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.
Glasses in the MeF2-Na2B4O7 (Me = Mg, Ca, Sr, and Ba) system have been synthesized. It is shown that the glass formation is observed at a MeF2 content of up to 40 mol %. The influence of the MeF2 content on the electrical conductivity and the fluorine concentration in the glass bulk is examined. From the analysis of the concentration dependence of the electrical conductivity with due regard for the fluorine content, it is concluded that the glass structure is predominantly built up of the polar groupings Na+[BO4/2]-, Na+[F-BO3/2], Me 1/2 2+ [BO4/2], Me 1/2 2+ [FBO3/2], [MeF4/2], and [MeF6/3] and the BO3/2 nonpolar structural-chemical units. The electricity transport is governed by the migration of sodium ions formed upon dissociation of the Na+[BO4/2]-and Na+[F-BO3/2] groupings. An increase in the MeF2 content leads to a decrease in the total concentration of sodium ions, a decrease in the Na+[BO4/2]- concentration, and an increase in the Na+[F-BO3/2] concentration. Upon introduction of MeF2 up to ∼20 mol %, the fluorine losses during the synthesis are caused by the dehydration of glass melt. An addition of 20–25 mol % MeF2 brings about the saturation of the glass by the [F-BO3/2]-type structural units, so that the fluorine concentration reaches a saturation in the structures of calcium-, strontium-, and barium-containing glasses and increases in magnesium-containing glasses, owing to the formation of the [MgF+6/3] groupings.  相似文献   

2.
The temperature–concentration dependences of the electrical conductivity and the activation energy for electrical conduction of glasses in the Na2O–B2O3 and Na2O–2PbO · B2O3 systems are studied. The investigation into the nature of the electrical conduction in these glasses reveals that the contribution from the electronic component (10–3%) of the conductivity is within the sensitivity of the Liang–Wagner technique. A considerable alkali conductivity is observed upon introduction of more than 12 mol % Na2O. The true transport number of sodium Na is as large as unity at [Na2O] 15 mol %. It is shown that the observed temperature–concentration dependences of the electrical and transport properties are governed by the ratio between the concentrations of polar and nonpolar structural–chemical units of the Na+[BO4/2], Na+[OBO2/2] Na+[OBO2/2], Pb2+ 1/2[BO4/2], Pb2+ 1/2[OBO2/2], and [BO3/2] types.  相似文献   

3.
The temperature–concentration dependence of the electrical conductivity of glasses in the NaPO3–NaF system has been investigated. The regularities revealed are interpreted from the standpoint of the structural microinhomogeneity of glasses, which is due to the formation of polar structural units of the Na+[OPOO2/2], Na2 +[O 2POO1/2], Na+[FPOO2/2], and Na+F types. It is shown that the concentration dependence of the electrical conductivity is governed by the ratio between the concentrations of these structural units.  相似文献   

4.
This study presented the melt structure evolution of mold flux during the substitution of fluorine by B2O3, and a computational model for the degree of polymerization (DOP) for borosilicate structure was developed. The results showed that the reduction of fluorine content would promote the replacement of F in [SiF6]-octahedral unit by the dissociative free oxygen ions (O2−), and release F ions into the melt to compensate the reduction of F ions. With the 2 mass% addition of B2O3, the original Si–O–Si bond would be disrupted, and connect with [BO3]-trihedral to form boroxol ring structure containing [BO2O]-trihedral and [BO3]-trihedral structural units. Then, the Si–O–B bond that [BO3]-trihedral links [SiO4]-tetrahedral in boroxol ring was destroyed with the further addition of B2O3, and then the [BO3]-trihedral could link with the dissociative Q1(Si) and Q0(Si) structural units to transform into [BO4]-tetrahedral and form a borosilicate long chain. Finally, with 6 mass% addition of B2O3, the borosilicate chain would combine with simple borate and borosilicate structures, and a complex borosilicate structure containing boroxol ring with certain symmetry was formed ultimately. Besides, the calculated result of DOP suggested that the DOP of the melt structure improved during the process of fluorine replacement by B2O3.  相似文献   

5.
The concentration dependence of the electrical conductivity of glasses in the Tl2O-B2O3 system is studied. The nature of charge carriers in this system is experimentally investigated for the first time. It is demonstrated using the Hittorf, Tubandt, and Hebb-Liang-Wagner techniques and the Faraday law that neither Tl+ ions nor electrons are involved in the electricity transport. The verification of the Faraday law does not reveal the presence of thallium in the amalgam of the cathode or a change in the sample weight after electrolysis, to within the experimental error. This allows one to make the inference that protons can be charge carriers in glasses of the Tl2O-B2O3 system. It is shown using extended X-ray absorption fine structure (EXAFS) spectroscopy that Tl3+ ions and thallium Tl0 reduced to the metallic state are absent in the structure of the glasses under investigation. This means that thallium in glasses of the Tl2O-B2O3 system occurs only in the form of Tl+ ions. The analysis of the IR spectroscopic data leads to only a qualitative conclusion that the water content in the glasses insignificantly increases with an increase in the thallium oxide content. An increase in the electrical conductivity of glasses in the Tl2O-B2O3 system with an increase in the thallium oxide content is explained by the increase in the number of protons formed upon dissociation of H+[BO4/2]? structural-chemical units, because their concentration increases with increasing Tl2O content. In the structure of boron oxide, impurity hydrogen enters predominantly into the composition of H+[O2/2BO?] structural-chemical units, for which the dissociation energy is higher than that for the H+[BO4/2]? structural-chemical units. The increase in the concentration of H+[BO4/2]? structural-chemical units is accompanied by the increase in the number of dissociated protons, which are charge carriers in glasses of the Tl2O-B2O3 system.  相似文献   

6.
The effect of Na2O and temperature on the thermal conductivity of the Na2O–B2O3 binary system has been measured using the hot‐wire method to examine the relationship between the thermal conductivity and structure in high‐temperature melts. The thermal conductivity of the binary melt is measured from 1173 to 1473 K in the fully liquid state. The thermal conductivity slightly increases with Na2O content up to 20 wt%. Above 20 wt% Na2O, the thermal conductivity decreases with increasing Na2O. The network structure of molten glass was analyzed using Fourier transform infrared (FTIR), Raman spectroscopy, and XPS. The FTIR analysis shows that 3‐D complex borate structures, such as tri‐, tetra‐, and pentaborate are made by [BO4] tetrahedral units interconnected with 2‐D structure boroxol rings in the low Na2O region. Above 20 wt% Na2O content, nonbridged oxygen in [BO2O?] units and diborate groups increase with increase in Na2O. The same tendency is shown by the Raman spectroscopy and XPS analyses. The Raman analysis shows that boroxol rings disappeared with large [BO4] groups, such as tri‐, tetra‐, and pentaborate structures, which increase at low Na2O content. Isolated diborate groups and nonbridged oxygen in [BO2O?] units increase at high Na2O content. It can be inferred that single structure units, such as isolated diborate groups, interfere with conduction. The XPS analysis results show that free oxygen produced by the interconnection of Na2O in the borate structure does not cause significant changes to O2? in the low Na2O region, but increases the Oo and decreases the O?. Above 20 wt% Na2O, O? slightly increases and Oo shows a decreasing trend.  相似文献   

7.
The fundamental natures of the interaction between thiophene and ionic liquids of 1-n-butyl-3-methylimidazolium hexafluorophosphate ([BMIM]+[PF6]) and 1-n-butyl-3-methylimidazolium tetrafluoroborate ([BMIM]+[BF4]) were investigated using ab initio calculations and correlated with previous experimental results. The optimized structures show that the anions of the ionic liquids are situated outside the ring plane of the thiophene, with the fluorine atoms interacting with the hydrogen atoms of the thiophene, and the cation of the ionic liquids approaches the thiophene with its positively charged atoms approaching the negatively charged atoms of TS. It is concluded that thiophene molecules interact with the ionic liquids mainly via Coulombian attraction. Further analysis explained the results obtained from an absorption experiment that the molar ratios of the absorbed thiophene to the ionic liquids were approximately 3.5/1 and 2.4/1 for [BMIM]+[PF6] and [BMIM]+[BF4], respectively. The strong electron donation of the phosphorus atom to the fluorine atoms in the PF6 cluster is believed to be the major factor resulting in the higher molar ratio of thiophene/[BMIM]+[PF6]. The other factor is the difference of the compactness between the cation and the anion in the two ionic liquids.  相似文献   

8.
Low- and high-alkali cement pastes were made with or without LiNO3 or a Li-bearing glass. The [Li]/[Na+K] molar ratio was kept constant to 0.74. The specimens were stored at 23, 38, and 60 °C in sealed containers. After 3, 7, 28, and 91 days, their pore solutions were extracted and analysed, and their residual water contents were obtained by drying. The Li glass was found to react quite slowly, and the corresponding [Li+] in solution progressively increased with time, temperature, fineness (as-received glass vs. ground glass), and the [Na++K+] concentration in solution. This glass increased the pH by about 0.1, and by about 0.2 after it was finely ground. In contrast, LiNO3 decreased the pH by about 0.1, despite significantly increasing the [Na++K+] in the pore solution. The higher the total %Na2Oe content (including Li) in the original mixtures, the higher the total alkali content incorporated in the cement hydrates. The [Li+]-[Na++K+] ratio in solution was about half of the initial ratio (0.74), while this ratio in the cement hydrates was always over 1.1. Li is the alkali most preferentially incorporated into the cement hydrates, while K is the least.  相似文献   

9.
The effect of introduction of aluminum oxide into the composition of sodium silicate glasses has been studied by IR absorption and reflection spectroscopy. The change in the spectroscopic characteristics of glasses after their treatment with HNO3 and AgNO3 aqueous solutions is analyzed. The concentration profiles of Na+ and Ag+ ions in the surface layers of these glasses are determined by the HF-sectioning technique. It is found that silver ions predominantly interact with the [AlO4/2]- groups in the glass. The leaching of sodium ions, formation of amorphous silica in the surface layers of the treated glass samples, and exchange of sodium ions by hydrogen ions are revealed from changes in the spectra.  相似文献   

10.
The effect of introduction of aluminum oxide into the composition of sodium silicate glasses has been studied by IR absorption and reflection spectroscopy. The change in the spectroscopic characteristics of glasses after their treatment with HNO3 and AgNO3 aqueous solutions is analyzed. The concentration profiles of Na+ and Ag+ ions in the surface layers of these glasses are determined by the HF-sectioning technique. It is found that silver ions predominantly interact with the [AlO4/2]- groups in the glass. The leaching of sodium ions, formation of amorphous silica in the surface layers of the treated glass samples, and exchange of sodium ions by hydrogen ions are revealed from changes in the spectra.  相似文献   

11.
The occlusion of lithium, sodium, potassium and silver nitrates from the anhydrous melts in the respective zeolites has been studied by measuring the gain in weight and by direct analysis. The results in moles of occluded nitrate per formula weight of zeolite M12I[(AlO2)12(SiO2)12] are 11.7 for LiNO3, 10 for NaNO3 and 10 for AgNO3 and essentially zero for KNO3. A model is proposed for the structure of the occlusion compound Na22[(AlO2)12(SiO2)12(NO3)10], in terms of four NaI+ ions and four [NaI - NO3 - Na]+ groups at the apexes of the cubic unit cell and four [NaII - NO3 - Na]+1/3 groups coordinated with two [NaNO3]?2/3 groups at the faces, to explain the properties of the occluded zeolites.  相似文献   

12.
The labile side-on α-carbonyl diazoalkane platinum complex (dtbpm-κ2P)Pt[N2C(CO2Me)2-κ2N,N′] (2) displays nucleophilic as well as electrophilic reactivity patterns. H2O is added stoichiometrically, forming the hydrazonido hydroxo platinum(II) complex (dtbpm-κ2P)Pt(OH)[NHNC(CO2Me)2] (3). Protonation of 2 with BF3·OEt2 and MeOH yields [(dtbpm-κ2P)Pt{NHNC(CO2Me)2-κN,κO}]+[BF4] ([4]+[BF4]). Both new complexes have been fully characterized spectroscopically and by single crystal X-ray diffraction.  相似文献   

13.
We studied the solution fluorescence quenching of poly(methylphenethylsilane) (1#), poly(dimethylsilane‐co‐methylphenethylsilane) (2#), poly(n‐hexylmethylsilane) (3#), and poly(dimethylsilane‐con‐hexylmethylsilane) (4#) by such quenchers as CCl4, CHCl3, Cl2CHCHCl2, and methyl benzoate. We treated the fluorescence quenching data using the equations F0/F = 1 +KSV[Q],F0/F exp(−NV[Q]) = 1 +KSV[Q], and ln(F0/F) =NV[Q], where F and F0 are the fluorescence intensity with and without the addition of a quencher, respectively; KSV, the Stern–Volmer constant; [Q], the quencher concentration; N, Avogadro's constant; and V, the volume of the active sphere. For the systems with both static quenching and dynamic quenching, we calculated their contributions and the critical quencher concentration [Q]C and determined the nature of the fluorescence quenching in different quencher concentration ranges. We observed that, under the condition of the same quencher, the fluorescence quenching of the polysilane homopolymer is smaller than that of its corresponding polysilane copolymer, that is, 1# < 2# and 3# < 4#, and that for the fluorescence quenching of the same polysilane by different chlorohydrocarbons the fluorescence quenching ability of CCl4 is larger than that of CHCl3 and Cl2CHCHCl2. © 2000 John Wiley & Sons, Inc. J Appl Polym Sci 78: 133–139, 2000  相似文献   

14.
Tb3+-doped 25Na2O-23CaO-6P2O5-44B2O3-2ZrO2 glass was fabricated by conventional melt-quenching technique. Glass-ceramics containing NaCaPO4 crystals were then obtained by heating the as-prepared glasses. Their optical and luminescence properties were studied by FT-IR spectroscopy, photoluminescence (PL), absorption spectra, thermoluminescence (TL), and optically stimulated luminescence in continuous wave modality (CW-OSL). The glasses were composed of [PO4], [BO3], and [BO4] basic structural units. The PL excitation and emission spectra exhibited Tb3+-related transitions, as well as the strongest excitation and emission wavelengths at 370 and 454 nm, respectively. We further investigated the CW-OSL properties as a function of dopant concentration and time elapsed after irradiation (signal fading). Results indicated that the CW-OSL intensity reached a maximum when the Tb4O7 concentration was 0.25 mol%. The fading of the OSL signal showed that the OSL signal of Tb3+-doped NaCaPO4 glass-ceramics was approximately 65% in 8 days, after which the intensity remained stable. The TL glow curves had a broad peak feature peaking at 180 ± 5ºC. The samples also exhibited good signal reusability and a broad linear dose-response range (0.03-1000 Gy). The excellent luminescent and dosimetric properties of these Tb3+-doped NaCaPO4 glass-ceramics indicated their potential applications in radiation dosimetry.  相似文献   

15.
Jinqing Qu  Toshio Masuda 《Polymer》2007,48(22):6491-6500
Pyrene-functionalized chiral methylpropargyl esters, (R)-3-butyn-2-yl-1-pyrenebutyrate [(R)-1], (S)-3-butyn-2-yl-1-pyrenebutyrate [(S)-1], (R)-3-butyn-2-yl-1-pyrenecarboxylate [(R)-2], and 3-butyn-2-yl-1-pyrenecarboxylate [(R,S)-2] were polymerized with (nbd)Rh+[η6-C6H5B(C6H5)3] to obtain the corresponding polymers with moderate molecular weights (Mn: 10?500-66?500) in good yields (82-97%). All the polymers were soluble in CHCl3, CH2Cl2, and THF. The polarimetric and CD spectroscopic data indicated that poly[(R)-1], poly[(S)-1], and poly[(R)-2] existed in a helical structure with predominantly one-handed screw sense in these solvents. The helical structure of poly[(R)-1] and poly[(S)-1] was stable upon heating and addition of MeOH, while that of poly[(R)-2] changed upon MeOH addition. The copolymerization of (R)-1 with (S)-1 was also conducted to obtain the copolymers satisfactorily. Poly[(R)-1], poly[(S)-1], and poly[(R)-2] emitted fluorescence smaller than the corresponding racemic copolymers. The fluorescence intensity was tuned by the addition of MeOH to THF solutions of the polymers.  相似文献   

16.
Single-crystal of fully dehydrated, Mg2+-exchanged zeolite Y, |Mg34.5Na6|[Si117Al75O384]-FAU (Si/Al = 1.56), was successfully prepared from undried methanol solution (water concentration 0.02 M). A crystal of Na-Y was treated with 0.05 M MgCl2 ·6H2O in the solvent at 333 K, followed by vacuum dehydration at 723 K and 1 × 10?6 Torr for 2 days. Its structure was determined by single-crystal synchrotron X-ray diffraction techniques, in the cubic space group \(Fd\overline{3} m\) at 100 K. It was refined to the final error indices R 1/wR 2 = 0.0587/0.2210 with 1,294 reflections for which F o > 4σ(F o). In the structure of |Mg34.5Na6|[Si117Al75O384]-FAU, 34.5 Mg2+ ions per unit cell are found at four different crystallographic sites: 15 per unit cell are located at site I at the center of the hexagonal prism [Mg–O = 2.216(2) Å], two are at site I’ in the sodalite cavity near the hexagonal prism [Mg–O = 2.20(3) Å], only one is located at site II’ in the sodalite cavity [Mg–O = 2.197(23) Å], and the remaining 16.5 are at site II near single 6-oxygen rings in the supercage [Mg–O = 2.103(3) Å]. The residual 6 Na+ ions per unit cell are found at site II [Na–O = 2.218(7) Å]. No water molecules are found in this structure.  相似文献   

17.
Na‐ion conducting Na1+x[SnxGe2?x(PO4)3] (x = 0, 0.25, 0.5, and 0.75 mol%) glass samples with NASICON‐type phase were synthesized by the melt quenching method and glass‐ceramics were formed by heat treating the precursor glasses at their crystallization temperatures. XRD traces exhibit formation of most stable crystalline phase NaGe2(PO4)3 (ICSD‐164019) with trigonal structure. Structural illustration of sodium germanium phosphate [NaGe2(PO4)3] displays that each germanium is surrounded by 6 oxygen atom showing octahedral symmetry (GeO6) and phosphorous with 4 oxygen atoms showing tetrahedral symmetry (PO4). The highest bulk Na+ ion conductivities and lowest activation energy for conduction were achieved to be 8.39 × 10?05 S/cm and 0.52 eV for the optimum substitution levels (x = 0.5 mol%, Na1.5[Sn0.5Ge1.5(PO4)3]) of tetrahedral Ge4+ ions by Sn4+ on Na–Ge–P network. CV studies of the best conducting Na1.5[Sn0.5Ge1.5(PO4)3] glass‐ceramic electrolyte possesses a wide electrochemical window of 6 V. The structural and EIS studies of these glass‐ceramic electrolyte samples were monitored in light of the substitution of Ge by its larger homologue Sn.  相似文献   

18.
The kinetic investigation of Hg(II)-promoted reaction between [Fe(CN)6]4− and 2,2′-bipyridine (Bipy) has been performed in anionic sodium dodecyl sulfate (SDS) micellar medium by recording the surge in absorbance at 400 nm, corresponding to ultimate reaction product [Fe(CN)4 Bipy]2− using UV–visible spectrophotometer. Pseudo-first-order condition has been used to examine the progress of reaction as a function of temperature, [Fe(CN)64−], ionic strength, [SDS], pH, [Hg2+], and [Bipy] by changing one parameter at a time. The results exhibit that [Hg2+], [SDS], and pH are the decisive parameter showing maximum reaction rate at 1.5 × 10−4 mol dm−3, 6.0 × 10−3 mol dm−3, and 3.8, respectively. [Fe(CN)6]4− does not show any appreciable effect on the critical micellar concentration (CMC) of SDS as the polar head of SDS and [Fe(CN)6]4− both are negatively charged. Variable order kinetics was observed for [Fe(CN)6]4− and Bipy in their examined concentration range. The reverse response observed in the reaction rate with [KNO3] shows a negative salt effect. The K+ provided by K4[Fe(CN)6] and KNO3 decreases the repulsion between the negatively charged heads of the surfactant molecules thereby decreasing the CMC of SDS. The negative value for the entropy of activation also supports the interchange dissociative (Id) mechanism recommended by us.  相似文献   

19.
Sodium rare earth borate glasses (Na2O)35.7(RE2O3)7.2(B2O3)57.1 (RE = Sm, Gd, Dy, Ho, Y, Er, and Yb), were prepared from a mixture of Na2CO3, RE2O3 and B2O3, and their properties as an Na+ ionic conductor were investigated. Density increased with increasing atomic weight of RE. Crystallization temperature and crystal melting temperature of the present borate system was lower than that of the previously reported silicate and germanate system. Results of the 11B NMR measurement suggested that half of all boron atoms are coordinated by four oxide ions to give a [BO4] tetrahedral unit and the others are coordinated by three oxide ions to give a [BO3] planar triangular unit. The electrical conductivity slightly decreased with increasing the ionic radius of RE3+. (Na2O)35.7(Y2O3)7.2(B2O3)57.1 glass exhibited the electrical conductivity which is about one order of magnitude lower than those of the previously reported (Na2O)35.7(Y2O3)7.2(SiO2)57.1 and (Na2O)35.7(Y2O3)7.2(GeO2)57.1 glasses. It was assumed that this lower electrical conductivity is due to the lower content of Na+ ions as conduction species in the former glass, compared with the latter two glasses.  相似文献   

20.
《Ceramics International》2022,48(11):15737-15747
Zn4O(BO2)6 based on the [B24O48] sodalite-cage structure fixed by the inside [Zn4O13] clusters is expected to be a new class of solid-state lighting material with perfect thermal and mechanical stability. Herein, in the current work, we have respectively introduced non-equivalent rare-earth cations Eu3+ and Pr3+ into Zn4O(BO2)6 host to design white and green emission materials by a novel solid-phase sintering method at lower temperatures. Zn2B6O11 replacing B2O3 or H3BO3 as raw materials can effectively avoid the impure products caused by the uncontrollable volatilization of B2O3 or H3BO3. The newly designed light-emitting materials of Zn4(1-x)O(BO2)6: xRe3+ (ReEu or Pr), including Zn4O(BO2)6 host, have good absorption capacity in the ultraviolet region. Under ultraviolet irradiation, Zn4O(BO2)6, Zn4(1-x)O(BO2)6: xEu3+and Zn4(1-x)O(BO2)6: xPr3+ emit the blue, white and green lights, respectively. In addition, all these materials can effectively degrade methylene blue, in which Zn4(1-x)O(BO2)6: xPr3+ has the highest efficiency. The luminescence and degradation mechanisms of Zn4O(BO2)6, Zn4(1-x)O(BO2)6: xEu3+and Zn4(1-x)O(BO2)6: xPr3+ have been adequately explained by their electronic structures based on the first principle calculations. The current study confirms that the doping of Eu3+/Pr3+ in Zn4O(BO2)6 can broaden its applications as photoluminescent and photocatalytic materials.  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