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.
The neutral gas shielding model and neutral-gas-plasma shielding model are analyzed qualitatively. The main physical processes that govern the formation of the shielding gas cloud and, consequently, the ablation rate are considered. For the neutral gas shielding model, simple formulas relating the ablation rate and cloud parameters to the parameters of the pellet and the background plasma are presented. The estimates of the efficiency of neutral gas shielding and plasma shielding are compared. It is shown that the main portion of the energy flux of the background electrons is released in the plasma cloud. Formulas for the ablation rate and plasma parameters are derived in the neutral-gas-plasma shielding model. The question is discussed as to why the neutral gas shielding model describes well the ablation rate of the pellet material, although it does not take into account the ionization effects and the effects associated with the interaction of ionized particles with the magnetic field. The reason is that the ablation rate depends weakly on the energy flux of hot electrons; as a result, the attenuation of this flux by the electrostatic shielding and plasma shielding has little effect on the ablation rate. This justifies the use of the neutral gas shielding model to estimate the ablation rate (to within a factor of about 2) over a wide range of parameters of the pellet and the background plasma.  相似文献   

2.
Laser interferometry methods were used to measure the density of free electrons and degree of plasma ionization in a hydrogen target intended for experiments on determining energy losses of heavy ion beams in an ionized matter. It is shown that the linear electron density can be varied in the range from 3.3 × 1017 to 1.3 × 1018 cm?2 by varying the initial plasma parameters (the hydrogen pressure in the target and the discharge current). The error in measuring the linear electron density in the entire range of the varied plasma parameters was less than 1%. The maximum degree of plasma ionization achieved at the initial gas pressure of 1 mbar was 0.62 ± 0.05.  相似文献   

3.
The phenomenon of subnanosecond electrical breakdown in a strong electric field observed in an open discharge in helium at pressures of 6–20 Torr can be used to create ultrafast plasma switches triggering into a conducting state for a time shorter than 1 ns. To evaluate the possible repetition rate of such a subnanosecond switch, it is interesting to study the decay dynamics of the plasma remaining in the discharge gap after ultrafast breakdown. In this paper, a kinetic model based on the particle-in-cell Monte Carlo collision method is used to study the dynamics of the plasma afterglow in the discharge gap of a subnanosecond switch operating with helium at a pressure of 6 Torr. The simulation results show that the radiative, collisional-radiative, and three-body collision recombination mechanisms significantly contribute to the afterglow decay only while the plasma density remains higher than 1012 cm?3; the main mechanism of the further plasma decay is diffusion of plasma particles onto the wall. Therefore, the effect of recombination in the plasma bulk is observed only during the first 10–20 μs of the afterglow. Over nearly the same time, plasma electrons become thermalized. The afterglow time can be substantially reduced by applying a positive voltage Uc to the cathode. Since diffusive losses are limited by the ion mobility, the additional ion drift toward the wall significantly accelerates plasma decay. As Uc increases from 0 to +500 V, the characteristic time of plasma decay is reduced from 35 to 10 μs.  相似文献   

4.
Plasma membranes were isolated after binding liver and hepatoma cells to polylysine-coated polyacrylamide beads, and the effect of concanavalin A on the membrane-bound Mg2+-ATPase and the Mg2+-ATPase solubilized by octaethylene glycol monododecyl ether (C12E8) was studied. In the experiment of membranebound Mg2+-ATPase, plasma membranes were pretreated with Concanavalin A and the activity was assayed. Concanavalin A stimulated the activity of both liver and hepatoma enzymes assayed above 20°C. Concanavalin A abolished the negative temperature dependency characteristic of liver plasma membrane Mg2+-ATPase. On the other hand, Concanavalin A prevented the rapid inactivation due to storage at ?20°C, which was characteristic of hepatoma plasma membrane Mg2+-ATPase. With solubilized Mg2+-ATPase from liver plasma membranes, the negative temperature dependency was not observed. Concanavalin A, which was added to the assay medium, stimulated the activity of the enzyme solubilized in C12E8 at a high ionic strength. However, Concanavalin A failed to show any effect on the enzyme solubilized in C12E8 at a low ionic strength. With solubilized Mg2+-ATPase from hepatoma plasma membranes, Concanavalin A could not prevent the inactivation of the enzyme during incubation at ?20°C.  相似文献   

5.
Pulse radiolysis-kinetic spectrometry has been used to investigate the reaction of hydrated electrons with ferricytochrome c in dilute aqueous solution at pH 6.5–7.0. Time resolutions from 2·10?7 to 1 s were employed. Transient spectra from 320 to 580 nm were characterized with a wavelength resolution of ±0.5 nm. 1 In neutral salt-free solution, k(ferricytochrome c+e?aq)=(6.0±0.9)·1010 M?1·s?1 and k(ferricytochrome c+H)=(1.2±0.2)·1010 M?1·s?1. The reaction of ferricytochrome c with hydrated electrons is sensitive to ionic strength; in 0.1 M NaClO4, k(ferricytochrome c+e?aq)=(2.4±0.4)·1010 M?1·s?1. In contrast, k(ferricytochrome c+H) is insensitive to ionic strength. Time resolution of three spectral stages has been accomplished. The primary spectrum is the first observable spectrum detectable after irradiation and is formed in a second-order process. Its rate of formation is indisting-uishable from the rate of disappearance of the electron spectrum. The secondary spectrum is generated in a true first order intramolecular process, k(p→s)=(1.2±0.1)·105 s?1. The tertiary spectrum is also generated in a true first-order process, k(s→t)=(1.3±0.2)·102 s?1. The specific rates of both transformations are independent of the wavelength of measurement. The tertiary spectrum, observable 50 ms after initial reaction and remaining unchanged thereafter for at least 1 s, shows that relaxed ferrocytochrome c is the only detectable product. This product is not autoxidizable, as expected for native reduced enzyme. It is more probable that the intramolecular changes responsible for the p→s and s→t spectral transformations involve the influence of conformational relaxation of ferrocytochrome c upon electronic energy states then that they are intramolecular transmission of reducing equivalents from primary sites of electron attachment.  相似文献   

6.
Emission of xenon excited by a 120-keV electron beam at gas pressures of 100, 200, 500, and 760 Torr nm was studied experimentally and theoretically. More than 30 spectral lines were identified in the wavelength range of 750–1000 nm. A self-consistent kinetic model is developed to calculate the emission intensity of xenon atoms in the near IR range. The model includes balance equations for the number densities of electrons, ions and excimer molecules; equations for the populations of electron levels; and the Boltzmann equation for the low-energy part of the electron energy distribution function with a source of slow electrons. Excitation and ionization rates of xenon by the beam electrons and the energy spectrum of slow electrons are calculated by the Monte Carlo method. It is shown that, under these conditions, the main mechanism of xenon atom excitation is dissociative recombination of Xe3 + ions.  相似文献   

7.
8.
The L-type Ca2+ channel pore-forming α subunit, α1C can be detected in brain and heart as two proteins with molecular masses of ∼240 kDa and ∼190 kDa known as α1C-long and α1C-short, respectively. In brain, the α1C-short is thought to be the product of a ∼50 kDa C-terminus calpain-mediated proteolytic deletion. We now show that uterine smooth muscle also possesses α1C-long and α1C-short isoforms, and that the relative expression of these two forms is regulated by sex hormones in a tissue-specific manner. Protein expression of α1C L-type Ca2+ channels was examined in uterine smooth muscle, brain and heart, comparing non-pregnant (NP) estrus vs. late-pregnant (21 days) rats. The two forms of α1C were detected in all studied tissues. In late-pregnant uterus, α1C-long doubled the expression of α1C-short; in NP uterus the opposite occurred. However, these changes were restricted to the uterine muscle, with no changes in brain and heart. To investigate the mechanism of such regulation, ovariectomized rats were treated with sex hormones, progesterone (P4) and/or 17β-estradiol (estrogen, E2). P4 treatment, which yielded P4 plasma levels of 5±1 ng/ml and a high P4/E2 ratio (3±1.5×103) similar to the ratio in late-pregnant uterus (1.5±0.3 ×103), facilitated α1C-long expression. In contrast, E2 or E2+P4 treatment that increased E2 plasma levels to 60±8 pg/ml and 75±24 pg/ml, produced low P4/E2 ratios of 0.03±0.006 and 0.2±0.1, respectively. These low P4/E2 ratios also found in NP rats at estrus (0.3±0.1) favored the expression of α1C-short form in myometrium. Neither hormone treatment altered α1C expression in brain or heart. Our results indicate that expression of α1C isoforms depends on P4/E2 ratios. Plasma P4/E2 ratios <1×103 favor the expression of the α1C-short; whereas ratios >1×103 facilitate the expression of the α1C-long form. This regulation is tissue-specific for myometrium since it did not occur in heart and brain tissues.  相似文献   

9.
In the present investigation, intracellular sodium ([Na+]i) levels were determined in GH4C1 cells using the fluorescent probe SBFI. Fluorescence was determined by excitation at 340 nm and 385 nm, and emission was measured at 500 nm. Intracellular free sodium ([Na+]i) was determined by comparing the ratio 340/385 to a calibration curve. The ratio was linear between 10 and 60 mM Na+. Resting [Na+]i in GH4C1 cells was 26 ± 6.2 mM (mean ± SD). In cells incubated in Na+-buffer [Na+]i decreased to 3 ± 3.6 mM. If Na+/K+ ATPase was inhibited by incubating the cells with 1 mM ouabain, [Na+]i increased to 47 ± 12.8 mM in 15 min. Stimulating the cells with TRH, phorbol myristyl acetete, or thapsigargin had no effect on [Na+]i. Incubating the cells in Ca2+-buffer rapidly increased [Na+]i. The increase was not inhibited by tetrodotoxin. Addition of extracellular Ca2+, nimodipine, or Ni2+ to these cells immediately decreased [Na+]i, whereas Bay K 8644 enhanced the influx of Na+. In cells where [Na+]i was increased the TRH-induced increase in intracellular free calcium ([Ca2+]i) was decreased compared with control cells. Our results suggest that Na+ enters the cells via Ca2+ channels, and [Na+]i may attenuate TRH-induced changes in [Ca2+]i in GH4C1 cells. © 1993 Wiley-Liss, Inc.  相似文献   

10.
We examined the involvement of membrane microdomains during human luteinizing hormone (LH) receptor recovery from receptor desensitization after removal of bound hormone. Lateral motions of individual desensitized LH receptors expressed on the surface of Chinese hamster ovary cells and transient association of these receptors with detergent-resistant membrane (DRM) microdomains isolated using isopycnic sucrose gradient ultracentrifugation were assessed. Single particle tracking experiments showed untreated individual LH receptors to be confined within cell-surface membrane compartments with an average diameter of 199 ± 17 nm and associated with membrane fractions characteristic of bulk plasma membrane. After brief exposure to human chorionic gonadotropin (hCG), LH receptors remained for several hours desensitized to hCG challenge. Throughout this period, significantly increased numbers of LH receptors were confined within smaller diameter (<120 nm) membrane compartments and associated with DRM fragments of characteristically low density. By 5 h, when cells again produced cAMP in response to hCG, unoccupied LH receptors were found in larger 169 ± 22 nm diameter cell-surface membrane compartments and >90 % of LH receptors were again found in high-density membrane fragments characteristic of bulk plasma membrane. Taken together, these results suggest that, during recovery from LH receptor desensitization, LH receptors are both located with DRM lipid environments and confined within small, mesoscale (80–160 nm) cell-surface compartments. This may reflect hormone-driven translocation of receptors into DRM and formation there of protein aggregates too large or too rigid to permit effective signaling. Once bound hormone is removed, receptor structures would have to dissociate before receptors can again signal effectively in response to hormone challenge. Moreover, such larger protein complexes would be more easily constrained laterally by membrane structural elements and so appear resident in smaller cell-surface compartments.  相似文献   

11.
Excitation energy transfer in monomeric and trimeric forms of photosystem I (PSI) from the cyanobacterium Synechocystis sp. PCC 6803 in solution or immobilized on FTO conducting glass was compared using time-resolved fluorescence. Deposition of PSI on glass preserves bi-exponential excitation decay of ~4–7 and ~21–25 ps lifetimes characteristic of PSI in solution. The faster phase was assigned in part to photochemical quenching (charge separation) of excited bulk chlorophylls and in part to energy transfer from bulk to low-energy (red) chlorophylls. The slower phase was assigned to photochemical quenching of the excitation equilibrated over bulk and red chlorophylls. The main differences between dissolved and immobilized PSI (iPSI) are: (1) the average excitation decay in iPSI is about 11 ps, which is faster by a few ps than for PSI in solution due to significantly faster excitation quenching of bulk chlorophylls by charge separation (~10 ps instead of ~15 ps) accompanied by slightly weaker coupling of bulk and red chlorophylls; (2) the number of red chlorophylls in monomeric PSI increases twice—from 3 in solution to 6 after immobilization—as a result of interaction with neighboring monomers and conducting glass; despite the increased number of red chlorophylls, the excitation decay accelerates in iPSI; (3) the number of red chlorophylls in trimeric PSI is 4 (per monomer) and remains unchanged after immobilization; (4) in all the samples under study, the free energy gap between mean red (emission at ~710 nm) and mean bulk (emission at ~686 nm) emitting states of chlorophylls was estimated at a similar level of 17–27 meV. All these observations indicate that despite slight modifications, dried PSI complexes adsorbed on the FTO surface remain fully functional in terms of excitation energy transfer and primary charge separation that is particularly important in the view of photovoltaic applications of this photosystem.  相似文献   

12.
A variety of metal ions can bind to the iron-transport protein, transferrin, at two specific sites. For each metal ion, a carboxylate anion is concomitantly bound. Six metal ions which were examined fall into two classes based on proton release and ultraviolet spectral changes which accompany binding to the protein. Class II ions, which include Cu2+ and Zn2+, release approximately 2 H+/metal bond. Class III ions, which include Fe3+, Ga3+, Al3+, and VO2+, release approximately 3 H+/metal bound. The increase in absorbance near 242 nm, characteristic of tyrosine ionization, has the ratio 0.55–0.75 for class II:class III ions. Both Fe3+ and Cu2+ form metal-transferrin-oxalate complexes in the presence of excess C2O42?. Fe3+ releases close to 3 H+/metal whether forming oxalate or bicarbonate complexes with transferrin. Binding of Cu2+ to transferrin releases 2 H+/metal in the presence of C2O2?4 or HCO3?. Since equal numbers of H+/metal are released for both anions, it is likely that the bicarbonate ion does not lose its proton, and remains as HCO3? in transferrin. These results are interpreted in terms of possible combinations of ligands at the metal binding sites.  相似文献   

13.

The effect of dipole-dipole interactions of free electrons on the spectral characteristics of simple metals and their nanoparticles is analyzed using Drude theory and the model of the Lorentz local field. It is established that accounting for the dispersion of a local field under conditions of one-dimensional (1D) confinement based on the optical constants of the bulk metal allows the determination of its spectral micro-characteristics in the frequency region of the longitudinal collective motions of the free electrons. This corresponds to the spectra of the dielectric losses of bulk plasma oscillations. A similar procedure for three-dimensional (3D) confinement produces the spectrum of dielectric losses at the frequency of localized plasma oscillations. Using a number of simple metal examples, viz., Li, Na, and K, and also Al, Be, and Mg, it is shown that the frequencies of volume and localized plasma oscillations obtained from a model of dispersion of the local field in the long-wave limit are in good agreement with the actual frequencies of the plasma oscillations of the corresponding metals and the absorption maxima of their nanoparticles with a radius of 2–20 nm. It is shown that the frequencies of the main mode of longitudinal plasma oscillations and the absorption frequency of localized plasmons are well described using the dynamic theory of crystal lattice vibrations.

  相似文献   

14.
Alkaline isomerization of ferricytochrome C from Euglena gracilis   总被引:1,自引:0,他引:1  
Euglena gracilis ferricytochrome c has a small absorption maximum at about 700 nm having an extinction of 850 ± 10 M?1cm?1. This absorption band is analogous to the more commonly found maximum at 695 nm which is observed in ferricytochromes from other sources and which is characteristic of ligation of methionine 80 with the heme ion. The 700 nm band disappears upon raising the pH to 11 giving a transition involving a single proton having an apparent pK of about 10. These results demonstrate that the phenolic ionization of tyrosine 67 is not required to trigger the alkaline isomerization of ferricytochromes c since Euglena cytochrome has a phenylalanine residue at position 67.  相似文献   

15.
Abstract: Serotonin 5-HT2C receptor-mediated intracellular Ca2+ mobilization was investigated in Chinese hamster ovary (CHO) cells transfected with 5-HT2C receptors. Fura-2 acetoxymethyl ester was used to investigate the regulation of 5-HT2C receptor function. CHO cells, transfected with a cDNA clone for the 5-HT2C receptor, expressed 287 fmol/mg of the receptor protein as determined by mianserin-sensitive [3H]mesulergine binding (KD = 0.49 nM). The addition of 5-HT mobilized intracellular Ca2+ in a dose-dependent fashion, ranging from a basal level of 99 ± 1.8 up to 379 ± 18 nM, with an EC50 value for 5-HT of 0.029 µM. Exposure to 5-HT, 1-(3-chlorophenyl)piperazine dihydrochloride (a 5-HT2C agonist), and 1-(4-iodo-2,5-dimethoxyphenyl)-2-aminopropane (a 5-HT2C and 5-HT2A agonist) resulted in increased intracellular Ca2+ levels. Mianserin, mesulergine, ritanserin, and ketanserin each blocked 5-HT-mediated intracellular Ca2+ mobilization more effectively than spiperone. The receptor was rapidly desensitized by preexposure to 5-HT in a time- and concentration-dependent manner. Mezerein and phorbol 12-myristate 13-acetate, protein kinase C activators, weakly inhibited the intracellular Ca2+ mobilization induced by 10 µM 5-HT. Furthermore, the protein kinase C inhibitor H-7 partially prevented the protein kinase C activator-induced inhibition of the 5-HT-mediated increase in intracellular Ca2+ concentration. The desensitization induced by pretreatment with 5-HT was blocked by W-7, added in conjunction with 5-HT, and partially inhibited by W-5, a nonselective inhibitor of protein kinases and weak analogue of W-7. Therefore, the 5-HT2C receptor may be connected with protein kinase C and calcium/calmodulin turnover. These results suggest that 5-HT2C receptor activation mobilizes Ca2+ in CHO cells and that the acute desensitization of the receptor may be due to calmodulin kinase-mediated feedback.  相似文献   

16.
A simple, rapid, and low-cost technique was developed to allow reliable analysis of the anti-hepatitis C drug sofosbuvir in bulk, tablet form, and spiked human plasma. This method depends on the ability of sofosbuvir to quench the fluorescence of the newly synthesized 2-amino-3-cyano-4,6-dimethylpyridine (reagent 3 ). Elemental analysis and spectral data were used to validate the structure of the synthesized reagent. The newly synthesized reagent exhibited a satisfactory level of fluorescence emission at 365 nm after excitation at 247 nm. All experimental variables that might affect the quenching process were analyzed and optimized. Linearity, range, accuracy, precision, limit of detection (LOD), and limit of quantitation (LOQ) were all validated in accordance with the International Council for Harmonization of Technical Requirements for Pharmaceuticals for Human Use (ICH) guidelines. The concentration range was shown to be linear between 0.1 and 1.5 μg/mL. The technique was effectively utilized for sofosbuvir analysis in both its tablet dosage form and spiked human plasma, with mean percentage recoveries of 100.13 ± 0.35 and 94.26 ± 1.69, respectively.  相似文献   

17.
A sensitive time‐resolved luminescence method for the determination of amlodipine (AM) in methanol and in aqueous solution is described. The method is based on the luminescence sensitization of terbium (Tb3+) by formation of a ternary complex with AM in the presence of tri‐n‐octylphosphine oxide (TOPO) as co‐ligand, dodecylbenzenesulfate as surfactant and europium ion as a co‐luminescence reagent. The signal for Tb–AM–TOPO is monitored at λex = 242 nm and λem = 550 nm. Optimum conditions for the formation of the complex in aqueous system were 0.015 m Tris (hydroxylmethyl) amino methane buffer, pH 9.0, TOPO (1.0 × 10–4 m ), Eu3+ (2.0 × 10–7 m ), dodecylbenzenesulfate (0.14%) and 6.0 × 10–5 m of Tb3+, which allows the determination of 10–50 ppb of AM with a limit of detection of 1.2 ppb. The relative standard deviations of the method range between 0.1 and 0.2% indicated excellent reproducibility of the method. The proposed method was successfully applied for the assay of AM in pharmaceutical formulations and in plasma samples. Average recoveries of 98.5 ± 0.2% and 95.2 ± 0.2% were obtained for AM in tablet and plasma samples respectively. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

18.
The conformation of human placental alkaline phosphatase (EC 3.1.3.1) has been studied using the spectroscopic structural probes of pH difference spectroscopy, solvent perturbation difference spectroscopy, and circular dichroism. Of the 37 ± 1 tyrosine residues in placental alkaline phosphatase (PAP), 5 ± 1 residues are observed by pH difference spectroscopy to be “free” and presumed to be located on the surface of the enzyme molecule. The ionization of these 5 “free” tyrosyl groups is not time dependent and is reversible with a pKapp of 10.29. The remaining 32 ± 1 tyrosines are considered “buried” and ionization is observed to be both time dependent and irreversible. Treatment of the enzyme with 4 m guanidine-hydrochloride normalizes all 37 ± 1 tyrosine residues (pKapp = 10.08). The difference pH titration studies thus provide spectrophotometric evidence for a change in molecular conformation of PAP in the pH region of 10.5. Using solvent perturbation difference spectroscopy and circular dichroism, the local environments of tyrosine and tryptophan residues were elucidated for the native enzyme and the enzyme in the presence of ligands that influence catalytic function: inorganic phosphate (competitive inhibitor), l-phenylalanine (uncompetitive inhibitor), d-phenylalanine (noninhibitor). and Mg2+ ion (activator). The spectral observations from these studies led to the following interpretations: (i) the binding of inorganic phosphate, a competitive inhibitor, induces a conformational change in the enzyme that may alter the active site and thereby decrease enzyme catalytic function; (ii) perturbation with l-phenylalanine gives spectral results indicating a conformational change consistent with the postulate that this uncompetitive inhibitor prevents the dissociation of the phosphoryl enzyme intermediate; and (iii) Mg2+ ion causes a slight separation of the enzyme subunits, which could increase accessibility to the active site and, thus, enzyme activity.  相似文献   

19.
Abstract— The effects of divalent metal ions, sulfhydryl reagents, carbonyl trapping reagents, substrate analogs, and organic solvents on purified mouse brain 4-aminobutyrate-2-ketoglutarate transaminase (EC 2.6.1.19) and the subunit structure of this enzyme were studied. Of the metal ions tested, Hg2+ was found to be the most potent inhibitor inhibiting the enzyme 50 percent at a concentration of 0-7 μM. The order of decreasing inhibitory potency for the divalent metal ions was: Hg2+± Cd2+± Zn2+± Cu2+± Co2+± Ba2+± Sr2+± Ni2+± Mn2+± Ca2+± Mg2+. p-Chloromercuribenzoale was the most potent inhibitor among the sulfhydryl reagents tested inhibiting the enzyme to the extent of 50 per cent at 0-5 μM 3-Mercaptopropionic acid was found to be a competitive inhibitor for GABA and non-competitive for 2-ketoglutarate. The Ki, value was estimated to be 13 μM. Aminooxyacetic acid was the most potent inhibitor of the carbonyl trapping agents with a K, value of 0-06 μM. being competitive with GABA and non-competitive with 2-ketoglutarate. Hydroxylamine and hydrazine were the next most potent compounds in this group. Of a series of substrate analogs and metabolites tested, only acetic acid, propionic acid, butyric acid, glutamic acid, adipic acid, pimelic acid and 2-ketoadipic acid inhibited the enzyme to a significant extent. Dioxan inhibited the enzyme 50 per cent at a concentration of 5 per cent (v/v) whereas methanol and ethanol only inhibited 5-10 per cent at 10 per cent (v/v) concentration. A spectrum of the native enzyme at pH 7-2 showed maxima at 278 nm. 330 nm and 411 nm. Treatment of the enzyme with aminooxyacetic acid or 3-mercaptopropionic acid caused the maximum at 411 nm to disappear. Sodium dodecyl sulfate polyacrylamide gel electrophoresis of the enzyme revealed two protein bands. The molecular weights of these two subunits were determined to be 53.000 and 58,000, respectively.  相似文献   

20.
Results of experimental and theoretical study of plasma decay in the afterglow of high-voltage nanosecond discharges in gaseous ethylene and dimethyl ether at room temperature and pressures from 2 to 20 Torr are presented. Using a microwave interferometer, the time behavior of the electron density in the range from 2 × 1010 to 3 × 1012 cm–3 during plasma decay is investigated. By processing the experimental data, the effective coefficients of electron–ion recombination as functions of the gas pressure are obtained. It is found that these coefficients substantially exceed the recombination coefficients of simple hydrocarbon ions. This distinction, as well as the increase in the effective recombination coefficient with pressure, is explained by the formation of cluster ions in three-body collisions, which recombine with electrons more efficiently than simple molecular ions. The coefficients of three-body conversion of simple molecular ions into cluster ions in the plasmas of ethylene and dimethyl ether, as well as the coefficients of recombination of electrons with cluster ions in these gases, are determined by analyzing the experimental data.  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