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 62 毫秒
1.
Two‐dimensional (2D) AA′n?1MnX3n+1 type halide perovskites incorporating straight‐chain symmetric diammonium cations define a new type of structure, but their optoelectronic properties are largely unexplored. Reported here is the synthesis of a centimeter‐sized AA′n?1MnX3n+1 type perovskite, BDAPbI4 (BDA=NH3C4H8NH3), single crystal and its charge‐transport properties under X‐ray excitation. The crystal shows a staggered configuration of the [PbI6]4? layers, a band gap of 2.37 eV, and a low trap density of 3.1×109 cm?3. The single‐crystal X‐ray detector exhibits an excellent sensitivity of 242 μC Gyair?1 cm?2 under the 10 V bias (0.31 V μm?1), a detection limit as low as 430 nGyair s?1, ultrastable response current, a stable baseline with the lowest dark current drift of 6.06×10?9 nA cm?1 s?1 V?1, and rapid response time of τrise=7.3 ms and τfall=22.5 ms. These crystals are promising candidates for the next generation of optoelectronic devices.  相似文献   

2.
Inorganic Bi-based perovskites have shown great potential in X-ray detection for their large absorption to X-rays, diverse low-dimensional structures, and eco-friendliness without toxic metals. However, they suffer from poor carrier transport properties compared to Pb-based perovskites. Here, we propose a mixed-halogen strategy to tune the structural dimensions and optoelectronic properties of Cs3Bi2I9−nBrn (0≤n≤9). Ten centimeter-sized single crystals are successfully grown by the Bridgman technique. Upon doping bromine to zero-dimensional Cs3Bi2I9, the crystal transforms into a two-dimensional structure as the bromine content reaches Cs3Bi2I8Br. Correspondingly, the optoelectronic properties are adjusted. Among these crystals, Cs3Bi2I8Br exhibits negligible ion migration, moderate resistivity, and the best carrier transport capability. The sensitivities in 100 keV hard X-ray detection are 1.33×104 and 1.74×104 μC Gyair−1 cm−2 at room temperature and 75 °C, respectively, which are the highest among all reported bismuth perovskites. Moreover, the lowest detection limit of 28.6 nGyair s−1 and ultralow dark current drift of 9.12×10−9 nA cm−1 s−1 V−1 are obtained owing to the high ionic activation energy. Our work demonstrates that Br incorporation is an effective strategy to enhance the X-ray detection performance by tuning the dimensional and optoelectronic properties.  相似文献   

3.
The kinetics of the reactions of ground state oxygen atoms with 1-pentene, 1-hexene, cis-2-pentene, and trans-2-pentene was investigated in the temperature range 200 to 370 K. In this range the temperature dependences of the rate constants can be represented by k = A′ Tn exp(− E′a/RT) with A′ = (1.0 ± 0.6) · 10−14 cm3 s−1, n = 1.13 ± 0.02, E′a = 0.54 ± 0.05 kJ mol−1 for 1-pentene: A′ = (1.3 ± 1.2) · 10−14 cm3 s−1, n = 1.04 ± 0.08, E′a = 0.2 ± 0.4 kJ mol−1 for 1-hexene; A′ = (0.6 ± 0.6) · 10−14 cm3 s−1, n = 1.12 ± 0.05, E′a = − 3.8 ± 0.8 kJ mol−1 for cis-2-pentene; and A′ = (0.6 ± 0.8) · 10−14 cm3 s−1, n = 1.14 ± 0.06, E′a = − 4.3 ± 0.5 kJ mol−1 for trans-2-pentene. The atoms were generated by the H2-laser photolysis of NO and detected by time resolved chemiluminescence in the presence of NO. The concentrations of the O(3P) atoms were kept so low that secondary reactions with products are unimportant. © 1997 John Wiley & Sons, Inc.  相似文献   

4.
Although the development of single-molecule magnets (SMMs) is rapid, there are only two families of high energy barrier (Ueff) dysprosium(III) SMMs known so far: the cyclopentadienyl (Cp) family with a sandwich structure and the pentagonal-bipyramidal (PB) family with D5h symmetry. These high-barrier SMMs, which usually possess Ueff>500 cm−1 allow the separate study of the four magnetic relaxation paths, namely, direct, quantum tunnelling, Raman and Orbach processes, in detail. Whereas the first family is chemically more challenging to modify the Cp rings, it is shown herein that the latter family, with the common formulae [DyX1X2(Leq)5]+, such as X1/X2=OCMe3, OSiMe3, OPh, Cl or Br; Leq=THF/pyridine/4-methylpyridine, can be readily fine-tuned with a range of axial and equatorial ligands by simple substitution reactions. This allows unambiguous confirmation that the Ueff mainly depends on the identity of X1 and X2, rather than on Leq. More importantly, the fitted parameters are barrier dependent. If X1 is an O donor and X2 is a halide, 500<Ueff<600 cm−1, log τ0avg (s)=−10.66, log Cavg (s−1 Kn)= −5.05, navg=4.1 and TH=9 K (in which τ0 is the pre-exponential factor for the Orbach relaxation process, C and n are parameters used to describe Raman relaxation, and TH is the highest temperature at which magnetic hysteresis is observed). For cases in which both X1 and X2 are O donors, 900<Ueff<1300 cm−1, log τ0avg (s)=−11.63, log Cavg (s−1 Kn)= −6.03, navg=4.1 and 18<TH<25 K. Based on these results, it can be further concluded that Ueff not only has a linear correlation to the axial Dy−X bond lengths, but also to TH for these PB SMMs. This represents the first systematic study of a family of lanthanide SMMs and derives the first magneto-structural correlation in Dy SMMs.  相似文献   

5.
CH3NH2 thermal decomposition is shown to provide a suitable NH2 radical source for spectroscopic and kinetic shock tube studies. Using this precursor, the absorption coefficient of the NH2 radical at a detection wavelength of 16739.90 cm−1 has been determined. In the temperature range 1600–2000K the low‐pressure absorption coefficient is described by the polynominal equation: kNH2=3.953×1010/T 3+7.295×105/T 2−1.549×103/T [atm−1 cm−1] The uncertainty of the determined absorption coefficient is estimated to be ±10%. The rate of the thermal decomposition reaction CH3NH2+M → CH3+NH2+M is determined over the temperature range 1550–1900 K and at pressures near 1.6 atm. The rate coefficient was found to be: k1=2.51×1016 exp(−28430/T) [cm3 mol−1 s−1] The uncertainty of the determined rate coefficients is estimated to be ±20%. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 323–330, 1999  相似文献   

6.
《Chemical physics letters》1986,127(4):347-353
Infrared multiphoton photooxidation of NH2D in NH3 mixtures was observed to produce exclusively HDO, suggesting a single step deuterium separation efficiency of [D2O]/([D20]+[H2O]) ⩾ 50% which is significantly higher than that of the theoretical value, 33%. The results are explained by the large rate differences in the radical scavenging steps, i.e. k(D+O2) = 2.2 × 109M−1 s−1, k(NH2+O2) ⩽ 5 × 106 M−1 s−1 and k(NH2+NH2)=1.6 × 1010 M−1 s−1. With Ti solid powder as a catalyst, we observed that the formation yields of HDO are at least three to four times higher than those without a catalyst.  相似文献   

7.
The chlorine transfer reaction between 3-azabicyclo[3,3,0]octane “AZA” and chloramine was studied over pH 8–13 in order to follow both the amination and halogenation properties of NH2Cl. The results show the existence of two competitive reactions which lead to the simultaneous formation of N-amino- and N-chloro- 3-azabicyclo[3,3,0]octane by bimolecular kinetics. The halogenation reaction is reversible and the chlorine derivative obtained, which is thermolabile and unstable in the pure state, was identified by electrospray mass spectrometry. These phenomena were quantified by a reaction between neutral species according to an apparent SN2-type mechanism for the amination process and a ionic mechanism involving a reaction between chloramine and protonated amine for the halogenation process. Amination occurs only in strongly basic solutions (pH ≥ 13) while chlorination occurs at lower pH's (pH ≤ 8). At intermediate pH's, a mixture of these two compounds is obtained. The relative proportions of the products are a function of intrinsic rate constants, pH and pKa of the reactants. The rate constants and thermodynamic activation parameters are the following: k1 = 45.5 × 10−3 M−1 s−1; ΔH10# = 59.8 kJ mol−1; ΔS10# = − 86.5 J mol−1 K−1 for amination; k2 = 114 × 10−3 M−1 s−1; ΔH20# = 63.9 kJ mol−1; and ΔS20# = − 48.3 J mol−1 K−1 for chlorination. The ability of an interaction corresponding to a specific (NH3Cl+/RR′NH) or general (NH2Cl/RR′NH) acid catalysis has been also discussed. © 1997 John Wiley & Sons, Inc.  相似文献   

8.
The low-dimensional halide perovskites have attracted increasing attention due to their improved moisture stability, reduced defects, and suppressed ions migration in many optoelectronic devices such as solar cells, light-emitting diodes, X-ray detectors, and so on. However, they are still limited by their large band gap and short charge carriers’ diffusion length. Here, we demonstrate that the introduction of metal ions into organic interlayers of two-dimensional (2D) perovskite by cross-linking the copper paddle-wheel cluster-based lead bromide ([Cu(O2C−(CH2)3−NH3)2]PbBr4) perovskite single crystals with coordination bonds can not only significantly reduce the perovskite band gap to 0.96 eV to boost the X-ray induced charge carriers, but can also selectively improve the charge carriers’ transport along the out-of-plane direction and blocking the ions motion paths. The [Cu(O2C−(CH2)3−NH3)2]PbBr4 single-crystal device can reach a record charges/ions collection ratio of 1.69×1018±4.7 % μGyair−1 s, and exhibit a large sensitivity of 1.14×105±7% μC Gyair−1 cm−2 with the lowest detectable dose rate of 56 nGyair s−1 under 120 keV X-rays irradiation. In addition, [Cu(O2C−(CH2)3−NH3)2]PbBr4 single-crystal detector exposed to the air without any encapsulation shows excellent X-ray imaging capability with long-term operational stability without any attenuation of 120 days.  相似文献   

9.
Tin organic–inorganic halide perovskites (tin OIHPs) possess a desirable band gap and their power conversion efficiency (PCE) has reached 14 %. A commonly held view is that the organic cations in tin OIHPs would have little impact on the optoelectronic properties. Herein, we show that the defective organic cations with randomly dynamic characteristics can have marked effect on optoelectronic properties of the tin OIHPs. Hydrogen vacancies originated from the proton dissociation from FA [HC(NH2)2] in FASnI3 can induce deep transition levels in the band gap but yield relatively small nonradiative recombination coefficients of 10−15 cm3 s−1, whereas those from MA (CH3NH3) in MASnI3 can yield much larger nonradiative recombination coefficients of 10−11 cm3 s−1. Additional insight into the “defect tolerance” is gained by disentangling the correlations between dynamic rotation of organic cations and charge-carrier dynamics.  相似文献   

10.
Perovskite single crystals and polycrystalline films have complementary merits and deficiencies in X-ray detection and imaging. Herein, we report preparation of dense and smooth perovskite microcrystalline films with both merits of single crystals and polycrystalline films through polycrystal-induced growth and hot-pressing treatment (HPT). Utilizing polycrystalline films as seeds, multi-inch-sized microcrystalline films can be in situ grown on diverse substrates with maximum grain size reaching 100 μm, which endows the microcrystalline films with comparable carrier mobility-lifetime (μτ) product as single crystals. As a result, self-powered X-ray detectors with impressive sensitivity of 6.1×104 μC Gyair−1 cm−2 and low detection limit of 1.5 nGyair s−1 are achieved, leading to high-contrast X-ray imaging at an ultra-low dose rate of 67 nGyair s−1. Combining with the fast response speed (186 μs), this work may contribute to the development of perovskite-based low-dose X-ray imaging.  相似文献   

11.
Perylene diimides (PDIs) and their derivatives are excellent semiconductors, while conjugated polymers based on PDIs have limited applications because of their low electron mobility (μe) derived from low molecular weight. The reported maximum number‐average molecular weight (Mn) of related polymers is only 21 kDa because PDIs have very poor solubility due to strong π–π stacking of their big planar conjugated cores. Herein, it is found that suitable semi‐perfluoroalkyl groups could enhance the solubility of PDIs significantly, and a series of semi‐perfluoroalkyl modified conjugated polymers with high molecular weight and electron mobility were synthesized. The maximum Mn reaches 94.8 kDa [P(4CF8CH‐PDI‐T2)HW]. In their space‐charge‐limited current (SCLC) devices, all polymers exhibit typical characters of electron transporting semiconductors, and the highest μe is up to 8.40 × 10−3 cm2 V−1 s−1 [P(4CF8CH‐PDI‐T2)HW], which is similar as that of widely used electron transporting semiconductor PC61BM (6.41 × 10−3 cm2 V−1 s−1). © 2017 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2018 , 56, 116–124  相似文献   

12.
A novel sulfonated diamine, 4,4′‐bis(4‐amino‐3‐trifluoromethylphenoxy) biphenyl 3,3′‐disulfonic acid (F‐BAPBDS), was successfully synthesized by nucleophilic aromatic substitution of 4,4′‐dihydroxybiphenyl with 2‐chloro‐5‐nitrobenzotrifluoride, followed by reduction and sulfonation. A series of sulfonated polyimides of high molecular weight (SPI‐x, x represents the molar percentage of the sulfonated monomer) were prepared by copolymerization of 1,4,5,8‐naphathlenetetracarboxylic dianhydride (NTDA) with F‐BAPBDS and nonsulfonated diamine. Flexible and tough membranes of high mechanical strength were obtained by solution casting and the electrolyte properties of the polymers were intensively investigated. The copolymer membranes exhibited excellent oxidative stability due to the introducing of the CF3 groups. The SPI membranes displayed desirable proton conductivity (0.52×10−1–0.97×10−1 S·cm−1) and low methanol permeability (less than 2.8×10−7 cm2·s−1). The highest proton conductivity (1.89×10−1 S·cm−1) was obtained for the SPI‐90 membrane at 80°C, with an IEC of 2.12 mequiv/g. This value is higher than that of Nafion 117 (1.7×10−1 S·cm−1). Furthermore, the hydrolytic stability of the obtained SPIs is better than the BDSA and ODADS based SPIs due to the hydrophobic CF3 groups which protect the imide ring from being attacked by water molecules, in spite of its strong electron‐withdrawing behaviors.  相似文献   

13.
The interaction between the ground and excited states of 1,4-bis[2-(5-phenyloxazolyl)]-benzene and bromomethanes such as CBr4, CHBr3 and CH2Br2 were investigated in benzene. Distinct complex formation was not observed either in the ground state or in the excited states. The excited singlet and triplet states are deactivated by these bromomethanes. The triplet yield is increased on the addition of CHBr3 or CH2Br2, whereas it is decreased on the addition of CBr4. The fluorescence quenching rate constants kq at 23 °C were determined to be 1.6 × 1010 M−1 s−1, 3.6 × 108M−1s−1 and 2.4 × 107M−1s−1 for CBr4, CHBr3 and CH2Br2 respectively. The rate constants kST′ of the enhanced intersystem crossing associated with the fluorescence quenching were evaluated from emission—absorption flash photolysis experiments as 3.0 × 108 M−1s−1, 1.9 × 108 M−1s−1 and 5.1 × 107 M−1s−1 for CBr4, CHBr3 and CH2Br2 respectively. kST′ increases with increasing number of bromine atoms contained in the quencher, so that the enhanced intersystem crossing is due to the external heavy-atom effect of the quencher. The apparent triplet yield for the quenching system depends not only on kST′ but also on the rates of the other non-radiative processes. This is the reason why the apparent triplet yield does not necessarily increase on fluorescence quenching by bromomethanes.  相似文献   

14.
Using a relative rate method, rate constants for the gas-phase reactions of 2-methyl-3-buten-2-ol (MBO) with OH radicals, ozone, NO3 radicals, and Cl atoms have been investigated using FTIR. The measured values for MBO at 298±2 K and 740±5 torr total pressure are: kOH=(3.9±1.2)×10−11 cm3 molecule−1 s−1, kO3=(8.6±2.9)×10−18 cm3 molecule−1 s−1, k=(8.6±2.9)×10−15 cm3 molecule−1 s−1, and kCl=(4.7±1.0)×10−10 cm3 molecule−1 s−1. Atmospheric lifetimes have been estimated with respect to the reactions with OH, O3, NO3, and Cl. The atmospheric relevance of this compound as a precursor for acetone is, also, briefly discussed. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet: 30: 589–594, 1998  相似文献   

15.
The reactions between OH radicals and hydrogen halides (HCl, HBr, HI) have been studied between 298 and 460 K by using a discharge flow-electron paramagnetic resonance technique. The rate constants were found to be kHCl(298 K) = (7.9 ± 1.3) × 10−13 cm3 molecule−1 s−1 with a weak positive temperature dependence, kHBr (298-460 K) = (1.04 ± 0.2) × 10−11 cm3 molecule−1 s−1, and kHI(298 K) = (3.0 ± 0.3) × 10−11 cm3 molecule−1 s−1, respectively. The homogeneous nature of these reactions has been experimentally tested.  相似文献   

16.
Multiarm star‐branched polymers based on poly(styrene‐b‐isobutylene) (PS‐PIB) block copolymer arms were synthesized under controlled/living cationic polymerization conditions using the 2‐chloro‐2‐propylbenzene (CCl)/TiCl4/pyridine (Py) initiating system and divinylbenzene (DVB) as gel‐core‐forming comonomer. To optimize the timing of isobutylene (IB) addition to living PS⊕, the kinetics of styrene (St) polymerization at −80°C were measured in both 60 : 40 (v : v) methyl cyclohexane (MCHx) : MeCl and 60 : 40 hexane : MeCl cosolvents. For either cosolvent system, it was found that the polymerizations followed first‐order kinetics with respect to the monomer and the number of actively growing chains remained invariant. The rate of polymerization was slower in MCHx : MeCl (kapp = 2.5 × 10−3 s−1) compared with hexane : MeCl (kapp = 5.6 × 10−3 s−1) ([CCl]o = [TiCl4]/15 = 3.64 × 10−3M; [Py] = 4 × 10−3M; [St]o = 0.35M). Intermolecular alkylation reactions were observed at [St]o = 0.93M but could be suppressed by avoiding very high St conversion and by setting [St]o ≤ 0.35M. For St polymerization, kapp = 1.1 × 10−3 s−1 ([CCl]o = [TiCl4]/15 = 1.82 × 10−3M; [Py] = 4 × 10−3M; [St]o = 0.35M); this was significantly higher than that observed for IB polymerization (kapp = 3.0 × 10−4 s−1; [CCl]o = [Py] = [TiCl4]/15 = 1.86 × 10−3M; [IB]o = 1.0M). Blocking efficiencies were higher in hexane : MeCl compared with MCHx : MeCl cosolvent system. Star formation was faster with PS‐PIB arms compared with PIB homopolymer arms under similar conditions. Using [DVB] = 5.6 × 10−2M = 10 times chain end concentration, 92% of PS‐PIB arms (Mn,PS = 2600 and Mn,PIB = 13,400 g/mol) were linked within 1 h at −80°C with negligible star–star coupling. It was difficult to achieve complete linking of all the arms prior to the onset of star–star coupling. Apparently, the presence of the St block allows the PS‐PIB block copolymer arms to be incorporated into growing star polymers by an additional mechanism, namely, electrophilic aromatic substitution (EAS), which leads to increased rates of star formation and greater tendency toward star–star coupling. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 1629–1641, 1999  相似文献   

17.
The reaction of Cl atoms with a series of C2–C5 unsaturated hydrocarbons has been investigated at atmospheric pressure of 760 Torr over the temperature range 283–323 K in air and N2 diluents. The decay of the hydrocarbons was followed using a gas chromatograph with a flame ionization detector (GC‐FID), and the kinetic constants were determined using a relative rate technique with n‐hexane as a reference compound. The Cl atoms were generated by UV photolysis (λ ≥ 300 nm) of Cl2 molecules. The following absolute rate constants (in units of 10−11 cm3 molecule−1 s−1, with errors representing ±2σ) for the reaction at 295 ± 2 K have been derived from the relative rate constants combined to the value 34.5 × 10−11 cm3 molecule−1 s−1 for the Cl + n‐hexane reaction: ethene (9.3 ± 0.6), propyne (22.1 ± 0.3), propene (27.6 ± 0.6), 1‐butene (35.2 ± 0.7), and 1‐pentene (48.3 ± 0.8). The temperature dependence of the reactions can be expressed as simple Arrhenius expressions (in units of 10−11 cm3 molecule−1 s−1): kethene = (0.39 ± 0.22) × 10−11 exp{(226 ± 42)/T}, kpropyne = (4.1 ± 2.5) × 10−11 exp{(118 ± 45)/T}, kpropene = (1.6 ± 1.8) × 10−11 exp{(203 ± 79)/T}, k1‐butene = (1.1 ± 1.3) × 10−11 exp{(245 ± 90)/T}, and k1‐pentene = (4.0 ± 2.2) × 10−11 exp{(423 ± 68)/T}. The applicability of our results to tropospheric chemistry is discussed. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 478–484, 2000  相似文献   

18.
Rate coefficients have been measured for the reactions of Cl atoms with methanol (k1) and acetaldehyde (k2) using both absolute (laser photolysis with resonance fluorescence) and relative rate methods at 295 ± 2 K. The measured rate coefficients were (units of 10−11 cm3 molecule−1 s−1): absolute method, k1 = (5.1 ± 0.4), k2 = (7.3 ± 0.7); relative method k1 = (5.6 ± 0.6), k2 = (8.4 ± 1.0). Based on a critical evaluation of the literature data, the following rate coefficients are recommended: k1 = (5.4 ± 0.9) × 10−11 and k2 = (7.8 ± 1.3) × 10−11 cm3 molecule−1 s−1 (95% confidence limits). The results significantly improve the confidence in the database for reactions of Cl atoms with these oxygenated organics. Rate coefficients were also measured for the reactions of Cl2 with CH2OH, k5 = (2.9 ± 0.6) × 10−11 and CH3CO, k6 = (4.3 ± 1.5) × 10−11 cm3 molecule−1 s−1, by observing the regeneration of Cl atoms in the absence of O2. Based on these results and those from a previous relative rate study, the rate coefficient for CH3CO + O2 at the high pressure limit is estimated to be (5.7 ± 1.9) × 10−12 cm3 molecule−1 s−1. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 776–784, 1999  相似文献   

19.
NH2 radicals were generated by laser pulsed photolysis of O3 in the presence of NH3 and their subsequent kinetics were monitored by LIF. The room-temperature rate constant for O(1D)+NH3 was measured to be (3.3±0.1)×10−1u cm−3 molecule−1 s−1 by studying NH2 accumulation profiles. The nascent energy distribution over NH2 vibrational levels was studied and the fraction of NH2 radicals generated in the ground vibrational state was found to be 0.42±0.08. Acceleration of the reaction NH2+O3 caused by vibrational excitation of NH2 radicals was observed - the rate constants were (1.5±0.4)×10−12 and (1.5±0.3)×10−13 cm−3 molecule−1 s−1, for vibrationally excited and non-excited NH2, respectively. The reactive channel for removal of vibrationally excited NH2 by ozone appears to be dominant.  相似文献   

20.
A combination of microvolumetry, the rotating sector method, ESR, 1H NMR, and IR allowed to establish a detailed mechanism of liquid‐phase oxidation of vinyl compounds X1CH=CHX2 and X1CH=CH–CH=CHX2 (X1 and X2—a polar substitute: С6Н5–, CO–, СOO–) initiated by azobisisobutyronitrile. A distinctive feature of the mechanism is the fact that the oxidation chain is carried out by a low‐molecular hydroperoxide radical joining the π‐bond. For nine compounds in the temperature range of 303–353 K, relative chain propagation and termination rate constants were measured (k 2k 3−0.5). Absolute values of k 2 were obtained for diphenylethylene (110 L·mol−1·s−1), ethyl ether of trans‐phenyl‐pentadiene acid (13 L·mol−1·s−1), and methyl ether of trans‐phenyl‐pentadiene acid (14.2 L·mol−1·s−1) at T = 323 K. For the same conditions, 10−8k 3 were calculated for diphenylethylene (0.87 L·mol−1·s−1) and methyl ether of trans‐phenyl‐pentadiene acid (1.21 L·mol−1·s−1). A cyclic mechanism of the oxidation chain termination on introduced antioxidants (stable nitroxyl radicals of the piperidine series ( > NO) and the transition metal compounds (Men )) was established. The inhibition factor (f ) showing how many reaction chains are terminated by the one particle of the antioxidant is equal to 102. The cyclic chain termination is caused by the following reactions: HO2 + > NO → NOH + O2, HO2● + NOH → >NO + H2O2 (for >NO) and HO2 + Men → Men +1 + HO2, HO2 + Men +1 → Men + H+ + O2 (for Men ).  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