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.
Transetherification of methyl ethers of melamine‐formaldehide resins (MER) with monophenyl ethers of ethylene glycol or propylene glycol (ROH) and competing reaction of self‐condensation are studied depending on MER composition (amounts of CH3O? , ? CH2OH, and NH2? groups), ROH type, MER/ROH molar ratio, presence or absence of acid catalysts, and temperature. High rates of self‐condensation processes prevent a complete conversion of CH3O? into RO‐groups. It turned out MER free of methylol groups were not able to be transetherified with high yields due to a premature gelation taking place prior to attaining 50% conversion of methoxy groups (~4 mol/kg) even at low MER/ROH ratios. In contrast, transetherification of MER with methylol groups content up to 3 mol/kg affords the incorporation of RO‐groups into the resin up to 8 mol/kg owing to direct etherification of ? CH2OH groups. The following factors are responsible for the growth of etherified product yield: presence of methylol groups in MER in some amounts without deterioration of MER–ROH compatibilization; CH3O? /ROH molar ratio no higher than 1; primary alcohols (ROH) is more preferable compared to secondary ones; thermal activation of the process is more efficient in comparison with acidic catalysis. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 101: 2977–2985, 2006  相似文献   

2.
The kinetics of dehydrogenation reaction and the structural evolution in polyacrylonitrile precursor fibers during thermal stabilization in air have been studied by Fourier transform infrared spectroscopy. The results indicate that, with the progress of dehydrogenation, the absorbance of methylene groups (? CH2? ) gradually decreases, whereas that of methine groups (?CH? ) gradually increases. The dehydrogenation reaction in the fibers is basically completed after 20‐min stabilization above 255°C. According to the Beer–Lambert law, the values of the absorbance for both ? CH2? groups and the resulting ?CH? groups have been calculated and converted into the concentration fractions of ? CH2? groups via the Lorentzian multipeak fitting. According to the principles of chemical kinetics, the dehydrogenation reaction has been determined as a pseudo‐second‐order reaction with an activation energy of 107.6 kJ mol?1. © 2012 Wiley Periodicals, Inc. J. Appl. Polym. Sci., 2013  相似文献   

3.
Melamine–formaldehyde–polyvinylpyrrolidone (MFP) polymer resin was prepared with 1 : 16 : 1 ratios of melamine, formaldehyde (CH2O), and polyvinylpyrrolidone amounts, respectively, by condensation polymerization at 6.9 pH. Structures were determined with IR, 1H‐NMR, and 13C‐NMR spectroscopies. Chemical shifts (δ, ppm) were analyzed with singlet at δ 4.5, duplet from 3.13 to 3.17 and a quartet at 1.5 to 2.2 ppm for methylene (? CH2? ) bridging group, pyrrolidone, and polyvinyl constituents. The 3389.25, 1290.38, and 1655.28 cm?1 stretching frequencies of ? N?, ? CH? and ? C? O? O? groups, respectively, were noted on FTIR spectrum. The ? C?N? melamine units reacted with CH2O to adjoin with polyvinylpyrrolidone (PVP). An average viscosity molecular weight ( v) 57,000 g mol?1 was obtained with Mark–Houwink–Sakurada equation. The chemical shift of ? N(CH2O)2? C? pyrrolidone ring on 13C‐NMR spectra was shifted toward lower magnetic field at 175.18 ppm. The resin was partially miscible with water thereby densities and viscosities of aqueous solutions were measured at 298.15 K temperature. It showed higher densities and viscosities than those of water. The resin developed exceptionally higher adhesive strengthen when its 62.29‐μm uniform thin film was applied on surfaces of wooden strips. The resin showed micellar behavior at about 0.009 g/100 mL aqueous solution. © 2009 Wiley Periodicals, Inc. J Appl Polym Sci, 2009  相似文献   

4.
The structure of ZnO-catalyzed phenol–urea–formaldehyde (PUF) resin at different synthesis stages was analyzed by liquid 13C nuclear magnetic resonance spectroscopy. The results showed that the general structure of ZnO-catalyzed PUF resin was almost the same as the control PUF resin. Addition reaction between phenol and formaldehyde mainly occurred at the first stage. Total methylol groups amount between phenols of the control resin was a little lower than that of the ZnO-catalyzed PUF resin. Co-condensation and self-condensation reaction occurred at the second stage. The preparation method of ZnO-catalyzed PUF resin favored the co-condensation reaction between phenol methylol groups and urea units, while self-condensation reaction dominated the control resin at the second stage. Formaldehyde completely reacted for both the control and ZnO-catalyzed PUF final resin. The total amount of methylol and methylene groups between urea units and phenols, respectively, was almost the same for the two final resins. The total quantity of methylol groups between phenols represented a continuing downward trend from the first stage to the final stage, and the amount of methylol group (p-Ph–CH2OH) of ZnO-catalyzed PUF resin was 30% more than that of the control resin. Total co-condensed methylene groups amount of ZnO-catalyzed PUF resin was greater than that of the control resin, which indicated that ZnO could make the urea units well incorporated into the co-condensed PUF resin.  相似文献   

5.
Oxotris(dimethyl dithiocarbamato) vanadium(V) [VO(S2CN(CH3)2)3] sensitizes the polymerization of styrene when irradiated by light of λ = 365 nm at 25°C. Under the experimental conditions employed, no retardation occurs, and the rate of initiation is independent of monomer concentration. The mean values of the quantum yield of iniiation (?i) and polymerization (?o) are 2.85 × 10?3 and 6.72 respectively. Spectroscopic analysis shows that initiation occurs predominatly through scission of the N,N-dimethyl dithiocarbamate ligand (—SC(S)N(CH3)2) with reduction of vanadium(V) to (IV), and VO (S2CN(CH3)2)2 is the final photolytic product. A reaction mechanism is proposed based on an intramolecular photoredox reaction which leads to the primary formation of SC(S)N(CH3)2 radicals and a vanadium(IV) chelate complex. The rellevant kinetic parameters are evaluated. The polystyrene produced shows a photoactivity when irradiated with UV-light.  相似文献   

6.
This study focuses on methylol functional benzoxazines as precursors to build a network structure utilizing both benzoxazine and resole chemistry. The first part is a review of systems that contain methylol groups which play a role on their crosslinking formation. The polymerization mechanism and properties of resoles will be highlighted as the most abundant polymers that are characterized by polymerization through condensation reaction of methylol group. In the second part, the effect of incorporating methylol group into benzoxazine monomers is studied. Differential scanning calorimetry (DSC) is used to study the effect of methylol group on the rate of polymerization. Kissinger and Ozawa methods using non-isothermal DSC at different heating rates show that methylol monomer exhibits lower average activation energy compared to the un-functionalized monomer. The effect of adding catalysts into the monomers is also studied. p-Toluene sulfonic acid (PTSA) is found to be more efficient than 1-methyl-imidazole (IMD) and lithium iodide (LiI) in the case of methylol monomer due to its ability of accelerating both the methylol condensation and ring-opening polymerization. Additionally, thermal behavior of the monomers is studied using thermogravimetric analysis (TGA).  相似文献   

7.
Chitosan (Chi; 0.5 g) in 69.66 mM aqueous acetic acid was mixed with 312.4 mM methionine (methi) at 0.01 mL/s to disperse and cause optimum collisions for supporting condensation reactions through ? NH2 of Chi and ? COOH groups of methi. The functionalized chitosan (f‐Chi) product with methi developed an amide bond, which was represented as methi‐functionalized chitosan [Chi–NH? C(?O)–methi]. Both the 1‐Ethyl‐3‐(3‐dimethylaminopropyl)carbodiimide (EDC) and Dean–Stark methods were followed for Chi functionalization. Sulfonation with chlorosulfonic acid in a dimethylformamide medium was conducted at 90 °C and 750 rpm with an approximately 72% yield. The Chi–NH? C(?O)–methi was characterized by 1H‐NMR spectroscopy and Fourier transform infrared stretching frequencies. The onset temperature of 280 °C recorded by thermogravimetric analysis/differential scanning calorimetry analysis, confirmed the high stability of the covalent bonds in Chi–NH? C(?O)–methi. The synthesis was repeated with other series members of sulfur (S) atoms containing α‐amino acids: homocysteine, ethionine, and propionine. The shielding of terminal ? CH3 was enhanced on elongation of the terminal alkyl chain in the case of propionine. The peak for the ? NH2 of Chi at a δ value of 4.73 ppm shifted to 5.36 ppm in Chi–NH? C(?O)–methi because of the involvement of ? NH2 in ? NH? C(?O)? . Theoretically, the value of ? NH2 of Chi was 5.11 ppm, with a difference of 0.38 ppm as compared to the experimentally determined value of 4.73 ppm. Additionally, a new peak at a δ value of 3.26 ppm also confirmed Chi functionalization. © 2017 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2018 , 135, 46000.  相似文献   

8.
Reactions of N‐(2,4‐dinitrophenyl)‐4‐arylpyridinium chlorides (aryl (Ar) = phenyl and 4‐biphenyl) with piperazine or homopiperazine caused opening of the pyridinium ring and yielded polymers that consisted of 5‐piperazinium‐3‐arylpenta‐2,4‐dienylideneammonium chloride (? N(CH2CH2)2N+ (Cl?)?CH? CH?C(Ar)? CH?CH? ) or 5‐homopiperazinium‐3‐arylpenta‐2,4‐dienylideneammonium chloride (? N(CH2CH2CH2)(CH2CH2)N+ (Cl?)?CH? CH?C(Ar)? CH?CH? ) units. 1H NMR spectral analysis suggested that the π‐electrons of the penta‐2,4‐dienylideneammonium group of the polymers were delocalized. UV‐visible spectral measurements revealed that the π‐conjugation system expanded along the polymer chains because of the orbital interaction between electrons of the two nitrogen atoms of the piperazinium and homopiperazinium rings. However, the π‐conjugation length depended on the distance between the two nitrogen atoms; that is, the polymers containing the piperazinium ring had a longer π‐conjugation length than those containing the homopiperazinium ring. Conversion of the piperazinium and homopiperazinium rings from the boat to the chair form led to a decrease in the π‐conjugation length. The surface of pellets that were molded from the polymers exhibited metallic luster, and these polymers underwent electrochemical oxidation in solution. Copyright © 2010 Society of Chemical Industry  相似文献   

9.
A preceramic polymer for Si? B? N? C fiber, polyborosilazane, has been synthesized by one‐step condensation reaction of dichloromethylsilane, BCl3, and hexamethyldisilazane with high yield. The reaction mainly involves the condensation of Si? Cl and B? Cl with N? SiMe3 followed by SiMe3Cl evaporation and dehydrogenation between N? H and Si? H. The resulted polymer is a soluble colorless transparent solid with melting point of 70°C and molecular weight of 10,800. The backbone of the polymer is mainly composed of ? Si? N? B‐bridge with some borazine rings. The polymer exhibits good processability and flexible polymer fibers with diameter of 15–20 μm were obtained by melt spinning. Pyrolysis of the as‐synthesized polymer to 1000°C under nitrogen atmosphere results in a ceramic yield of 63 wt %, and the obtained Si? B? N? C ceramic remains fully amorphous up to 1700°C, and only small amount of poorly crystallized BN, Si3N4, and SiC phases were observed upon heating at 1850°C. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

10.
A proton magnetic nuclear resonance study was performed on co-condensation reactions of resorcinol, 5-methylresorcinol and 2,5-dimethylresorcinol with methylol compounds, including ortho- and para-methylolphenol, N-methylol-caprolactam, methylol-N,N-diethylurea, methylolurea and N,N′-dimethylolurea. Spectral assignments, reaction kinetics and composition of products are discussed. The reaction in melt (120°C) with methylolphenols occurs as co-condensation in the presence of all catalysts studied. In resorcinols C4C6 substitution is favored. The rate constants of methylol disappearance clearly indicate the preferable influence of acid and alkaline catalyst (not zinc acetate) on para-methylol. The reaction with N-containing methylol compound does not give any co-condensate in the presence of alkaline catalyst. The optimum conditions for co-condensation mainly depend on reactivity of co-reagents with formaldehyde and stability of methylol compound. The quantitative amount of co-condensate with methylolcaprolactam can be obtained in melt-condensation (70°C) in the presence of acid catalyst. Because of the higher reactivity of urea, the reaction in melt (100°C) in the presence of acid catalyst does not lead to quantitative co-condensation. The condensation of methylol compound or resorcinols with formaldehyde can be avoided substantially by performing the reaction in aqueous solution at lower temperature.  相似文献   

11.
Linear asymmetrical poly(propylene oxide) was synthesized through four‐step reactions: selective benzylation, alcohol exchange reaction, propylene oxide anionic polymerization, debenzylation. One terminal of the asymmetrical polymer chains is alcohol hydroxyl and the other is phenol hydroxyl. It was characterized with infrared (IR) and 1H Nuclear Magnetic Resonance (1H‐NMR). Peaks at 1.11, 3.38, and 3.53 ppm were attributed to side groups (? OCH2CH(CH3)? ), backbone units (? OCH2CH(CH3)? ) and (? OCH2CH(CH3)? ) of poly(propylene oxide), respectively. Molecular weight and molecular weight distribution were measured with 1H‐NMR and laser light scattering (LLS), which showed that the linear asymmetrical poly(propylene oxide) was mono‐disperse (PDI = 1.02–1.07). Then, its carbamate reaction with phenyl isocyanate was studied; the reaction rate constants for phenol hydroxyl and alcohol hydroxyl of poly(propylene oxide) were k1 = 0.209 mol L?1 min?1 and k2 = 0.051 mol L?1 min?1. There was a great reactivity difference for two types of hydroxyls in asymmetrical poly(propylene oxide), contrasting to the single carbamate reaction rate constant of symmetrical poly(propylene oxide) (k3 = 0.049 mol L?1 min?1). © 2010 Wiley Periodicals, Inc. J Appl Polym Sci, 2010  相似文献   

12.
Three alkyds of high hydroxyl numbers with oil lengths of 21, 30, and 41 were prepared from crude palm oil. The excess hydroxyl groups provided good compatibility with melamine resins, and also served as sites for crosslinking reactions. They were made into baking enamels by blending with 20% of a commercial melamine resin. All of them could be cured at temperatures below 200°C although thermodecomposition might occur above 290°C. The dry‐hard time of these enamels cured at temperatures between 140–180°C ranged from 10 to 180 min. Fourier transform infrared spectroscopy could be used to follow the major curing reactions. The absorbance of ? O? H and ? N? CH2? OCH3 groups showed significant reduction. The changes in the absorbance of these peaks with time and temperature were investigated. The predominant reaction was identified as the formation of methylene ether linkages. However, the self‐condensation reactions of the amino resin and ester linkages did not occur to any noticeable extent. © 2001 John Wiley & Sons, Inc. J Appl Polym Sci 80: 2309–2315, 2001  相似文献   

13.
Thermoplastic resins have been prepared by the reaction of formaldehyde and aroyl propionic acids derived from the succinoylation of aromatic hydrocarbons. A study has been made of the condensation reaction, and nuclear magnetic resonance (n.m.r.) spectroscopy has been found to be extremely useful in the elucidation of the reaction mechanism. Two types of initial reaction with formaldehyde have been shown to occur: firstly, attack at the aromatic nucleus to give a substituted benzyl alcohol (which may condense further); and, secondly, reaction at the CH2 groups α to C = O groups leading to the formation of a γ-lactone. The relative extents of the two reactions are determined by the activity of the aromatic nucleus to formaldehyde attack. In the case of resins from 2,4,6-trimethylbenzoylpropionic acid, γ-lactone formation is virtually absent and attempts have been made to consider the further condensation reactions of the substituted benzyl alcohol formed initially.  相似文献   

14.
Lot-to-lot variations in a phenol/cresol-formaldehyde resin were fully characterized. Resin property-product hardness relationships were examined to elucidate the cause of variable hardness in product insulators and to establish acceptance criteria for the resin. A high methylol content (?CH2OH), determined by quantitative carbon-13 nuclear magnetic resonance (13C-NMR), and a narrow molecular weight distribution (MWD), determined by size exclusion chromatography, were found necessary for resin to produce insulator meeting the minimum Shore D durometer specification of 40.  相似文献   

15.
The reaction rate of crosslinking of PVC with dithioltriazine has been studied by following gel formation and changes in the molecular weight distribution (MWD). Compounding was performed on a roll mill at 145°C and crosslinking by heat treatment at 180 or 90°C. In this system crosslinking is executed by the thiolate anion, formed in situ by reaction with MgO. We have studied the catalyzing effect of several polyols in order to achieve a more efficient reaction. Most likely, these catalysts work by chelating the Mg2+ ions, thus increasing the nucleophilic character of the thiolate. With the most efficient ones, ditrimethylolpropane and HO(CH2CH2)6-7H, complete crosslinking can be obtained in 3 min at 180°C, i.e., at processing temperatures. We also followed the changes in the MWD before gelation at a considerably lower temperature, 145°C, and found an extensive molecular enlargement even after 5-10 min. Most surprisingly, μMn increased up to 100% without formation of insoluble material. By 1H-NMR measurements on low molecular weight extracts, we have shown this to be due to a fast and selective reaction with allylic chlorine in the unsaturated end groups, ~ CH2? CH?CH? CH2Cl, formed in the mechanism of chain transfer to monomer. Due to this reaction, formulations with too high reactivity may crosslink during processing, which calls for a careful balancing of the reactivity for each processing case.  相似文献   

16.
The structure of the cyclic urea influenced the rate of reaction with cotton cellulose and the mechanism by which reaction occurred. Reaction of N, N′-dimethylolethylene-urea (DMEU) and N, N′-dimethylolpropyleneurea (DMPU) with cellulose in presence of inorganic salt catalysts proceeded through methylol hydroxyls and at the same rate; but reaction mechanism differed. With DMEU, N → metal ion coordination occurred and SN2 mechanism prevailed. With DMPU, O → metal ion coordination resulted. Reaction of dihydroxyethylene urea (DHEU), N, N′-dimethyldihydroxyethyleneurea (DMeDHEU), and N, N′-dimethyloldihydroxyethyleneurea (DMDHEU) with cotton cellulose proceeded through ring hydroxyls with the formation of a carbonium ion, indicating an SN1 mechanism. The much faster rate of reaction with DMeDHEU than with DHEU was attributed to the more electronegative environment of its ring hydroxyl, while the much slower rate of reaction of DMDHEU was attributed to hydrogen bonding between its methylol and ring hydroxyls.  相似文献   

17.
Homogeneous periodate oxidation of cellulose was achieved through methylol cellulose. The dissolution of methylol cellulose into aqueous periodate solution was followed by the gradual decomposition of methylol groups at random sites along the methylol cellulose chain. The recovery of glycol hydroxyl groups at the C2 and C3 positions on the glucopyranose ring during the above decomposition process caused uniform cleavage of C2? C3 bonds by the periodate ion. The oxidation level reached nearly 100% in 10 h. The reduced product of the resulting dialdehyde cellulose, i.e., dialcohol cellulose, resulted in mechanical properties quite different from those of conventional dialcohol cellulose. Examination of the thermal deformation and tensile properties revealed that no notable cellulose degradation occurred during the reaction. Our dialcohol cellulose gave a clear and transparent film with a flexible nature.  相似文献   

18.
Zinc substituted zeolite Y is a potential catalyst in the CH2Cl2 formation in a one step gas phase reaction of CH2O with CHl. Methanol and dimetoxymethane are suggested as possible intermediates in the formation of CH3Cl and CH2Cl2. The decomposition pathway of CH2O on zeolites seems to play an important role in the chlorination reaction mechanism. A certain combination of such factors as temperature, catalyst acidity and cation action must occur in order to activate CH2O molecules for the reaction with HCl. The chlorination reaction is accompanied by side reactions resulting in the formation of nonchlorinated compounds such as methyl formate, ethers, aldehyde.  相似文献   

19.
By polymerizing unsaturated oligomers of siloxane derivatives (I) and by copolymerizing them with the usual vinyl monomers a variety of soluble polymers can be obtained which possess different organic structural units and di- or polysiloxane substituents. Hereby the siloxane content can vary greatly and in case of some homopolymers may amount to as much as 50 yo, When using polysiloxanes with two unsaturated end groups instead of monofunctional siloxane compounds analogous products are obtained which are however cross-linked and insoluble. Linear organic polymers with Functional silicon-substituents (such as Cl? Si? , HO ? Si? , H ? Si? , CH2 ? CH? Si? , CH2?CH? CH2? Si? , CH?C? Si, n? C4H9O? Si? , ClCH2? Si? groups) are formed by homo-, and copolymerization of silanes and siloxanes which contain a functional moiety besides the unsaturated group. Macromolecules which carry the functional substituents on the ends of the chains instead of exhibiting them as side groups can be prepared by termination of living polymers with chlorosilanes or by initiation of anionic vinyl polymerization with organometallic starter molecules. High molecular weight polymers with two reactive end groups may also be obtained by cohydrolysis or equilibration reactions of the corresponding chlorosilanes or disiloxanes. By these synthetic routes a great number of organic or silicone polymer molecules with functional groups in side or end positions are available by suitable block, graft, and crosslinking reactions in polymer systems with definite structural units, and tailor made micro structures. To achieve this, polymer molecules of different structure are linked together by suitable reactions of their functional groups or one uses polymer molecules with reactive centers as poly-functional initiators for the polymerization reaction. By addition polymerization of organic diisocyanates with oligomer siloxane or silmethylene diols, silicon-modified polyurethanes are obtained. The reaction of the partners in molar ratio 2/1 leads to low molecular weight polyurethanes (degree of addition polymerization m 2) with two isocyanate end groups, which could be employed in the diisocaynate addition polymerization as easily available silicon-containing diisocyanates. Oligomer siloxanes or silmethylenes with two H? Si-end groups (I), (R and R1?H) and ω,ω′-divinyl- or ω,ω′-diallylpolysiloxanes (I), (R and R1 ? vinyl-, allyl-) or several other unsaturated molecules, such as p-divinylbenzene and acetylene undergo addition polymerization to give heat stable polysilcarbanes, in which oligomer siloxane or silmethylene groups are linked via ethylene or n-propylene bridges. If the siloxane or silmethylene groups are replaced by ferrocenyl-groups, polyadducts result which are stable up to 400°C for several hours in presence of oxygen. Polycarbonates and polyesters are the condensation polymers of siloxane-containing diphenols and phosgene or terephthalic acid chloride. Polyimides, which are soluble, fusible and stable above 300°C, are the condensation products of silicon-containing dianhydrides and diamines or of organosilicon derivatives of diamines and pyromellitimide, and several organosilicon compounds are also the condensation polymers of dichloro-Si-compounds and N,N′-bis-(3-dimethylhydro-xisilylpropyl)-pyromellitimide, which is a very easily available disilanol. A comparatively lowering of melting points and glass transition points of organic polymer molecules possessing polysiloxane derivatives (approximately 5 wt.-% onwards) is due to the covalent bonds of polysiloxanes. Moreover, one notices also great improvement in solubility as well as in their hydrophobic character. A great improvement in heat stability can be achieved by introducing more siloxane-content (approximately 15 wt.-% onwards).  相似文献   

20.
Imidovanadium complexes with cyclopentadienyl (Cp) ligands—(Cp)V(?NC6H4Me‐4)Cl2 (1), (Cp)V(?NtBu)Cl2 (2), and (tBuCp)V(?NtBu)Cl2 (3; tBuCp = tert‐butylcyclopentadienyl)—were synthesized through the reaction of imidovanadium trichloride with (trimethylsilyl)cyclopentadiene derivatives. The molecular structure of 3 was determined by X‐ray crystallography. The monocyclopentadienyl complex 1 exhibited moderate activity in combination with methylaluminoxane [MAO; 10.3 kg of polyethylene (mol of V)?1 h?1 atm?1], whereas similar complexes with bulky tBu groups, 2 and 3, were less active. (2‐Methyl‐8‐quinolinolato)imidovanadium complexes, V(?NR)(O ?N)Cl2 (R = C6H3iPr2‐2,6 (4) or n‐hexyl (5), O ?N = 2‐methyl‐8‐quinolinolato), were obtained from the reaction of imidovanadium trichloride with 2‐methyl‐8‐quinolinol. Upon activation with modified MAO, complex 4 showed moderate activities for the polymerization of ethylene at room temperature. The complex 5/MAO system also exhibited moderate activity at 0°C. The polyethylenes obtained by these complexes had considerably high melting points, which indicated the formation of linear polyethylene. Moreover, the 5/dried MAO system showed propylene polymerization activities and produced polymers with considerably high molecular weights and narrow molecular weight distributions. © 2005 Wiley Periodicals, Inc. J Appl Polym Sci 97: 1008–1015, 2005  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