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 78 毫秒
1.
Nanostructured Pt electrodes were prepared by electrodeposition of Pt nanoparticles on different substrates (GC, Pt and Au) under cyclic voltammetric conditions and with various number (n) of potential cycling, and were denoted as nm-Pt/S(n) (S = GC, Pt and Au). Adsorption of (bi)sulfate on the nm-Pt/S(n) was studied by in situ FTIR reflection spectroscopy. It has been revealed that the nanostructured Pt electrodes exhibit anomalous IR properties for (bi)sulfate adsorption regardless of the different reflectivity of substrate, i.e. the IR absorption of (bi)sulfate species adsorbed on all the nm-Pt/S(n) electrodes is significantly enhanced and the IR band direction is completely inverted in comparison with the same species adsorbed on a bulk Pt electrode. The two IR bands around 1200 and 1110 cm−1 attributed to adsorbed (bi)sulfate species are shifted linearly with increasing electrode potential, yielding Stark tuning rates () of 152.1 and 21.1 cm−1 V−1 on nm-Pt/GC(20), respectively. Along with increasing n, the Stark tuning rate of the IR band around 1200 cm−1 decreases quickly and declined to 7.6 cm−1 V−1 on nm-Pt/GC(80), while the Stark tuning rate of the IR band near 1100 cm−1 is fluctuated between 23.0 and 16.2 cm−1 V−1. It has determined that the enhancement of IR absorption of (bi)sulfate adsorbed on nanostructured Pt electrode is varied with substrate material and n, and a maximal 16-fold enhancement of the IR band near 1200 cm−1 has been measured on the nm-Pt/GC(30) electrode. The in situ FTIR studies illustrated that the adsorption of (bi)sulfate occurs mainly in the double layer potential region, and reaches a maximum around 0.80 V. The results demonstrated also that the competitive adsorption of CO and oxygen species can inhibit completely (bi)sulfate adsorption, which has evidenced a weak interaction of (bi)sulfate with nm-Pt/S(n) electrode surface.  相似文献   

2.
The electrochemical reduction of nitrate ion was studied by cyclic voltammetry on Pt(1 1 1) and [n(1 1 1) × (1 1 1)] stepped Pt surfaces, where n (=14, 10, 7, 6, 5, 4, 3, 2) is the number of terrace atoms, in 0.1 M HClO4 + 10 mM KNO3. The electrocatalytic nitrate reduction was found to hardly proceed on Pt(1 1 1) in the hydrogen adsorption region, while the electrocatalytic activity was improved with the increase in the step density. Inactivation was observed in the presence of adsorbed hydrogen or nitrate-derived reduced adsorbate, i.e. adsorbed NO, on (1 1 1) step sites. It was, therefore, concluded that the electrocatalytically active NO3 species does not adsorb on the (1 1 1) terraces but on the (1 1 1) monoatomic steps. The nitrate reduction current increased with the step density in a non-linear relationship. The overall current density at 0.21 V (RHE) corresponding to the peak potential of the main electrocatalytic nitrate reduction wave which was maximum at n = 2, abruptly increased with short terraces, i.e. n < 5, where the current wave of adsorbed hydrogen on the Pt stepped surface with comparatively narrow (1 1 1) terraces, denoted as Hnt, also appeared unmodified for n < 5 on voltammograms recorded in 0.1 M HClO4 in the absence of nitrate.  相似文献   

3.
Elecrochemical ATR-FTIRAS measurements were conducted for the first time to investigate nature of CO adsorbed under potential control on a highly dispersed Pt catalyst with average particle size of 2.6 nm supported on carbon black (Pt/C) and carbon un-supported Pt black catalyst (Pt-B). Each catalyst was uniformly dispersed by 10 μg Pt/cm2 and fixed by Nafion® film of 0.05 μm thick on a gold film chemically deposited on a Si ATR prism window. Adsorption of CO was conducted at 0.05 V on the catalysts in 1 and 100% CO atmospheres, for which CO coverage, θCO, was 0.69 and 1, respectively. Two well-defined ν(CO) bands free from band anomalies assigned to atop CO (CO(L)) and symmetrically bridge bonded CO (CO(B)sym.) were observed. It was newly found that the CO(L) band was spitted into two well-defined peaks, particularly in 1% CO, from very early stage of adsorption, which was interpreted in terms of simultaneous occupation of terrace and step-edge sites, denoted as CO(L)terrace and CO(L)edge, respectively. This simultaneous occupation was commonly observed in our work both on Pt/C and Pt-B. A new band was also observed around 1950 cm−1 in addition to the bands of CO(L) and CO(B)sym., which was assigned to asymmetric bridge CO, CO(B)asym., adsorbed on (1 0 0) terraces, based on our previous ECSTM observation of CO adsorption structures on (1 0 0) facet. The CO(B)asym. on the Pt/C, particularly in 100% CO atmosphere, results in growth of a sharp band at 3650 cm−1 accompanied by a concomitant development of a band around 3500 cm−1. The former and the latter are assigned to ν(OH) vibrations of non-hydrogen bonded and hydrogen bonded water molecules adsorbed on Pt, respectively, interpreted in term of results from a bond scission of the existing hydrogen bonded networks by CO(L)s and from a promotion of new hydrogen bonding among water molecules presumably by CO(B)asym..It was found that the frequency ν(CO) of CO(L) both on Pt/C and Pt-B is lower than that on bulky polycrystalline electrode Pt(poly) or different crystal planes of Pt single-crystal electrodes by 30-40 cm−1 at corresponding potentials, which implies a stronger electronic interaction between CO and Pt nano-particles and/or an increased contribution of step-edge sites on the particles. Determination of the band intensities of CO(L), CO(B)asym. and CO(B)sym. has led us to conclude a much higher bridged occupation of sites at Pt nano-particles than Pt(poly) electrodes.  相似文献   

4.
The electrochemical behavior of germanium irreversibly adsorbed at stepped surfaces vicinal to the Pt(1 0 0) pole is reported. The process taking part on the (1 0 0) terraces is evaluated from charge density measurements and calibration lines versus the terrace dimension are plotted. On the series Pt(2n − 1,1,1) having (1 1 1) monoatomic steps, the charge involved in the redox process undergone by the irreversibly adsorbed germanium is able to account for (n − 0.5) terrace atoms, thus suggesting some steric difficulties in the growth of the adlayer on the (1 0 0) terraces. Conversely, no steric problems are apparent in the series Pt(n,1,0) in which more open (1 0 0) steps are present on the (1 0 0) terraces. In this latter case the charge density under the germanium redox peaks is proportional to the number of terrace atoms. Some comparison is made with other stepped surfaces to understand the behavior and stability of germanium irreversibly adsorbed on the different platinum surface sites.  相似文献   

5.
IR optical properties of Pd nanoparticles with different size and aggregation state were studied in the current paper. The dispersed Pd nanoparticles () stabilized with poly(N-vinylpyrrolidone) (PVP) were synthesized by the seeding growth method, in which the seeds were formed step by step through reducing H2PdCl4 with ethanol. The dispersed Pd nanoparticles of much large size () were grown from the by keeping the colloid of undisturbed for 150 days at room temperature around 20 °C. The aggregates of () were prepared through an agglomeration process induced during a potential cyclic scanning between −0.25 V and 1.25 V for 20 min at a scan rate of 50 mV s−1. Scanning electron microscope (SEM) patterns confirmed such aggregation of . Fourier transform infrared (FTIR) spectroscopy together with CO adsorption as probe reaction was employed in studies of IR optical properties of the prepared Pd nanoparticles. The results demonstrated that CO adsorbed on films substrated on CaF2 IR window or glassy carbon (GC) electrode yielded two strong IR absorption bands around 1970 cm−1 and 1910 cm−1, which were assigned to IR absorption of CO bonded on asymmetric and symmetric bridge sites, respectively. Similar IR bands were observed in spectra of CO adsorbed on films, except the IR bands were much weak, whereas CO adsorbed on film produced an IR absorption band near 1906 cm−1, and an anomalous IR absorption band whose direction has been completely inverted around 1956 cm−1. The direction inversion of the IR band of CO bonded to asymmetric bridge sites on was ascribed to the interaction between Pd nanoparticles inside the aggregates. Based on FTIR spectroscopic and cyclic voltammetric results, the aggregation mechanism of Pd nanoparticles from to has been suggested that the agglomeration of Pd nanoparticles was driven by the alteration of electric field across electrode-electrolyte interface, when the PVP stabilizer was stripped via oxidation during cyclic voltammetry.  相似文献   

6.
The electrooxidation of adsorbed and bulk solution of 10−2 M ethanol and D6-ethanol at polycrystalline platinum, smooth, roughened and Ru modified Pt(3 3 2), Pt(3 3 1) and Pt(1 1 1) electrodes was studied by on-line differential electrochemical mass spectroscopy (DEMS) using a dual thin layer flow through cell.On polycrystalline Pt, the main (or even single) product is acetaldehyde; due to the flow through conditions the amount of acetaldehyde further oxidized to acetic acid is negligible. At stepped single crystals with (1 1 1) terraces (Pt(s)[n(1 1 1) × (1 1 1)], acetic acid is produced at a lower potential than acetaldehyde. This demonstrates that in addition to the reaction path involving C-C bond splitting leading to CO2 (via adsorbed CO and CHx) and the reaction path leading to acetaldehyde there is a third, direct reaction path leading to the formation of acetic acid.Step decoration by Ru does not lead to an increased reactivity. This is different from the strong cocatalytic effect of Ru at step sites on the oxidation of CO. Furthermore, Ru does not influence the relative amount of acetaldehyde formed.  相似文献   

7.
Electrode-potential-dependent activation energies for electron transfer have been calculated using a local reaction center model and constrained variation theory for the oxygen reduction reaction on platinum in base. Results for four one-electron transfer steps are presented. For the first, O2(ads) is predicted to be reduced to adsorbed superoxide, O2(ads), which dissociates with a low activation barrier to O(ads) + O(ads). Then a proton transfer form H2O(ads) to O(ads) takes place, forming OH(ads) + OH(aq). The second electron transfer reacts O(ads) with H2O(aq) to form a second OH(ads) + OH(aq). The third and fourth electron transfers react the two OH(ads) with two H2O(aq) to form two H2O(ads) + two OH(aq). All three different surface reduction reactions are predicted to have reversible potentials in the −0.24 V(SHE) to −0.29 V(SHE) range for 0.1 M base and activation energies for the superoxide formation step are close to the experimentally observed range in 0.1 M base for the overall four-electron to water over the three low index (1 1 0) (1 0 0) and (1 1 1) surfaces: 0.38-0.49 eV at 0.35 eV respectively at 0.88 V(RHE). Predicted reversible potentials for forming O2(ads) are compared with estimates from the experimental literature. The difference between the acid mechanism, where the peroxyl radical, OOH(ads) is the first reduction intermediate, and the base mechanism, where superoxide, O2(ads) is the first reduction intermediate, is discussed.  相似文献   

8.
The coverage of Sn on Pt(1 1 1) which is obtained by electrochemical deposition from 5×10−5 M Sn2+ in 0.5 M H2SO4 has been determined by XPS for different deposition times. Complete suppression of hydrogen adsorption corresponds to a coverage of ?max=0.35 (Sn to surface Pt atoms).Co-adsorption of CO with Sn on Pt(1 1 1) has been studied by FTIR spectroscopy. The IR spectra of the stretching vibration of CO can be interpreted in terms of the vibrational signature of the Pt(1 1 1)/CO system and no vibrational bands associated with CO on Sn are detected. At high Sn coverages, the 1840 cm−1 band associated with bridge-bonded CO and the 2070 cm−1 band assigned to on-top CO are present, however, no hollow site adsorption which is characterized by the 1780 cm−1 band is revealed within the resolution of the experiment. This vibrational signature corresponds to a less compressed adlayer compared to the (2×2)-3CO saturation structure on Pt(1 1 1). At lower Sn coverages, signatures from both the compressed and the less compressed CO adlayer structures are seen in the spectra. From earlier structural and electrochemical studies it is known that Sn is adsorbed in 2D islands and influences CO molecules in its neighbourhood electronically. This leads to a disappearance of the IR band from CO adsorbed in the hollow site at high Sn coverages and to higher population of the weakly adsorbed state of CO for all Sn-modified surfaces, i.e. a relative increase of the amount of CO oxidised at low potentials. In addition to this electronic effect, Sn also exerts a co-catalytic effect at low Sn coverages on that part of CO which is adsorbed at a larger distance from Sn due to a bi-functional mechanism. The IR spectra shows for the Sn-modified Pt(1 1 1) surface that the transition from the compressed CO adlayer which is characterized by the hollow site adsorption of CO to the less compressed one which exhibits a characteristic band associated with bridge-bonded CO occurs already at 250 mV instead of 400 mV.  相似文献   

9.
The kinetics of electrocatalytic reduction of nitrate on Pt(1 1 0) in perchloric acid was studied with cyclic voltammetry at a very low sweep rate of 1 mV s−1, where pseudo-steady state condition was assumed to be achieved at each electrode potential. Stationary current-potential curves in perchloric acid in the absence of nitrate showed two peaks at 0.13 V and 0.23 V (RHE) in the so-called adsorbed hydrogen region. The nitrate reduction proceeded in the potential region of the latter peak in the pH range studied. The reaction orders with respect to NO3 and H+ were observed to be close to 0 and 1, respectively. The former value means that the adsorbed NO3 at a saturated coverage is one of the reactants in the rate-determining step (rds). The latter value means that hydrogen species is also a reactant above or on the rds. The Tafel slope of nitrate reduction was −66 mV per decade, which is taken to be approximately −59 mV per decade, indicating that the rds is a pure chemical reaction following electron transfer. We discuss two possible reaction schemes including bimolecular and monomolecular reactions in the rds to explain the kinetics and suggest that the reactants in the rds are adsorbed hydrogen and adsorbed NO3 with the assistance of the results in our recent report for nitrate reduction on Pt(S)[n(1 1 1) × (1 1 1)] electrodes: the nitrate reduction mechanism can be classified within the framework of the Langmuir-Hinshelwood mechanism.  相似文献   

10.
Polyiodides (Ix, x = 3 and 5) and 2I…I2 adducts were established from the Raman spectra study of 1-methyl-3-propylimidazolium iodide (MPIm+Ix; 1 ≤ x ≤ 5) ionic liquids containing various amounts of iodine (0 mol ≤ I2 ≤ 2 mol). The existence of I3 and 2I…I2 was established for 1 ≤ x ≤ 2.5, symmetric I3 ions for x = 3, while linear and discrete I5 was substantiated for 3 ≤ x ≤ 5. The presence of polyiodide species in MPIm+Ix (1 ≤ x ≤ 5) was correlated with an enhanced ionic conductivity, attributed to the established relay-type Grotthus mechanism. Two-step conductivity increase was also reflected in decrease of the hydrogen bond interactions between the CH ring groups and polyiodides. While in the concentration range 1 ≤ x ≤ 3 (triiodides and tetraiodides) IR bands changed only slightly in intensity, in the concentration range x > 3 the CH stretching bands (3040-3170 cm−1) split and the new band at 1585 cm−1 appeared in the IR spectra beside the already existing Im+ ring stretching mode at 1566 cm−1.  相似文献   

11.
A formylporphyrin has been covalently bound to Poly (Allylamine Hydrochloride) (PAH) and electrostatically self-assembled polyelectrolyte films, containing the attached metalloporphyrin, have been constructed. The UV-vis absorption band at 390 nm has been followed as core porphyrin marker. The reflection-absorption IR spectra of the gold films modified with layer-by-layer (LBL) polyelectrolytes were recorded after 6 and 12 layers. Characteristic infrared absorbance bands of porphyrin, PAH and PVS became more evident on increasing the number of bilayers. The absorption bands at 750, 1214 and 2960 cm−1, attributed at ν(S-O), νs(SO3) and ν(NH2+), respectively, showed a linear growth (R2 > 0.99) with the number of adsorbed layers. A lower correlation coefficient was observed for the band at 1585 cm−1 attributed to Fe-protoporphyrin. In order to evaluate the electron transfer (ET) rate, the ΔEp of the [Fe(CN)6]4−/[Fe(CN)6]3− couple in solution was measured after covering the electrode. A proportional increase of the ΔEp with the number of layers is observed up to the 4th layer. After the second bilayer, the magnitude of the peak separation is highly related to the charge of the topmost layer. The method allowed controlling the film thickness via the number of deposited layers (LBL). The electrode described, resulted in a good catalyst for O2 reduction and sulfite oxidation.  相似文献   

12.
In this study we report the characterization of a prototype solid-state electrochromic device based on poly(ethylene oxide) (PEO)/siloxane hybrid networks doped with lithium bis(trifluoromethanesulfonyl)imide (LiTFSI). The polymer networks prepared, designated as di-ureasils and represented as d-U(2000), were produced by a sol-gel procedure and are composed of a siliceous framework to which both ends of polyether chains containing about 40 CH2CH2O units are covalently bonded through urea linkages. Samples with compositions of 200 ≥ n ≥ 0.5 (where n is the molar ratio of CH2CH2O to Li+) were characterized by thermal analysis, complex impedance measurements and cyclic voltammetry at a gold microelectrode. Electrolyte samples were obtained as self-supporting, transparent, amorphous films and at room temperature the highest conductivity was observed with the d-U(2000)35LiTFSI composition (3.2 × 10−5 Ω−1 cm−1). We report the results of preliminary evaluation of these polymer electrolytes as multi-functional components in prototype electrochromic displays. Device performance parameters such as coloration efficiency, optical contrast and image stability were also evaluated. The electrolytes with n > 8 presented an optical density above 0.56 and display assemblies exhibited good open-circuit memory and stable electrochromic performances.  相似文献   

13.
You-Jun Fan 《Electrochimica acta》2004,49(26):4659-4666
The dissociative adsorption of ethylene glycol (EG) on Pt(1 0 0) electrode surface cooled in air after flame annealing was investigated by using programmed potential step technique and in situ FTIR spectroscopy. The stable adsorbates derived from EG dissociative adsorption on Pt(1 0 0) were determined by in situ FTIR spectroscopy as linear- and bridge-bonded CO. The quantitative results demonstrated that the average rate of dissociative adsorption of EG on Pt(1 0 0) surface varies with electrode potential, yielding a volcano-type distribution with a maximum value located near 0.10 V versus SCE. From the variation of the quantity of CO adsorbates generated in EG dissociative adsorption with the adsorption time tad, the initial rate (νi) of this surface reaction was evaluated quantitatively. The maximum value of νi has been determined to be 2.64 × 10−11 mol cm−2 s−1 in a solution containing 2 × 10−3 mol L−1 EG. The influence of the surface structure of Pt(1 0 0) electrode obtained by different pretreatment as well as of the specific adsorption of (bi)sulfate anions on the kinetics of EG dissociative adsorption has been also investigated and discussed. In comparison with a Pt(1 0 0) surface cooled in air atmosphere after flame treatment, the Pt(1 0 0) surface cooled in an Ar-H2 stream or subjected to a treatment of fast potential cycling decreased significantly the initial rate νi of EG dissociative adsorption. Similar effect was also observed for the specific adsorption of (bi)sulfate anions. However, the maximum attainable coverage () of adsorbates derived from EG dissociative adsorption is not affected either by the surface structure of Pt(1 0 0) or by (bi)sulfate anions adsorption.  相似文献   

14.
Electron transfer (ET) kinetics through n-dodecanethiol (C12SH) self-assembled monolayer on gold electrode was studied using cyclic voltammetry (CV), scanning electrochemical microscopy (SECM) and electrochemical impedance spectroscopy (EIS). An SECM model for compensating pinhole contribution, was used to measure the ET kinetics of solution-phase probes of ferrocyanide/ferricyanide (Fe(CN)64−/3−) and ferrocenemethanol/ferrociniummethanol (FMC0/+) through the C12SH monolayer yielding standard tunneling rate constant () of (4 ± 1) × 10−11 and (3 ± 1) × 10−10 cm s−1 for Fe(CN)64−/3− and FMC0/+ respectively. Decay tunneling constants (β) of 0.97 and 0.96 Å−1 for saturated alkane thiol chains were obtained using Fe(CN)64− and FMC respectively. Also, it was found that methylene blue (MB) molecules are effectively immobilized on the C12SH monolayer and can mediate the ET between the solution-phase probes and underlying gold substrate. SECM-mediated model was used to simultaneously measure the bimolecular ET between the solution-phase probes and the monolayer-immobilized MB molecules, as well as tunneling ET between the monolayer-immobilized MB molecules and the underlying gold electrode, allowing the measurement of kBI = (5 ± 1) × 106 and (4 ± 2) × 107 cm3 mol−1 s−1 for the bimolecular ET and and (7 ± 3) × 10−2 s−1 for the standard tunneling rate constant of ET using Fe(CN)64−/3− and FMC0/+ probes respectively.  相似文献   

15.
The changes of the surface topography of float zone (FZ) n-Si(1 1 1) upon conditioning of the electrodes at potentials slightly anodic of the rest potential are monitored with atomic force microscopy (AFM) in the contact mode. The influence of the composition of the used 0.1 and 0.2 M NH4F electrolyte at pH 4, of the potential and of the charge passed on the topography is investigated. The dissolution charges Qdiss ranged from 0.28 to 10.6 mC cm−2 corresponding to ∼0.5 and ∼21 bilayers (BL), respectively. The root mean square roughness Rq changes from Rq=0.2 nm for the H-terminated surface to 2.9 nm for a charge passed of 10.6 mC cm−2 at an electrode potential of 0.1 V positive of the rest potential. The evaluation of height, deflection and line scan AFM data shows pitting to originate at edges of oriented steps which separate atomically smooth terraces. Upon increased dissolution charge, island-type smooth and rather circular features form. Only for the highest Qdiss, these islands are beginning to show corrosion. An exponential relation between Rq and Qdiss is found by evaluation of the three-dimensional roughness. The slope, i.e. the increase of ln Rq with Qdiss depends on the composition of the electrolyte and is higher for the 0.1 M NH4F solution. From these data, a branching of the dissolution reaction between charge going into terrace removal or pit formation is obtained. Synchrotron radiation photoelectron spectroscopy (SRPES) is used to identify chemical products of the dissolution process.Comparison of data obtained at 0.15 V anodic of the electrode rest potential with an elaborate model of Gerischer and coworkers (which, however, only describes terrace dissolution) yields partial agreement with the predictions.  相似文献   

16.
Ionically conducting materials based on a poly(?-caprolactone) (PCL)/siloxane organic/inorganic host framework doped with magnesium triflate (Mg(CF3SO3)2) were synthesized by the sol-gel process. In this matrix short PCL chains are covalently bonded to the siliceous network via urethane linkages. In this study the salt content of samples was identified using the conventional notation n, where n indicates the number of (C(O)(CH2)5O) PCL repeat units per Mg2+ ion. Xerogels with compositions ranging from n = ∞ to 1 were prepared. The only composition studied that was not entirely amorphous was that prepared with n = 1. Xerogels with n ≥ 7 are thermally stable up to at least 200 °C. The composition with the highest conductivity of the series is that with n = 34 (5.9 × 10−9 and 9.8 × 10−7 S cm−1 at 24 and 104 °C, respectively).  相似文献   

17.
Thin films of carbonate or sulphate green rusts were synthesised from potentiostatic oxidation of solutions containing ferrous species and bicarbonate or sulphate ions at slightly alkaline pHs and ambient temperature. The thin films were characterised by means of electrochemical quartz crystal microbalance, scanning electron microscopy, X-ray diffraction and infrared reflection-absorption spectroscopy. The composition of carbonate or sulphate green rusts was studied through chemical titration, inductively coupled plasma-optical emission spectroscopy (ICP-OES) and gravimetry and is as follows:
[FeII(2R)FeIII2(OH)(4R−2R′+6)(H2O)(2R′−2)]2R′+·[R′CO3,(2R-{3 or 4}R′ + 2)·H2O]2R′− and [FeII(2R)FeIII2(OH)(4R−2R″+6)(H2O)(2R″−2)]2R″+·[R″SO4,(4R − 4R′ + 4)·H2O]2R″−  相似文献   

18.
Li[Ni(1/3−z)Co(1/3−z)Mn(1/3−z)Mgz]O2 (z = 0, 0.04) positive electrode materials were synthesized via a co-precipitation method. These materials have α-NaFeO2 () structure, as confirmed by X-ray diffraction (XRD) studies. Cation mixing in Li layer seemed to be decreased by Mg substitution as examined by Rietveld refinements of XRD data. Spherical morphologies were observed for the as-synthesized final products by scanning electron microscopy. Their electrochemical properties during charge and discharge were discussed. When magnesium ions are substituted, the initial reversible capacity reduced. However, the substitution for Mn sites in Li[Ni1/3Co1/3Mn1/3]O2 did not decrease the capacity because Mn sites substitution did not result in loss of electroactive elements in the compound. Differential scanning calorimetric studies showed the exothermic peaks of the charged electrode Li[Ni(1/3−z)Co(1/3−z)Mn(1/3−z)Mgz]O2 (z = 0.04) were significantly smaller than that of Li[Ni1/3Co1/3Mn1/3]O2, which means that thermal stability was greatly improved by Mg substitution even at highly delithiated state.  相似文献   

19.
The electrochemical reactivity of crotyl alcohol (CA) on Pt in acid media has been studied applying electrochemical methods combined with in situ Fourier transform IR spectroscopy (FTIRS) and on-line mass spectrometry (DEMS). Mass spectrometric cyclic voltammograms (MSCVs) and IR spectra, acquired with the alcohol in the solution, point out the formation of CO2 and crotonaldehyde as oxidation products.On the other hand, spectra recorded during the electroreduction display negative features at 2964, 2934 and 2875 cm−1, which are assigned to the CH3 and CH2 asymmetric stretchings and the CH2 symmetric stretching, respectively. The intensity of these signals confirms that a massive hydrogenation process occurs at E < 0.05 VRHE with a strong production of hydrocarbons. Accordingly, several reduction compounds (2-butene, butane, propane, propene, ethane and methane) were detected by DEMS.The same products, with the exception of the aldehyde, were established from CA adsorbed species, isolated applying an electrolyte exchange procedure. A feature at 2020 cm−1 due to lineal adsorbed CO develops in the IR spectra. These results allow us to propose a general overview of the electrochemical reactions of CA at Pt.  相似文献   

20.
The spherulite growth behavior and mechanism of l-lactide copolymers, poly(l-lactide-co-d-lactide) [P(LLA-DLA)], poly(l-lactide-co-glycolide) [P(LLA-GA)], and poly(l-lactide-co-ε-caprolactone) [P(LLA-CL)] have been studied using polarization optical microscopy in comparison with poly(l-lactide) (PLLA) having different molecular weights to elucidate the effects of incorporated comonomer units. The incorporation of comonomer units reduced the radius growth rate of spherulites (G) and increased the induction period of spherulite formation (ti), irrespective of the kind of comonomer unit. Such effects became remarkable with the content of comonomers. At a crystallization temperature (Tc) of 130 °C, the disturbance effects of comonomers on the spherulite growth decreased in the following order: d-lactide>glycolide>ε-caprolactone, when compared at the same comonomer unit or reciprocal of averaged l-lactyl unit sequence length (ll). The ti estimation indicated that the glycolide units have the lowest disturbance effects on the formation of spherulite (crystallite) nuclei. The PLLA having the number-average molecular weight (Mn) exceeding 3.1×104 g mol−1 showed the transition from regime II to regime III at Tc=120 °C, whereas PLLA with the lowest Mn of 9.2×103 g mol−1 crystallized solely in regime III kinetics and the copolymers excluding P(LLA-DLA) with 3% of d-lactide units crystallized solely according to regime II kinetics. The nucleation and front constant for regime II and III [Kg(II), Kg(III), G0(II), and G0(III), respectively] estimated with each (not with a fixed for high-molecular-weight PLLA) decreased with increasing the amount of defects per unit mass of the polymer for crystallization, i.e. with increasing the comonomer content and the density of terminal group through decreasing the molecular weight.  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