首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到18条相似文献,搜索用时 171 毫秒
1.
作为表生土壤环境中易生成且分布广泛的氧化锰矿物,水锰矿(γ-MnOOH)能参与铁氧化物的生成过程,影响Fe_(2+)的迁移、转化和归趋。本文考察了pH值为3.0~7.0的模拟水溶液体系中水锰矿与Fe_(2+)的相互作用及其生成铁氧化物的过程,分析了Fe_(2+)浓度、pH值和空气(氧气)对Fe(Ⅲ)氧化物晶体结构类型、化学组成和反应速率的影响。研究结果表明,水锰矿氧化Fe_(2+)产物主要为针铁矿和纤铁矿;pH值为3.0~5.0时产物为针铁矿,而pH值为7.0时产物为针铁矿与纤铁矿的混合物,且高浓度Fe_(2+)会促使纤铁矿生成;引入空气利于针铁矿生成;反应速率随着pH值升高、氧气分压的增大而加快。本工作丰富了对铁氧化物在常见锰氧化物表面形成和转化过程的认识。  相似文献   

2.
氧化铁矿物对五氯苯酚表面吸附实验及其反应模式   总被引:9,自引:2,他引:7  
批量法实验研究表明,五氯苯酚 (PCP)在合成的针铁矿、纤铁矿和赤铁矿表面吸附的浓度等温线为饱和型特征,可用 Langmuir公式拟合,而 pH吸附等温线为峰型曲线,峰值在 pH≈ 5. 矿物表面化合态和 PCP溶液离子化合态分析与计算表明, PCP在氧化铁矿物表面存在两种吸附反应模式,即表面静电吸附反应和表面络合吸附反应,表面静电吸附反应常数比表面络合吸附反应常数高 1~ 2个数量级.并进一步从反应机制上证明了憎水型可离解有机化合物 (HIOCs)在矿物表面吸附的模式与趋向性.  相似文献   

3.
骆少勇  周跃飞  刘星 《地学前缘》2020,27(5):218-226
通过在滇池开展原位实验,研究探讨了湖泊沉积物中磷灰石制约水铁矿分解和转化的机制,以及二者共存时的环境效应。结果表明:将水铁矿放置到沉积物中1个月,矿物保持稳定;放置时间达到3个月时,添加磷灰石实验中水铁矿发生了显著物相转变。冬天(12—2月)实验中,转化产物随深度的变化趋势为针铁矿+磁(赤)铁矿→针铁矿+纤铁矿→针铁矿;夏天(6—9月)实验中,转化产物随深度的变化趋势为针铁矿+纤铁矿+磁(赤)铁矿→针铁矿+纤铁矿→未转化。透射电镜分析结果显示冬天实验中生成的磁性铁氧化物为纳米磁铁矿和磁赤铁矿,夏天实验中产生的则主要为纳米磁铁矿。X射线光电子能谱分析结果显示冬天表层实验样品具有较高P含量。分析表明的湖泊沉积物中磷灰石促进水铁矿转化的过程为:(1)微生物促进磷灰石溶解;(2)磷灰石溶解释放的P促进铁还原菌生长;(3)铁还原菌促进水铁矿还原;(4)水铁矿还原产生的溶解态Fe2+催化水铁矿向针铁矿、纤铁矿和磁铁矿转化。冬天及沉积氧化-还原界面最适宜磷灰石分解菌和铁还原菌生长,水铁矿的转化和P释放能力也更强,相应地内源磷释放的风险也更大。  相似文献   

4.
磁黄铁矿催化分解苯酚反应动力学及其产物特征   总被引:2,自引:0,他引:2  
本文用批量实验法研究了天然磁黄铁矿在过氧化氢参与下催化分解苯酚的动力学反应。结果表明,在pH=3.81~5.88时,苯酚都可被有效的分解,分解速率k=4.0~212h-1(g/L)-1,高于氧化铁矿物类催化剂,可与Fenton试剂相比拟。同时它具有反应速率可控、催化剂与反应产物易于分离和回收等优点。对反应液中Fe(II)、Fe(Ⅲ)浓度和紫外光谱分析表明,反应过程与Fenton反应类似,苯酚先是被快速转化成多酚类化合物,接着被分解成羧酸类化合物。溶液pH不同,产物不同,总有机碳(TOC)矿化率也不同,一般可达50%~58%。  相似文献   

5.
磁铁矿和赤铁矿是自然界铁氧化物的两种主要存在形式,也是弓长岭铁矿区的主要矿石矿物,二者之间的转化曾经 被认为是氧化还原反应的结果。文中根据近几年提出的非氧化还原反应成矿理论,对弓长岭铁矿区内磁铁矿/赤铁矿之间的 转化关系进行新的解释。通过对弓长岭矿区矿石样品进行偏光显微镜和扫描电镜背散射等实验研究,发现了赤铁矿交代磁 铁矿、针铁矿交代赤铁矿、黄铁矿与磁铁矿、赤铁矿共生等现象。结合前人研究成果,从矿物组合、矿石结构以及矿物转 化前后体积变化等方面,论证了部分后生的赤铁矿是在缺氧的环境下由磁铁矿经非氧化反应转变而成,为该区后生赤铁矿 的形成现象提供了一种新的解释。  相似文献   

6.
本研究以石英砂为载体,在其颗粒表面合成铁胶膜,并探讨合成体系不同初始pH(5,6,7)及铁摩尔比(R=[Fe(Ⅱ)]/[Fe(Ⅲ)])等条件对铁胶膜形成的影响。研究表明,在初始pH相同的情况下,当R为0时,铁胶膜的矿物成分为弱晶质的水铁矿;R为0.01时的矿物成分为赤铁矿;R分别为0.02、0.04、0.06和0.10时为针铁矿,且随着R增加,X射线衍射(XRD)图谱中针铁矿的峰强度逐渐增加,扫描电镜(SEM)可观察到针铁矿晶形逐渐变大,且在R为0.10时晶体形貌最大;当R为0.50和体系只加入Fe(Ⅱ)时合成的铁矿物主要为针铁矿与磁铁矿的混合物。当R一定时,随着合成体系初始pH的增加,胶膜中针铁矿的XRD峰强度逐渐增强,在初始pH为7时其峰最强,且晶形逐渐变大。  相似文献   

7.
本文实验研究了希瓦氏奥奈达菌株(Shewanella oneidensis MR-1,以下简称MR-1)在pH为中性的厌氧条件下还原针铁矿的过程,探讨了MR-1菌异化还原针铁矿的动力学特征。采用邻菲罗啉分光光度法检测了反应前后溶液中铁含量的变化,利用扫描电子显微镜、粉晶X射线衍射和激光拉曼光谱分析了针铁矿及其还原产物的形貌特征和物相组成。结果表明,针铁矿在厌氧条件下可被MR-1还原,生成磁铁矿、菱铁矿等次生矿物。本文认为针铁矿的微生物异化还原过程以直接接触机制为主,同时存在间接还原机制;溶液中的Fe2+与CO32-、SO42-等沉淀生成菱铁矿等次生产物,同时部分Fe2+、Fe3+离子可吸附于矿物表面,甚至能引起矿物相的转化,两者共同构成了针铁矿的次生分解路径。  相似文献   

8.
制备了不同阳离子掺杂改性的针铁矿,并采用XRD、红外、热重差热分析以及TEM等手段对其进行了表征,结果显示,掺入Mn~(2+)、Cr~(3+)、Al~(3+)后并未明显改变α-FeOOH的晶体结构类型,说明部分阳离子掺杂进入α-FeOOH晶格,从而分别形成因Fe~(3+)被Mn~(2+)、Cr~(3+)、Al~(3+)部分取代的固溶体;红外分析也得到类似结果;热重和差热分析以及TEM观察结果均表明几种金属阳离子没有形成氧化物相,可以确定这些离子已经进入针铁矿的晶格;紫外-可见漫反射光谱分析发现Mn~(2+)、Cr~(3+)掺杂者的能带隙较之纯相针铁矿的稍小,依次为2.18 e V和2.24 e V;而Al~(3+)掺杂者的能带隙与纯相针铁矿相比有所提高,增加至2.34 e V。此外还考查了不同阳离子掺杂及纯相针铁矿与Fe(Ⅱ)构成的复合系统对邻硝基苯酚(2-NP)的还原效果,研究表明,在溶液pH=6和25℃等的条件下,上述还原转化反应符合准一级动力学方程。其中,Fe(Ⅱ)/Al~(3+)掺杂针铁矿复合系统对2-NP的降解效果最好,在120 min时就达到了100%去除率,在一定浓度范围内,随着2-NP浓度的升高,2-NP的还原速率降低。而且,Fe(Ⅱ)/Al~(3+)-针铁矿复合系统速率常数(k)随着溶液的pH值增大而升高。  相似文献   

9.
为了考察铁锰氧化物对酚类污染物的氧化降解能力,采用天然以及合成的铁锰氧化物对苯酚的氧化降解进行对比实验研究。土壤中铁锰氧化物样品分别为天然针铁矿及氧化锰,合成铁锰氧化物样品分别为合成针铁矿及软锰矿。结果表明:苯酚与铁锰氧化物发生氧化还原作用时,还可能与土壤中杂质发生吸附等作用;铁锰氧化物还原反应强度随着反应介质pH值的升高而迅速下降;可用零级反应动力学方程拟合铁氧化物还原溶解反应,针铁矿溶解反应的强度与介质的pH值呈负相关关系;天然针铁矿对酚类污染物的氧化降解能力明显高于合成针铁矿,pH值对天然针铁矿溶解反应影响较大;可用一级指数衰减方程拟合锰氧化物还原溶解反应,锰氧化物溶解反应的强度与介质的pH值呈指数衰减关系;pH值对软锰矿还原溶解反应的影响大于对土壤中氧化锰的影响,pH值越小,影响越显著;对比pH值对铁和锰还原作用的影响发现,在pH=6.5时,锰氧化物仍有较强的氧化性能。  相似文献   

10.
以合成针铁矿为原料,通过煅烧法获得比表面积分别为85.74和22.65 m2/g的多孔纳米赤铁矿和磁赤铁矿,通过静态实验探究了针铁矿和煅烧产物的Sb(Ⅲ)吸附性能.结果 表明,Sb(Ⅲ)吸附效率为赤铁矿>磁赤铁矿>针铁矿,其前二者效率显著高于针铁矿.Sb(Ⅲ)在3种矿物表面的吸附均为快速的化学吸附,吸附在2h内即可接近平衡,符合准二级动力学反应和Freundlich等温吸附模型,为自发进行的吸热反应,升高温度有利于反应的进行.在45 ℃、pH=7的条件下,针铁矿、赤铁矿和磁赤铁矿的最大吸附量分别可达16.04、50.44和33.53 mg/g.pH对赤铁矿和磁赤铁矿的Sb(Ⅲ)吸附效率影响不大,但pH升高会导致针铁矿的吸附能力降低.CO32-、SiO44-、PO43-和胡敏酸会与Sb(Ⅲ)竞争吸附位,抑制3种矿物对Sb(Ⅲ)的吸附,但这种抑制作用只在阴离子浓度较高的条件下有效.研究认为磁赤铁矿具有更多的表面活性位和较强的磁回收能力,是优于针铁矿和赤铁矿的含Sb(Ⅲ)废水处理材料.  相似文献   

11.
Ferrihydrite (2.5 Fe2O2-4.5 H2O) is an unstable colloidal mineral. It dissolves in highly alkaline solutions and is precipitated from them in the form of goethite. Jarosite is stable at very low pH but is decomposed at higher values of pH with separation of iron oxides. Experiments show that in rapid decomposition of jarosite a protohematite substance, ferrihydrite, is formed. This transformation occurs at moderate pH values when solutions percolate through the aggregates of jarosite. Ferrihydrite, an unstable colloidal hydrated oxide of ferric iron, changes spontaneously to stable hematite with time. Very slow decomposition of jarosite results in its replacement by iron hydroxide, goethite. Under laboratory conditions in alkaline solutions lepidocrocite may be obtained from jarosite. The synthesis of this iron hydroxide passes through a stage of intermediate products: ferrihydrite and hydrated ferric oxide - ferriprotolepidocrocite, formed by solution of ferrihydrite in strongly alkaline solutions. The transformation of ferriprotolepidocrocite into lepidocrocite may be regarded as a topotactic reaction. —Authors.  相似文献   

12.
The reaction between dissolved sulfide and synthetic iron (oxyhydr)oxide minerals was studied in artificial seawater and 0.1 M NaCl at pH 7.5 and 25°C. Electron transfer between surface-complexed sulfide and solid-phase Fe(III) results in the oxidation of dissolved sulfide to elemental sulfur, and the subsequent dissolution of the surface-reduced Fe. Sulfide oxidation and Fe(II) dissolution kinetics were evaluated for freshly precipitated hydrous ferric oxide (HFO), lepidocrocite, goethite, magnetite, hematite, and Al-substituted lepidocrocite. Reaction kinetics were expressed in terms of an empirical rate equation of the form:
  相似文献   

13.
The behaviour of trace amounts of arsenate coprecipitated with ferrihydrite, lepidocrocite and goethite was studied during reductive dissolution and phase transformation of the iron oxides using [55Fe]- and [73As]-labelled iron oxides. The As/Fe molar ratio ranged from 0 to 0.005 for ferrihydrite and lepidocrocite and from 0 to 0.001 for goethite. For ferrihydrite and lepidocrocite, all the arsenate remained associated with the surface, whereas for goethite only 30% of the arsenate was desorbable. The rate of reductive dissolution in 10 mM ascorbic acid was unaffected by the presence of arsenate for any of the iron oxides and the arsenate was not reduced to arsenite by ascorbic acid. During reductive dissolution of the iron oxides, arsenate was released incongruently with Fe2+ for all the iron oxides. For ferrihydrite and goethite, the arsenate remained adsorbed to the surface and was not released until the surface area became too small to adsorb all the arsenate. In contrast, arsenate preferentially desorbs from the surface of lepidocrocite. During Fe2+ catalysed transformation of ferrihydrite and lepidocrocite, arsenate became bound more strongly to the product phases. X-ray diffractograms showed that ferrihydrite was transformed into lepidocrocite, goethite and magnetite whereas lepidocrocite either remained untransformed or was transformed into magnetite. The rate of recrystallization of ferrihydrite was not affected by the presence of arsenate. The results presented here imply that during reductive dissolution of iron oxides in natural sediments there will be no simple correlation between the release of arsenate and Fe2+. Recrystallization of the more reactive iron oxides into more crystalline phases, induced by the appearance of Fe2+ in anoxic aquifers, may be an important trapping mechanism for arsenic.  相似文献   

14.
Iron oxides may undergo structural transformations when entering an anoxic environment. These transformations were investigated using the isotopic exchange between aqueous Fe(II) and iron oxides in experiments with 55Fe-labelled iron oxides. 55Fe was incorporated congruently into a ferrihydrite, two lepidocrocites (#1 and #2), synthesised at 10°C and 25°C, respectively, a goethite and a hematite. The iron oxides were then submerged in Fe2+ solutions (0-1.0 mM) with a pH of 6.5. In the presence of aqueous Fe2+, an immediate and very rapid release of 55Fe was observed from ferrihydrite, the two lepidocrocites and goethite, whereas in the absence of Fe2+ no release was observed. 55Fe was not released from hematite, even at the higher Fe2+ concentration. The release rate is mainly controlled by characteristics of the iron oxides, whereas the concentration of Fe2+ only has minor influence. Ferrihydrite and 5-nm-sized lepidocrocite crystals attained complete isotopic equilibration with aqueous Fe(II) within days. Within this timeframe ferrihydrite transformed completely into new and more stable phases such as lepidocrocite and goethite. Lepidocrocite #2 and goethite, having larger particles, did not reach isotopic equilibrium within the timeframe of the experiment; however, the continuous slow release of 55Fe suggests that isotopic equilibrium will ultimately be attained.Our results imply a recrystallization of solid Fe(III) phases induced by the catalytic action of aqueous Fe(II). Accordingly, iron oxides should properly be considered as dynamic phases that change composition when exposed to variable redox conditions. These results necessitate a reevaluation of current models for the release of trace metals under reducing conditions, the sequestration of heavy metals by iron oxides, and the significance of stable iron isotope signatures.  相似文献   

15.
Bioreduced anthraquinone-2,6-disulfonate (AH2DS; dihydro-anthraquinone) was reacted with a 2-line, Si-substituted ferrihydrite under anoxic conditions at neutral pH in PIPES buffer. Phosphate (P) and bicarbonate (C); common adsorptive oxyanions and media/buffer components known to effect ferrihydrite mineralization; and Fe(II)aq (as a catalytic mineralization agent) were used in comparative experiments. Heterogeneous AH2DS oxidation coupled with Fe(III) reduction occurred within 0.13-1 day, with mineralogic transformation occurring thereafter. The product suite included lepidocrocite, goethite, and/or magnetite, with proportions varing with reductant:oxidant ratio (r:o) and the presence of P or C. Lepidocrocite was the primary product at low r:o in the absence of P or C, with evidence for multiple formation pathways. Phosphate inhibited reductive recrystallization, while C promoted goethite formation. Stoichiometric magnetite was the sole product at higher r:o in the absence and presence of P. Lepidocrocite was the primary mineralization product in the Fe(II)aq system, with magnetite observed at near equal amounts when Fe(II) was high [Fe(II)/Fe(III)] = 0.5 and P was absent. P had a greater effect on reductive mineralization in the Fe(II)aq system, while AQDS was more effective than Fe(II)aq in promoting magnetite formation. The mineral products of the direct AH2DS-driven reductive reaction are different from those observed in AH2DS-ferrihydite systems with metal reducing bacteria, particularly in presence of P.  相似文献   

16.
We examine the possibility that crystalline hematite (α-Fe2O3) deposits on Mars were derived from the precursor iron oxyhydroxide minerals akaganéite (β-FeOOH) or lepidocrocite (γ-FeOOH) and compare them to an earlier study of goethite (α-FeOOH) and magnetite (Fe3O4) precursors. Both the mid-infrared and visible/near infrared spectra of hematite are dependent upon the hematite precursor mineral and the temperature of transformation. Laboratory spectra are compared to spectra from the Mars Global Surveyor Thermal Emission Spectrometer (MGS-TES) and the Mars Exploration Rover (MER) Opportunity Mini-TES and Pancam experiments, allowing us to infer the formation environment of Martian crystalline hematite deposits. Akaganéite and lepidocrocite readily transform to hematite at temperatures of 300 and 500°C, respectively. The visible/near-infrared and mid-infrared spectra of akaganéite-derived hematite are poor matches to data returned from TES, Mini-TES, and Pancam. The spectra of lepidocrocite-derived hematite are slightly better fits, but previously published spectra of goethite-derived hematite still represent the best match to MGS and MER spectral data. The experiments demonstrate that hematite precursor mineralogy, temperature of formation, and crystal shape exert a strong control on the hematite spectra.  相似文献   

17.
Due to the strong reducing capacity of ferrous Fe, the fate of Fe(II) following dissimilatory iron reduction will have a profound bearing on biogeochemical cycles. We have previously observed the rapid and near complete conversion of 2-line ferrihydrite to goethite (minor phase) and magnetite (major phase) under advective flow in an organic carbon-rich artificial groundwater medium. Yet, in many mineralogically mature environments, well-ordered iron (hydr)oxide phases dominate and may therefore control the extent and rate of Fe(III) reduction. Accordingly, here we compare the reducing capacity and Fe(II) sequestration mechanisms of goethite and hematite to 2-line ferrihydrite under advective flow within a medium mimicking that of natural groundwater supplemented with organic carbon. Introduction of dissolved organic carbon upon flow initiation results in the onset of dissimilatory iron reduction of all three Fe phases (2-line ferrihydrite, goethite, and hematite). While the initial surface area normalized rates are similar (∼10−11 mol Fe(II) m−2 g−1), the total amount of Fe(III) reduced over time along with the mechanisms and extent of Fe(II) sequestration differ among the three iron (hydr)oxide substrates. Following 16 d of reaction, the amount of Fe(III) reduced within the ferrihydrite, goethite, and hematite columns is 25, 5, and 1%, respectively. While 83% of the Fe(II) produced in the ferrihydrite system is retained within the solid-phase, merely 17% is retained within both the goethite and hematite columns. Magnetite precipitation is responsible for the majority of Fe(II) sequestration within ferrihydrite, yet magnetite was not detected in either the goethite or hematite systems. Instead, Fe(II) may be sequestered as localized spinel-like (magnetite) domains within surface hydrated layers (ca. 1 nm thick) on goethite and hematite or by electron delocalization within the bulk phase. The decreased solubility of goethite and hematite relative to ferrihydrite, resulting in lower Fe(III)aq and bacterially-generated Fe(II)aq concentrations, may hinder magnetite precipitation beyond mere surface reorganization into nanometer-sized, spinel-like domains. Nevertheless, following an initial, more rapid reduction period, the three Fe (hydr)oxides support similar aqueous ferrous iron concentrations, bacterial populations, and microbial Fe(III) reduction rates. A decline in microbial reduction rates and further Fe(II) retention in the solid-phase correlates with the initial degree of phase disorder (high energy sites). As such, sustained microbial reduction of 2-line ferrihydrite, goethite, and hematite appears to be controlled, in large part, by changes in surface reactivity (energy), which is influenced by microbial reduction and secondary Fe(II) sequestration processes regardless of structural order (crystallinity) and surface area.  相似文献   

18.
硫酸盐是大气颗粒物的重要组分,SO2与矿质颗粒物的非均相反应可能是硫酸盐和水溶性铁形成的重要途径之一,但目前对该反应途径的研究比较有限.本研究开展了不同相对湿度条件下SO2((7.14±0.29)μg/L)、NO2((5.13±0.21)μg/L)与针铁矿、磁铁矿、赤铁矿的非均相反应,定量分析了产物硫酸盐、硝酸盐以及水...  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