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.
王晨  黄吉玲 《中国化学》2006,24(10):1397-1401
Two new complexes[η~5-C_5H_4CMe_2-(p-fluorophenyl)]TiCl_3(1)and[μ~5-C_5H_4C(cyclo-C_5H_(10))-(p-fluoro-phenyl)]TiCl_3(2)were synthesized and characterized.Their activities and selectivities for trimerization of ethylenewere investigated.The introduction of fluorine atom greatly weakened the arene coordination,but this disadvanta-geous factor can be eliminated by introduction of a bulky substituent,such as cyclo-C_5H_(10),to the bridging carbonlinked to the Cp ring.The combinative effect of the fluorine substitute and the bridging unit can make complex 2 asa highly active and selective catalyst for ethylene trimerization.Its productivity and selectivity for 1-hexene canreach 1024.0 kg·mol~(-1)·h(-1) and 99.3% respectively.  相似文献   

2.
The molecules of (2RS,4SR)‐2‐exo‐(5‐bromo‐2‐thienyl)‐7‐chloro‐2,3,4,5‐tetrahydro‐1H‐1,4‐epoxy‐1‐benzazepine, C14H11BrClNOS, (I), are linked into cyclic centrosymmetric dimers by C—H...π(thienyl) hydrogen bonds. Each such dimer makes rather short Br...Br contacts with two other dimers. In (2RS,4SR)‐2‐exo‐(5‐methyl‐2‐thienyl)‐2,3,4,5‐tetrahydro‐1H‐1,4‐epoxy‐1‐benzazepine, C15H15NOS, (II), a combination of C—H...O and C—H...π(thienyl) hydrogen bonds links the molecules into chains of rings. A more complex chain of rings is formed in (2RS,4SR)‐7‐chloro‐2‐exo‐(5‐methyl‐2‐thienyl)‐2,3,4,5‐tetrahydro‐1H‐1,4‐epoxy‐1‐benzazepine, C15H14ClNOS, (III), built from a combination of two independent C—H...O hydrogen bonds, one C—H...π(arene) hydrogen bond and one C—H...π(thienyl) hydrogen bond.  相似文献   

3.
(Z)‐3‐(1H‐Indol‐3‐yl)‐2‐(3‐thienyl)­acrylo­nitrile, C15H10N2S, (I), and (Z)‐3‐[1‐(4‐tert‐butyl­benzyl)‐1H‐indol‐3‐yl]‐2‐(3‐thienyl)­acrylo­nitrile, C26H24N2S, (II), were prepared by base‐catalyzed reactions of the corresponding indole‐3‐carbox­aldehyde with thio­phene‐3‐aceto­nitrile. 1H/13C NMR spectral data and X‐ray crystal structures of compounds (I) and (II) are presented. The olefinic bond connecting the indole and thio­phene moieties has Z geometry in both cases, and the mol­ecules crystallize in space groups P21/c and C2/c for (I) and (II), respectively. Slight thienyl ring‐flip disorder (ca 5.6%) was observed and modeled for (I).  相似文献   

4.
The unsymmetrical bis (arylimino)pyridines, 2‐[CMeN{2,6‐{(4‐FC6H4)2CH}2–4‐t‐BuC6H2}]‐6‐(CMeNAr)C5H3N (Ar = 2,6‐Me2C6H3 L1 , 2,6‐Et2C6H3 L2 , 2,6‐i‐Pr2C6H3 L3 , 2,4,6‐Me3C6H2 L4 , 2,6‐Et2–4‐MeC6H2 L5 ), each containing one N‐aryl group bedecked with ortho‐substituted fluorobenzhydryl groups, have been employed in the preparation of the corresponding five‐coordinate cobalt (II) chelates, LCoCl2 ( Co1 – Co5 ); the symmetrical comparator [2,6‐{CMeN(2,6‐(4‐FC6H4)2CH)2–4‐t‐BuC6H2}2C5H3N]CoCl2 (Co6) is also reported. All cobaltous complexes are paramagnetic and have been characterized by 1H/19F NMR spectroscopy, FT‐IR spectroscopy and elemental analysis. The molecular structures of Co3 and Co6 highlight the different degrees of steric protection given to the metal center by the particular N‐aryl group combination. Depending on the aluminoxane co‐catalyst employed to activate the cobalt precatalyst, distinct variations in thermal stability and activity of the catalyst towards ethylene polymerization were exhibited. In particular with MAO, the resultant catalysts reached their optimal performance at 70 °C delivering high activities of up to 10.1 × 106 g PE (mol of Co)?1 h?1 with Co1  >  Co4  >  Co2  >  Co5  >  Co3 >>  Co6 . On the other hand, using MMAO, the catalysts operate most effectively at 30 °C but are by comparison less productive. In general, the polyethylenes were highly linear, narrowly disperse and displayed a wide range of molecular weights [Mw range: 18.5–58.7 kg mol?1 (MAO); 206.1–352.5 kg mol?1 (MMAO)].  相似文献   

5.
Sigma‐ versus Pi‐Coordination in Bis‐indenyl‐ and Bis‐2‐methallyl Imido Complexes of Hexavalent Molybdenum and Tungsten: DF‐Calculations and Crystal Structure Analysis Bis‐indenyl and bis‐2‐methallyl imido complexes [(C9H7)2M(NR)2] (M = Mo, W; R = tert‐butyl, mesityl) 1 — 4 and [(H3C‐C3H4)2M(NtBu)2] (M = Mo, W) 6 , 7 have been prepared starting from [Mo(NtBu)2Cl2] or [M(NR)2Cl2L2] (M = W, R = tBu, L = py; M = Mo, W, R = Mes, L2 = dme) and indenyl lithium or 2‐methallyl magnesium bromide, respectively. According to spectroscopic data and the crystal structure of 4 there are two different coordination modes of the indenyl ligands, [(η3‐C9H7)M(NR)21‐C9H7)], in solution as well as in the solid state. These compounds show fluxional rearrangements in solution, namely σ, π‐exchange of η1‐ and η3‐coordinated ligands. Similar behavior has been observed for the 2‐methallyl complexes 6 and 7 in solution. In agreement with experimental observations, DF calculations on models of 6 strongly suggest a (σ+π)‐coordination mode of the η3‐coordinated ligand.  相似文献   

6.
Tandem catalysis offers a promising synthetic route to the production of linear low‐density polyethylene. This article reports the use of homogeneous tandem catalytic systems for the synthesis of ethylene/1‐hexene copolymers from ethylene stock as the sole monomer. The reported catalytic systems employ the tandem action between an ethylene trimerization catalyst, (η5‐C5H4CMe2C6H5)TiCl3 ( 1 )/modified methylaluminoxane (MMAO), and a copolymerization metallocene catalyst, [(η5‐C5Me4)SiMe2(tBuN)]TiCl2 ( 2 )/MMAO or rac‐Me2Si(2‐MeBenz[e]Ind)2ZrCl2 ( 3 )/MMAO. During the reaction, 1 /MMAO in situ generates 1‐hexene with high activity and high selectivity, and simultaneously 2 /MMAO or 3 /MMAO copolymerizes ethylene with the produced 1‐hexene to generate butyl‐branched polyethylene. We have demonstrated that, by the simple manipulation of the catalyst molar ratio and polymerization conditions, a series of branched polyethylenes with melting temperatures of 60–128 °C, crystallinities of 5.4–53%, and hexene percentages of 0.3–14.2 can be efficiently produced. © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 4327–4336, 2004  相似文献   

7.
The P,O‐chelated shell higher olefin process (SHOP) type nickel complexes are practical homogeneous catalysts for the industrial preparation of linear low‐carbon α‐olefins from ethylene. We describes that a facile synthetic route enables the modulation of steric hindrance and electronic nature of SHOP‐type nickel complexes. A series of sterically bulky SHOP‐type nickel complexes with variable electronic nature, {[4‐R‐C6H4C(O) = C‐PArPh]NiPh (PPh3); Ar = 2‐[2′,6′‐(OMe)2C6H3]C6H4; R = H ( Ni1 ); R = OMe ( Ni2 ); R = CF3 ( Ni3 )}, were prepared and used as single component catalysts toward ethylene polymerization without using any phosphine scavenger. These nickel catalysts exhibit high thermal stability during ethylene polymerization and result in highly crystalline linear α‐olefinic solid polymer. The catalytic performance of the SHOP‐type nickel complexes was significantly improved by introducing a bulky ortho‐biphenyl group on the phosphorous atom or an electron‐withdrawing trifluoromethyl on the backbone of the ligand, indicating steric and electronic effects play critical roles in SHOP‐type nickel complexes catalyzed ethylene polymerization.  相似文献   

8.
Linear low‐density polyethylene (LLDPE) can be prepared by addition of ethylene to a mixture of two catalysts. In this “tandem catalysis” scheme one catalyst dimerizes or oligomerizes ethylene to α‐olefins while the second site incorporates these α‐olefins into a growing polyethylene chain. A variety of classical catalyst combinations are available for this purpose. Better control over the polymerization process, and therefore product properties, is attained by the use of homogenous “single site” catalysts. The best‐behaved tandem processes take advantage of well‐defined catalysts that require stoichiometric quantities of activators. One such system employs [(C6H5)2PC6H4C(OB(C6F5)3)O‐κ2P,O]Ni(η3‐CH2CMeCH2) and {[(η5‐C5Me4)SiMe2(η1‐NCMe3)]TiMe}{MeB(C6F5)3}. The nickel sites are responsible for converting ethylene to 1‐butene or mixtures of 1‐butene with 1‐hexene. These olefins are copolymerized with ethylene at the titanium sites. It is possible to obtain a linear correlation between the branching content in the polymer product and the Ni/Ti ratio. The effect of ligand substitution at nickel has also been investigated. When the benzyl derivative [(C6H5)2PC6H4C(O‐B(C6F5)3)O‐κ2P,O]Ni(η3‐CH2C6H5) is used instead of the methallyl counterpart [(C6H5)2PC6H4C(OB(C6F5)3)O‐κ2P,O]Ni(η3‐CH2CMeCH2), one obtains, at a constant Ni/Ti ratio, considerably more branching in the final polymer structure. These results are rationalized in terms of a more efficient initiation when the more labile benzyl ligand is used.  相似文献   

9.
Five examples of unsymmetrical 1,2‐bis (arylimino) acenaphthene ( L1 – L5 ), each containing one N‐2,4‐bis (dibenzocycloheptyl)‐6‐methylphenyl group and one sterically and electronically variable N‐aryl group, have been used to prepare the N,N′‐nickel (II) halide complexes, [1‐[2,4‐{(C15H13}2–6‐MeC6H2N]‐2‐(ArN)C2C10H6]NiX2 (X = Br: Ar = 2,6‐Me2C6H3 Ni1 , 2,6‐Et2C6H3 Ni2 , 2,6‐i‐Pr2C6H3 Ni3 , 2,4,6‐Me3C6H2 Ni4 , 2,6‐Et2–4‐MeC6H2 Ni5 ) and (X = Cl: Ar = 2,6‐Me2C6H3 Ni6 , 2,6‐Et2C6H3 Ni7 , 2,6‐i‐Pr2C6H3 Ni8 , 2,4,6‐Me3C6H2 Ni9 , 2,6‐Et2–4‐MeC6H2 Ni10 ), in high yield. The molecular structures Ni3 and Ni7 highlight the extensive steric protection imparted by the ortho‐dibenzocycloheptyl group and the distorted tetrahedral geometry conferred to the nickel center. On activation with either Et2AlCl or MAO, Ni1 – Ni10 exhibited very high activities for ethylene polymerization with the least bulky Ni1 the most active (up to 1.06  ×  107 g PE mol?1(Ni) h?1 with MAO). Notably, these sterically bulky catalysts have a propensity towards generating very high molecular weight polyethylene with moderate levels of branching and narrow dispersities with the most hindered Ni3 and Ni8 affording ultra‐high molecular weight material (up to 1.5  ×  106 g mol?1). Indeed, both the activity and molecular weights of the resulting polyethylene are among the highest to be reported for this class of unsymmetrical 1,2‐bis (imino)acenaphthene‐nickel catalyst.  相似文献   

10.
Compounds (Bu4N)[2-B10H9{NH=C(NHR)CH3}] are obtained by reactions of the tetrabutylammonium salt of the [2-B10H9(N≡CCH3)] anion with aliphatic and aromatic primary amines RNH2 (R = n-C3H7, n-C4H9, cyclo-C5H9, C6H5, cyclo-C6H11, n-C6H13, C7H7, C8H8NH2, C6H4NO2, and C18H37) and identified by IR, ESI/MS, and NMR (1H, 11B, and 13C) spectroscopy. The structures of the amidine-type derivatives [2-B10H9{Z-NH=C(NH-cyclo-C5H9)CH3}] and [2-B10H9{Z-NH=C(NH-C7H7)CH3}] are determined by X-ray diffraction.  相似文献   

11.
The asymmetric unit of the racemic form of the title compound, C12H15NOS, contains four crystallographically independent molecules. The olefinic bond connecting the 2‐thienyl and 1‐azabicyclo[2.2.2]octan‐3‐ol moieties has Z geometry. Strong hydrogen bonding occurs in a directed co‐operative O—H...O—H...O—H...O—H R44(8) pattern that influences the conformation of the molecules. Co‐operative C—H...π interactions between thienyl rings are also present. The average dihedral angle between adjacent thienyl rings is 87.09 (4)°.  相似文献   

12.
Ethylene/styrene copolymerizations using Cp′TiCl2(O‐2,6‐iPr2C6H3) [Cp′ = Cp* (C5Me5, 1 ), 1,2,4‐Me3C5H2 ( 2 ), tert‐BuC5H4 ( 3 )]‐MAO catalyst systems were explored under various conditions. Complexes 2 and 3 exhibited both high catalytic activities (activity: 504–6810 kg‐polymer/mol‐Ti h) and efficient styrene incorporations at 25, 40°C (ethylene 6 atm), affording relatively high molecular weight poly (ethylene‐co‐styrene)s with unimodal molecular weight distributions as well as with uniform styrene distributions (Mw = 6.12–13.6 × 104, Mw/Mn = 1.50–1.71, styrene 31.7–51.9 mol %). By‐productions of syndiotactic polystyrene (SPS) were observed, when the copolymerizations by 1 – 3 ‐MAO catalyst systems were performed at 55, 70 °C (ethylene 6 atm, SPS 9.0–68.9 wt %); the ratios of the copolymer/SPS were affected by the polymerization temperature, the [styrene]/[ethylene] feed molar ratios in the reaction mixture, and by both the cyclopentadienyl fragment (Cp′) and anionic ancillary donor ligand (L) in Cp′TiCl2(L) (L = Cl, O‐2,6‐iPr2C6H3 or N=CtBu2) employed. Co‐presence of the catalytically‐active species for both the copolymerization and the homopolymerization was thus suggested even in the presence of ethylene; the ratios were influenced by various factors (catalyst precursors, temperature, styrene/ethylene feed molar ratio, etc.). © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 4162–4174, 2008  相似文献   

13.
A series of ethylene‐bridged C1‐symmetric ansa‐(3‐R‐indenyl)(fluorenyl) zirconocene complexes ( 1 , 2 , 3 , 4 , 5 , 6 , 7 , 8 , 9 ) incorporating a pendant arene substituent on the 3‐position of indenyl ring have been synthesized. The structure of complex 4 was further confirmed by X‐ray diffraction analysis. When activated with methylaluminoxane, four sterically less encumbered complexes 1 , 2 , 4 and 5 could catalyze the dimerization of propylene in toluene at 100°C to afford 2‐methyl‐1‐pentene with high selectivities up to 95.7–98.4% and moderate activities of 2.00 × 104 to 7.89 × 104 g (mol‐Zr?h)?1. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

14.
Olefin polymerizations catalyzed by Cp′TiCl2(O‐2,6‐iPr2C6H3) ( 1 – 5 ; Cp′ = cyclopentadienyl group), RuCl2(ethylene)(pybox) { 7 ; pybox = 2,6‐bis[(4S)‐4‐isopropyl‐2‐oxazolin‐2‐yl]pyridine}, and FeCl2(pybox) ( 8 ) were investigated in the presence of a cocatalyst. The Cp*TiCl2(O‐2,6‐iPr2C6H3) ( 5 )–methylaluminoxane (MAO) catalyst exhibited remarkable catalytic activity for both ethylene and 1‐hexene polymerizations, and the effect of the substituents on the cyclopentadienyl group was an important factor for the catalytic activity. A high level of 1‐hexene incorporation and a lower rE · rH value with 5 than with [Me2Si(C5Me4)(NtBu)]TiCl2 ( 6 ) were obtained, despite the rather wide bond angle of Cp Ti O (120.5°) of 5 compared with the bond angle of Cp Ti N of 6 (107.6°). The 7 –MAO catalyst exhibited moderate catalytic activity for ethylene homopolymerization and ethylene/1‐hexene copolymerization, and the resultant copolymer incorporated 1‐hexene. The 8 –MAO catalyst also exhibited activity for ethylene polymerization, and an attempted ethylene/1‐hexene copolymerization gave linear polyethylene. The efficient polymerization of a norbornene macromonomer bearing a ring‐opened poly(norbornene) substituent was accomplished by ringopening metathesis polymerization with the well‐defined Mo(CHCMe2Ph)(N‐2,6‐iPr2C6H3)[OCMe(CF3)2]2 ( 10 ). The key step for the macromonomer synthesis was the exclusive end‐capping of the ring‐opened poly(norbornene) with p‐Me3SiOC6H4CHO, and the use of 10 was effective for this polymerization proceeding with complete conversion. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 4613–4626, 2000  相似文献   

15.
A series of imino‐indolate half‐titanocene chlorides, Cp′Ti(L)Cl2 ( C1 – C7 : Cp′ = C5H5, MeC5H4, C5Me5, L = imino‐indolate ligand), were synthesized by the reaction of Cp′TiCl3 with sodium imino‐indolates. All complexes were characterized by elemental analysis, 1H and 13C NMR spectroscopy. Moreover, the molecular structures of two representative complexes C4 and C6 were confirmed by single crystal X‐ray diffraction analysis. On activation with methylaluminoxane (MAO), these complexes showed good catalytic activities for ethylene polymerization (up to 7.68 × 106 g/mol(Ti)·h) and ethylene/1‐hexene copolymerization (up to 8.32 × 106 g/mol(Ti)·h), producing polyolefins with high molecular weights (for polyethylene up to 1808 kg/mol, and for poly(ethylen‐co‐1‐hexene) up to 3290 kg/mol). Half‐titanocenes containing ligands with alkyl substituents showed higher catalytic activities, whereas the half‐titanocenes bearing methyl substituents on the cyclopentadienyl groups showed lower productivities, but produced polymers with higher molecular weights. Moreover, the copolymerization of ethylene and methyl 10‐undecenoate was demonstrated using the C1 /MAO catalytic system. The functionalized polyolefins obtained contained about 1 mol % of methyl 10‐undecenoate units and were fully characterized by several techniques such as FT‐IR, 1H NMR, 13C NMR, DSC, TGA and GPC analyses. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 47: 357–372, 2009  相似文献   

16.
The crystal and molecular structures of bis(η5‐2,4,7‐tri­methyl­indenyl)­cobalt(II), [Co(C12H13)2], (I), and rac‐2,2′,4,4′,7,7′‐hexamethyl‐1,1′‐biindene, C24H26, (II), are reported. In the crystal structure of (I), the Co atom lies on an inversion centre and the structure represents the first example of a bis(indenyl)cobalt complex exhibiting an eclipsed indenyl conformation. The (1R,1′R) and (1S,1′S) enantiomers of the three possible stereoisomers of (II), which form as by‐products in the synthesis of (I), cocrystallize in the monoclinic space group P21/c. In the unit cell of (II), alternating (1R,1′R) and (1S,1′S) enantiomers pack in non‐bonded rows along the a axis, with the planes of the indenyl groups parallel to each other and separated by 3.62 and 3.69 Å.  相似文献   

17.
N‐sulfinylacylamides R‐C(=O)‐N=S=O react with (CF3)2BNMe2 ( 1 ) to form, by [2+4] cycloaddition, six‐membered rings cyclo‐(CF3)2B‐NMe2‐S(=O)‐N=C(R)‐O for R = Me ( 2 ), t‐Bu ( 3 ), C6H5 ( 4 ), and p‐CH3C6H4 ( 5 ) while N‐sulfinylcarbamic acid esters R‐O‐C(=O)‐N=S=O react with 1 to yield mixtures of six‐membered (cyclo‐(CF3)2B‐NMe2‐S(=O)‐N=C(OR)‐O) and four‐membered rings (cyclo‐(CF3)2B‐NMe2‐S(=O)‐N(C=O)OR) for R = Me ( 6 and 9 ), Et ( 7 and 10 ), and C6H5 ( 8 and 11 ). The structure of 5 has been determined by X‐ray diffraction.  相似文献   

18.
New Ti and Zr complexes that bear imine–phenoxy chelate ligands, [{2,4‐di‐tBu‐6‐(RCH=N)‐C6H4O}2MCl2] ( 1 : M=Ti, R=Ph; 2 : M=Ti, R=C6F5; 3 : M=Zr, R=Ph; 4 : M=Zr, R=C6F5), were synthesized and investigated as precatalysts for ethylene polymerization. 1H NMR spectroscopy suggests that these complexes exist as mixtures of structural isomers. X‐ray crystallographic analysis of the adduct 1 ?HCl reveals that it exists as a zwitterionic complex in which H and Cl are situated in close proximity to one of the imine nitrogen atoms and the central metal, respectively. The X‐ray molecular structure also indicates that one imine phenoxy group with the syn C?N configuration functions as a bidentate ligand, whereas the other, of the anti C?N form, acts as a monodentate phenoxy ligand. Although Zr complexes 3 and 4 with methylaluminoxane (MAO) or [Ph3C]+[B(C6F5)4]?/AliBu3 displayed moderate activity, the Ti congeners 1 and 2 , in association with an appropriate activator, catalyzed ethylene polymerization with high efficiency. Upon activation with MAO at 25 °C, 2 displayed a very high activity of 19900 (kg PE) (mol Ti)?1 h?1, which is comparable to that for [Cp2TiCl2] and [Cp2ZrCl2], although increasing the polymerization temperature did result in a marked decrease in activity. Complex 2 contains a C6F5 group on the imine nitrogen atom and mediated nonliving‐type polymerization, unlike the corresponding salicylaldimine‐type complex. Conversely, with [Ph3C]+[B(C6F5)4]?/AliBu3 activation, 1 exhibited enhanced activity as the temperature was increased (25–75 °C) and maintained very high activity for 60 min at 75 °C (18740 (kg PE) (mol Ti)?1 h?1). 1H NMR spectroscopic studies of the reaction suggest that this thermally robust catalyst system generates an amine–phenoxy complex as the catalytically active species. The combinations 1 /[Ph3C]+[B(C6F5)4]?/AliBu3 and 2 /MAO also worked as high‐activity catalysts for the copolymerization of ethylene and propylene.  相似文献   

19.
The synthesis and characterization of the novel zirconium (IV) tris(pyrazolyl)borate compound {TpMs*}ZrCl3 ( 1 ) (TpMs* = hydridobis(3‐mesitylpyrazol‐1‐yl)(5‐mesitylpyrazol‐1‐yl)), as well as its performance in polymerizing ethylene are described. The reaction of ZrCl4 with 1 equivalent of TlTpMs* in toluene at room temperature affords 1 as a white solid in 62% yield. Compound 1 in the presence of MAO showed remarkable productivity using a low Al : Zr molar ratio (6.79×104 kg of PE/(mol Zr·h·[C2H4]); toluene, 60°C, Al/Zr = 100). Under identical polymerization conditions, compound 1 and Cp2ZrCl2 showed comparable productivities. Compound 1 displayed similar productivities at temperatures in the range of 0–75°C and noticeable productivity at 105°C. The viscosity‐average molecular weight of the polyethylenes depends on the Al : Zr molar ratio and polymerization temperature and varied between 1.09 and 8.98×105 g·mol–1.  相似文献   

20.
Defects were created on the surface of highly oriented pyrolytic graphite (HOPG) by sputtering with an Ar+ ion beam, then characterized using X‐ray photoelectron spectroscopy (XPS) and time‐of‐flight secondary ion mass spectrometry (ToF‐SIMS) at 500°C. In the XPS C1s spectrum of the sputtered HOPG, a sp3 carbon peak appeared at 285.3 eV, representing surface defects. In addition, 2 sets of peaks, the Cx and CxH ion series (where x = 1, 2, 3...), were identified in the ToF‐SIMS negative ion spectrum. In the positive ion spectrum, a series of CxH2+• ions indicating defects was observed. Annealing of the sputtered samples under Ar was conducted at different temperatures. The XPS and ToF‐SIMS spectra of the sputtered HOPG after 800°C annealing were observed to be similar to the spectra of the fresh HOPG. The sp3 carbon peak had disappeared from the C1s spectrum, and the normalized intensities of the CxH and CxH2+• ions had decreased. These results indicate that defects created by sputtering on the surface of HOPG can be repaired by high‐temperature annealing.  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