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 140 毫秒
1.
钙钛矿材料化学组分是决定钙钛矿太阳能电池效率和稳定性的关键,纯无机钙钛矿CsPbI3具有相对较好的热稳定性和光稳定性,但由于Cs+具有较小的离子半径而导致无机钙钛矿相不稳定。最近研究发现富铯FAxCs1?xPbI3钙钛矿具有相对稳定的相结构,且可以很大程度上保持无机钙钛矿材料的热稳定性和光照稳定性,是一种非常具有前景的钙钛矿材料体系。目前这种富铯的FAxCs1?xPbI3材料合成是通过引入过量有机组分FAI实现的,其中FAI一方面充当钙钛矿的掺杂剂,另一方面过量的FAI充当添加剂。由于其具有较高的升华温度,后续需要较高的温度使过量的FAI升华,实际上这在实验上很难实现对FAI升华量的精确控制。本文重点研究具有低升华温度的胺类,如碘甲胺(MAI)、碘化二甲胺(DMAI)、碘化乙胺(EAI)、碘化胺(NH4I)和醋酸甲脒(FAAC),作为添加剂制备富铯FAxCs1?xPbI3钙钛矿材料体系的可行性,这一方面可以有效降低钙钛矿薄膜的热处理温度;另一方面可拓宽的制备纯相钙钛矿成分的窗口期,这对大面积制备纯相富铯FAxCs1?xPbI3钙钛矿薄膜尤为重要。结果表明MAI和DMAI可以作为合成FAxCs1?xPbI3钙钛矿材料的有效添加剂,其与PbI2间较强的作用力可以促进Cs4PbI6的形成并有效抑制δ-CsPbI3副产物的生成。合适的升华温度可以使薄膜在保持钙钛矿相结构的同时在较低温度升华去除过量的添加剂,最终实现在相对温和的条件下制备纯相富铯FAxCs1?xPbI3钙钛矿材料。  相似文献   

2.
以MnO2, Ca(OH)2和La(OH)3为反应原料, 在惰性气氛、 低温(500 ℃)熔融KOH体系中合成了具有菱形钙钛矿结构Ca, K共掺杂的La0.64Ca0.25K0.11MnO3纳米材料, 并对Mn的价态及磁学性能等进行了讨论. X射线光电子能谱(XPS)分析结果表明, La0.64Ca0.25K0.11MnO3纳米材料中的Mn具有三重混合价态, 其零场冷却低温磁化率曲线表现出顺磁-铁磁转变, 居里温度(Tc)为280 K.  相似文献   

3.
采用静电纺丝法结合溶胶-凝胶技术制备了钙钛矿型La0.67Ba0.33MnO3微纳米纤维, 并利用差示扫描量热-热失重分析(DSC-TGA)、 X射线衍射(XRD)、 傅里叶变换红外光谱(FTIR)和扫描电子显微镜(SEM)等技术对产物进行了表征, 利用IR-2红外发射率测试仪测试了La0.67Ba0.33MnO3在280~370 K范围内的红外发射率. 结果表明, La0.67Ba0.33MnO3在600 ℃时已形成钙钛矿结构. 随着煅烧温度的升高, La0.67Ba0.33MnO3的形貌由纤维状向三维网络状转变, 并最终失去纤维形态. 在280~370 K范围内, La0.67Ba0.33MnO3微纳米纤维的红外发射率随温度升高而升高, 由0.564增加至0.689. 利用钙钛矿材料双交换理论解释了这一现象, 并进一步探讨了其在红外发射率可变材料中的应用前景.  相似文献   

4.
利用水热法制备了新颖的、漂浮型的膨胀珍珠岩(EP)负载BiFeO3(BiFeO3/EP)复合光催化材料。采用X射线衍射(XRD)、傅里叶变换红外光谱(FT-IR)、扫描电子显微镜(SEM)、透射电子显微镜(TEM)、X射线光电子能谱仪(XPS)和紫外可见漫反射(UV-Vis-DRS)对制备的复合材料进行了表征与分析。SEM和TEM结果清楚地表明BiFeO3纳米粒子已被负载到EP表面。与纯BiFeO3相比,BiFeO3/EP复合材料明显提高了对可见光的吸收能力,减小了带隙宽度,在可见光下对亚甲基蓝(MB)的降解表现出更强的光催化活性。其中,70%BiFeO3/EP复合材料对MB染料废水的光催化活性最高,其光催化反应一级速率常数是纯BiFeO3的2.2倍。由于质轻中空的特点,制备的BiFeO3/EP颗粒漂浮在液面上,有利于相分离和反应后光催化剂的回收。材料的重复性试验表明,复合材料在MB光降解过程中是相当稳定的。  相似文献   

5.
铁酸铋(BiFeO3)是一种典型的钙钛矿型氧化物,具有一定的可见光催化和多相类芬顿催化性能。在可见光照射且存在过氧化氢的情况下,BiFeO3可活化过氧化氢并产生强氧化性物种,这些物种会攻击污染物分子从而使其降解。但BiFeO3量子效率不高,光生电子和空穴容易复合;其活化过氧化氢的能力也有待提高。本文综述了近二十年来铁酸铋及其改性物作为可见光催化剂及多相类芬顿催化剂的研究进展,重点介绍了在制备过程中对其形貌的调控、贵金属沉积、离子掺杂、半导体复合、或负载于其他材料表面等改善其环境催化性能方法与效果,改性后发现BiFeO3的可见光催化和多相类芬顿催化性能均得到提升。最后,对铁酸铋复合催化剂未来的发展方向进行了展望。  相似文献   

6.
La1-xLixMnO3体系的输运特性及EPR研究   总被引:1,自引:1,他引:0  
利用固相反应法合成了掺杂锂的La1-xLixMnO3(x=0,0.10,0.15,0.20,0.25,0.30)钙钛矿氧化物.XRD测试表明,所有样品均为菱方结构.除x=0,0.3外,其它样品均随温度的降低在液氮温区可观察到从绝缘态到金属态的转变,其中x=0.15样品的转变温度最高为167K.在H=1T的磁场下,出现了负磁阻现象.EPR谱上的g=2.00信号与Mn3+和Mn4+组成的复合团簇有关.  相似文献   

7.
采用溶胶凝胶法制备了4种稀土莫来石型复合氧化物REMn2O5 (RE=Pr, Sm, Eu, Y)并负载贵金属Pd,对比了其结构特点和甲烷催化氧化性能。结果表明,Pd/SmMn2O5和Pd/YMn2O5具有优异的甲烷催化氧化性能,其T50分别为328和308℃,优于目前有报道的贵金属类甲烷氧化催化剂。XRD分析和DFT计算结果表明,稀土元素的离子半径对莫来石结构有明显影响,只有半径合适的离子(Sm, r=0.1079 nm; Eu, r=0.1066 nm)才容易形成莫来石相。Pr3+离子半径较大(r=0.1126 nm), PrMn2O5样品出现较多的钙钛矿型氧化物;Y3+离子半径较小(r=0.1019nm), YMn2O5样品中除莫来石相外同时还存在钙钛矿型氧化物和锰氧化物。稀土莫来石型复合氧化物中含...  相似文献   

8.
有机-无机杂化卤化铅钙钛矿因具有独特的电子和光学特性,已经成为光电领域最有前途的材料。但是,有机-无机钙钛矿材料及器件稳定性差,限制了其实际应用。与杂化钙钛矿相比,全无机卤化物钙钛矿CsPbX3(X=Cl,Br,I)显示出更强的热稳定性。全无机卤化物钙钛矿CsPbX3具有多个晶型,在不同的温度下呈不同相结构。目前,关于CsPbX3的结构和物理性质仍存在争议。本文我们针对三个晶相α-,β-和γ-CsPbX3的结构,热力学稳定性和电子性质进行了全面的理论研究。第一性原理计算表明,从高温α相到低温β相,然后再到γ相的相变伴随着PbX6八面体的畸变。零温形成能计算表明,γ相最稳定,这与实验中γ相为低温稳定相的结论一致。电子性质计算表明,所有CsPbX3钙钛矿都表现出直接带隙性质,并且带隙值从α相到β相再到γ相逐渐增加。这是由于相变发生时,Pb-X成键强度逐渐减弱,使价带顶能量降低,进而带隙增加。在所有相中,α相结构中较强的Pb-X相互作用,导致了较强的带边色散,使其具有较小的载流子有效质量。  相似文献   

9.
近年来,混合铅卤钙钛矿材料在光电领域引发的研究热潮引人注目。然而,钙钛矿材料对水和氧气的敏感性严重的阻碍了其实用化进程。在众多的稳定钙钛矿的方法中,利用简单的原子层沉积方法(Atomic layer deposition,ALD)在钙钛矿表面沉积一层保护层的技术具有极大的潜力。而ALD应用的困难在于,在常规的ALD过程中,做为氧源的H2O和O3对铅卤钙钛矿有着腐蚀作用。在本文,我们提出将双官能团的5-氨基戊酸(5-Aminovaleric acid,AVA)引入到CH3NH3PbBr3(MAPbBr3)钙钛矿晶格层中,形成稳定的铰链结构的2D/3D钙钛矿AVA(MAPbBr3)2。AVA的引入可以钝化并防止ALD过程中水对钙钛矿的侵蚀,从而成功地直接在钙钛矿表面沉积了Al2O3保护层。覆盖了保护层的AVA(MAPbBr3)2钙钛矿薄膜获得了优异的热稳定性和抗水性。  相似文献   

10.
钙钛矿具有优异的光学和电学性质, 近年来成为太阳能电池领域的研究热点. 大量实验报道钙钛矿热载流子弛豫时间变化顺序为CsPbBr3>MAPbBr3(MA=CH3NH3)>FAPbBr3[FA=HC(NH2)2], 但A位阳离子(Cs +, MA +, FA +)对弛豫快慢的影响机制仍不明确. 采用基于含时密度泛函理论的非绝热动力学方法研究了上述3种钙钛矿热电子和热空穴的能量弛豫动力学, 计算得到的热载流子弛豫时间与实验结果吻合. 结果表明, A位阳离子通过静电和氢键作用影响其与无机Pb—Br骨架的电子-振动耦合, 使非绝热耦合强度遵从FAPbBr3>MAPbBr3>CsPbBr3的变化趋势, 进而使热载流子弛豫时间尺度变化趋势与之相同, 表明合理选择A位阳离子可以优化钙钛矿太阳能电池的性能.  相似文献   

11.
The pK 2 * for the dissociation of sulfurous acid from I=0.5 to 6.0 molal at 25°C has been determined from emf measurements in NaCl solutions with added concentrations of NiCl2, CoCl2, McCl2 and CdCl2 (m=0.1). These experimental results have been treated using both the ion pairing and Pitzer's specific ion-interaction models. The Pitzer parameters for the interaction of M2+ with SO 3 2? yielded $$\begin{gathered} \beta _{NiSO_3 }^{(0)} = - 5.5, \beta _{NiSO_3 }^{(1)} = 5.8, and \beta _{NiSO_3 }^{(2)} = - 138 \hfill \\ \beta _{CoSO_3 }^{(0)} = - 12.3, \beta _{CoSO_3 }^{(1)} = 31.6, and \beta _{CoSO_3 }^{(2)} = - 562 \hfill \\ \beta _{MnSO_3 }^{(0)} = - 8.9, \beta _{MnSO_3 }^{(1)} = 18.7, and \beta _{MnSO_3 }^{(2)} = - 353 \hfill \\ \beta _{CdSO_3 }^{(0)} = - 7.2, \beta _{CdSO_3 }^{(1)} = 13.8, and \beta _{CdSO_3 }^{(2)} = - 489 \hfill \\ \end{gathered} $$ The calculated values of pK 2 * using Pitzer's equations reproduce the measured values to within ±0.01 pK units. The ion pairing model yielded $$\begin{gathered} logK_{NiSO_3 } = 2.88 and log\gamma _{NiSO_3 } = 0.111 \hfill \\ logK_{CoSO_3 } = 3.08 and log\gamma _{CoSO_3 } = 0.051 \hfill \\ logK_{MnSO_3 } = 3.00 and log\gamma _{MnSO_3 } = 0.041 \hfill \\ logK_{CdSO_3 } = 3.29 and log\gamma _{CdSO_3 } = 0.171 \hfill \\ \end{gathered} $$ for the formation of the complex MSO3. The stability constants for the formation of MSO3 complexes were found to correlate with the literature values for the formation of MSO4 complexes.  相似文献   

12.
Kinetics and equilibria for the formation of a 1:1 complex between palladium(II) and chloroacetate were studied by spectrophotometric measurements in 1.00 mol HClO4 at 298.2 K. The equilibrium constant, K, of the reaction
was determined from multi-wavelength absorbance measurements of equilibrated solutions at variable temperatures as log 0.006 with and , and spectra of individual species were calculated. Variable-temperature kinetic measurements gave rate constants for the forward and backward reactions at 298.2 K and ionic strength 1.00 mol as and , with activation parameters and , respectively. From the kinetics of the forward and reverse processes, and were derived in good agreement with the results of the equilibrium measurements. Specific Ion Interaction Theory was employed for determination of thermodynamic equilibrium constants for the protonation of chloroacetate () and formation of the PdL+ complex (). Specific ion interaction coefficients were derived.  相似文献   

13.
The temperature dependencies of europium carbonate stability constants were examined at 15, 25, and 35°C in 0.68 molal Na+(ClO 4 ? , HCO 3 ? ) using a tributyl phosphate solvent extration technique. Our distribution data can be explained by the equilibria $$\begin{gathered} Eu^{3 + } + H_2 O + CO_2 (g)_ \leftarrow ^ \to EuCO_3^ + + 2H^ + \hfill \\ - log\beta _{12} = 9.607 + 496(t + 273.16)^{ - 1} \hfill \\ Eu^{3 + } + 2H_2 O + 2CO_2 (g)_ \leftarrow ^ \to Eu(CO_3 )_2^ - + 4H^ + \hfill \\ - log\beta _{24} = 21.951 + 670(t + 273.16)^{ - 1} \hfill \\ Eu^{3 + } + H_2 O + CO_2 (g)_ \leftarrow ^ \to EuHCO_3^{2 + } + H^ + \hfill \\ - log\beta _{11} = 1.688 + 1397(t + 273.16)^{ - 1} \hfill \\ \end{gathered}$$   相似文献   

14.
The equilibrium constant for the hydrolytic disproportionation of I2
has been determined at 25°C and at ionic strength 0.2 M(NaClO4) in buffered solution. The reaction was followed in the pH range where the equilibrium concentration of I2, I, and IO3 are commensurable, i.e., the fast equilibrium
is also established. The equilibrium concentrations of I2and I3 were determined spectrophotometrically, and the concentrations of all the other species participating in process (1) were calculated from the stoichiometric constraints. The constants determined are \log K_1 = -47.61\pm 0.07 and \log K_2 = 2.86 \pm 0.01.  相似文献   

15.
The inorganic behavior of most divalent metals in natural waters is affected by the formation of carbonate complexes. The acidification of the oceans will lower the carbonate concentration in the oceans. This will increase the concentration of free copper that is toxic to marine organisms. To be able to determine the effect of this acidification, reliable stability constants are needed for the formation of copper carbonate complexes. In this paper, the speciation of Cu(II) with bicarbonate and carbonate ions
${rcl}&&\mathrm{Cu}^{2+}+\mathrm{HCO}_{3}^{-}\rightleftharpoons \mathrm{CuCO}_{3(\mathrm{aq})}+\mathrm{H}^{+}\\[4pt]&&\mathrm{Cu}^{2+}+2\mathrm{HCO}_{3}^{-}\rightleftharpoons \mathrm{Cu}(\mathrm{CO}_{3})_{2}^{2-}+2\mathrm{H}^{+}\\[4pt]&&\mathrm{Cu}^{2+}+\mathrm{CO}_{3}^{2-}\rightleftharpoons \mathrm{CuCO}_{3(\mathrm{aq})}\\[4pt]&&\mathrm{Cu}^{2+}+2\mathrm{CO}_{3}^{2-}\rightleftharpoons \mathrm{Cu}(\mathrm{CO}_{3})_{2}^{2-}\\[4pt]&&\mathrm{Cu}^{2+}+\mathrm{HCO}_{3}^{-}\rightleftharpoons \mathrm{CuHCO}_{3}^{+}$\begin{array}{rcl}&&\mathrm{Cu}^{2+}+\mathrm{HCO}_{3}^{-}\rightleftharpoons \mathrm{CuCO}_{3(\mathrm{aq})}+\mathrm{H}^{+}\\[4pt]&&\mathrm{Cu}^{2+}+2\mathrm{HCO}_{3}^{-}\rightleftharpoons \mathrm{Cu}(\mathrm{CO}_{3})_{2}^{2-}+2\mathrm{H}^{+}\\[4pt]&&\mathrm{Cu}^{2+}+\mathrm{CO}_{3}^{2-}\rightleftharpoons \mathrm{CuCO}_{3(\mathrm{aq})}\\[4pt]&&\mathrm{Cu}^{2+}+2\mathrm{CO}_{3}^{2-}\rightleftharpoons \mathrm{Cu}(\mathrm{CO}_{3})_{2}^{2-}\\[4pt]&&\mathrm{Cu}^{2+}+\mathrm{HCO}_{3}^{-}\rightleftharpoons \mathrm{CuHCO}_{3}^{+}\end{array}  相似文献   

16.
The oxidation of H2NOH is first-order both in [NH3OH+] and [AuCl4 ]. The rate is increased by the increase in [Cl] and decreased with increase in [H+]. The stoichiometry ratio, [NH3OH+]/[AuCl4 ], is 1. The mechanism consists of the following reactions.
The rate law deduced from the reactions (i)–(iv) is given by Equation (v) considering that [H+] K a.
The reaction (iii) is a combination of the following reactions:
The activation parameters for the reactions (ii) and (iii) are consistent with an outer-sphere electron transfer mechanism.  相似文献   

17.
The luminescence spectra of the polycrystalline compounds [Cr(CH2NH2COO)3 · H2O] and [Cr2(OH)2(CH2NH2COO)4] are investigated in the temperature range of 120K – 4.2K. From the known crystal structure (P21/c =D 2h /5 ) of the mononuclear compound assignment of the zero-phonon bands based on crystal field theory becomes possible. Both of the highly intense phosphorescence transitions are observed at \(P_1 = 14493 cm^{ - 1} ({}^2A'' \xrightarrow{{0.0}} {}^4A) and P_2 = 14428 cm^{ - 1} ({}^2A' \xrightarrow{{0.0}} {}^4A)\) . Assignment of the accompanying vibronic bands is made from the measured infrared data. Crystal field parameters Dq, B and C are determined from the luminescence and reflectance spectra. In the case of the binuclear compound the Cr3+-Cr3+ interaction via hydroxyl brides may be described by an axchange operator \(H_{ex} = - 2 \sum\limits_{ij} {J_{ij} S_i^a \cdot S_j^a } \) and from this the energy level diagram is calculated. Both observed strong phosphorescence bands at 14369 cm?1 and 14184 cm?1 are assigned to \(\left| {{}^2E \cdot {}^4A_2 \rangle _{s = 2} \xrightarrow{{0.0}}} \right| {}^4A_2 \cdot {}^4A_2 \rangle _{s = 2} and \left| {{}^2E \cdot {}^4A_2 \rangle _{s = 1} \xrightarrow{{0.0}}} \right| {}^4A_2 \cdot {}^4A_2 \rangle _{s = 1} \) transitions.  相似文献   

18.
The extraction kinetics of uranium(VI) and thorium(IV) with Tri-iso-amyl phosphate (TiAP) from nitric acid medium has been investigated using a Lewis Cell. Especially, dependences of the extraction rate on stirring speed, temperature, interfacial area were firstly measured to elucidate the extraction kinetics regimes. The experimental results demonstrated that extraction kinetic of U(VI) is governed by chemical reactions at interface with an activation energy, Ea, of 43.41 kJ/mol, while the rate of Th(IV) extraction is proved to be intermediate controlled, of which the Ea is 23.20 kJ/mol. Reaction orders with respect to the influencing parameters of the extraction rate are determined, and the rate equations of U(VI) and Th(IV) at 293 K have been proposed as $$ {\text{r}} = - {\text{dcUO}}_{ 2} \left( {{\text{NO}}_{ 3} } \right)_{ 2} /{\text{dt}} = 1. 80 \times 10^{ - 3} \left[ {{\text{UO}}_{ 2} \left( {{\text{NO}}_{ 3} } \right)_{ 2} } \right]^{ 1.0 1} \left[ {\text{TiAP}} \right]^{0. 5 5} , $$ $$ {\text{r}} = - {\text{dcTh }}\left( {{\text{NO}}_{ 3} } \right)_{ 4} /{\text{dt}} = 1. 8 8\times 10^{ - 3} \left[ {{\text{Th }}\left( {{\text{NO}}_{ 3} } \right)_{ 4} } \right]^{ 1.0 4} \left[ {\text{TiAP}} \right]^{ 1. 7 7} \left[ {{\text{HNO}}_{ 3} } \right]^{0. 3 8} , $$ respectively.  相似文献   

19.
The heat of solution of GaCl3 and heats of dilution of single GaCl3 solutions in water and of mixed GaCl3−HCl solutions in HCl solutions (with a fixed HCl concentration of 0.1337 mol-kg−1 HCl) up to 4 mol-kg−1 GaCl3 were measured at 25°C. While in the acid solutions hydrolysis is suppressed to below 0.5% of total gallium concentration, the measurements in water allow evaluation of the effect of hydrolysis on the relative enthalpy. The Pitzer interaction model for excess properties of aqueous electrolytes was used to interpret the change in relative enthalpy with concentration. Pitzer parameters were derived by statistical inference using ridge regression. Their physical significance is supported by the heat of solution data. The measurements yield the following results for standard heats of formation and Pitzer parameters for the relative molar enthalpy at 25°C: With these parameters the overall variance in the partial molar heat of solution at infinite dilution, extrapolated from the present experiments, is minimized to 0.35 kJ2-mol−2, while the experimental apparent molar heats of dilution are reproduced on average within 2.7 kJ-mol−1.  相似文献   

20.
The reaction between VV and TlI was studied in 4.0 mol dm–3 HCl at an ionic strength of 4.1 mol dm–3 at 25° C. The main active species under the reaction conditions were found to be VO inf2 sup+ and TlCl inf3 sup2– for the oxidant and reductant, respectively. A probable mechanism in terms of these species is given, and follows the rate law:
  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