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.
Thermal stability of [(CH3SiH)30(C6H5SiCH3)70]n a hydropolysilane copolymer, in vacuum and its crosslinking reactions with vinylic silanes as crosslinking agents was evaluated in order to obtain high yields of oxygen-free silicon carbide ceramics. It was found that the polymer was thermally stable in vacuum up to 140 °C for 20 hrs based on Fourier transform infrared spectroscopy analysis. The crosslinking reactions of the polymer occurred to various extents depending on the type of vinylic silanes used as evidenced by Fourier transform infrared spectroscopy, ultraviolet spectroscopy, gel permeation chromatography, thermogravimetry and solubility data. The additions of vinylic silanes to Si-H in the hydropolysilane were found to obey anti-Farmer's rule, despite Farmer's addition of unsaturated hydrocarbons to Si-H.  相似文献   

2.
Platinum(III) is no longer an uncommon oxidation state. Numerous binuclear platinum(III) complexes have been prepared and structurally characterized over the past eight years. These include sulfate-bridged dimers of D4h symmetry, [Pt2(SO4)4L2]2−, L = H2O, DMSO; phosphate-bridged complexes [Pt2(HPO4)4-(H2O)2]2− and [Pt2(H2PO4)(HPO4)3(py)2]; POP (H2 P2 O2−5)-bridged ions [Pt2-(POP)4X2]2−, X = halide; an extensive series of α-pyridonate (C5H4NO)-bridged head-to-head and head-to-tail complexes, [Pt2(NH3)4 (C5H4NO)2XY]n, X, Y = NO3, NO2, H2O, Cl, Br; n = 2, 3; and organometallic derivatives such as [Pt2(CH3)4-(CF3CO2)2(4-Mepy)2]. In all cases there is a Pt–Pt single bond of length 2.47–2.7 Å, pseudo-octahedral geometry about platinum, and two or more bridging ligands. The complexes are stable in solution and some undergo quasi-reversible two-electron redox reactions. Mononuclear platinum(III) complexes are less well characterized structurally, but have been stabilized in diamagnetic host lattices in the solid state and by macrobicyclic cage ligands in solution following pulse radiolytic or γ-irradiation of precursor platinum(II) complexes. The first unequivocal, crystallographically characterized mononuclear platinum(III) complex, [Pt(C6Cl5)4], has just been reported.  相似文献   

3.
The effect of the nature of the anion on the performance of ionic rhodium catalysts has received little attention. Herein it is shown that the use of highly fluorous tetraphenylborate anions can enhance catalyst activity in both conventional and fluorous media. For hydrogenation catalysts of the type [Rh(COD)(dppb)][X] {COD=1,5‐cis,cis‐cyclooctadiene; dppb=1,4‐bis(diphenylphosphino)butane; X=BF4 ( 1a ), [BPh4] ( 1b ), [B{C6H4(SiMe3)‐4}4] ( 1c ), [B{C6H3(CF3)2‐3,5}4] ( 1d ), [B{C6H4(SiMe2CH2CH2C6F13)‐4}4] ( 1e ), [B{C6H4(C6F13)‐4}4] ( 1f ) and [B{C6H3(C6F13)2‐3,5}4] ( 1 g )} the activity towards the hydrogenation of 1‐octene in acetone increased in the order 1c < 1b < 1e < 1a < 1d ~ 1f < 1g with 1g being twice as active as the commonly applied 1a . Despite the fluorophilic character introduced by the substituted tetraarylborate anions, the presence of some perfluoroalkyl‐substituents in the cation was still required for achieving high partition coefficients. Therefore, [Rh(COD)(Ar2PCH2CH2PAr2)][X] {Ar=C6H4(SiMe2CH2CH2C6F13)‐4, X=[B{C6H3(C6F13)2‐3,5}4] ( 3f ); Ar=C6H4(SiMe(CH2CH2C6F13)2)‐4 and X=[B{C6H4(C6F13)‐4}4] ( 2g )} were prepared, which were active in the hydrogenation of 1‐octene, 2g even more so than 3f . Both these highly fluorous catalysts could be recycled with 99% efficiency through fluorous biphasic separation, whereas the corresponding BF4 complex of 2g ( 2a ) did not show any affinity for the fluorous phase.  相似文献   

4.
The carbonylation of CH4 with carbon monoxide in superacids HF/SbF5 or HSO3F/SbF5 leads to the exclusive and quantitative formation of the acylium ion ([CH3CO]+[SbF6]) with the concomitant stoichiometric formation of SbF3.  相似文献   

5.
Polysiloxanes [RSiO1.5]n with R=CH3 (PMS) and C6H5 (PPS), respectively, were transformed to Si–O–C ceramics of variable composition and structure upon pyrolysis in inert atmosphere at 800–1500°C. The electrical conductivities of the Si–O–C ceramics in air were measured at room temperature by using a shielded two point configuration. In situ measurements of the dc-conductivity during the pyrolytic conversion from the polymer to the ceramic phase were carried out up to 1500°C with four point contacted carbon electrodes in inert atmosphere. During polymer-ceramic conversion excess carbon precipitates above 400°C (PPS)–700°C (PMS). At temperatures above 800°C (PPS) and 1400°C (PMS) coagulation and growth of the carbon clusters results in a percolation network formation. While below the percolation threshold electrical conductivity can be described according to Motts mechanism by variable-range-hopping of localized charge carriers, regular electron band conduction due to the instrinsic conductivity of turbostratic carbon (8×10−4 (Ωcm)−1) predominates above. Thus, the in situ measurement of non-linear electrical property changes can be used as a microprobe of high sensivity to detect microstructural transformations during the pyrolysis of preceramic polymers.  相似文献   

6.
The performance of jet fuel depends on the density (ρ), condensed phase heat of formation (▵fH°(c)), and specific impulse (ISP). Exo‐tricyclo[5.2.1.0(2,6)]decane (C10H16) or JP‐10 is now used as a suitable synthetic liquid jet fuel because it has the approximated values of ρ=1.1 g cm−3 and ▵fH°(c)=− 123 kJ mol−1 and a broad range between the melting and boiling points, i.e. TbpTmp=196.2 K. This work introduces a suitable pathway for calculation of the values of ρ, ▵fH°(c), and ISP of 13 well‐known isomers of JP‐10 and a series of saturated polycyclic hydrocarbons with general formula of CnHn (5≤n≤12) in order to specify high performance jet fuels. Although 13 compounds have larger values of ISP*ρ than JP‐10, only two compounds, tetraspiro[2.0.0.0.2.1.1.1]undecane and tetracyclo[3.2.0.0(2,7).0(4,6)]heptane, are suitable as jet fuels.  相似文献   

7.
Three novel Si-C-B-N ceramic compositions, namely Si2.9B1.0C14N2.9, Si3.9B1.0C11N3.2 and Si5.3B1.0C19N3.4, were synthesized using the polymer-to-ceramic transformation of the polyorganoborosilazanes [B(C2H4Si(Ph)NH)3]n, [B(C2H4Si(CH3)NH)2–(C2H4Si(CH3)N(SiH2Ph))]n, and [B(C2H4Si(CH3)–N(SiH2Ph))3]n, where Ph is phenyl (C6H5), at 1050°C in argon. The Si-B-C-N ceramics exhibited significant stability with respect to composition and mass change in the temperature range between 1000° and 2200°C, including isothermal annealing of the samples at the final temperature for 30 min in argon. The mass loss rate at 2200°C was as low as 1.4 wt%·h−1 for Si5.3B1.0C19N3.4, 1.7 wt%·h−1 for Si2.9B1.0C14N2.9, and 2.4 wt%·h−1 for Si3.9B1.0C11N3.2. The measured amount of mass loss rate was comparable to that of pure SiC materials. As crystalline phases, β-Si3N4 and β-SiC were found exclusively in the samples annealed at 2200°C at 0.1 MPa in argon. For thermodynamic reasons, β-Si3N4 should have decomposed into the elements silicon and nitrogen at that particular temperature and gas pressure. However, the presence of β-Si3N4 in our materials indicated that carbon and boron kinetically stabilized the Si3N4-based composition.  相似文献   

8.
Cationic heterobimetallic {[P(C6H4CH2NMe2-2)3]Ag-bipy-Cu[P(C6H4CH2NMe2-2)3]}2+(OTf,PF6) (5) (bipy = 4,4-bipyridine, OTf=OSO2CF3) is accessible by the subsequent reaction of [P(C6H4CH2NMe2-2)3]AgOTf (1) with bipy (2) to afford {[P(C6H4CH2NMe2-2)3]Ag(bipy)}+(OTf) (3), which on treatment with equimolar amounts of {[P(C6H4CH2NMe2-2)3]Cu}+(PF6) (4) produces 5. In 5 the respective group-11 metal(I) centres are bridged by the π-conjugated organic ligand bipy. The electrochemical behaviour of 5 is reported.  相似文献   

9.
Two asymmetric alkylidene‐bridged dinuclear titanocenium complexes (CpTiCl2)25‐η5‐C9H6(CH2)nC5H4), 1 (n = 3) and 2 (n = 4) have been prepared by treating two equivalents of CpTiCl3 with the corresponding dilithium salts of the ligands C9H7(CH2)nC5H5 (n = 3, 4). Additionally, Ti(η55n‐BuC5H4C5H5)Cl2 (3) and Ti(η55n‐BuC9H6C5H5)Cl2 (4) were synthesized as corresponding mononuclear complexes. All complexes were characterized by 1H, 13C NMR, and IR spectroscopy. Homogenous ethylene polymerization catalyzation using those complexes has been conducted in the presence of methylaluminoxane (MAO). The influences of reaction parameters, such as [MAO]/[Cat] molar ratio, catalyst concentration, ethylene pressure, temperature, and time have been studied in detail. The results showed that the catalytic activities of both dinuclear titanocenes were higher than those of the corresponding mononuclear titanocenes. Although the two dinuclear complexes were different in only one [CH2] unit, the catalytic activity of 2 was about 50% higher than that of 1; however, the molecular weight of polyethylene (PE) obtained by 2 was lower than that obtained from 1. The molecular weight distribution of PE produced by these dinuclear complexes reached 6.9 and 7.3, respectively. © 2006 Wiley Periodicals, Inc. J Appl Polym Sci 101: 3317–3323, 2006  相似文献   

10.
Photolysis of Ternary Fe(III)-Complexes with Oxalate and Phenolate Ligands The overall iron(II) quantum yields of the photolysis of mixed ligand complexes [Fe(C2O4)3−nLn]3− (L = salicylate, sulfosalicylate or the dianion of catechol, sodium 3,5-sulfocatechol, tetrachlorcatechol) in aqueous solution within the spectral region 436 nm ≤ λirr. ≤ 546 nm are much smaller than those of the complexes [Fe(C2O4)3−n(H2O)2n]3−2n (n = 0; 1; 2). This means that the intramolecular photoredoxreaction between iron(III) and the oxalate ligand is not sensitized by the L→Fe(III) charge transfer excitation. Moreover, filter effects through photoinactive complexes [FeL2(H2O)2] or [FeL(H2O)4]+ probably give rise to the strong decrease of the quantum yields.  相似文献   

11.
The isostructural complexes [C5H5N+(CH2)nCOO]2HX and [C6H5(CH2)nCOO]2HK (n = 1–4), which differ in their counterions and charge on the ring, were synthesized, and their powder FT-IR spectra analyzed. All complexes containing a charged pyridine ring are of Hadži type iii, characterized by an intense broad (continuum) absorption below 1600 cm−1 typical of a short-strong hydrogen bond (SSHB) with a delocalized proton and a single vC=O band. The positively charged nitrogen atoms interact electrostatically with the X ion and, additionally, with one of the oxygen atoms of the carboxylic group, producing a more or less symmetric environment of the H-bonded proton, and stabilizing the SSHB. The broad absorption of [C6H5CH2COO]2HK is very similar to that of other pyridine complexes. Upon addition of methylene groups the broad absorption moves to higher wavenumbers, the O···O distance is elongated, and the H-bonded proton becomes more localized. In the spectrum of [C6H5(CH2)4COO]2HK the vC=O and vasCOO bands were found at 1704 and 1641 cm−1, respectively, which shows that the H-bonded proton is asymmetrically located. The observed variation of absorption with the number of CH2 groups reflects changes of contacts between the K+ ion and COO groups.  相似文献   

12.
The reaction of the sodium salt of the monoanion, nido-[2,3-(Si(CH3)3)2-2,3-C2B4H5], with (chloromethyl)dimethylchlorosilane in a 1:1 molar ratio produced the B(cage)-substituted cluster, nido-5-ClCH2Si(CH3)2-2,3-(Si(CH3)3)2-2,3-C2B4H5 (1), in 81% yield. This product (1) was reacted further with the lithium salt of [closo-1-R-1,2-C2B10H10] monoanion (R=Me, Ph) to give the novel linked and mixed C2B4/C2B10 carborane species, 1-Me-2-[5-SiMe2CH2-2,3-(SiMe3)2-2,3-C2B4H5]-1,2-C2B10H10 (2), 1-Ph-2-[5-SiMe2CH2-2,3-(SiMe3)2-2,3-C2B4H5]-1,2-C2B10H10 (3), in yields of 76% and 81%, respectively.  相似文献   

13.
The reaction of the Cu(II) bis N,O‐chelate‐complexes of L‐2,4‐diaminobutyric acid, L‐ornithine and L‐lysine {Cu[H2N–CH(COO)(CH2)nNH3]2}2+(Cl)2 (n = 2–4) with terephthaloyl dichloride or isophthaloyl dichloride gives the polymeric complexes {‐OC–C6H4–CO–NH–(CH2)n–CH(nh2)(COO)Cu(OOC)(NH2)CH–CH2)n–NH‐}x 1 – 5 . From these the metal can be removed by precipitation of Cu(II) with H2S. The liberated ω,ω′‐N,N′‐diterephthaloyl (or iso‐phthaloyl)‐diaminoacids 6 – 10 react with [Ru(cymene)Cl2]2, [Ru(C6Me6)Cl2]2, [Cp*RhCl2]2 or [Cp*IrCl2]2 to the ligand bridged bis‐amino acidate complexes [Ln(Cl)M–(OOC)(NH2)CH–(CH2)nNH–CO]2–C6H4 11 – 14 .  相似文献   

14.
Rates of hydrogen atom attack on o-fluorotoluene (o-FTOL) and m-fluorotoluene (m-FTOL) at temperatures of 988–1144 K and pressures of 2–2.5 bar have been determined in a single-pulse shock tube study. Hydrogen atoms, generated from the decomposition of hexamethylethane, were allowed to react with the substrates and the characteristic products observed. Rate constants for two reaction channels, displacement of fluorine or methyl, were determined relative to displacement of methyl from 1, 3,5-trimethylbenzene (135TMB). Evidence is presented that abstraction of F is unimportant over the studied temperature range. With k(H + 135TMB → m-xylene + CH3) = 6.7 × 1013 exp(–3255/T) cm3 mol−1s−1, the following rate expressions have been derived: k(H + o-FTOL → C6H5CH3 + F) = 8.38 × 1013 exp(–6041/T) cm3 mol−1s−1; (1012–1142 K) k(H + o-FTOL → C6H5F + CH3) = 2.37 × 1013 exp(–2938/T) cm3 mol−1s−1; (988–1142 K) k(H + m-FTOL → C6H5CH3 + F) = 1.33 × 1014 exp(–6810/T) cm3 mol−1s−1; (1046–1144 K) k(H + m-FTOL → C6H5F + CH3) = 2.04 × 1013 exp(–3104/T) cm3 mol−1s−1; (1008–1144 K) Uncertainties in the relative rate constants are estimated to be factors of about 1.1, while the above absolute values have estimated expanded uncertainties of about a factor of 1.4 in rate, 10 kJ mol−1 in the activation energy, and a factor of 3 in the A-factor. The present data are compared with relevant literature data. From our data and the thermochemistry, a model of the elementary steps comprising displacement of F is developed. On the basis of the model fit to our data, rate constants for the addition of atomic fluorine to toluene at 1100 K are derived. Rate expressions for fluorination reactions of toluene are also determined. The significance of the present results is discussed in the context of the formation of fluorinated byproducts in high-temperature systems.  相似文献   

15.
On Pt(111) at 110 K, 1-iodopropane, C3H7I, adsorbs molecularly, but for doses below 1.7 × 1014 molecules cm−2, only H2 and I appear in thermal desorption. C–I bond cleavage occurs between 160 and 220 K, forming adsorbed n-propyl, C(a)H2CH2CH3, and atomic iodine, based on temperature-programmed desorption (TPD), high-resolution electron energy loss spectroscopy (HREELS), and X-ray photoelectron spectroscopy (XPS). n-Propyl undergoes β-hydride elimination forming propylene, with desorption peaks at 185 and 240 K. At 240 K, hydrogenation to propane also occurs. Some di-σ bonded propylene, C(a)H2C(a)HCH3, remains at 240 K and it rearranges to propylidyne near 300 K. Atomic H, bound to Pt, recombines and desorbs at ca. 260 K. Further desorption of H2 is limited by C–H bond breaking and occurs over a broad temperature range with local maxima at ca. 280, 320, and 420 K, typical of propylidyne fragments on Pt. Atomic iodine desorbs in a broad feature at 825 K.  相似文献   

16.
Substituted Alkinyles as Axial Ligands at Hemine Like Bound Iron(III) - Incorporation into a Spectrochemical Series . Substituted lithium alkynyles Li CC R (R = tBu, Ph, p-Cl C6H4, Me3Si, iPr3Si, Ph3Si) react with the hemine like macrocyclic iron(III) complex 6,13-di(ethoxycarbonyl)-5, 14-dimethyl-1, 4, 8, 11-tetraazatetradeca-4,6,12,14-tetraenato[2]iron(III)-iodide (formula 2 ;) in tetrahydrofuran to form anionic low-spin di-adducts [fe(CC R)2]. The incorporation of the alkynyles into a spectrochemical series of the axial ligands (studied by the sharp equatorial-ligand-to-metal CT absorption band) results in the wavelength-sequence (nm): OH (≈︁ 510) « N3 (≈︁ 625) < tBu CC (664) < NH3 (666) < Ph CC (692) < Ph NH2 (695) < Me3Si CC (698) < SCN (713) < Ph3Si C  C (716) < CN (739) < 4-picoline (759) < pyridine (765) < nicotinamide (776) < methylnicotinat (788) < pyrazine (798) and points to a significant π-acceptor ability of the silyl substituents.  相似文献   

17.
The few clusters [B?nA+n+1]+ (n = 0,1) with resolvable mobilities formed in electrosprays of large salts have been used for nanoparticle instrument testing and calibration at sizes smaller than 2 nm. Extensions of this modest size range by charge reduction with uncontrolled gas phase ions has resulted in impure singly charged clusters. Here, we combine two oppositely charged electrosprays of solutions of the same salt B?A+, including: (CnH2n+1)4N+Br? (n = 4,7,12,16), the large phosphonium cation (C6H13)3(C16H33)P+ paired with the anions Im? [(CF3SO2)2N?] or FAP? [(C2F5)3PF3?], and the asymmetric pair [1-methyl-3-pentylimidazolium+FAP?]. Both polarities are simultaneously produced by this source in comparable abundances, primarily as singly charged A+nB?n±1, with tiny contributions from higher charge states. Some but not all of these clusters produce narrow mobility peaks typical of pure ions, even beyond n = 43. Excellent independent stable control of the positive and the negative sprays brought very close to each other is achieved by isolating them electrostatically with a symmetrically interposed metallic screen. Two nanoDMAs covering the size range up to 30 nm (Halfmini and Herrmann DMAs, with classification lengths of 2 and 10 cm) are characterized with these standards, revealing resolving powers considerably higher than previously seen with unipolar electrospray sources. The bipolar source of pure and chemically homogeneous clusters described permits studying size and charge effects in a variety of aerosol instruments in the 1–4 nm size range.

Copyright © 2017 American Association for Aerosol Research  相似文献   

18.
Irradiation of surfactant vesicles prepared from (C18H37)2 N+(CH3)C6H4-CHCH2p,CT, 1, [C15H31CO2(CH2)2]2N+(CH3)CH2C6H4CHCH2,CL 2, and (C18-H37)2N+(CH3)CH2CH2OCOC6H4CHCH2p,Br, 3, by ultraviolet light or by bursts of 266 nm laser pulses have resulted in the loss of styrene absorbances. This process has been accounted for in terms of a model which considers intravesicular surface reactions to give polymers with average chainlength of 22. Degreees of photopolymerization have been determined in vesicles prepared from 3 subsequent to separating the polystryrene, formed in the photolysis, from the surfactants. Vesicle surface photopolymerizations result in aqueous cleft formation and in enhanced stabilities. Polymerized vesicles provide media for in situ generation of colloidal catalysts and semiconductors.  相似文献   

19.
Two Q[n]-based porous compounds, Q[5]·[PtCl6]2 ·2H3O·12H2O and 2Q[8]·[PtCl6]2 ·2H3O·74H2O in cooperating the hexachloroplatinate(IV) anion ([PtCl6]2 ) as an inorganic structure directing agent are demonstrated. The driving forces for the formation of such novel Q[n]-based porous compounds are considered to be the outer surface interactions of Q[n]s, including dipole interactions between Q[n]s and ion-dipole interactions between [PtCl6]2  anions and Q[n]s. Moreover, the Q[5]-based porous compound displays absorption distinctness for tetrachloromethane, whereas the Q[8]-based porous compound displays absorption distinctness for methanol.  相似文献   

20.
Ionic liquids have been projected as the best solvent for extraction and separation of bioactive compounds from various origins. This review offers a collection of the published results, using ionic liquids for the extraction and purification of biomolecules. Ionic liquids have been studied as solvents, co-solvents and supported materials for separation of bioactive compounds. The ionic liquids-based extraction procedures were previously reported, such as ionic liquids-based solid-liquid extraction, liquid-liquid extraction and ionic liquids-modified materials are reviewed and compared to their performance. In this review, the main activities and future challenges are discussed, with major gaps identified using ionic liquids in extraction procedures and by advancing few steps to overcome these drawbacks.

Abbreviation: [(HSO3)C4MIM]+: 1-(4-sulfonylbutyl)-3-methylimidazolium; [(C6H3OCH2)2im]+: 1,3-dihexyloxymethylimidazolium; [CnC1MIM]+: 1-alkyl-2,3-dimethylimidazolium; [CnMIM]+; [Cn, 2, 3, 4, 6, 8, 10, 12]: 1-alkyl-3-methylimidazolium; [CnC1pyr]+: 1-alkyl-3-methylpyridinium; [Cnim]+: 1-alkylimidazolium; [Cnpyr]+: 1-alkylpyridinium; [aCnim]+: 1-allyl-3-alkylimidazolium; [C7H7MIM]+: 1-benzyl-3-methylimidazolium; [C4(C1C1C1Si)im]+: 1-butyl-3-trimethylsilylimidazolium; [(HOOC)C2MIM]+: 1-carboxyethyl-3-methylimidazolium; [(OH)CnMIM]+: 1-hydroxyalkyl-3-methylimidazolium; [(C2H5O)3SiC3MIM]+: 1-methyl-3-(triethoxy)silypropyl imidazolium; [(NH2)C3MIM]+: 1-propylamine-3-methylimidazolium; [CwHxNyOz]+: Chirally functionalized methylimidazolium; [P10(3OH)(3OH)(3OH)]+: Decyltris(3-hydrox- ypropyl) phosphonium; [N111(2OH)]+: N,N,N-trimethyl-N-(2-hydroxyethyl) ammonium (cholinium); [N00nn]+: N,N-dialkylammonium; [N0nn(2OH)]+: N,N-dialkyl-N-(2-hydroxyethyl) ammonium; [C10C10C1gluc]+: N,N-didecyl-N-methyl-d-glucaminium; [N11(2(O)1)0]+: N,N-dimethyl(2-methoxyethyl) ammonium; [N11(2OH)(C7H7)]+: N-benzyl-N,N-dimethyl-N-(2-hydroxyethyl) ammonium; [P66614]+: Trihexyltetradecylph- osphonium; [Pi(444)1]+: Triisobutyl (methyl) phosphonium; P.minus: Polygonum minus; NPs: Nanoparticle; ZnO : Zinc oxide nanoparticles ; Ni NPs: Nickel nanoparticles; MO: Methyl orange; UAE: Ultrasonic-assisted extraction; LLE: Liquid-liquid extraction; ABS: Aqueous biphasic system ; [Ace]?: Acesulfamate; [Ala]?: alalinate; [TMPP]?: bis(2,4,4-trimethylpentyl)phosphinate; : ; [NTf2]?: bis(trifluoromethylsulfonyl)imide; [[Br]–]: [Br]omide; [Calc]: calkanoate; [Cl]: chloride; [Bz]?: benzoate; [PF6]?: hexafluorophosphate; [HSO4]?: hydrogenosulfate; [OH]?: hydroxide; I: iodide; [Lac]?: lactate; [NO3]?: nitrate; [[Cl]O4]?: perchlorate; [Phe]?: phenilalaninate; [BF4]?: tetrafluoroborate; [SCN]?: thiocyanate; [C(CN)3]?: tricyanomethanide; [CF3CO2]?: trifluoroacetate; [CF3SO3]?: trifluoromethanesulfonate; [FAP]?: tris(pentafluoroethyl)trifluorophosphate; ILs: Ionic liquids; Ag NPs: Silver nanoparticle; Cu NPs: Copper nanoparticle; MB: Methylene blue; MR: Methyl red ; MAE: Microwave-assisted extraction; SLE: solid-liquid extraction.  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