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 15 毫秒
1.
In order to determine whether Li+ cations penetrate into the octahedral layers of montmorillonites upon mild heating (Hofmann-Klemen effect) 57Fe Mössbauer spectra of Na+ and Li+ exchanged montmorillonite were obtained before and after treatment at 220 ° C. The 57Fe nucleus was used as a remote probe to detect electronic perturbations which would occur if a Li cation was to move into the octahedral layer from the interlayer after heating. The ambient Mössbauer spectra showed that a high charge density interlayer cation such as Li+ is effective in reducing the phonon energy of VIFe2+. In addition the EFG at octahedral sites can be significantly modified by interlayer cations as evidenced by the larger quadrupole splitting value measured for the Li+-exchanged sample with respect to the Na+-sample. Interlayer collapse and migration of exchange cations into the montmorillonite lattice after heating to 220 ° C resulted in the oxidation of the VIFe2+ and a decrease in site distortion for IVFe3+. Similar spectral parameters for the Fe3+ resonances of both Na+ — and Li+-heated samples suggested the interlayer cations do not penetrate as far as the octahedral layers. In order to utilize the enhanced sensitivity of VIFe2+ Δ values to changes in EFG the Fe3+ in the heated montmorillonites was reduced to Fe2+ with hydrazine. Similar spectral parameters for both the Na+ — and Li+-exchanged montmorillonite were observed giving further evidence that Li cations do not migrate into vacant octahedral sites.  相似文献   

2.
The crystal structures of synthetic 7 angstrom and 10 angstrom manganates, synthetic birnessite and buserite, substituted by mono- and divalent cations were investigated by X-ray and electron diffractions. The monoclinic unit cell parameters of the subcell of lithium 7 angstrom manganate, which is one of the best ordered manganates, were obtained by computing the X-ray powder diffraction data: a = 5.152 angstroms, b = 2.845 angstroms, c = 7.196 angstroms, beta = 103.08 degrees. On the basis of the indices obtained by computing the X-ray diffraction data of Li 7 angstrom manganate, monovalent Na, K and Cs and divalent Be, Sr and Ba 7 angstrom manganates were interpreted as the same monoclinic structure with beta = 100-103 degrees as that of Li 7 angstrom manganate, from their X-ray diffraction data. In addition, divalent Mg, Ca and Ni 10 angstrom manganates were also interpreted as the same monoclinic crystal system with beta=90-94 degrees. The unit cell parameters, especially a, c and beta change possibly with the type of substituent cation probably because of the different ionic radius, hydration energy and molar ratio of substituent cation to manganese. However, these diffraction data, except for those of Sr and Ba 7 angstrom and Ca and Ni 10 angstrom manganates, reveal only some parts of the host manganese structure with the edge-shared [MnO6] octahedral layer. On the other hand, one of the superlattice reflections observed in the electron diffractions was found in the X-ray diffraction lines for heavier divalent cations Sr and Ba 7 angstrom and Ca and Ni 10 angtrom manganates. The reflection presumably results from the substituent cation position in the interlayer which is associated with the vacancies in the edge-shared [MnO6] layer and indicates that the essential vacancies are linearly arranged parallel to the b-axis. Furthermore, the characteristic superlattice reflection patterns for several cations, Li, Mg, Ca, Sr, Ba and Ni, manganates were interpreted that the substituent cations are regularly distributed in the interlayer according to the exchange percentage of substituent cation to Na+. In contrast, the streaking in the a-direction observed strongly in the electron diffractions for heavier monovalent cations, K and Cs, manganates probably results from the disordering of their cations in the a-direction in the interlayer.  相似文献   

3.
锂改性蒙脱石矿物中锂的化学状态研究   总被引:2,自引:0,他引:2  
采用XRD、IR、XPS等测试方法,对锂改性蒙脱石矿物中锂 的化学状态进行分析,发现锂改性过程中,Li+离子不仅与蒙脱石层间吸附阳离子进行交 换,还能置换蒙脱石晶格八面体中的阳离子和进入间隙位置。Li+离子是以三种化学状态 存在于蒙脱石晶格结构中。  相似文献   

4.
蒙脱石热处理产物的微结构变化研究   总被引:9,自引:0,他引:9       下载免费PDF全文
本文对广东和平蒙脱石及其热处理产物进行了化学分析、差热和热重分析、X射线粉末衍射分析、红外吸收光谱分析、扫描电镜、原子力显微镜及魔角旋转核磁共振等研究。结果表明,蒙脱石在热处理温度为126℃-148℃时,主要脱出吸附水和层间水,这一脱水过程是可逆的。当热处理温度达到659℃时,蒙脱石八面体片中的羟基开始脱失,但层状结构仍然保持,这种羟基的脱失过程对应着八面体片中Al向Al的转变。当温度达到900℃时,蒙脱石的层状结构完全被破坏,并有新的矿物相μ-堇青石产生。当温度为1200℃时,则出现方英石及莫来石相。当热处理温度达到1350℃时,方英石及莫来石的含量略有减少,并出现较多的含铁堇青石相。  相似文献   

5.
 Ni-saturated montmorillonite from Camp-Bertaux heated at different temperatures has been studied by X-ray powder diffraction, X-ray absorption (EXAFS) and vibration IR spectroscopy. Analysis of the experimental data has shown that heating of samples at temperatures higher than 150° C was accompanied by migration of Ni cations into vacant cis-octahedra of 2:1 layers. In the octahedral sheet the Ni cation has two “heavy” (Fe) and four “light” (Al and Mg) nearest octahedral cations. A model for the octahedral cation distribution in Camp-Bertaux montmorillonite was proposed in which Fe and Mg octahedral cations are segregated in small clusters. Received July 7, 1996 / Revised, accepted August 23, 1996  相似文献   

6.
蒙脱石晶体反常膨胀的电解质浓度效应   总被引:2,自引:0,他引:2  
谭罗荣 《矿物学报》2002,22(4):371-374
用不同浓度电解质溶液调制蒙脱石试样进行X射线衍射试验,发现蒙脱石晶体的d001晶面间距在试样失水至极低含水量过程中,存在着反向膨胀的现象,这种反常膨胀与制样时所用电解质溶液浓度有关,但不是因晶层间隔离子的代换造成的,也不是温度变化引起。  相似文献   

7.
蒙脱石层电荷与有机改性蒙脱石凝胶性能关系研究   总被引:2,自引:2,他引:2  
利用自然沉降法提取了山东两不同产地膨润土中的钙基蒙脱石矿物M1和M2,利用结构式推算法计算了两提纯蒙脱石的层电荷:M1单位半晶胞的层电荷为0.38,M2单位半晶胞的层电荷为0.59;利用十八烷基三甲基氯化铵对蒙脱石进行有机改性,并对有机蒙脱石凝胶性能和层电荷的相关关系进行研究。研究发现:蒙脱石单位半晶胞的层电荷越低,在水中的分散性越好,越有利于十八烷基三甲基铵阳离子的插层,相应有机改性蒙脱石在二甲苯和乙醇体系中的凝胶性能越好(如M1);蒙脱石单位半晶胞的层电荷越高,在水中的分散性越差,越不利于十八烷基三甲基铵阳离子的插层,相应有机改性蒙脱石产品的凝胶性能较差(如M2)。  相似文献   

8.
ABSTRACT

Previous study of subducted continental crust within the Luliang Shan terrane in Northwest China has documented metasomatic formation of thick, hydrated phengite + garnet-rich selvages at the interface between mafic eclogite blocks and quartzofeldspathic host gneiss. Whole rock concentrations of Cs and Ba within the selvage are enriched by two orders of magnitude relative to the eclogite blocks and host gneiss. We performed in situ ion microprobe analyses of Li, Be, B, Rb, Sr, Cs and Ba and δ11B of phengite within the Luliang Shane terrane to better constrain the source(s) of the infiltrating fluid. The phengite within the selvage are enriched in Li, Cs and Ba and yield δ11B values between ?30‰ and ?9‰, values that are lower than mantle values. High Ba/Rb, Cs/Rb coupled with low B/Be, B/Li and highly negative δ11B values indicate that the high-pressure fluid that formed the selvage was derived from highly devolatilized rocks within the subduction channel. In contrast, muscovite, which crystallized in the adjacent host gneiss during a subsequent lower pressure phase of fluid infiltration at approximately 0.9 GPa depths, has much lower Li, Cs and Ba relative to the high-pressure phengite. These retrograde muscovite have very high concentrations of B (up to 5500 ppm) and Be (up to 50 ppm) and high (?2 to +8‰) δ11B values that are consistent with crystallization from a fluid derived from shallower and less devolatilized regions of the subduction zone. Additional host gneiss samples, regionally distributed and kilometres away from the studied area lack the B-rich signature and indicate that the late stage fluids were likely localized to the region near the studied traverse.  相似文献   

9.
The radioactive fission product, 137Cs, has been observed to mobilize from bottom sediments of two South Carolina reservoirs during summer thermal stratification and hypolimnetic anoxia. Mobilization is attributed to ion-exchange displacement of 137Cs from sediments by cations such as NH+4, Fe+2 and Mn+2 released under anaerobic conditions.Three types of 137Cs binding sites to sediment clay minerals are identified: 1) surface and planar sites from which 137Cs is generally exchangeable by all cations studied (Na+, NH+4, H+, Cs+, Ca+2, Mg+2, Fe+2, and Mn+2); 2) wedge sites where 137Cs exchange is sterically limited to cations of similar size and charge (NH+4, Cs+, K+, and perhaps H3O+); 3) interlayer sites from which 137Cs is not readily exchanged. More than 15 years after final 137Cs inputs, the reservoir sediments we studied showed the following percentage distribution of sites: 2 to 9% surface sites, 6 to 13% wedge sites, and 78 to 85% interlayer sites. In contrast, lake and stream sediments near Oak Ridge, Tennessee receiving 137Cs inputs more than 20 years earlier had greater than 99% of their 137Cs associated with non-exchangeable interlayer sites. The difference is attributed to the paucity in the South Carolina sediments of weathered micaceous clay minerals with their abundant interlayer sites. Such interlayer deficient clays are dominant in the Atlantic and Gulf coastal plains of the United States and elsewhere. This suggests that 137Cs will be physically and chemically more mobile in such areas as well as more biologically available. Mobility will be enhanced in regimes where cation inputs favoring 137Cs exchange occur. Subsurface waste disposal sites where anaerobic conditions develop with NH+4 production and Fe+2 and Mn+2 release might be such a regime.  相似文献   

10.
采用γ-氨丙基三乙氧基硅烷(APTES),在乙醇-水混和溶剂中对预先经不同温度煅烧的蒙脱石进行改性。采用XRD、FTIR、热分析、元素分析、比表面积及孔分析等多种手段对产物进行分析。结果表明:硅烷主要赋存于蒙脱石层间,呈双层排布,少量嫁接于片层端面。热处理温度通过影响蒙脱石层间含水量,进而影响硅烷在层间的水解缩合。硅烷改性蒙脱石的过程为:硅烷分子通过阳离子交换插层至蒙脱石层间;随后水解生成的硅醇分子相互缩合;最终由于乙醇对硅烷水解的抑制作用以及蒙脱石片层相邻电荷位点存在一定的间距,与蒙脱石片层结合较弱的硅烷分子被洗脱,剩余的Si-O-Si交联网络在层间形成类似"柱子"结构。该硅烷改性蒙脱石与原始蒙脱石的比表面积相差不大,但其结构中微孔的比重增加。  相似文献   

11.
聚苯胺/蒙脱石纳米复合材料的实验制备研究   总被引:1,自引:0,他引:1       下载免费PDF全文
对聚苯胺 /蒙脱石纳米复合材料的制备方法与制备产物进行了研究。将新疆某地蒙脱石在钠化改型后进行了季铵盐改性处理。以苯胺作为客体物质 ,利用季铵盐 /蒙脱石插层复合物与有机物良好的相容性 ,将季铵盐 /蒙脱石插层复合物分散在苯胺单体中后 ,苯胺单体进入了蒙脱石层间域 ,通过原位聚合法制备了聚苯胺 /蒙脱石纳米复合材料。XRD、SEM、TG_DTA分析结果表明 ,苯胺单体的进入使蒙脱石的层间距大大增加 ,经聚合后聚苯胺 /蒙脱石复合物中蒙脱石及其插层复合物的X射线衍射特征、形态特征和热学性质特征完全消失 ,表明蒙脱石晶层已被剥离分散在聚苯胺中形成聚苯胺 /蒙脱石纳米复合材料。  相似文献   

12.
Drillcores and waters from Wairakei and Broadlands geothermal areas New Zealand have been analyzed for Li, Rb, Cs, Na, K, Mg, Ca, Al, Ti, Mn, and Be. The drillcores were altered to various degrees at temperatures below 300°C in slightly alkaline chloride water, probably derived from rock-water interaction in untapped horizons at higher temperature. It changes its composition as it leaches Ca and Na from the rock and adds K, Rb, Cs and Li. Evaluation of these changes in relation to the dimensions of the altered zone under observation suggests that a high mass ratio of water to rock (e.g., 100) and a period of up to 1 million years are responsible for the present stage of alteration.Increase of K and Rb in the altered rocks is a result of the formation of abundant adularia in addition to illite. The KRb ratio of the rock decreases during alteration but remains higher than that of the fluid. Only clay materials and zeolites that preferentially absorb Rb give slightly lower KRb ratios than the fluid.The mineral phases responsible for the uptake of lithium during alteration are chlorite (300 ppm Li) and quartz (up to 430 ppm Li). Li uptake in quartz is considered to be the mechanism by which Al-rich quartz crystallises from alumino-silicates. LiAl atomic ratios of 0.3–0.57 and Al concentrations up to 3000 ppm have been observed.Relatively small concentrations of Cs are found in potassic minerals (e.g. 10 ppm Cs in adularia, 44 ppm Cs in illite). However, 240 ppm Cs are found in wairakite from Wairakei equilibrated at 235°C. Lower equilibration temperatures may lead to higher cesium concentrations. This effect, in conjunction with a more concentrated hydrothermal fluid, could explain a content of 4500 ppm Cs in wairakite extracted from a drillcore taken in the El-Tatio geothermal field in Chile.  相似文献   

13.
 Cation partitioning data for coexisting muscovite and biotite are shown to be useful indicators of relative interlayer bond length/strength in these minerals. These data therefore provide a useful crystal-chemical perspective on relative mass-transfer kinetics of radiogenic isotopes, and account for the observation that biotite is generally less retentive of 40Ar and 87Sr than coexisting muscovite. Partitioning behavior of trace elements underscores three reasons why overall interlayer bonding in biotite is weaker than in muscovite. First, the preferences of large (Rb, Cs)+ in biotite and of small La3+ and Na+ in muscovite indicate a relatively spacious interlayer volume in biotite (suggesting a longer mean K−O bond). Second, the preference of interlayer vacancies in biotite (with some/all possibly H2O/H3O+-filled) suggests that its adjacent 2:1 sheets are connected by fewer interlayer bonds per unit cell than those of muscovite. Third, the relative exclusion of large Ba2+ from biotite despite its large interlayer sites is attributed to O−H bonds pointing into the interlayer cavity sub-normal to (001); (K+, Ba2+)-H+ repulsion thereby induced by the bare proton both destabilizes Ba2+ and weakens K−O bonds. In contrast, muscovite offers a more favorable electrostatic environment for Ba2+ substitution since its O−H bonds are directed into the vacant M 1 octahedral site sub-parallel to (001). This hypothesis is supported by the observation that progressive F(OH)−1 exchange enhances Ba2+ partitioning into biotite/phlogopite relative to coexisting muscovite. These crystal-chemical differences between biotite and muscovite are mirrored in calculated values of “ionic porosity”, Z i , defined here as the percentage of their interlayer unit-cell volume not occupied by ions. A monitor of ionic packing density and geometry, Z i is inversely correlated with K−O bond strength, which appears to be the rate-determining “kinetic common denominator” for a variety of processes affecting micas – including those responsible for loss of radiogenic isotopes in biotite and muscovite. Accordingly, the relatively longer/weaker K−O bonds of biotite are envisioned as being more easily stretched (during volume diffusion) or broken (during recrystallization or retrograde alteration). This in turn accounts for common observations of enhanced radiogenic Ar/Sr loss and younger 40Ar/39Ar and Rb/Sr ages in natural biotite (high Z i ) relative to coexisting muscovite (lower Z i ). Significantly, this pattern may arise irrespective of isotopic loss mechanism (diffusion or recrystallization, etc.), and it follows that any age discordance observed between muscovite and biotite cannot be ascribed uniquely to one mechanism or the other without appropriate field, petrographic, and petrologic constraints. Extension of this partitioning/porosity-based synthesis leads to prediction of corollary age-retentivity-composition effects among chemically diverse trioctahedral and dioctahedral micas, which are best field tested in terranes that cooled slowly under dry, static conditions. Pressure effects on argon retention are also inferred from the porosity model. Received: 9 February 1995 / Accepted: 8 September 1995  相似文献   

14.
 The crystal structure of a synthetic Rb analog of tetra-ferri-annite (Rb–TFA) 1M with the composition Rb0.99Fe2+ 3.03(Fe3+ 1.04 Si2.96)O10.0(OH)2.0 was determined by the single-crystal X-ray diffraction method. The structure is homooctahedral (space group C2/m) with M1 and M2 occupied by divalent iron. Its unit cell is larger than that of the common potassium trioctahedral mica, and similar lateral dimensions of the tetrahedral and octahedral sheets allow a small tetrahedral rotation angle α=2.23(6)°. Structure refinements at 0.0001, 1.76, 2.81, 4.75, and 7.2 GPa indicate that in some respects the Rb–TFA behaves like all other micas when pressure increases: the octahedra are more compressible than the tetrahedra and the interlayer is four times more compressible than the 2:1 layer. However, there is a peculiar behavior of the tetrahedral rotation angle α: at lower pressures (0.0001, 1.76, 2.81 GPa), it has positive values that increase with pressure [from 2.23(6)° to 6.3(4)°] as in other micas, but negative values −7.5(5)° and −8.5(9)° appear at higher pressures, 4.75 and 7.2 GPa, respectively. This structural evidence, together with electrostatic energy calculations, shows that Rb–TFA has a Franzini A-type 2:1 layer up to at least 2.81 GPa that at higher pressure yields to a Franzini B-type layer, as shown by the refinements at 4.75 and 7.2 GPa. The inversion of the α angle is interpreted as a consequence of an isosymmetric displacive phase transition from A-type to B-type structure between 2.81 and 4.75 GPa. The compressibility of the Rb–TFA was also investigated by single-crystal X-ray diffraction up to a maximum pressure of 10 GPa. The lattice parameters reveal a sharp discontinuity between 3.36 and 3.84 GPa, which was associated with the phase transition from Franzini-A to Franzini-B structure. Received: 21 October 2002 / Accepted: 25 February 2003  相似文献   

15.
The structure of deuterated portlandite, Ca(OD)2, has been investigated using time-of-flight neutron diffraction at pressures up to ∼4.5 GPa and temperatures up to ∼823 K. Rietveld analysis of the data reveals that with increasing pressure, unit-cell parameter c decreases at a rate about 4.5 times larger than that for a, which is largely due to rapid contraction of the interlayer spacing in this pressure range. Fitting of the determined cell volumes to the third-order Birch–Murnaghan equation of state yields a bulk modulus (K 0) of 32.2 ± 1.0 GPa and its first derivative (K 0′) of 4.4 ± 0.6. Moreover, on compression, hydrogen-mediated interatomic interactions within the interlayer become strengthened, as reflected by decreases in interlayer D···O and D···D distances with increasing pressure. Correspondingly, D–D, the distance between the three equivalent sites over which D is disordered, increases, suggesting a pressure-induced hydrogen disorder. This behavior is similar to that reported in brucite at elevated pressure. On heating at ∼2.1 GPa, cell parameter c increases more rapidly than a, as expected. However, because of the pressure effect, the thermal expansion coefficients, particularly along c, are much smaller than those at ambient pressure. With increasing temperature, the three partially occupied D sites become further apart, and the D-mediated interactions, mainly the interlayer D···D repulsion, become weakened.  相似文献   

16.
在水体系中,蒙脱石矿物表面结合水膜厚度大小,取决于矿物层间阳离子及矿物的晶体化学特征.当层间阳离子种类相同时,首先取决于水与矿物结构参数的不适应程度.结合水膜厚度,影响蒙脱石层间阳离子在土—水体系中的电导率.本文通过对不同地区、不同类型蒙脱石—水体系电导率测定,讨论水与蒙脱石晶胞参数不适应度对体系电导率的影响规律.  相似文献   

17.
颗石藻是海洋中广泛分布的超微型浮游藻,经生物矿化作用形成的碳酸钙质颗石,在古海洋学研究中具有重要意义。海洋粘土矿物与有机质的有机-无机相互作用在全球碳循环中扮演着重要角色。本文选取广泛分布于海洋的赫氏颗石藻Emiliania huxleyi与海洋粘土矿物中具有代表性的伊利石和蒙脱石共培养。通过对颗石藻生长曲线和Sr/Ca、Mg/Ca元素比值、颗石藻与粘土矿物样品的紫外可见光吸收光谱、红外吸收光谱和矿物物相等分析,研究海洋粘土矿物与颗石藻的相互作用规律。通过研究表明伊利石对颗石藻的影响较小,蒙脱石因对营养元素的吸附和颗石藻的絮凝作用对颗石藻的生长和Sr/Ca、Mg/Ca元素比值影响较大。颗石藻代谢分泌的生物分子未能通过层间插层作用进入伊利石层间,颗石藻分泌的生物分子可通过插层作用进入并储存于蒙脱石层间,海洋粘土矿物中的蒙脱石与海洋微生物的相互作用值得地球微生物家关注,可能有助于对古海洋环境的认识。  相似文献   

18.
为明确可尔因矿田岩浆演化过程中锂的迁移与富集,指明稀有金属矿床的找矿方向,文章通过野外地质观察、室内岩矿鉴定及云母LA-ICP-MS原位测试,发现随着岩浆分异程度的增强,云母由镁铁质黑云母演化为硅铝白云母,以云母Mg~#/n(Li)值作为岩体分异程度指示,显示其与亲铁元素组合(V、Cr、Co、Ni)、稀碱金属(Li、Rb、Cs)、稀有金属(Nb、Ta、Sn)及稀散元素(Ga)和指示岩浆演化的元素对(Nb/Ta、K/Rb)均存在显著的相关性。根据岩体结晶环境、源区类型及云母结晶时对应熔(流)体中的Li含量,将可尔因矿田岩浆-热液活动划分为3期次7阶段,进而指出矿田内第三期次一、二阶段岩浆热液活动为Li迁移与富集的主要阶段。该过程中深源Li成矿元素和热量持续供给、早阶段岩体黑云母分解、三叠系围岩混染和上侵流体圈闭促成了富Li熔(流)体的形成与就位,表明矿田内及外围寻找岩浆型锂矿和三叠系层控富锂矿体的潜力。结合Li和其他稀有金属的相关性,指明了Sn、Nb、Ta等稀有金属元素的找矿方向。  相似文献   

19.
近年来美国和智利等国家新发现了大量赋存在粘土岩中的锂资源,该类锂资源主要是由火山灰蚀变而成。火山灰沉积水解形成的绿豆岩,广泛分布于我国西南地区,其是否存在锂的富集现象值得研究。笔者对重庆铜梁地区绿豆岩开展了调查研究,通过X射线荧光光谱仪和等离子质谱仪分析,发现绿豆岩中钾的平均含量为8.77%,锂的含量可达663×10~(-6),相当于0.14%的Li_2O含量,远高于固体、露采盐类矿的边界品位0.06%;稀土元素总量可以达到500×10~(-6),接近离子吸附型矿的边界品位。X射线衍射分析显示绿豆岩主要含有石英、长石和粘土矿物,其中粘土的主要成分为伊蒙混层和少量伊利石。伊蒙混层含量占比高的样品较占比低的样品Li的含量偏高。结合粘土及锂的物理化学特性,推断绿豆岩中的Li呈离子形态被粘土矿物吸附。如果这些锂资源能够被综合利用,势必将会产生巨大的经济效益。  相似文献   

20.
TDTMA+-柱撑蒙脱石吸附硝基苯的实验研究   总被引:3,自引:0,他引:3  
以阳离子表面活性剂(十四烷基三甲基溴化铵,TDTMBA)为柱撑剂,在不同浓度条件下(0.2 CEC~2.5 CEC)制备了一系列柱撑蒙脱石,并通过XRD对有机蒙脱石进行表征,研究季铵盐离子在层间的排列结构.同时针对TDTMA -柱撑蒙脱石对硝基苯的吸附进行研究.实验结果表明,吸附去除率受硝基苯的初始浓度、反应时间等因素的影响,而与pH值没有太大的关系.经过柱撑改性后,粘土对硝基苯的吸附能力较原土有了明显的提高.随着层间柱撑浓度的增大,柱撑蒙脱石对硝基苯的吸附量也增大.另外钙基蒙脱石和平卧双层,倾斜单层以及倾斜双层的结构的TDTMA -柱撑蒙脱石对硝基苯的吸附等温线符Freundlich和Langmuir等温方程,假三层结构的TDTMA -柱撑蒙脱石对硝基苯的吸附等温线符合Linear等温方程,这表明前者是以表面吸附为主,后者是以分配作用为主.  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