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 171 毫秒
1.
6FDA-pMDA polyimide membranes were implanted with 140 keV N+ ions to fluences between 2 × 1014 and 5 × 1015 cm−2. Variable energy positron annihilation spectra were taken and spectral features compared to previously reported changes in gas permeability and permselectivity of these membranes as a function of ion fluence. Positron data corroborate the explanation of these changes in terms of molecular damage caused by the implant: for fluences up to about 1 × 1015 cm−2, the concentration of irradiation-induced defects merely increases with implant fluence; while fluences exceeding this threshold value create a second type of positron annihilation site, thereby marking a distinct change in the structure of the polymer, which is responsible for the vast improvement of gas permselectivity data found at the same threshold fluence. PACS codes: 78.70.Bj—positron annihilation; 61.82.Pv—polymers, organic compounds; 61.72.Ww—doping and impurity implantation. © 1998 John Wiley & Sons, Inc. J. Polym. Sci. B Polym. Phys. 36: 2413–2421, 1998  相似文献   

2.
The yield of three photon positron annihilation is measured using semiconductor and scintillation detectors in a comparison for applications in positron emission tomography, particularly in the exploitation of three photon positron annihilation imaging where good energy resolution and good efficiency are required. In this experimental study four detectors, High-purity Germanium (HPGe), Sodium Iodide (NaI(Tl)), Lanthanum Chloride (LaCl3:10%Ce3+) and Lanthanum Bromide (LaBr3:5%Ce3+) were used. The peak-to-peak method was used with a 22Na source to determine these yields. Aluminium was employed as a reference material as its high electron density reduces positronium formation and lifetimes. Teflon was also used in order to enhance the formation of ortho-positronium, since quenching is low, leading to increased three photon positron annihilation. The relative 3γ/2γ yields obtained were (3.04±0.11)·10−2, (2.17±0.11)·10−2, (3.26±0.10)·10−2 and (2.03±0.11)·10−2 for LaBr3:Ce, LaCl3:Ce, NaI(Tl) and HPGe detectors, respectively. Among these detectors LaBr3:Ce proved to be the detector of choice for three photon imaging applications as it has both good energy resolution and efficiency.  相似文献   

3.
The size and shape of free-volume holes available in membrane materials control the rate of gas diffusion and its permeability. Based on this principle, two segmented thermo-sensitive polyurethane (TSPU) membranes with functional gates, i.e. the ability to sense and respond to external thermo-stimuli, were synthesized and used for water vapor controllable permeation. Differential scanning calorimetry (DSC), positron annihilation lifetimes (PAL), water swelling and water vapor permeability (WVP) were used to evaluate how the structure of the polyurethane (PU) and the temperature influence the free-volume holes size and the water vapor permeability (WVP) of the PU membranes. DSC study reveals that TSPU with a glass transition or a crystalline transition reversible phase shows an obvious phase-separated structure and a phase transition temperature (defined as switch temperature, Ts). PAL study indicates that the free-volume holes size of TSPU is closely related to the Ts. When the temperature is higher than the Ts, the ortho-positronium (o-Ps) lifetime (τ3) and the average radius (R) of free-volume holes of TSPU membrane increase dramatically. As a result, the WVP of TSPU membrane shows a dramatic increase. Additionally, the water swelling and the WVP of TSPU membrane are found to depend on the inner structure of the polymer, and they also give different responses to temperature variation. When the temperature is higher than the Ts, there is a significant increase of WVP from 3.80 kg/m2 day to 7.63 kg/m2 day for TSPU(a) and from 4.30 kg/m2 day to 8.58 kg/m2 day for TSPU(b), respectively. Phase transition accompanying significant changes in free-volume holes size and WVP can be used to develop “smart membranes” with functional gates and controllable gas permeation.  相似文献   

4.
Determination of the size distribution of free-volume holes in solids, in particular, polymers, is an important physicochemical problem. The positron annihilation technique has been proposed for this purpose. The central point in this technique is the quantitative interpretation of data, especially, for substances with a high specific surface area. A developed free-volume system in open-pore membrane materials, such as poly(trimethylsilylpropyne) PTMSP and the spirocyclically bound benzodioxane polymer PIM-1, and polymeric sorbents (hypercrosslinked polystyrenes) makes it possible for the first time to compare the sorption characteristics and positron annihilation data on the character of size distribution of nanopores in these polymers. In combination with the results of mathematical simulation of the structure and radiothermoluminescence measurements, the array of data indicate the structural inhomogeneity of the test amorphous materials. It was shown that this inhomogeneity in relation to the positron annihilation technique is expressed in the insufficiency of the representation of the orthopositronium decay curve by one component that takes into account the Gaussian lifetime distribution (symmetrical pore size distribution) and in the necessity of use of several decay components. The feasibility of revealing a nonrandom character of pore size distribution gives the positron annihilation technique an advantage over other approaches (inverse gas chromatography, 129Xe NMR) to investigation of nanopores in polymers.  相似文献   

5.
Polyvinylidene fluoride (PVDF) membranes were prepared via the phase inversion method from casting solutions containing PVDF, dimethylformamide (DMF), and polyvinylpyrrolidone (PVP) as pore former. PVP was used in the casting solution in a range of 0–5 wt % and extracted. The effect on membranes of using PVP in the casting process was analyzed by X-ray diffraction, differential scanning calorimetry, scanning electron microscopy, viscosity, and water permeability techniques. With an increase of PVP from 0 to 5 wt %, the PVDF casting solution viscosities increased from 858 to 1148 cP; the resulting PVDF membrane thickness increased; and the crystallinity of PVDF membranes decreased from 40.0 to 33.3%, which indicates that the addition of PVP inhibits the degree of crystallization in the PVDF membranes. SEM results revealed the shape and size of macropores in the membranes; these macropores changed after PVP addition to the casting solutions. The impact of structural changes on free-volume properties was evaluated using positron annihilation lifetime spectroscopy (PALS) studies. PALS analysis indicated no effect on the average radius (~3.4 Å) of membrane free-volume holes from the addition of PVP to the casting solution. However, the percentage of o-Ps pick-off annihilation intensity, I3, increased from 1.7 to 5.1% with increased PVP content. Further, increasing the PVP content from 0.5 to 5% resulted in an increased final pure water permeability flux. For instance, the 210 min flux for a 14% PVDF + 0.5% PVP membrane was found to be 3.3 times greater than a control membrane having the same PVDF concentration. © 2020 Wiley Periodicals, Inc. J. Polym. Sci. 2020 , 58, 589–598  相似文献   

6.
We report the results of a combined study of the local structure and the reorientation dynamics in a series of five amorphous polymers of different fragility: cis-trans-1,4-poly(butadiene) (c-t-1,4-PBD), cis-1,4-poly(isoprene) (cis-1,4-PIP), poly(isobutylene) (PIB), poly(vinyl methylether)(PVME) and poly (propylene glycol) (PPG) by using two different probe methods. The reorientation dynamics of the molecular spin probe 2,2,6,6-tetramethyl-1-piperidinyloxy (TEMPO) from electron spin resonance (ESR) is related to the annihilation behaviour of the atomic ortho-positronium (o-Ps) one as obtained by positron annihilation lifetime spectroscopy (PALS). It was found that a slow to fast transition in the spin probe rotation mobility at the operationally defined spectral temperature parameter, T50G, is connected with the mean o-Ps lifetime, τ3 (T50G) = (2.04 ± 0.26) ns. Consequently, using the free-volume concept of the o-Ps annihilation in terms of a quantum-mechanical model of o-Ps lifetime this transition can be connected with the occurrence of the mean free volume hole, Vh (T50G) = (102 ± 17) Å3, nearly independent of the chemical composition and the basic structural relaxation parameters of the amorphous polymers investigated. Finally, the free volume hole distribution aspect of the slow to fast transition indicates the presence of a sufficient free volume fluctuation at T50G for both typical fragile PVME and strong PIB polymer and emphasizes the essential role of free volume in the spin probe dynamics.  相似文献   

7.
The temperature dependence of positron annihilation characteristics, 3 andI 3, has been studied on sample of poly(butadiene), poly(isobutylene) and poly(chloroprene). The temperature range was between 15 and 470 K. The rate of expansion of holes or free-volume in all samples was deduced belowT g as well aboveT g as appr. 3·10–3 K–1 and 2·10–2 K–1, respectively. These values are very close to the rate of the mean squared displacement of scatterer<r 2>observed in neutron scattering experiments. A possibility to use an inverse value of free-volume,V f –1 for study of viscoelastic state of polymers is demonstrated.  相似文献   

8.
Diffusion coefficients between 5 × 10-14 and 1 × 10-16 cm2 s-1 have been measured for diffusion of gold and silver in the glassy state of bisphenol trimethylcyclohexanen polycarbonate in the temperature range between Tg = 507 K and 420 K using the radiotracer technique in combination with ion-beam sputtering for serial sectioning. The Arrhenius plot exhibits a downward curvature, which is interpreted within an extension of the free-volume theory to the glassy state by Vrentas and Duda. The very small metal diffusivities in comparison to values for gas molecules of comparable size suggest substantial interaction energies. © 1997 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 35 :1043–1048, 1997  相似文献   

9.
The diffusion coefficients of water vapor in poly(vinyl alcohol)–fumed silica (PVA–FS) nano-composite membranes were determined using the gravimetric method. Water vapor was observed to diffuse more rapidly in membranes with increased FS content. The vapor diffusion coefficient was determined as 1.2 × 10−13 m2/s in pure PVA and was observed to increase to 3.0 × 10−13 m2/s in PVA composites containing 30% FS nano-particles. The free-volumes of PVA–FS membranes were characterized using positron annihilation lifetime (PAL) spectroscopy. PAL results showed that both the ortho-positronium (o-Ps) lifetime and intensity increased with the addition of FS. The intensity (I3) was found to be higher than the estimated value determined from the linear combination of the data from pristine PVA and FS, and correlated excellently with the polymer amorphous content. The PAL results indicate that a higher FS content in PVA increases the free-volume hole size (a volume increase from 40 to 55 Å3) and free-volume hole density (an I3 increase from 23 to 28%), resulting in a higher fractional free-volume in the nano-composites. The increase in the relative polymer free-volume with higher FS content was associated with a decrease in the PVA crystallinity, as determined from differential scanning calorimetry measurements. It is postulated that the incorporated FS nano-particles interrupt polymeric chain packing and retard crystallization during membrane formation. More crystalline segments were transformed into amorphous regimes in the nano-composites containing more FS. A correlation between water diffusivity and the fractional free-volume was obtained, and the water diffusivity was successfully expressed by the free-volume theory.  相似文献   

10.
Molecular geometries and heats of formation have been calculated, using MINDO/3, for mass spectral fragment pairs (A+ + B) derived from formamide. There are five stable isomeric forms of the molecular ion: [H2NC(OH)]+, (H3NCO)+, [HNC(OH2)]+, [NCH(OH2)]+, and (NCOH3)+ (in order of increasing
, but no isomer (H2NCHO)+. There are three isomeric forms of (M—H)+: (H2NCO)+ (HNCOH)+, and (NCOH2)+: the only stable form of (M—2H)+ is (NCOH)+. Other (A/B)+ fragment pairs calculated are (CO/NH3)+, (HCO/NH2)+, (H2O/HCN)+, (H2O/HNC)+ [HO. + (HCNH)+], and [HO. + (H2NC)+]. The structure of the doubly charged ion M2+ is also reported.  相似文献   

11.
The photo‐degradation of polymer coating systems due to irradiation by UV and Xenon light sources is studied using positron annihilation spectroscopy and electron spin resonance (ESR). Doppler broadened spectra of positron annihilation, as a function of slow positron implantation energy and ESR spectra, are measured in two types of polyurethane which were exposed, ex situ, to UV irradiation for up to 800 h. The UV irradiation systematically decreases the S parameter as a function of exposure duration and increases the ESR signals. Thus, significant S parameter decrease is correlated with the ESR signal increase resulting from photo‐degradation of polymers due to UV irradiation. Parallel in situ positron annihilation and ESR experiments are performed as a function of Xenon light exposure for up to 100 min. These results show that the photo‐degradation of the polyurethane coatings involves initial free‐radical formation, which is correlated with the subnanometer defects detected by positron annihilation spectroscopy. © 1999 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 37: 1289–1305, 1999  相似文献   

12.
Positron annihilation measurements as a function of temperature and time have been carried out on a poly(butadiene). The measurements were performed at several temperature points from 14 to 225 K. The measurement time was several hours to four days. The analysis of data shows the following features:
(i) the value of τ3 does not depend on the rate of cooling or time,
(ii) the value of I3 depends on the rate of cooling and the history of thermal treatment,
(iii) the dependence of I3 on time can be described by Debye function. But the rise in I3 is observed at very low temperatures,
(iv) the I3 decays to value of I3 observed during very slow cooling.

Article Outline

1. Introduction
2. Experiments
3. Results
4. Discussion
5. Conclusions
6. Uncited Reference
Acknowledgements
References

1. Introduction

If a glass is formed by rapid cooling of a super-cooled liquid to a temperature below the glass–liquid transition temperature, Tg, its properties will not be static, but will relax toward values characteristic of the corresponding “equilibrium” supercooled liquid as extrapolated from above to below Tg. This process named as structural relaxation or “physical aging” is of great practical importance because of its relevance to the designing and engineering of amorphous materials with desired properties. The relaxation property and transport phenomena of disordered polymers can be explained within the free-volume concept (Ferry, 1980). However, an unsettled problem is a way of quantifying the free-volume properties, such as the free-volume fraction, the average and the distribution of the free-volume size. In the last decade, the positron annihilation lifetime spectroscopy (PALS) technique has been recognised as a useful method to detect atomic scale free-volume holes of polymers ( Schrader and Jean, 1988). This technique involves using a positron source, mostly 22Na, to emit positrons into the sample. But these positrons and the accompanying gamma–quanta have sufficient energy (average positron energy 200 keV, gamma 890 keV) to induce radiation effects, and the positron probe can thus affect the sample being investigated during PALS experiments.The basic assumption of positron annihilation lifetime spectroscopy (PALS) data interpretation in terms of the free-volume concept is the proportionality of the intensity of long-lived ortho-positronium (o-Ps) component, I3, to the concentration of free-volume holes (Kobayashi et al., 1989). However, there are different findings regarding the influence of external factors on the “true” intrinsic value of I3. Its variation with the measurement time is regarded as a manifestation of the relaxation of free-volume fraction. On the other hand, the decrease in I3 with PALS measurement time is related to the activity of the positron source and the chemical processes in the positron spur, e.g., formation of free radicals. There are PALS measurements on semi-crystalline samples (Suzuki et al., 1996), observing the I3 increase with elapsed time when the temperature of the sample is below Tg.All these reports indicate that the o-Ps formation in polymers is more complicated and the basic assumption of PALS interpretation may be questionable.In this work, PALS results will be presented on the amorphous cistrans-1,4-poly(butadiene), cistrans-1,4-PBD, in a wide temperature range from 14 to 350 K. The aim of this paper is the study of the influence of temperature, time and sample history on the intensity I3, life time of o-Ps, τ3, as well as the S-parameter from Doppler broadening measurements.

2. Experiments

The PALS experiments were conducted using a conventional fast–fast coincidence system having a time resolution of ca. 320 ps (FWHM). Cistrans-1,4-PBD has a molecular weight of Mw = 2 × 104, the glass transition temperature Tg = 178 K (Zorn et al., 1995). The isomer composition was 41% cis, 52% trans and 7% vinyl form. This isomer composition was chosen to avoid a crystallisation process on the PBD sample (Zorn et al., 1995).The positron source, consisting of 2 MBq 22N a sealed between two 3.5 μm Ni foils, was sandwiched between polymer discs, each of about 3 mm thick and with a diameter of 10 mm. At a chosen temperature, each spectrum was accumulated for 1 h, resulting in a total number of counts of about 1.14 mil. At least, two such spectra were recorded at each temperature point.The 22Na source–sample assembly was mounted on a closed cycle helium gas refrigerator. The assembly was kept in a rotary pump vacuum of about 4 Pa. Automatic temperature regulation was used during all the measurements and the temperature was controlled within ±1 K. Several different temperature scans on the specimens were performed. The first sequence (heating) was the following: I3, τ3 were first evaluated at room temperature of 300 K immediately after the source installation. Then, fast cooling to the temperature of 40 K at a rate 4 K/min was performed and the temperature increased in steps of 10 K. The second sequence (cooling) started at 300 K, then the temperature decreased to 14 K in steps of 10 K.For the PALS measurement as a function of time, the PBD was annealed in the chamber at 300 K for several hours, then cooled to the measurement temperature and the measurement began immediately.The positron life-time spectra were measured as a function of the elapsed time at 14 different temperature points below and above Tg.The PALS data were also accumulated during heating of the samples to 300 K and cooling of PBD to chosen temperature below 300 K. The total irradiation time of 1080 h was divided between PALS and calibration (Bi) measurements. To clearly describe the thermal history of the experiment, the time dependence of I3 and τ3 is shown in Fig. 1 and Fig. 2, respectively. The values of I3 and τ3 at room temperature were the same despite the long irradiation time and complicated thermal history. This indicates that a possible irradiation damage does not influence the annihilation observables.  相似文献   

13.
Two new Macroacyclic Schiff base chemosensors (L1 and L2) were synthesized by the one pot condensation reaction of 2-[3-(2-formyl phenoxy)propoxy]benzaldehyde and aminophenol in a 1:2 molar ratio and were characterized by IR, NMR spectroscopy. Both Schiff bases displayed high selectivity and sensitivity towards Fe3+ over other metal ions in H2O-DMF solution (Ag+,Cu2+, Ni2+, Zn2+, Mg+2, Mn+2, Pb+2, Co+2, Hg+2, Cr+3, Na+, Ba+2 and Cd2+) due to their structure including oxygen donor atoms. The test results showed fluorescence quenching of the fluorophores when Fe3+ was bound to the recognition units. From test results, a high selectivity for Fe3+ were discovered in this type of sensors, especially, the probe based on 2-aminophenol exhibited more significant quenching in fluorescence intensity compared with 4-aminophenol-based due to its rigidity structure. In addition, the structure of ligands and their antibacterial properties was investigated.  相似文献   

14.
A new poly(p‐phenylene ethynylene) derivative with pendant 2,2′‐bipyridyl groups and glycol units (PPE‐bipy) has been prepared, and its metal ion sensing properties were investigated. The polymer of PPE‐bipy exhibited high selectivity for Hg2+ as compared with Li+, Na+, K+, Ba2+, Ca2+, Mg2+, Al3+, Mn2+, Ag+, Zn2+, Pb2+, Ni2+, Cd2+, Cu2+, Co,2+ and Fe3+ in THF/EtOH (1:1, v/v) solution. The fluorescence of PPE‐bipy was efficiently quenched by Hg2+ ions, and the detection limit was found to be 8.0 nM in a THF/EtOH (1:1, v/v) solvent system. PPE‐bipy also showed a selective chromogenic behavior toward Hg2+ ions by changing the color of the solution from slight yellow to colorless, which can be detected with the naked eye. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 1998–2007, 2008  相似文献   

15.
In this review article, we discuss and analyze the validities of centrifugal sudden (CS) approximations in chemical reactions, with emphasis on the recent progress in the comparison studies of close‐coupling and CS approximations in chemical dynamics both adiabatically and nonadiabatically. All these relevant studies are performed using the time‐dependent wave packet approach, focusing on several typical and benchmark chemical reactions, for example, the triatomic adiabatic ion–molecule reactions of Ne + , He + HeH+, O+ + H2, O+ + D2, and O+ + HD, the triatomic nonadiabatic reactions of N + NH and O + N2, and the tetraatomic and polyatomic adiabatic reactions of H2 + D2 and H + CHD3. © 2015 Wiley Periodicals, Inc.  相似文献   

16.
We report novel liquid crystalline (LC) polymers containing pendant azobenzene moieties with n‐dodecyl substituents and ethyleneoxy spacers of different lengths and describe their selective detection behaviors to alkali metal ions. The new azopolymers produce homogenous smectic phases with a typical fan‐shaped texture. UV‐Vis and 1H NMR studies confirm that the azopolymers selectively bind to Li+ and Na+, but do not complex with K+, Ba2+, Mg2+, or Ca2+. Both the ethyleneoxy spacer and azobenzene units participate in binding to Li+ and Na+ cations in solution. Interestingly, after formation of the complexed structure, the ratio of cis to trans conformer is considerably increased suggesting stronger interactions of the cis conformer with alkali metal ions. Irradiation of the complexed structure with 365 nm UV induces conversion of the uncomplexed trans to the cis. These findings suggest a great potential of the LC azopolymers as selective sensors or separation membranes for alkali metal ions. © 2016 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2016 , 54, 1713–1723  相似文献   

17.
The role of hydrogen bonding in the chemistry of transition‐metal complexes remains a topic of intense scientific and technological interest. Poly(acrylo‐amidino diethylenediamine) was synthesized to study the effects of hydrogen bonding on complexes at different pHs. The polymer was synthesized through the coupling of diethylene triamine with polyacrylonitrile fiber in the presence of AlCl3 · 6H2O addition. The adsorption capacity of this polymer was 11.4 mequiv/g. The ions used for the adsorption test were CrO, PO, Cu2+, Ni2+, Fe2+, and Ag+. All experiments were confirmed with Fourier transform infrared. In the study of anion adsorption, at low pHs, only ionic bonds existed, whereas at high pHs, no bonds existed. However, in the middle pH region, both ionic bonds and hydrogen bonds formed between poly(acrylo‐amidino diethylenediamine) and the chromate ion or phosphate ion. When poly(acrylo‐amidino diethylenediamine) and metal ions (Cu2+, Ni2+, Fe2+, and Ag+) formed complexes, a hydrogen‐bonding effect was not observed with Fourier transform infrared. The quantity of metal ions adsorbed onto poly(acrylo‐amidino diethylenediamine) followed the order Ag+ > Cu2+ > Fe2+ > Ni2+. © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 2010–2018, 2004  相似文献   

18.
The ultrafiltration technique evaluates the interactions of water‐soluble polymers with metal ions. Aqueous solutions containing poly(sodium 4‐styrenesulfonate) (PSS), Cu(NO3)2, NaNO3, and iminodiacetic acid (IDAA) are examined by this technique. Cu2+ undergoes complex formation with IDAA and intreracts electrostatically with PSS. On the other hand, Na+ ions are in competition with Cu2+ for the electrostatic binding to PSS. The solutions are ultrafiltered keeping the ionic strength constant, so their compositions are allowed to change continuously. The concentration of Cu2+ bound to the polymer showed an exponential decay during filtration. The concentration of Cu2+ bound to the polymer before ultrafiltration is calculated by extrapolation. The concentration of the different species in solution is proposed as a function of the filtration factor. © 2002 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 40: 2587–2593, 2002  相似文献   

19.
Correlations of transport parameters (diffusion coefficients D and permeabilities P of gases) and thermodynamic parameters (solubility coefficients S and parameters C H of sorption isotherms) with the sizes of free-volume elements, v h, as estimated via positron annihilation lifetime spectroscopy are analyzed for the first time on the basis of the data array obtained for glassy polymers. Correlations of logD and logP with 1/v h that agree with the free-volume model under the condition of a weak change in the concentration of free-volume elements in different polymers are ascertained. Certain deviations from linear correlations with 1/v h for polymers with high free volumes are interpreted as evidence that the connectivity (openness) of pores increases with the sizes of free-volume elements. For solubility coefficients and Langmuir parameters of sorption capacity C H , good linear correlations with the value of v h are demonstrated.  相似文献   

20.
All 5,5′‐hydrazinebistetrazoles reported in the literature are sensitive to oxidation and react with atmospheric oxygen to yield the corresponding 5,5′‐azobistetrazolates on time. Herewith, we report on the synthesis of the free acid 5,5′‐hydrazinebistetrazole (HBT) which showed to be stable on air for extended periods of time. The compound was fully characterized by analytical and spectroscopic methods and its X‐ray structure was determined by diffraction techniques. Besides, we determined its explosive properties by BAM methods and calculated its heat of formation (+414 kJ mol?1), detonation velocity (8523 m s?1) and detonation pressure (27.7 GPa). HBT proved to be very safe to handle (impact sensitivity: >30 J, friction sensitivity: ~108 N) and was used as a starting material for the synthesis of some already reported 5,5′‐azobistetrazolates: NH4+, NH2NH3+, Li+, Na+, K+, Rb+, Cs+, Mg2+, Ca2+, Sr2+ and Ba2+.  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