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 46 毫秒
1.
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   

2.
Two environmentally friendly succinic acid monofluoroalkyl sulfonate surfactants were synthesized from maleic anhydride and polyethylene glycol mono (1H,1H,7H‐dodecafluoroheptyl) ether, i.e. H(CF2)6CH2OCH2CH2OCOCH(SO3Na)CH2COOH (FEOS‐1) and H(CF2)6CH2(OCH2CH2)3OCOCH(SO3Na)CH2COOH (FEOS‐3). The obtained surfactants were characterized by FT‐IR, 1H NMR, 13C NMR and 19F NMR in detail. The synthesized fluorinated surfactants have a high thermal stability on the basis of thermogravimetric analysis. Their surface properties were examined and the results show that FEOS‐1 and FEOS‐3 surfactants can reduce the surface tension of water to 25.55 mN m?1 at 10.25 mmol L?1 and 21.63 mN m?1 at 8.33 mmol L?1, respectively; meanwhile, the introduction of oxyethylene groups enhances the hydrophilicity and micellar forming ability and the longer oxyethylene chains the better surface properties. The Krafft points (Kp) of FEOS‐1 and FEOS‐3 were both below 0 °C, which was lower than perfluoro‐n‐heptanesulfonic acid sodium salt (n‐C7F15SO3Na, Kp = 56.5 °C) at a similar length of fluorocarbon chains. Comparison studies on two surfactants above and the conventional fluorocarbon surfactants, perfluorooctanoate of ammonium (PFOA) show that the surfactants have comparable properties to PFOA, thus offering an environmentally friendly synthesizing alternatives to PFOA.  相似文献   

3.
Block copolymers having a pendant trichlorogermyl group as a part of polyamide segment? (CO? R′? CO? NH? Ar? NH? )xCO? R′? CO? and polydimethylsiloxane of general formula [(? CO? R′? CO? HN? Ar? NH)x? CO? R′? CO? NH(CH2)3SiO(CH3)2 ((CH3)2SiO)ySi(CH3)2(CH2)3 NH? ]n (where R′ = CH2CH(GeCl3), CH(CH3)CH(GeCl3), CH(GeCl3)CH(CH3); Ar = C6H4, (? C6H3? CH3)2, (? C6H3? OCH3)2, 2,5‐(CH3)2? C6H2, C6H4? O? C6H4) were prepared by a polycondensation reaction and characterized using CHN and Ge analysis, Fourier transform infrared (FTIR) and 1H NMR spectroscopy, thermogravimetric analysis (TGA) and molecular weight determination. They have a lamellar structure with weight‐average molecular weight in the range 1.21 × 105–4.79 × 105 g mol?1. These copolymers display two glass transition temperatures and have an average decomposition temperature of 489 °C. TGA, FTIR and gas chromatography/mass spectrometry studies indicate that degradation of these block copolymers results in carbon monoxide, oligomeric siloxanes and polyamide fragments. They are thermally stable due to the hydrogen bonded interlinked chains of polyamide, while they absorb water due to the presence of Ge? Cl bonding. Copyright © 2010 Society of Chemical Industry  相似文献   

4.
Branched polyethylene (PE) was prepared with a novel (α‐diimine)nickel(II) complex of 2,3‐bis(2,6‐dimethylphenyl)‐butanediimine nickel dichloride {[2,6‐(CH3)2C6H3? N?C(CH3)C(CH3)?N? 2,6‐(CH3)2C6H3]NiCl2} activated by methylaluminoxane in the presence of a single ethylene monomer. The influences of various polymerization conditions, including the temperature, Al/Ni molar ratio, Ni catalyst concentration, and time, on the catalytic activity, molecular weight, degree of branching, and branch length of PE were investigated. According to gel permeation chromatography, the weight‐average molecular weights of the polymers obtained ranged from 1.7 × 105 to 6.0 × 105, with narrow molecular weight distributions of 2.0–3.5. The degree of branching in the polymers rapidly increased with the polymerization temperature increasing; this led to highly crystalline to totally amorphous polymers, but it was independent of the Al/Ni molar ratio and catalyst concentration. At polymerization temperatures greater than 20°C, the resultant PE was confirmed by 13C‐NMR to contain significant amounts of not only methyl but also ethyl, propyl, butyl, amyl, and long branches (longer than six carbons). The formation of the branches could be illustrated by the chain walking mechanism, which controlled their specific spacing and conformational arrangements with one another. © 2002 Wiley Periodicals, Inc. J Appl Polym Sci 84: 1123–1132, 2002; DOI 10.1002/app.10398  相似文献   

5.
Effect of temperature (5°–65°C) on the separation of 11 inorganic anions by ion interaction chromatography (IIC) was studied employing RP C18 and C PhenylHexyl columns and aqueous mobile phase: 2.8 mM NaHCO3 + 0.7 mM TBAOH (tetra-n-butylammonium hydroxide). The apparent enthalpy changes, ΔH for hydrophobic ions like I?, SCN?, and ClO4 ? largely exceeded 3 kcal/mole suggesting that added to ion exchange they are retained by hydrophobic adsorption. Unlike conventional strongly basic anion exchangers, our system can be used at elevated temperatures with alkaline eluents without irreversible damaging the column.  相似文献   

6.
A new series of low-melting, low-viscosity, hydrophobic ionic liquids based on relatively small tertiary sulfonium cations ([R1R2R3S]+, wherein R1, R2, R3 = CH3 or C2H5, R3 = CH2CH2OCH3, CH2CH2COOCH3, CH2CH2CN) and bis(fluorosulfonyl)imide (FSI) anion have been prepared and characterized. The important physicochemical and electrochemical properties of these salts, such as melting point, glass transition, viscosity, density, ionic conductivity, thermal and electrochemical stability, have been determined. The influence of structure variation in the tertiary sulfonium cations on the above physicochemical properties is discussed. Among these new salts, some of them show the desirable properties including low-melting points, low viscosities, and high conductivities, to be selected as potential candidates as electrolytes in energy devices, and two salts are ionic plastic crystals.  相似文献   

7.
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   

8.
An asymmetric 3‐oxa‐pentamethylene bridged dinuclear titanocenium complex (CpTiCl2)25‐η5‐C9H6(CH2CH2OCH2CH2)C5H4) ( 1 ) has been prepared by treating two equivalents of CpTiCl3 with the corresponding dilithium salts of the ligand C9H7(CH2CH2OCH2 CH2)C5H5. The complex 1 was characterized by 1H‐, 13C‐NMR, and elemental analysis. Homogenous ethylene polymerization catalyzed using complex 1 has been conducted in the presence of methylaluminoxane (MAO). The influences ofreaction parameters, such as [MAO]/[Cat] molar ratio, catalyst concentration, ethylene pressure, temperature, and time have been studied in detail. The results show that the catalytic activity and the molecular weight (MW) of polyethylene produced by 1 /MAO decrease gradually with increasing the catalyst concentration or polymerization temperature. The most important feature of this catalytic system is the molecular weight distribution (MWD) of polyethylene reaching 12.4, which is higher than using common mononuclear metallocenes, as well as asymmetric dinuclear titanocene complexes like [(CpTiCl2)25‐η5‐C9H6(CH2)nC5H4)] (n = 3, MWD = 7.31; n = 4, MWD = 6.91). The melting point of polyethylene is higher than 135°C, indicating highly linear and highly crystalline polymers. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

9.
Oxidation of polyketones [poly(ethylene ketone) (? CH2? CH2? CO‐)n and poly(propylene ketone) (? CH2? CH(CH3)? CO‐)n] in the presence of antioxidant [2,2′‐methylene‐bis(4‐methyl‐6‐tert‐butylphenol)] was studied in the temperature range 120–190°C. The effectiveness of the antioxidant in polyketones is much lower than that in polyolefins. © 2003 Wiley Periodicals, Inc. J Appl Polym Sci 89: 1182–1185, 2003  相似文献   

10.
Interaction mechanisms between poly(methyl methacrylate) (PMMA) molecules and tricalcium silicate (C3S) occurring in a C3S-filled PMMA molecular structure before and after exposure to a simulated 25% geothermal brine solution at 240°C have been investigated by use of thermal analysis, infrared spectroscopy analysis, scanning electron microscopy, and compressive strength tests. For C3S-filled PMMA samples before exposure to hot brine, it was observed that an interaction between the C3S filler and the CH2 groups in the main chain of the PMMA occurred. After exposure to the hydrothermal environment, indications were that the brine solution induced ionic bonding between the carboxylate anion (? COO?) groups in the PMMA and the Ca2+ ions from the C3S. The reaction contributed to improvements in the thermal stability of the PMMA.  相似文献   

11.
Novel acid degradable polyacetal polyols and polyacetal polyurethanes able to controlled acid degradation were developed. Polyacetal polyols with various main‐chain structures were synthesized by polyaddition of various vinyl ethers with a hydroxyl group [4‐hydroxy butyl vinyl ether (CH2?CH? O? CH2CH2CH2CH2? OH), 2‐hydroxy ethyl vinyl ether (CH2?CH? O? CH2CH2? OH), diethylene glycol monovinyl ether (CH2?CH? O? CH2CH2OCH2CH2? OH), and cyclohexanedimethanol monovinyl ether (CH2?CH? O? CH2? C6H10? CH2? OH)] with p‐toluenesulfonic acid monohydrate (TSAM) as a catalyst in the presence of the corresponding diols [1,4‐butandiol (HO? CH2CH2CH2CH2? OH), ethylene glycol (HO? CH2CH2? OH), diethylene glycol (HO? CH2CH2OCH2CH2? OH), and 1,4‐cyclohexanedimethanol (HO? CH2? C6H10? CH2? OH)], respectively. Polyacetal polyurethanes were prepared by a two‐step polymerization, using the synthesized polyacetal polyols, 4,4′‐diphenylmethane diisocyanate (MDI), and 1,4‐butandiol (BD) as a chain extender. Depending on the main‐chain structures, these polyurethanes had different glass transition temperature (from ?44 to 19 °C) and properties such as hydrophobic or hydrophilic. Polyurethanes containing the hydrophilic main‐chain exhibited the thermoresponsiveness and had the certain volume phase transition temperature (VPTT). The polyacetal polyurethanes were flexible elastomers around room temperature (~25 °C) and thermally stable (Td ≥ 310 °C) and additionally exhibited smooth degradation with a treatment of aqueous acid in THF at room temperature to give the corresponding raw material diols. © 2016 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2016 , 133, 44088.  相似文献   

12.
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   

13.
The atom transfer radical polymerization (ATRP) of methyl methacrylate (MMA) is often carried out under homogeneous conditions, so the residual metal catalyst in the polymer often influences the quality of the polymer and causes environmental pollution in the long run. Novel CuBr/4,4′‐bis(RfCH2OCH2)‐2,2′‐bpy complexes (Rf = n‐C9F19, n‐C10F21, or n‐C11F23; 2,2′‐bpy = 2,2′‐bipyridine) are insoluble in toluene at room temperature yet readily dissolve in toluene at elevated temperatures to form homogeneous phases for use as catalysts in the ATRP reaction, and the Cu complexes precipitate again upon cooling. The CuBr/4,4′‐bis(n‐C9F19CH2OCH2)‐2,2′‐bpy system produced the best results (e.g., polydispersity index by gel permeation chromatography = 1.26–1.41), in that the residual Cu content in the polymer was as low as 19.3 ppm when the ATRP of MMA was carried out in the thermomorphic mode. © 2008 Wiley Periodicals, Inc. J Appl Polym Sci, 2008  相似文献   

14.
Reactions of N‐(2,4‐dinitrophenyl)pyridinium chloride (salt(Cl)) with H+MCl4?1 (M ≡ Fe and Bi) resulted in an anion exchange between Cl? and MCl4? to yield Zincke salts with metal chloride anions, namely salt(Fe) and salt(Bi), respectively. Reactions of the Zincke salts with piperazine resulted in ring‐opening of the pyridinium ring, yielding ionic polymers with 5‐piperazinium‐2,4‐dienylideneammonium metal chloride units, namely polymer(Fe) and polymer(Bi). The corresponding model compounds were synthesized via reactions using salt(Bi) or salt(Cl) as starting materials. The UV–visible spectra of the polymers had absorption maxima at longer wavelengths than those of the model compounds. This indicated that the π‐conjugation system is expanded along the polymer main chain. Superconducting quantum interference device measurements indicated that polymer(Fe) was paramagnetic. Cyclic voltammetry analysis suggested that the polymers underwent electrochemical oxidation. © 2019 Society of Chemical Industry  相似文献   

15.
A new class of liquid‐crystalline poly(ethylene imine)s (PEIs) having four differently substituted (? CN,? C4H9,? OCH3 and? NO2) azobenzene side‐chain groups attached through alkyl spacer groups were successfully synthesized using a solution polycondensation reaction. The synthesized polymers were characterized using differential scanning calorimetry, polarized optical microscopy and X‐ray diffraction. The photochemical, thermo‐optical and photo‐orientational behavior of the polymers were investigated in detail. Spin‐coated films of PEIs with azobenzene groups having? C4H9,? OCH3 and? NO2 substituents showed out‐of‐plane molecular orientation on annealing. Except for the PEI with an azobenzene group having ? NO2 substituent, all polymers exhibited good photoresponsive properties upon irradiation with UV and visible light. Films of PEIs with azobenzene side groups having? CN,? C4H9 and? OCH3 substituents showed reversible alignment behavior from random state to out‐of‐plane and from out‐of‐plane to random state on annealing and on irradiation with UV and non‐polarized visible light. The reversibility of the molecular orientation of PEIs from random state to out‐of‐plane and from out‐of‐plane to random state greatly depended on the substituent attached to the azobenzene side‐chain group. Copyright © 2010 Society of Chemical Industry  相似文献   

16.
Two zinc clusters Zn4(H3L)4(NO3)4?5H2O ( Zn4 , H4L=(1,2‐bis(1H‐benzo[d]imidazol‐2‐yl)ethane‐1,2‐diol) and [Zn5(H2L′)6](NO3)4]?8H2O?2CH3OH ( Zn5 , H3L′=(1,2‐bis(benzo[d] imidazol‐2‐yl)‐ethenol) have been obtained by the reaction of Zn(NO3)2?6H2O with H4L at 80 °C or 140 °C under solvothermal condition. Powder X‐ray Diffraction (PXRD) of precipitate and Electrospray Ionization Mass Spectrometry (ESI‐MS) of reaction solution revealed the existence of transformation behavior from Zn4 to Zn5 by increasing the temperature from 80 °C to 140 °C, or directly heating Zn4 at 140 °C via solvothermal reaction. Here we proposed a possible mechanism involves split process of Zn4 and reassembly to form Zn5 . ESI‐MS for single crystals revealed [Zn4(H3L)4?3H]+ splits to [Zn(H3L)]+ via [Zn2(H3L)2?H]+. Time dependent ESI‐MS of reaction solution revealed the [Zn(H2L′)]+→[Zn2(H2L′)2?H]+→[Zn5(H2L′)6?H]3+ stepwise assembly. It also has been captured the in situ reaction mainly occurs in the step of [Zn(H3L)]+ to [Zn(H2L′)]+.  相似文献   

17.
5‐Aminotetrazolium nitrate was synthesized in high yield and characterized using Raman and multinuclear NMR spectroscopy (1H, 13C, 15N). The molecular structure of 5‐aminotetrazolium nitrate in the crystalline state was determined by X‐ray crystallography: monoclinic, P 21/c, a=1.05493(8) nm, b=0.34556(4) nm, c=1.4606(1) nm, β=90.548(9)°, V=0.53244(8) nm3, Z=4, ϱ=1.847 g cm−3, R1=0.034, wR2 (all data)=0.090. The thermal stability of 5‐aminotetrazolium nitrate was determined using differential scanning calorimetry; the compound decomposes at 167 °C. The enthalpy of combustion (ΔcombH) of 5‐aminotetrazolium nitrate ([CH4N5]+[NO3]) was determined experimentally using oxygen bomb calorimetry: ΔcombH([CH4N5]+[NO3])=−6020±200 kJ kg−1. The standard enthalpy of formation (ΔfH°) of [CH4N5]+[NO3] was obtained on the basis of quantum chemical computations at the electron‐correlated ab initio MP2 (second order Møller‐Plesset perturbation theory) level of theory using a correlation consistent double‐zeta basis set (cc‐pVTZ): ΔfH°([CH4N5]+[NO3](s))=+87 kJ mol−1=+586 kJ kg−1. The detonation velocity (D) and the detonation pressure (P) of 5‐aminotetrazolium nitrate were calculated using the empirical equations by Kamlet and Jacobs: D([CH4N5]+[NO3])=8.90 mm μs−1 and P([CH4N5]+[NO3])=35.7 GPa.  相似文献   

18.
In this article, a redox-responsive poly(ionic liquid) (redox-PIL) hydrogel Poly(1-vinyl-3-propionate imidazole phenothiazine sulfonic acid)-chitosan [Poly(VPI+PTZ-(CH2)3SO3)-CS] was produced by using chitosan (CS) crosslinking with redox-PIL Poly(1-vinyl-3-propionate imidazole phenothiazine sulfonic acid [Poly(VPI+PTZ-(CH2)3SO3)]. The incorporation of redox-active counter anions 3-(phenothiazine-10-yl) propane 1-sulfonic acid anions (PTZ-(CH2)3SO3) into cationic PIL-polyimidazole rendered Poly(VPI+PTZ-(CH2)3SO3) with electron catalytic ability, ionic conductivity, and electron conductivity. Poly(VPI+PTZ-(CH2)3SO3)-CS combines the properties of hydrogel and redox-PIL, thus offering intrinsic porous conducting frameworks and promoting the transport of charges, ions, and molecules, leading hydrogel with excellent electrochemical properties. The crosslinking occurrence of Poly(VPI+PTZ-(CH2)3SO3) and CS resulting from the synthetic process of hydrogel was verified by differential scanning calorimetry and thermogravimetric analysis. A three-dimensional polymer network hydrogel with good biocompatibility and permeability was formed after crosslinking. In addition, only 64% weight loss within 600 °C was observed in Poly(VPI+PTZ-(CH2)3SO3)-CS representing its thermally stable performance. When used as an electrochemical sensor, the hydrogel-modified gold electrode improved the electrocatalytic oxidation of cysteine. Differential pulse voltammetry results indicated that the detection range was from 5 × 10−8 to 5 × 10−3 M and the limit of detection was 6.64 × 10−8 M. © 2019 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2019 , 136, 48051.  相似文献   

19.
Two different poly(organophosphazene)s-bearing single-substituents (two cinnamyloxide side groups per repeating unit) and phenoxide-co-substituents (one phenoxide and one cinnamyloxide group per repeating unit) were prepared to study their photochemical reaction behavior. Structural formulas confirmed by the NMRs (1H, 31P) and IR were ? [NP(OCH2CH = CH ? C6H5)2]n ? ( I ) and ? [NP(OCH2CH = CH ? C6H5)0.8 (O ? C6H5)1.2]n ? ( II ). Molecular characterization yielded $ \overline {M_w {\rm s}} $ of the order of 103 kg/mol for both the polymers. The onset temperature of decomposition found by TGA was about 250°C. DSC measurements gave Tgs as 1°C for I and 3°C for II . Their photolytic crosslinking behaviors were monitored by UV spectroscopy. The single-substituents polymer I showed a faster rate of photo reaction than the polymer II . The potential use for practical photosensitive application is considered to be greater for the polymer I .  相似文献   

20.
In order to improve the method of synthesis of poly(ethylene terephthalate) (PET), a series of ionic liquids (ILs) based on benzyl imidazolium ([YBMIM][X], Y = NO2, CH3, F; B = benzyl; X = Tf2N) were used to investigate the formation of PET at low temperature and pressure. High molecular weight PET (Mw up to 2.6 × 104 g mol?1) was obtained by two‐step polycondensation in these ILs at lower temperature (230–240 °C) than with traditional melt polycondensation (270–290 °C). Moreover, the molecular weight of the resulting PET was found to depend on the activities of the catalysts used in the ILs. The catalysts (Sb2(OCH2CH2O)3, Sb(OAc)3, Sb2O3) used in the preparation of PET have little effect on the thermostability of the ILs. The ILs can decrease the viscosity of the reaction system, and thus small molecules can be easily removed. Copyright © 2012 Society of Chemical Industry  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