the curve given by Mw/Mn = 1/(1‐xD), and the molecular weights fit the curve given by Mw = 4076/(1‐xD)2, where xD was the conversion of vinyl groups. DSC studies demonstrated a nonlinear relation of Tg values to the reciprocal of molecular weight (M), and Tg values decreased with the increase of molecular weight. For the Tg values of highly branched polymers in high molecular weight range, a relation of Tg = T + k/M was obtained, where T was obtained by extrapolating to infinite molecular weight and k was a constant. T was 136°C, and k = 2.9 for this work. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 97: 1941–1947, 2005  相似文献   

10.
Chemical defense in larvae of the plant bugHotea gambiae has been investigated. Results of analyses (GC, GC-MS) on the secretions from the three dorsally situated larval abdominal defense (scent) glands are reported. The secretion from the first abdominal gland consists of a mixture of C10 and C15 isoprenoids: (C10) -pinene, -pinene, limonene, -phellandrene; (C15) -caryophyllene, caryophyllene oxide, -humulene, and (the major component) humulene epoxide II. The secretions from the second and third abdominal glands are similar mixtures consisting of (E)-2-decenal, (E)-4-oxohex-2-enal, andn-tridecane together with lesser amounts of (E)-2-hexenal,n-dodecane, and other materials. Isoprenoid defense is now known from four species of plant bugs (Heteroptera) associated with Malvaceae.  相似文献   

11.
Propane ammoxidation to acrylonitrile over rutile-type vanadia catalysts is discussed regarding phase cooperation and site isolation effects. Compared with the pure phases, biphasic catalysts with both SbVO4and -Sb2O4are considerably more selective to the formation of acrylonitrile. It is demonstrated that cooperation between the phases during the calcination of the catalyst and the use in propane ammoxidation results in spreading of antimony species from free antimony oxide to the surface of SbVO4, forming Sb5+–V3+/V4+supra-surface sites being involved in the formation of acrylonitrile. Dilution and isolation of the vanadium centers in SbVO4through the partial replacement with, e.g., Al, Ti and W improves the catalytic properties. Structure–reactivity correlations using data for a nominal Sb0.9V0.9-x Ti x O y series indicate that the activation of propane occurs on a V3+site and the activation of ammonia requires an Sb5+site.  相似文献   

12.
The synthess of two new methacrylate esters containing 2-(4-benzyl piperazin-1-yl)-2-oksoethyl and 2-(4-methylpiperidin-1-yl)-2-oksoethyl group are described. The monomers produced from the reaction of corresponding 4-methylpiperidinechloroacetamide and 1-benzylpiperazinechloroacetamide with sodium methacrylate were polymerized in DMSO solution at 65°C using AIBN as an initiator. The monomers and their polymers were characterized by IR, 1H- and 13C-NMR spectroscopy. The glass transition temperature of the polymers were investigated by DSC and the apparent thermal decomposition activation energies (E d ) were calculated by Ozawa method using the SETARAM Labsys TGA thermobalance. By using gel permeation chromatography, weight average ( ) and number average ( ) molecular weights and polidispersity, indices of the polymers were determined.  相似文献   

13.
Summary Dilatometrically determined initial rate of alternating copolymerization of anethole (M1) and maleic anhydride (M2) in methyl ethyl ketone (MEK) is reported for the total monomer concentrations of 1M, 2M and 3M polymerized at 50°C with AIBN. The rate is larger than the first order with respect to the monomer concentration. The simplified complex model adopted by Shirota et al is found to be consistent with the data. The equilibrium constant of donor-acceptor complexation (C) of the comonomers is measured to be K=0.057 M–1 in MEK at 26°C. The ratios of the propagation rate constants are calculated to be k1c/k12=D=2 and k2c/k21=A=6.  相似文献   

14.
Preparatory to a detailed presentation of a large body of accumulated information, representative data have been selected to demonstrate that well-defined narrow molecular weight distribution (MWD) (¯Mw/¯Mn = 1.1) polyisobutylene (PIB) can be readily prepared in the Mn = 1000 – 100,000 range in the presence of certain types of electron pair donors (EDs) under a great variety of conditions. Specifically, many initiating systems, such as dicumyl chloride/BCl3, dicumyl alcohol/BCl3, 2-chloro-2,4,4-trimethylpentane/TiCl4, that induce nonliving polymerizations and/or give relatively broad MWD PIBs, in the presence of suitable EDs, e.g., dimethyl sulfoxide (DMSO), dimethyl acetamide (DMA), give rise to living polymerizations and yield narrow MWD products. Evidently, by the use of select EDs various undesirable side reactions, i.e., uncontrolled initiation, chain transfer, irreversible termination, indanyl end-group formation, etc., that plague carbocationic polymerizations and which therefore yield ill-defined relatively broad MWD products, can be eliminated and well-defined narrow MWD products can be obtained. The addition of EDs to otherwise extremely rapid carbocationic polymerizations results in lower controlled rates. All these observations and beneficial effects can be explained by controlled carbocation stabilization by EDs, a subject that will be explored and discussed in detail in this series of publications.Paper XXV in the series living carbocationic polymerization  相似文献   

15.
The effect of rare earths (Sm, Pr, Ce, Nd, and La) on the hydrogenation properties of chloronitrobenzene (CNB) over Pt/ZrO2 catalyst was studied in ethanol at 303K and normal pressure. The results show that the hydrogenation of CNB can be carried out over Pt/ZrO2 catalyst. The order of the hydrogenation rates of CNB is p>m>o, and the yield of chloroaniline (CAN) is p>o>m. The specific rate constant turnover frequency (TOF) expressed per surface Pt atom increases when the platinum catalyst is modified by rare earth. The conversion of CNB is >99% and CAN is the main product in the hydrogenation of CNB over PtM/ZrO2 catalysts. The PtPr/ZrO2 catalyst shows the best selectivity of CNB to CAN: 89.4mol% for o-CAN, 94.6mol% for m-CAN and 95.1mol% for p-CAN.  相似文献   

16.
Summary Living carbocationic polymerization (LCPzn) of isobutylene (IB) has been achieved by the 2-chloro-2, 4, 4-trimethylpentate (TMPCl)/TiCl4 initiating systems in the presence of KCl in conjunction with the 18-crown-6 ether in CH2Cl2/hexanes solvent mixture at –80°C. The rate of initiation is relatively slow and the molecular weight distribution (MWD) of the polyisobutylene (PIB) becomes narrower (Mw/Mn decreases from 1.8 to 1.2) in the course of incremental monomer addition (IMA). In the presence of the crown ether, and depending on its concentration, the charges become highly viscous rendering stirring difficult and preventing the synthesis of Mn's in excess of 15, 000 g/mole.  相似文献   

17.
Sodium polyphosphate and sodium–copper and sodium–nickel copolyphosphate glasses (with the ratio Na : M = 9 : 1 and 8 : 2, where M is Cu or Ni) are studied. The glass transition (T g) and melting (T m) temperatures are determined by differential scanning calorimetry (DSC). It is revealed that the T gand T mtemperatures depend on the molecular weight M t(determined from terminal groups) of polymeric glasses, the Na : M ratio, and the glass synthesis conditions. The activation energies are calculated, and the thermodynamic parameters H, S, and C pare measured.  相似文献   

18.
We synthesized a novel chiral cholesteryl-based N-propargylamide (Mch, HCCCH2NHCOCH2CH2COOch, ch = cholesteryl) from which homopolymers [P(Mch)] with different molecular weights (number-average molecular weight: 8600, 14100 and 30000) were prepared. The polymers formed helical structures with a preferential helicity. The three polymers increased in both helix content and specific rotation as the molecular weight increased. P(Mch)-8600 was studied in detail as the model polymer. P(Mch)-8600 adopted helical conformations in toluene, THF, CHCl3 and CH2Cl2, exhibited thermal stability with a decomposition temperature of 273 °C and formed a lyotropic liquid crystal under the studied conditions. Copolymers of different compositions of Mch and an achiral monomer (Met) were prepared. The copolymers formed helices to different degrees depending on the specific composition, indicating an effective approach for controlling the formation of helices in synthetic helical polymers.  相似文献   

19.
Three polysilane polymers, (n-PrSiMe) n , (i-PrSiMe) n , and (sec-BuSiMe) n , were synthesized and characterized by DSC. UV spectroscopy, wide-angle X-ray diffraction, and optical microscopy, all at variable temperatures. The known thermochromic transition of (n-PrSiMe) n at 48 C is associated with a change from an orthorhombic to an isotropic phase. (i-PrSiMe) n was examned as an insoluble and soluble (lowM w) fraction, both existing mainly in an orthohombic lattice at room temperature. (sec-BuSiMe) n has a mesophase structure at 25 C, undergoes a weak endothermic transition to a second (nematic) mesophase near 65 C, and becomes isotropic at 160 C.Dedicated to the memory of Professor Zygmunt Lasocki, a fine chemist and a kind and gentle person.  相似文献   

20.
Summary Various 4-methoxy-4-carbomethoxy--amino--cyanostilbenes were prepared as mixtures of E and Z isomers, and polymerized via condensation polymerization using dibutyltin diacetate as catalyst. The resulting low molecular weight homopolymers showed higher glass transition temperatures (168–183°C) than previously reported main chain nonlinear optical (NLO) homopolymers. A high molecular weight (Mn30, 200) copolymer possessed an even higher glass transition temperature of 187°C. The hyperpolarizability of the polymers and a model compound were found by EFISH measurements to be in the range of 61 to 79x10-48 esu.  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