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 109 毫秒
1.
TCE/PCE的DNAPL污染及零价铁墙防治技术   总被引:6,自引:0,他引:6  
介绍了一种国内环保领域鲜有提及的土壤及地下水中的重非水相液体污染即:DNAPL污染,并以三氯乙烯(TCE)和四氯乙烯(PCE)为例阐述了其产生来源、危害及污染行为。针对DNAPL污染,详细论述了零价铁墙防治原理、降解途径及其主要的影响因素。  相似文献   

2.
表面活性剂强化抽出处理含水层中DNAPL污染物的去除特征   总被引:1,自引:0,他引:1  
为明确表面活性剂强化抽出处理含水层中DNAPL污染物过程中表面活性剂的增强修复效果,及DNAPL自身理化性质和介质孔径的影响,利用数码图像分析技术对1,2-二氯乙烷和四氯乙烯2种DNAPL在石英砂填充的二维砂箱中的抽取迁移过程进行了实验模拟研究,并对抽出水样中DNAPL的浓度进行了测试分析。结果表明,实验条件下加入低浓度(0.18%)的十二烷基苯磺酸钠(SDBS)大幅提高了对弱透水层截留的2种DNAPL聚集体的抽出处理效率。1,2-二氯乙烷在该表面活性剂溶液中的表观溶解度远高于四氯乙烯,因此其短时间内的绝对去除率更高。SDBS强化抽出处理DNAPL的作用机理以增溶作用为主,而其增流作用使DNAPL迁移流动后分布面积增大,增加了与表面活性剂溶液接触的面积,对增溶作用起到促进效果。细粒介质中DNAPL迁移后的最大分布面积较大,因此体系中DNAPL的溶解速率较高。在DNAPL聚集体质量与水力梯度固定的条件下,油水界面张力越低,DNAPL的密度越大,DNAPL垂向迁移的风险就越大。本研究为修复工程中如何依据DNAPL种类与场地多孔介质的情况选择表面活性剂提供了参考。  相似文献   

3.
使用动态过滤法对聚偏氟乙烯(PVDF)超滤膜进行改性,得到PVDF/聚乙烯醇(PVA)改性膜。系统考察了改性前后PVDF超滤膜表面的微观结构变化,使用原子力显微镜结合自制的—COOH、—OH官能团探针,定量测定了改性前后超滤膜与官能团之间的作用力变化特点,并进行了富含—OH、—COOH官能团的典型有机物海藻酸钠(SA)及腐殖酸(HA)的膜污染实验。结果表明,与PVDF基膜相比,PVDF/PVA改性膜表面粗糙度增大,接触角有效降低;—COOH、—OH官能团与PVDF/PVA改性膜之间的作用力远远小于—COOH、—OH官能团与PVDF之间的作用力。在相同的膜污染运行时间内,改性膜的通量衰减速率及衰减幅度皆小于相应PVDF膜的通量衰减速率及衰减幅度。说明改性膜有效降低了膜-污染物之间相互作用力,提高了膜的抗污染能力,微观作用力评价方法可利用于揭示膜污染本质。  相似文献   

4.
西安市冬、夏两季PM2.5中碳气溶胶的污染特征分析   总被引:5,自引:0,他引:5  
为研究西安市冬、夏两季大气颗粒物PM2.5中碳组分的污染变化规律,利用TEOM系列RP1400a采样仪于2010年冬季和夏季进行采样,测定了样品中的有机碳(OC)、无机碳(EC)和水溶性有机碳(WSOA)的含量。结果显示,PM2.5中OC和EC的季节平均浓度值冬季较高,分别是夏季的2.62,1.75倍,这表明西安市冬季碳气溶胶污染严重。OC和EC日变化在不同季节均呈现双峰分布特征,这主要是由交通源的排放和不利的气象条件造成的。OC和EC在冬、夏两季都有较强的相关性(R2分别为0.823和0.543),且OC/EC平均值分别为5.36和3.58,均大于2,表明采样各时段有二次有机碳(SOC)生成。  相似文献   

5.
工业废水和环境污染中往往含有两种或以上的金属离子,因此筛选分离能吸附多种重金属的菌株是利用微生物技术治理环境重金属污染的关键。从剩余污泥中筛选获得一株能耐受重金属Cr和Cu的细菌,经鉴定为伯克氏菌(Burkholderia sp.),命名为Y-12,用于开展水体Cr(Ⅵ)/Cu(Ⅱ)复合污染微生物吸附转化研究。结果表明,Y-12对Cr(Ⅵ)和Cu(Ⅱ)的吸附规律不同。随时间延长,Cr(Ⅵ)的去除率逐渐增大,但是共存Cu(Ⅱ)的去除率在2.00h达到最大值之后却又逐渐被解吸出来;Y-12去除Cr(Ⅵ)/Cu(Ⅱ)复合污染的最佳pH为6,强酸条件下,Y-12对Cr(Ⅵ)的吸附受到抑制,而当溶液pH为9时,Y-12主要通过还原作用对Cr(Ⅵ)进行解毒;Cr(Ⅵ)浓度越大越不利于Cr(Ⅵ)的去除,溶液中Cu(Ⅱ)浓度变化对Cr(Ⅵ)去除没有显著影响。由此可见,在适宜环境条件下,Y-12能有效去除水体中Cr(Ⅵ)/Cu(Ⅱ)复合污染,具有广泛的应用前景。  相似文献   

6.
赵祥  姜春露  陈星  郑刘根  李畅 《环境工程学报》2021,15(12):3854-3864
还原稳定化修复技术是当前重金属污染修复领域的主要技术,其中,重金属去除率和环境扰动是评价修复效果的重要指标.以不同pH重铬酸钾溶液模拟Cr(Ⅵ)污染水体,分别研究了多硫化钙、硫酸亚铁和二者联用对水体重金属Cr(Ⅵ)的还原稳定性,同时考察了环境因素对药剂联用的影响并探究各药剂对Cr(Ⅵ)的去除机理.结果表明:当多硫化钙与硫酸亚铁投加比例为1:2时,药剂联用对Cr(Ⅵ)和总Cr的去除效果均优于单独施加药剂时的去除效果且还原产物较稳定,对体系扰动作用相对较小;低pH和较高温度有利于联用药剂对Cr(Ⅵ)的去除,HCO3-、Cl-、Mn2+的加入有利于去除Cr(Ⅵ),Fe3+对 Cr(Ⅵ)去除表现为低浓度抑制、高浓度促进;经反应产物成分分析,多硫化钙与硫酸亚铁反应生成了具有催化效果的FeS,提高了修复效果.以上研究结果可为后期Cr(Ⅵ)污染水体的控制和修复提供参考.  相似文献   

7.
高价铬及双酚A在铁-乳酸体系中的同时光处理   总被引:1,自引:0,他引:1  
研究了Fe(Ⅲ)-乳酸配合物体系同时对Cr(Ⅵ)的光还原及双酚A(BPA)的光氧化处理,考察了光源、初始pH值、Fe(Ⅲ)、乳酸盐、Cr(Ⅵ)及BPA初始浓度等因素对Cr(Ⅵ)及BPA光处理效率的影响。结果表明:光照条件下,铁-乳酸配合物能有效实现对六价铬及BPA的同时光处理。同一体系中,Cr(Ⅵ)的光还原快于BPA的光氧化,Fe(Ⅲ)初始浓度的增加可同时提高Cr(Ⅵ)的光还原效率和BPA的光氧化效率;Fe(Ⅲ)-乳酸盐配合物光解产生的Fe(Ⅱ)是Cr(Ⅵ)的主要还原剂,其次级光反应中产生的.OH是BPA的氧化剂。  相似文献   

8.
陈建  徐林 《环境工程学报》2013,7(1):191-195
为开发含Cr(Ⅵ)废水处理工艺提供必要资料,对不同条件下Fe(Ⅲ)催化有机酸光化学还原Cr(Ⅵ)进行了比较研究.研究结果表明,Cr(Ⅵ)的还原不仅受pH、Fe(Ⅲ)或有机酸的起始浓度以及共存阳离子的影响,而且还与有机酸种类有关.低pH的酸性条件有利于cr(Ⅵ)的光化学还原,在pH 3.0条件下经3h后的还原率达89.9%,在pH 5.0经3h后其还原率为37.3%.Fe(Ⅲ)或有机酸起始浓度增高会促进Cr(Ⅵ)的还原,在pH3.0和Fe(Ⅲ)浓度高于Cr(Ⅵ)浓度条件下导致在3h后Cr(Ⅵ)的光化学还原率达100%.共存Al(Ⅲ)或Cu(Ⅱ)会抑制Cr(Ⅵ)的光化学还原.由Fe(Ⅲ)催化3种有机酸对Cr(Ⅵ)的光化学还原作用大小次序为:酒石酸>柠檬酸>苹果酸.还对不同条件影响Cr(Ⅵ)的光化学还原可能机制作了讨论.  相似文献   

9.
为了提高介质阻挡放电(dielectric barrier discharge,DBD)等离子体的处理效率,研究了内外介质组成分别为:(1)石英 石英;(2)陶瓷 石英;(3)陶瓷 陶瓷这3种情况下苯的降解情况.试验结果表明,在处理低浓度含苯废气时,陶瓷 陶瓷效果最好;陶瓷(内管) 石英(外管)在处理高浓度含苯废气时显示出优势.通过对气相产物和固相结焦产物的分析验证了DBD能有效降解苯,降解产物不会带来新的污染.进一步分析了实验条件和介质材料的变化对DBD降解苯的影响机理.  相似文献   

10.
霉菌因其菌丝体生长快、吸附能力强、固液分离效果好在处理重金属污染废水中受到普遍关注.通过对霉菌HM6培养条件的研究考查了培养条件对霉菌菌丝球产量、菌球特征及菌球吸附性能的影响.实验结果表明,该霉菌在液体查氏培养基,250 mL三角瓶装液量为100 mL,1%接种量(孢子悬液浓度106),pH=5,30℃,150 r/min摇床培养60~72 h时,可形成直径在2.0~2.5 mm范围内的菌丝球,球体白色光滑均匀,具有一定的机械强度,对Pb(Ⅱ)具有一定的吸附能力.培养条件的改变对菌球的特征、干湿比、产量影响较大,对菌球吸附性能的影响相对较小.碱处理可以增大菌丝球的干湿比,提高其机械强度,但并不能提高其对Pb(Ⅱ)的吸附能力.  相似文献   

11.
The stable carbon isotope values of tetrachloroethene (PCE) and its degradation products were monitored during studies of biologically enhanced dissolution of PCE dense nonaqueous phase liquid (DNAPL) to determine the effect of PCE dissolution on observed isotope values. The degradation of PCE was monitored in a 2-dimensional model aquifer and in a pilot test cell (PTC) at Dover Air Force Base, both with emplaced PCE DNAPL sources. Within the plume down gradient from the source, the isotopic fractionation of dissolved PCE and its degradation products were consistent with those observed in biodegradation laboratory studies. However, close to the source zone significant shifts in the isotope values of dissolved PCE were not observed in either the model aquifer or PTC due to the constant input of newly dissolved, non fractionated PCE, and the small isotopic fractionation associated with PCE reductive dechlorination by the mixed microbial culture used. Therefore the identification of reductive dechlorination in the presence of PCE DNAPL was based upon the appearance of daughter products and the isotope values of those daughter products. An isotope model was developed to simulate isotope values of PCE during the dissolution and degradation of PCE adjacent to a DNAPL source zone. With the exception of very high degradation rate constants (>1/day) stable carbon isotope values of PCE estimated by the model remained within error of the isotope value of the PCE DNAPL, consistent with measured isotope values in the model aquifer and in the PTC.  相似文献   

12.
Natural remobilization of multicomponent DNAPL pools due to dissolution   总被引:1,自引:0,他引:1  
Mixtures of dense nonaqueous phase liquids (DNAPLs) trapped in the subsurface can act as long-term sources of contamination by dissolving into flowing groundwater. If the components have different solubilities then dissolution will alter the composition of the remaining DNAPL. We theorized that a multicomponent DNAPL pool may become mobile due to the natural dissolution process. In this study, we focused on two scenarios: (1) a DNAPL losing light component(s), with the potential for downward migration; and (2) a DNAPL losing dense component(s), with the potential for upward migration following transformation into a less dense than water nonaqueous phase liquid (LNAPL). We considered three binary mixtures of common groundwater contaminants: benzene and tetrachloroethylene (PCE), PCE and dichloromethane (DCM), and DCM and toluene. A number of physical properties that control the retention and transport of DNAPL in porous media were measured for the mixtures, namely: density, interfacial tension, effective solubility, and viscosity. All properties except density exhibited nonlinear relationships with changing molar ratio of the DNAPL. To illustrate the potential for natural remobilization, we modelled the following two primary mechanisms: the reduction in pool height as mass is lost by dissolution, and the changes in fluid properties with changing molar ratio of the DNAPL. The first mechanism always reduces the capillary pressure in the pool, while the second mechanism may increase the capillary pressure or alter the direction of the driving force. The difference between the rate of change of each determines whether the potential for remobilization increases or decreases. Static conditions and horizontal layering were assumed along with a one-dimensional, compositional modelling approach. Our results indicated that for initial benzene/PCE ratios greater than 25:75, the change in density was sufficiently faster than the decline in pool height to promote DNAPL breakthrough into the adjacent porous medium. In contrast, there was no potential for natural remobilization of a PCE-DCM mixture, primarily because the densities of the components are not sufficiently different. Dissolution of a DCM-toluene mixture decreased the density, reducing the tendency for downward displacement. However, the ultimate transformation from a DNAPL to an LNAPL may induce upward displacement. These results suggest that at sites with DNAPL pools containing a mix of components of sufficiently different densities and relative solubilities, natural remobilization may be an active mechanism, with implications for site evaluation and remediation.  相似文献   

13.
Mixtures of dense non-aqueous phase liquids (DNAPLs) trapped in the subsurface can act as long-term sources of contamination by dissolving into flowing groundwater. In general, the components of higher solubility are removed more quickly, thus altering the composition of the remaining DNAPL, and possibly leading to changes in its physical properties. Through the development of a simple compositional model, Roy et al. [J. Contam. Hydrol. 2002 (59) 163] showed that preferential dissolution of a mixed DNAPL could potentially result in changes in density and interfacial tension that could subsequently lead to remobilization of an initially static DNAPL pool. The laboratory experiments presented in this next paper provide a proof-of-concept for the previously presented theory, demonstrating and quantifying this process of remobilization. In addition, the experiments provide a data set for evaluation of the model presented by Roy et al. [J. Contam. Hydrol. 2002 (59) 163]. In the four experiments, a DNAPL pool comprised of tetrachloroethene and benzene was created as an open pool overlying glass beads within a water-saturated 2-D flow box. Experiments included rectangular and triangular pools. In each of the experiments, remobilization (as breakthrough) was observed more than 2 weeks after formation of the initial pool. During each experiment, the pool height declined as mass was lost by dissolution, while sampling indicated a decrease in the mole fraction of benzene, the more soluble component. Small protuberances formed along the bottom of the pool as its composition changed with time and the displacement pressure was achieved for various pore throats. Eventually one of the protuberances extended further, forming a finger (breakthrough). In general, the pool emptied as the finger proceeded further into the beads. It was also shown theoretically and experimentally that remobilization will occur sooner for pools with a triangular (pointing down), rather than rectangular, shape. The experimental results were simulated using the model developed by Roy et al. [J. Contam. Hydrol. 2002 (59) 163]. The model matched the observations well, suggesting that it accurately represents the primary mechanisms involved with natural remobilization under the conditions of the study.  相似文献   

14.
While the capability of nanoscale zero-valent iron (NZVI) to dechlorinate organic compounds in aqueous solutions has been demonstrated, the ability of NZVI to remove dense non-aqueous phase liquid (DNAPL) from source zones under flow-through conditions similar to a field scale application has not yet been thoroughly investigated. To gain insight on simultaneous DNAPL dissolution and NZVI-mediated dechlorination reactions after direct placement of NZVI into a DNAPL source zone, a combined experimental and modeling study was performed. First, a DNAPL tetrachloroethene (PCE) source zone with emplaced NZVI was built inside a small custom-made flow cell and the effluent PCE and dechlorination byproducts were monitored over time. Second, a model for rate-limited DNAPL dissolution and NZVI-mediated dechlorination of PCE to its three main reaction byproducts with a possibility for partitioning of these byproducts back into the DNAPL was formulated. The coupled processes occurring in the flow cell were simulated and analyzed using a detailed three-dimensional numerical model. It was found that subsurface emplacement of NZVI did not markedly accelerate DNAPL dissolution or the DNAPL mass-depletion rate, when NZVI at a particle concentration of 10g/L was directly emplaced in the DNAPL source zone. To react with NZVI the DNAPL PCE must first dissolve into the groundwater and the rate of dissolution controls the longevity of the DNAPL source. The modeling study further indicated that faster reacting particles would decrease aqueous contaminant concentrations but there is a limit to how much the mass removal rate can be increased by increasing the dechlorination reaction rate. To ensure reduction of aqueous contaminant concentrations, remediation of DNAPL contaminants with NZVI should include emplacement in a capture zone down-gradient of the DNAPL source.  相似文献   

15.
Microbial reductive dechlorination of trichloroethene (TCE) and perchloroethene (PCE) in the vicinity of their dense non-aqueous phase liquid (DNAPL) has been shown to accelerate DNAPL dissolution. A three-layer diffusion-cell was developed to quantify this bio-enhanced dissolution and to measure the conditions near the DNAPL interface. The 12 cm long diffusion-cell setup consists of a 5.5 cm central porous layer (sand), a lower 3.5 cm DNAPL layer and a top 3 cm water layer. The water layer is frequently refreshed to remove chloroethenes at the upper boundary of the porous layer, while the DNAPL layer maintains the saturated chloroethene concentration at the lower boundary. Two abiotic and two biotic diffusion-cells with TCE DNAPL were tested. In the abiotic diffusion-cells, a linear steady state TCE concentration profile between the DNAPL and the water layer developed beyond 21 d. In the biotic diffusion-cells, TCE was completely converted into cis-dichloroethene (cis-DCE) at 2.5 cm distance of the DNAPL. Dechlorination was likely inhibited up to a distance of 1.5 cm from the DNAPL, as in this part the TCE concentration exceeded the culture’s maximum tolerable concentration (2.5 mM). The DNAPL dissolution fluxes were calculated from the TCE concentration gradient, measured at the interface of the DNAPL layer and the porous layer. Biotic fluxes were a factor 2.4 (standard deviation 0.2) larger than abiotic dissolution fluxes. This diffusion-cell setup can be used to study the factors affecting the bio-enhanced dissolution of DNAPL and to assess bioaugmentation, pH buffer addition and donor delivery strategies for source zones.  相似文献   

16.
We have conducted well-controlled DNAPL remediation experiments within a 2-D, glass-walled, sand-filled chamber using surfactants (Aerosol MA and Tween 80) to increase solubility and an oxidant (permanganate) to chemically degrade the DNAPL. Initial conditions for each remediation experiment were created by injecting DNAPL as a point source at the top of the chamber and allowing the DNAPL to migrate downward through a water-filled, heterogeneous, sand-pack designed to be evocative of a fluvial depositional environment. This migration process resulted in the DNAPL residing as a series of descending pools. Lateral advection across the chamber was used to introduce the remedial fluids. Photographs and digital image analysis illustrate interactions between the introduced fluids and the DNAPL. In the surfactant experiments, we found that DNAPL configured in a series of pools was easily mobilized. Extreme reductions in DNAPL/water interfacial tension occurred when using the Aerosol MA surfactant, resulting in mobilization into low permeability regions and thus confounding the remediation process. More modest reductions in interfacial tension occurred when using the Tween 80 surfactant resulting in modest mobilization. In this experiment, capillary forces remained sufficient to exclude DNAPL migration into low permeability regions allowing the excellent solubilizing properties of the surfactant to recover almost 90% of the DNAPL within 8.6 pore volumes. Injection of a potassium permanganate solution resulted in precipitation of MnO2, a reaction product, creating a low-permeability rind surrounding the DNAPL pools. Formation of this rind hindered contact between the permanganate and the DNAPL, limiting the effectiveness of the remediation. From these experiments, we see the value of performing visualization experiments to evaluate the performance of proposed techniques for DNAPL remediation.  相似文献   

17.
Phase diagrams were used for the formulation of alcohol–surfactant–solvent and to identify the DNAPL (Dense Non Aqueous Phase Liquid) extraction zones. Four potential extraction zones of Mercier DNAPL, a mixture of heavy aliphatics, aromatics and chlorinated hydrocarbons, were identified but only one microemulsion zone showed satisfactory DNAPL recovery in sand columns. More than 90 sand column experiments were performed and demonstrate that: (1) neither surfactant in water, alcohol–surfactant solutions, nor pure solvent can effectively recover Mercier DNAPL and that only alcohol–surfactant–solvent solutions are efficient; (2) adding salts to alcohol–surfactant or to alcohol–surfactant–solvent solutions does not have a beneficial effect on DNAPL recovery; (3) washing solution formulations are site specific and must be modified if the surface properties of the solids (mineralogy) change locally, or if the interfacial behavior of liquids (type of oil) changes; (4) high solvent concentrations in washing solutions increase DNAPL extraction but also increase their cost and decrease their density dramatically; (5) maximum DNAPL recovery is observed with alcohol–surfactant–solvent formulations which correspond to the maximum solubilization in Zone C of the phase diagram; (6) replacing part of surfactant SAS by the alcohol n-butanol increases washing solution efficiency and decreases the density and the cost of solutions; (7) replacing part of n-butanol by the nonionic surfactant HOES decreases DNAPL recovery and increases the cost of solutions; (8) toluene is a better solvent than D-limonene because it increases DNAPL recovery and decreases the cost of solutions; (9) optimal alcohol–surfactant–solvent solutions contain a mixture of solvents in a mass ratio of toluene to D-limonene of one or two. Injection of 1.5 pore volumes of the optimal washing solution of n-butanol–SAS–toluene–D-limonene in water can recover up to 95% of Mercier DNAPL in sand columns. In the first pore volume of the washing solution recovered in the sand column effluent, the DNAPL is in a water-in-oil microemulsion lighter than the excess aqueous phase (Winsor Type II system), which indicates that part of the DNAPL was mobilized. In the next pore volumes, DNAPL is dissolved in a oil-in-water microemulsion phase and is mobilized in an excess oil phase lighter than the microemulsion (Winsor Type I system). The main drawback of this oil extraction process is the high concentration of ingredients necessary for DNAPL dissolution, which makes the process expensive. Because mobilization of oil seems to occur at the washing solution front, an injection strategy must be developed if there is no impermeable limit at the aquifer base. DNAPL recovery in the field could be less than observed in sand columns because of a smaller sweep efficiency related to field sand heterogeneities. The role of each component in the extraction processes in sand column as well as the Winsor system type have to be better defined for modeling purposes. Injection strategies must be developed to recover ingredients of the washing solution that can remain in the soil at the end of the washing process. ©1997 Elsevier Science B.V.  相似文献   

18.
Electrical impedance tomography (EIT) was used to monitor the movement of a fluorinated hydrocarbon dense nonaqueous phase liquid (DNAPL) through a saturated porous medium within a laboratory column. Impedance measurements were made using a horizontal plane of 12 electrodes positioned at regular intervals around the centre of the column. A 2D inversion algorithm, which incorporated the cylindrical geometry of the column, was used to reconstruct resistivity and phase images from the measured data. Differential time-lapse images of DNAPL movement past the plane of electrodes were generated by the cell-by-cell subtraction of resistivity and phase baseline models from those associated with the DNAPL release stage of the experiment.The DNAPL pulse was clearly delineated as resistive anomalies in the differential time-lapse resistivity images. The spatial extent of the resistive anomalies indicated that in addition to vertical migration, some lateral spreading of the DNAPL had occurred. Residual contamination could be detected after quasi-static conditions were reestablished. Residual DNAPL saturation was estimated from the resistivity model data by applying Archie's second equation.Despite significant measured phase changes due to DNAPL contamination, the differential phase images revealed only weak anomalies associated with DNAPL flow; these anomalies could be seen only in the initial stages of the experiment during peak flow through the plane of electrodes.  相似文献   

19.
Analytical solutions are developed for approximating the time-dependent contaminant discharge from DNAPL source zones undergoing dissolution and other decay processes. The source functions assume a power relationship between source mass and chemical discharge and can consider partial DNAPL source remediation (depletion) at any time after the initial DNAPL release. The source functions are used as a time-dependent boundary condition in an idealized chemical transport model to develop leading order approximations of the plume response to DNAPL source removal. The results suggest that partial DNAPL remediation does not tend to have a dramatic impact on the maximum extent of the plume if very low concentration values are used to define the plume boundaries. However, the solutions show that partial DNAPL removal from the source zone is likely to lead to large reductions in plume concentrations and mass, and it reduces the longevity of the plume. When the mass discharge from the source zone is linearly related to the DNAPL mass, it is shown that partial DNAPL depletion leads to linearly proportional reductions in the plume mass and concentrations.  相似文献   

20.
Two-dimensional chamber studies were conducted to determine qualitative and quantitative performance of cosolvents targeted at pooled dense non-aqueous phase liquid (DNAPL) (perchlorethylene, PCE) residing above a fine-grain capillary barrier. Downward mobilization of DNAPL, up gradient along an overriding cosolvent front, was observed. This produced significant pooling above a fine-grain layer that in some cases lead to entry into the capillary barrier beneath. Entry pressure calculations using physical and hydrogeologic parameters provided an excellent prediction of breakthrough of DNAPL into the capillary barrier. Calculations predict approximately 0.5 m of DNAPL would be necessary to enter a Beit Netofa clay, under extreme cosolvent flooding conditions (100% ethanol). Gradient injection of cosolvent did not appear to provide any benefit suggesting a rapid decrease in interfacial tension (IFT) compared to the rate of DNAPL solubilization. Use of a partitioning alcohol (tertiary butyl alcohol, TBA) resulted in DNAPL swelling and reduced entry into the capillary barrier. However, the trapping of flushing solution, containing PCE, could potentially lead to longer remediation times.  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