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 453 毫秒
1.
Low-energy CAD product-ion spectra of various molecular species of phosphatidylserine (PS) in the forms of [M−H] and [M−2H+Alk] in the negative-ion mode, as well as in the forms of [M+H]+, [M+Alk]+, [M−H+2Alk]+, and [M−2H+3Alk]+ (where Alk=Li, Na) in the positive-ion mode contain rich fragment ions that are applicable for structural determination. Following CAD, the [M−H] ion of PS undergoes dissociation to eliminate the serine moiety (loss of C3H5NO2) to give a [M−H−87] ion, which equals to the [M−H] ion of a phoshatidic acid (PA) and give rise to a MS3-spectrum that is identical to the MS2-spectrum of PA. The major fragmentation process for the [M−2H+Alk] ion of PS arises from primary loss of 87 to give rise to a [M−2H+Alk−87] ion, followed by loss of fatty acid substituents as acids (RxCO2H, x=1,2) or as alkali salts (e. g., RxCO2Li, x=1,2). These fragmentations result in a greater abundance of [M−2H+Alk−87−R2CO2H] than [M−2H+Alk−87−R1CO2H] and a greater abundance of [M−2H+Alk−87−R2CO2Li] than [M−2H+Alk−87−R1CO2Li]; while further dissociation of the [M−2H+Alk−87−R2(or 1)CO2Li] ions gives a preferential formation of the carboxylate anion at sn-1 (R1CO2) over that at sn-2 (R2CO2). Other major fragmentation process arises from differential loss of the fatty acid substituents as ketenes (loss of Rx′CH=CO, x=1,2). This results in a more prominent [M−2H+Alk−R2′CH=CO] ion than [M−2H+Alk−R1′CH=CO] ion. Ions informative for structural characterization of PS are of low abundance in the MS2-spectra of both the [M+H]+ and the [M+Alk]+ ions, but are abundant in the MS3-spectra. The MS2-spectrum of the [M+Alk]+ ion contains a unique ion corresponding to internal loss of a phosphate group probably via the fragmentation processes involving rearrangement steps. The [M−H+2Alk]+ ion of PS yields a major [M−H+2Alk−87]+ ion, which is equivalent to an alkali adduct ion of a monoalkali salt of PA and gives rise to a greater abundance of [M−H+2Alk−87−R1CO2H]+ than [M−H+2Alk−87−R2CO2H]+. Similarly, the [M−2H+3Alk]+ ion of PS also yields a prominent [M−2H+3Alk−87]+ ion, which undergoes consecutive dissociation processes that involve differential losses of the two fatty acyl substituents. Because all of the above tandem mass spectra contain several sets of ion pairs involving differential losses of the fatty acid substituents as ketenes or as free fatty acids, the identities of the fatty acyl substituents and their positions on the glycerol backbone can be easily assigned by the drastic differences in the abundances of the ions in each pair.  相似文献   

2.
The complexes [Bu4N]2+[PtBr6]2− (I), [Ph4P]2+[PtBr6]2− (II), and [Ph3(n-Am)P]2+ (III) are synthesized by the reactions of tetrabutylammonium bromide, tetraphenylphosphonium bromide, and triphenyl(n-amyl)-tetraphenylphosphonium bromide, respectively, with potassium hexabromoplatinate (mole ratio 2: 1). After recrystallization from dimethyl sulfoxide, complexes I, II, and III transform into [Bu4N]+[PtBr5(DMSO)] (IV), [Ph4P]+[PtBr5(DMSO)] (V), and [Ph3(n-Am)P]+[PtBr5(DMSO)] (VI). According to the X-ray diffraction data, the cations of complexes IVVI have a slightly distorted tetrahedral structure. The N-C and P-C bond lengths are 1.492(7)–1.533(6) and 1.782(10)–1.805(10) ?, respectively. The platinum atoms in the mononuclear anions are hexacoordinated. The dimethyl sulfoxide ligands are coordinated with the Pt atom through the sulfur atom (Pt-S 2.3280(18)–2.3389(11) ?). The Pt-Br bond lengths are 2.4330(6)–2.4724(6) ?.  相似文献   

3.
Mercury complexes [Ph3AlkP]2+[Hg2I6]2− and [Ph3AlkP]2+[Hg4I10]2− (R = Me, Et, Pr, iso-Pr, Bu, iso-Bu) are synthesized by the reactions of triphenylalkylphosphonium Ph3AlkPI with mercury iodide in acetone with the mole ratio 1: 1 and 1: 2, respectively. According to X-ray diffraction data, the phosphorus atom in the cations of the [Ph3(iso-Pr)P]2+[Hg2I6]2−, [Ph3BuP]2+[Hg2I6]2−, and [Ph3(iso-Pr)P]2+[Hg4I10]2− complexes has a distorted tetrahedral coordination. The CPC bond angles and P-C bond lengths vary within 107.3(4)°-112.0(4)° and 1.774(8)-1.827(7) ?. In the [Hg2I6]2− centrosymmetric binuclear anions, the mercury atoms of tetrahedral coordination lie in two near-perpendicular Hg2I6planes. Hg4I4 eight-membered cycles of the [Hg4I10]2− tetranuclear anion are joined into polymeric chains through Hg … I coordination bonds (3.334, 3.681 &OA) due to which Hg atoms have a trigonal bipyramidal coordination. Original Russian Text ? V.V. Sharutin, V.S. Senchurin, N.N. Klepikov, O.K. Sharutina, 2009, published in Zhurnal Neorganicheskoi Khimii, 2009, Vol. 54, No. 2, pp. 267–273.  相似文献   

4.
The behavior of the phosphine-phosphine sulfide complexes of silver, [Ph2P(S)(CH2) n PPh2] m ·AgNO3 (n=2 or 4;m=1 or 2), in pyridine was studied. Dissolution of the 1:1 complexes in pyridine leads to destruction of their dimeric structures Ag2[Ph2P(S)(CH2) n PPh2]2(NO3)2 (A) to form the complexes Agpy +−P(Ph2)(CH2) n Ph2P=S and Agpy +−S=PPh2(CH2) n PPh2. The solid complexes isolated from pyridine restore dimeric structure A. According to the data of X-ray diffraction analysis, the 1:2 complex isolated from pyridine has the structure [S=P(Ph2)(CH2)2(Ph2)P−(NO3)Ag(Py)−P(Ph2) (CH2)2(Ph2)P=S]Py. According to the data of IR spectroscopy, dissolution of this complex in chloroform leads to the formation of the dimeric structure Ag2Ph2P(S)(CH2)2PPh2]4(NO3)2. Deceased. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 9, pp. 1751–1758, September, 1998.  相似文献   

5.
Radical-ion salts bis(biphenyl)chromium(i) 1,4-di(2-cyanoisopropyl)-1,4-dihydrofulleride [(Ph2)2Cr][1,4-(CMe2CN)2C60]−· and bis(biphenyl)chromium(i) 1-(2-cyanoisopropyl)-1,2-dihydrofulleride [(Ph2)2Cr][1,2-(CMe2CN)(H)C60]−·, the salt bis(biphenyl)chromium(i) (2-cyanoisopropyl)fulleride [(Ph2)2Cr][(CMe2CN)C60], and neutral 1-(2-cyanoisopropyl)-1,2-dihydrofullerene 1,2-(CMe2CN)(H)C60 have been synthesized for the first time. The compounds [(Ph2)2Cr][1,4-(CMe2CN)2C60]−· and [(Ph2)2Cr][1,2-(CMe2CN)(H)C60]−· decompose in THF to form [(Ph2)2Cr][(CMe2CN)C60], whose protonation affords 1,2-(CMe2CN)(H)C60. 1,4-Di(2-cyanoisopropyl)-1,4-dihydrofullerene 1,4-(CMe2CN)2C60 and 1,2-(CMe2CN)(H)C60 are stable in vacuo up to 513 K. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 9, pp. 1935–1939, September, 2008.  相似文献   

6.
Complexes [Ph4P] 2 + [Hg4I10]2− (I) and [[Ph4P] 2 + [BiI5(Me2S=O)]2− (II) are synthesized by the reactions of tetraphenylphosphonium Ph4PI with mercury diiodide in acetone and with bismuth triiodide in dimethyl sulfoxide, respectively. According to X-ray diffraction analysis, structural units of these complexes are tetraphenylphosphonium cations and tetra- and mononuclear anions, respectively. The phosphorus atoms in the tetraphenylphosphonium cations have a distorted tetrahedral coordination. In the central fragment of the centrosymmetric anion [Hg4I10]2−, the distances between the terminal mercury atoms and iodine atoms are 3.503(2) Å. The mercury atoms in the central and terminal fragments of compound I have distorted tetrahedral and trigonal coordinations, respectively. The bismuth atom in the mononuclear octahedral anion of complex II contains a dimethyl sulfoxide molecule along with five iodine atoms in the coordination sphere. __________ Translated from Koordinatsionnaya Khimiya, Vol. 31, No. 10, 2005, pp. 791–795. Original Russian Text Copyright ? 2005 by Sharutin, Egorova, Sharutina, Dorofeeva, Molokov, Fukin.  相似文献   

7.
Complexes Ph3(n-Pr)P2+[CoI4]2− (I) and [Ph3(n-Am)P]2+ [CoI4]2− (II) were synthesized by reactions of triphenyl(alkyl)phosphonium iodide with cobalt(II) iodide in acetone. According to the X-ray diffraction data, complexes I and II consist of tetrahedral triphenyl(alkyl)phosphonium cations (for I, P-C is 1.787(4)–1.804(4) ? and CPC is 106.73(18)°–111.4(18)°; for II P-C is 1.786(6)–1.802(6) ? and CPC is 107.6(3)°–111.7(3)°) and [CoI4]2− anions (Co-I 2.5923(6)–2.6189(6) ?, ICoI 101.86(2)°–113.25(2)° for I; Co-I 2.5899(9)–2.6171(9) 107.01(3)°–110.47(3)° for II).  相似文献   

8.
Methods were developed for the controlled thermal synthesis of high-spin cubane-like pivalates {MII 43−OR)4} (M = Co or Ni; R = H or Me) starting from mono-and polynuclear complexes. The solid-state thermal decomposition of the known pivalate clusters [MII 43−OMe)4−(μ2−OOCBut)22−OOCBut)2(MeOH)4] and the new clusters [M4II3)−OH41−OOCBut)3−(μ−(NH2)2C6H2Me2)31−(NH2)2C6H2Me2)3]+(OOCBut)− (M = Co or Ni) was studied by differential scanning calorimetry and thermogravimetry. The thermolysis of cubane-like CoII and NiII pivalates is a destructive process. The phase composition of the decomposition products is determined by the nature of coordinated ligands and the structural features of the metal core.  相似文献   

9.
Four imidazolyl acetamido p-tert-butylcalix[4]arenes 5–8 have been prepared by reacting the corresponding methyl esters derivatives 1–4 with histamine in 1:1 mixture of methanol:toluene. The yields ranged from 56 to 68%. 5–8 have been shown to be in cone conformation. The complexation behaviour of 5–8 towards monovalent metal picrates M+Pic with M+ = Li+, Na+, K+, Rb+ and Cs+ and divalent metal picrates M2+(Pic)2 with M2+ = Mg2+, Ca2+, Sr2+, Ba2+, Pb2+, Cd2+, Zn2+ and Co2+ are given. Tentative localisation of the metal cations in the receptors is given. The binding properties towards these cations have been determined along with stoichiometries of the complexes.  相似文献   

10.
The [Ph3P+−CMe2−SiMe2−SEt]Br salt was prepared by the reaction of betaine Ph3P+−CMe2SiMeR−S (1a: R=Me) with EtBr. Acetylation of betaine1a or Et3P+−CHMeSiMe2−S (2a) afforded 2,2,6-trimethyl-1,3-dioxa-2-silacyclohex-5-ene-4-thione   相似文献   

11.
《Electroanalysis》2003,15(12):1043-1053
The redox chemistry of the stable tetracoordinated 16 valence electron d8‐[Ir+I(troppPh)2]+(PF6)? and pentacoordinated 18 valence d8‐[Ir+I(troppPh)2Cl] complexes was investigated by cyclic voltammetry (troppPh=dibenzotropylidenyl phosphine). The experiments were performed using a platinum microelectrode varying scan rates (100 mV/s–10 V/s) and temperatures (? 40 to 20 °C) in tetrahydrofuran, THF, or acetonitrile, ACN, as solvents. In THF, the overall two‐electron reduction of the 16 valence electron d8‐[Ir+I(troppPh)2]+(PF6)? proceeds in two well separated slow heterogeneous electron transfer steps according to: d8‐[Ir+I (troppPh)2]++e?→d9‐[Ir0(troppPh)2]+e?→d10‐[Ir?I(troppPh)2]?, [ks1=2.2×10?3 cm/s for d8‐Ir+I/d9‐Ir0 and ks2=2.0×10?3 cm/s for d9‐Ir0/d10‐Ir?I]. In ACN, the two redox waves merge into one “two‐electron” wave [ks1,2=7.76×10?4 cm/s for d8‐Ir+I/d9‐Ir0 and d9‐Ir0/d10‐Ir?I] most likely because the neutral [Ir0(troppPh)2] complex is destabilized. At low temperatures (ca. ? 40 °C) and at high scan rates (ca. 10 V/s), the two‐electon redox process is kinetically resolved. In equilibrium with the tetracoordianted complex [Ir+I(troppPh)2]+ are the pentacoordinated 18 valence [Ir+I(troppPh)2L]+ complexes (L=THF, ACN, Cl?) and their electrochemical behavior was also investigated. They are irreversibly reduced at rather high negative potentials (? 1.8 to ? 2.4 V) according to an ECE mechanism 1) [Ir+I(troppPh)2(L)]+e?→[Ir0(troppPh)2(L)]; 2) [Ir0(troppPh)2(L)]→[Ir(troppPh)2]+L, iii) [Ir0(troppPh)2]+e?→[Ir?I(troppPh)2]?. Since all electroactive species were isolated and structurally characterized, our measurements allow for the first time a detailed insight into some fundamental aspects of the coordination chemistry of iridium complexes in unusually low formal oxidation states.  相似文献   

12.
Binary excess molar volumes, V m E, have been evaluated from density measurements, using a vibrating tube densimeter over the entire composition range for binary liquid mixtures of ionic liquids 1-ethyl-3-methyl-imidazolium diethyleneglycol monomethylethersulphate [EMIM]+[CH3(OCH2CH2)2OSO3] or 1-butyl-3-methyl-imidazolium diethyleneglycol monomethylethersulphate [BMIM]+[CH3(OCH2CH2)2OSO3] or 1-methyl-3-octyl-imidazolium diethyleneglycol monomethylethersulphate [MOIM]+[CH3(OCH2CH2)2OSO3]+methanol and [EMIM]+[CH3(OCH2CH2)2OSO3]+water at 298.15, 303.15 and 313.15 K. The V m E values were found to be negative for all systems studied. The V m E results are explained in terms of intermolecular interactions and packing effects. The experimental data were fitted by the Redlich-Kister polynomial.  相似文献   

13.
As part of a mass spectrometric investigation of the binding properties of sulfonamide anion receptors, an atmospheric pressure chemical ionization mass spectrometric (APCI-MS) method involving direct infusion followed by thermal desorption was employed for identification of anionic supramolecular complexes in dichloromethane (CH2Cl2). Specifically, the dansylamide derivative of tris(2-aminoethyl)amine (tren) (1), the chiral 1,3-benzenesulfonamide derivatives of (1R,2S)-(+)-cis-1-amino-2-indanol (2), and (R)-(+)-bornylamine, (3), were shown to bind halide and nitrate ions in the presence of (n−Bu)4N+X (X = Cl, NO3, Br, I). Solutions of receptors and anions in CH2Cl2 were combined to form the anionic supramolecular complexes, which were subsequently introduced into the mass spectrometer via direct infusion followed by thermal desorption. The anionic supramolecular complexes [M+X], (M=13, X=Cl, NO3, Br, I) were observed in negative mode APCI-MS along with the deprotonated receptors [M−H]. Full ionization energy of the APCI corona pin (4.5 kV) was necessary for obtaining mass spectra with the best signal-to-noise ratios.  相似文献   

14.
3-(Diphenylphosphino)-1,3-diphenyltriazene Ph2P-NPh-N=NPh was synthesized. The reactions of this compound with bis(cycloocta-1,5-diene)nickel, (cod)2Ni, and nickel(I) bis(triphenylphosphino)bis(trimethylsilyl)amide, (Ph3P)2Ni-N(SiMe3)2, afforded the anionic nickel complex [Ph4P]+[Ni(PhNNNPh)3] in 15 and 78% yields, respectively. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 7, pp. 1585–1589, July, 2005.  相似文献   

15.
The single ratio kinetic method is applied to the discrimination and quantification of the thyroid hormone isomers, 3,5,3′-triiodothyronine and 3,3′,5′-triiodothyronine, in the gas phase, based on the kinetics of the competitive unimolecular dissociations of singly charged transition-metal ion-bound trimeric complexes [MII(A)(ref*)2-H]+ (MII = divalent transition-metal ion; A=T3 or rT3; ref* = reference ligand). The trimeric complex ions are generated using electrospray ionization mass spectrometry and the ions undergo collisional activation to realize isomeric discrimination from the branching ratio of the two fragment pathways that form the dimeric complexes [MII(A)(ref*)-H]+ and [MII(ref*)2-H]+. The ratio of the individual branching ratios for the two isomers Riso is found strongly dependent on the references and the metal ions. Various sets are tried by choosing the reference from amino acids, substituted amino acids, and dipeptides in combination with the central metal ion chosen from five transition-metal ions (CoII, CuII, MnII, NiII, and ZnII) for the complexes in this experiment. The results are compared in terms of the isomeric discrimination for the T3/rT3 pair. Calibration curves are constructed by relating the ratio of the branching ratios against the isomeric composition of their mixture to allow rapid quantitative isomer analysis of the sample pair. Furthermore, the instrument-dependence of this method is investigated by comparing the two sets of results, one obtained from a quadrupole ion trap mass spectrometer and the other from a quadrupole time-of-flight mass spectrometer.  相似文献   

16.
Aminomonosaccharides (glucosamine, galactosamine, and mannosamine) in H2O and D2O were ionized by atmospheric pressure chemical ionization (APCI) and their fragmentation patterns were investigated to identify them. All the aminomonosaccharides showed the same fragment ions but their relative ion intensities were different. Major product ions generated in H2O were [M + H]+, [M + H – H2O]+, and [2M + H – 3H2O]+, while in D2O were [MD6 + D]+, [MD6 + D – D2O]+, and [2MD6 + D – D2O – 2HDO]+. At a high fragmentor voltage above 120 V, the relative ion intensities of the major product ions showed different trends according to the aminomonosaccharides. For the use of H2O as solvent and eluent, the order of the ion intensity ratio of [M + H – H2O]+/[2M + H – 3H2O]+ was galactosamine > mannosamine > glucosamine. When using D2O as solvent and eluent, the order of the ion intensity ratios of [MD6 + D – D2O]+/[MD6 + D]+ and [2MD6 + D – D2O – 2HDO]+/[MD6 + D]+ was mannosamine > galactosamine > glucosamine. It was found that glucosamine, galactosamine, and mannosamine could be distinguished by the specific trends of the major product ion ratios in H2O and D2O. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

17.
Reactions of [Pt2(μ-S)2(PPh3)4] with Ph3PbCl, Ph2PbI2, Ph2PbBr2 and Me3PbOAc result in the formation of bright yellow to orange solutions containing the cations [Pt2(μ-S)2(PPh3)4PbR3]+ (R3 = Ph3, Ph2I, Ph2Br, Me3) isolated as PF6 or BPh4 salts. In the case of the Me3Pb and Et3Pb systems, a prolonged reaction time results in formation of the alkylated species [Pt2(μ-S)(μ-SR)(PPh3)4]+ (R = Me, Et). X-ray structure determinations on [Pt2(μ-S)2(PPh3)4PbMe3]PF6 and [Pt2(μ-S)2(PPh3)4PbPh2I]PF6 have been carried out, revealing different coordination modes. In the Me3Pb complex, the (four-coordinate) lead atom binds to a single sulfur atom, while in the Ph2PbI adduct coordination of both sulfurs results in a five-coordinate lead centre. These differences are related to the electron density on the lead centre, and indicate that the interaction of the heterometal centre with the {Pt2S2} metalloligand core can be tuned by variation of the heteroatom substituents. The species [Pt2(μ-S)2(PPh3)4PbR3]+ display differing fragmentation pathways in their ESI mass spectra, following initial loss of PPh3 in all cases; for R = Ph, loss of PbPh2 occurs, yielding [Pt2(μ-S)2(PPh3)3Ph]+, while for R = Me, reductive elimination of ethane gives [Pt2(μ-S)2(PPh3)3PbMe]+, which is followed by loss of CH4.  相似文献   

18.
Reactions of 5-(allyloxymethyl)- and 5-(methallyloxymethyl)-5-ethyl-1,3-dioxanes with methyl diazoacetate catalyzed by Rh2(OAc)4 or Cu(OTf)2 in the presence of [bmim]+Cl, [bmim]+BF4 , and [bmim]+PF6 proceed regioselectively at the C=C bond and lead to the formation of the corresponding cyclopropane-containing 1,3-dioxanes in yields up to 62%.  相似文献   

19.
    
The prominent “1/3” effect observed in the Hall effect plateaus of two-dimensional electron gas (2DEG) systems has been postulated to indicating 1/3 fractional charge quasiparticle excitations arising from electron-electron interactions. Tunneling shot-noise experiments on 2DEF exhibiting fractional quantum Hall effect (FQHE) shows evidence for tunnelling of particles with eand e/3 charges for a constant band mass. A “1/3” effect in the hydrogen molecule is seen in as much as its internuclear distance,d H-H = D + D+, with |D+/D| = 1/3. This is examined in terms of electron-electron interactions involving electron-and hole quasiparticles, (e-)and(h + ), equivalent to those observed in FQHE shot -noise experiments. The(e/m) ratio of the (e) and(h + ) quasiparticles is kept at 1: −3. Instead of a 2DEG, these particles are treated as being in flat Bohr orbits. A treatment in the language of charge-flux tube composites for the hydrogen atom as well as the hydrogen molecule is attempted. Such treatment gives important insights into changes in chemical potential and bond energy on crossing a phase boundary during the atom-bond transition as well as on models for FQHE itself.  相似文献   

20.
The synthesis and properties of [Ph 2 p-TolSi]2, [Ph 2 p-TolSi]2SiPh 2, [Ph 3 Si]2Si(p-Tol)2, [Ph 2 p-TolSi]2Si(p-Tol)2 and (SiPh 3)2SiH2 are described. The silanes are identified using IR,Ra- and29Si-NMR-spectroscopy.
  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