Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

In Situ Chemical Oxidation for Groundwater Remediation
In Situ Chemical Oxidation for Groundwater Remediation
In Situ Chemical Oxidation for Groundwater Remediation
Ebook1,542 pages16 hours

In Situ Chemical Oxidation for Groundwater Remediation

Rating: 0 out of 5 stars

()

Read preview

About this ebook

This volume provides comprehensive up-to-date descriptions of the principles and practices of in situ chemical oxidation (ISCO) for groundwater remediation based on a decade of intensive research, development, and demonstrations, and lessons learned from commercial field applications.
LanguageEnglish
PublisherSpringer
Release dateFeb 25, 2011
ISBN9781441978264
In Situ Chemical Oxidation for Groundwater Remediation

Related to In Situ Chemical Oxidation for Groundwater Remediation

Titles in the series (1)

View More

Related ebooks

Environmental Engineering For You

View More

Related articles

Related categories

Reviews for In Situ Chemical Oxidation for Groundwater Remediation

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    In Situ Chemical Oxidation for Groundwater Remediation - Robert L. Siegrist

    Robert L. Siegrist, Michelle Crimi and Thomas J. Simpkin (eds.)SERDP/ESTCP Environmental Remediation TechnologyIn Situ Chemical Oxidation for Groundwater Remediation10.1007/978-1-4419-7826-4_1© Springer Science+Business Media, LLC 2011

    1.  In Situ Chemical Oxidation: Technology Description and Status

    Robert L. Siegrist¹ , Michelle Crimi² and Richard A. Brown³

    (1)

    Colorado School of Mines, Golden, CO 80401, USA

    (2)

    Clarkson University, Potsdam, NY 13699, USA

    (3)

    Environmental Resources Management, Ewing, NJ 08618, USA

    Scope

    Overview of in situ chemical oxidation (ISCO) as a remediation technology, including history of development and use, field applications, performance expectations, and costs.

    Key Concepts

    ISCO has had a long history of development and use. While research and development still continue, ISCO is a relatively mature technology for the remediation of contaminated groundwater, including source zones and plumes.

    ISCO has primarily been applied for treatment of chlorinated organic solvents and petroleum hydrocarbons to achieve remediation objectives ranging from reducing contaminant mass in a source zone to achieving maximum contaminant levels in a plume. To achieve the more stringent remediation objectives, ISCO is almost always combined with another technology (e.g., bioremediation) or approach (e.g., monitored natural attenuation).

    The effectiveness of ISCO varies and is highly dependent on proper site characterization and design of the oxidant delivery system to achieve oxidative destruction of contaminants of concern in a target treatment zone.

    Typically ISCO applications require a targeted second or third oxidant delivery event since rebound in groundwater contaminant levels following cessation of active ISCO is a common occurrence.

    The median cost of an ISCO project appears to be on the order of $100 per cubic yard treated. However, costs can vary widely depending on contaminant characteristics, site conditions, and the oxidant used.

    When considering ISCO at a contaminated site, there are a number of frequently asked questions (Table 1.6) and key points to keep in mind to help support successful application (Table 1.7).

    1.1 Contaminated Sites and In Situ Remediation

    1.1.1 Introduction

    Subsurface contamination by chemicals that pose concerns for human health or environmental quality has been recognized as a widespread problem at industrial and military sites in the United States and abroad. Awareness of the significance of this contamination occurred in the 1970s when adverse health effects were observed at major sites across the United States (e.g., Times Beach, Love Canal, Valley of the Drums). During the past three decades, substantial progress has been made to identify, investigate, and remediate those sites where unacceptable levels of contamination are present. However, for the foreseeable future, contaminated sites, including thousands of known sites, newly discovered legacy sites, and new sites where recent spills and releases have occurred, will remain an ongoing challenge. Over the next decades, major efforts and funds will be spent cleaning up contaminated sites in the United States in an attempt to protect public health and preserve environmental quality (Figure 1.1).

    A978-1-4419-7826-4_1_Fig1_HTML.gif

    Figure 1.1

    Estimated (a) costs for remediation of contaminated sites and (b) total number of sites requiring remediation under current regulations and practices in the USA (during 2004–2033) (USEPA, 2004). Estimates for cleanup costs (total = $209 billion) and number of sites (total = 294,000) are based on reasonable assumptions regarding the average cleanup cost per site and the number of new site discoveries. Note: NPL - National Priorities List or Superfund, RCRA-CA - Resource Conservation and Recovery Act Corrective Action Program, UST - underground storage tanks, DoD - Department of Defense, DOE - Department of Energy; Civilian Agencies – non-DoD and non-DOE federal agencies; State and Private – state mandatory, voluntary, and brownfield sites, and private sites.

    1.1.2 Characteristics of Contaminated Sites

    Within the United States, contaminated sites are present at a wide range of facilities and operations spanning simple to complex activities over small to large areas. These sites include gasoline stations, drycleaners, manufacturing plants, energy generation facilities, mining operations, disposal facilities, and nuclear and military installations. Soil and groundwater continue to be the most prevalent contaminated media. The most frequent contaminants of concern (COCs) are volatile organic compounds (VOCs), which are present at a majority of sites. At many sites, semivolatile organic compounds (SVOCs) and metals are also prevalent. The Agency for Toxic Substances and Disease Registry (ATSDR) has identified 275 hazardous substances most commonly found at facilities on the U.S. NPL (ATSDR, 2009). Table 1.1 provides a list of 20 chemicals (or chemical groups), along with their prevalence and priority as COCs at NPL sites in the United States. Some of these same substances are also pervasive at thousands of non-NPL sites.

    Table 1.1.

    Prevalence of Chemicals at NPL Sites and Priority as Contaminants of Concern

    Note: data not available.

    aATSDR (2009).

    bPrevalence of COCs at NPL sites identified by the U.S. Environmental Protection Agency (USEPA). See ATSDR Toxicological Profiles at http://www.atsdr.cdc.gov/toxprofiles/index.asp (accessed June 23, 2010).

    cUSEPA Maximum Contaminant Levels (MCLs); http://www.epa.gov/ogwdw/contaminants/index.html (accessed June 23, 2010).

    Organic chemicals continue to be the primary COCs in soil and groundwater at contaminated sites in the United States and in other industrialized nations (USEPA, 1997, 2004; Siegrist et al., 2001; Kavanaugh et al., 2003; GAO, 2005). In 2004, there were an estimated 35,000 sites with petroleum contamination from USTs (USEPA, 2004). At most UST sites, the primary COCs are VOCs (e.g., benzene, toluene, ethylbenzene, xylenes). Excluding the UST sites, Kavanaugh et al. (2003) estimated there were 30,000–50,000 sites in the United States with groundwater contamination, of which about 80% were contaminated with organic chemicals. Kavanaugh et al. (2003) also estimated that about 60% of the sites with organic chemicals in groundwater likely have dense nonaqueous phase liquids (DNAPLs) present. Common DNAPLs include trichloroethene (TCE), perchloroethene (PCE; also termed perchloroethylene or tetrachloroethylene), chloroform, and carbon tetrachloride. DNAPLs present a long-term problem due to their chemical properties, including low solubility and high density, and their recalcitrance (stability) in the environment. The chemical properties and stability can contribute to high contaminant concentrations in groundwater over extensive time frames (Kavanaugh et al., 2003). Similarly, light nonaqueous phase liquids (LNAPLs) in the subsurface (e.g., gasoline and fuel products at UST sites) are also sources of long-term groundwater contamination.

    While there are many exposure scenarios that present serious current or future health risks at NPL sites and many other contaminated sites, baseline risk is commonly governed primarily by ingestion exposures to COCs that migrate into groundwater used for drinking water or inhalation exposures to VOCs that enter buildings through vapor intrusion (Figure 1.2). As a result, the main driving force for remediation of groundwater likely will continue to be protecting human health from these exposures. In addition, mitigating ecological impacts will become increasingly important.

    Despite general agreement concerning the value of reducing risk through remediation, there are challenges and topics of disagreement concerning the maximum level of contamination that can be left behind safely. There are also conflicting views about how to respond to sites where cleanup to protective levels is not now or likely to be in the future, practical or even possible.

    A978-1-4419-7826-4_1_Fig2_HTML.gif

    Figure 1.2.

    Illustration of an organic chemical release (shown in red) leading to long-term contamination of groundwater and potential risks to human health.

    1.1.3 Site Remediation Approaches

    Initially, the most common approach to the cleanup of groundwater contaminated by organic chemicals involved using pumping wells to extract contaminated groundwater, followed by treatment of the extracted groundwater aboveground (i.e., pump-and-treat). Over time, it became apparent that this approach had severe performance limitations and high costs (Mackay and Cherry, 1989; NRC, 1994; USEPA, 1999). Based on the cleanup costs for 15,000–25,000 U.S. sites with DNAPL contamination, Kavanaugh et al. (2003) estimated that the median cost to operate a groundwater pump-and-treat system is roughly $180,000 per year, with a range of $30,000–$4,000,000. The combined annual costs for all U.S. sites using pump-and-treat are $2.7–$4.5 billion a year. Assuming a 30-year life and a 5–10% interest rate, life cycle costs of cleanup of DNAPL sites could range from $50 to $100 billion (Kavanaugh et al., 2003). A widely held view that has emerged is that cleanup of contaminated groundwater using pump-and-treat alone is virtually impossible, though pump-and-treat can be used as a hydraulic containment technology. As a result, interest in, and development of, alternative in situ technologies and approaches has grown appreciably (NRC, 1994, 2005; Kavanaugh et al., 2003; GAO, 2005).

    Over the years, research and development efforts have been directed at advancing the scientific and technological understanding of the fate and transport of organic contaminants. These efforts have led to development of in situ technologies based on both engineered and natural attenuation processes (NRC, 1994, 1997, 2005). The major in situ technologies for groundwater remediation are listed in Table 1.2, along with their advantages and limitations (Stroo, 2010). In situ technologies are based on a number of different technical principles and can be applied for contaminated groundwater zones and, to varying extents, for treatment of near surface soil and vadose zones. Examples of technologies that exploit mass transfer and recovery methods include soil vapor extraction, air sparging, or surfactant/cosolvent flushing. Examples of those that use in place destruction include bioremediation or chemical oxidation/reduction. The viability of in situ technologies has been enhanced by developments in delivery methods and enabling technologies, such as soil mixing, hydraulic fracturing, and soil heating.

    Table 1.2.

    Advantages and Limitations of the Major In Situ Groundwater Remediation Technologies (Stroo, 2010)

    The strategy for remediating sites has also evolved. Early strategies based on pump-and-treat focused on containment. However, as discussed above, pump-and-treat did little to address the high residual mass of contaminants in nonaqueous phase liquid (NAPL) source zones. Currently, remediation of sites with source zones is increasingly viewed as best accomplished by combining remedies simultaneously or sequentially for different zones of contamination (NRC, 2005). For example, treatment of a DNAPL source zone with an aggressive technology like thermally enhanced extraction or in situ chemical oxidation (ISCO) might enable remediation of the associated groundwater plume using engineered bioremediation and permeable reactive barriers (Figure 1.3).

    A978-1-4419-7826-4_1_Fig3_HTML.gif

    Figure 1.3.

    Illustration of in situ technologies that might be viable for a combined remedy for a site with a DNAPL source zone and associated groundwater plume. (Note: DNAPL contamination is shown in red)

    ISCO is one of several technologies that have emerged with the potential for cost effective remediation of contaminated soil and groundwater. ISCO involves the subsurface delivery of a chemical oxidant to destroy organic COCs and thereby reduce potential risks to public health and environmental quality. The current chemical oxidants in relatively widespread use for ISCO are hydrogen peroxide (H2O2), potassium and sodium permanganate (KMnO4, NaMnO4), sodium persulfate (Na2S2O8), and ozone (O3).

    1.1.4 Organization of this Volume on ISCO

    This chapter presents an overview of ISCO, its evolution and current status, a summary of current costs and performance information, and the key points for practitioners and managers to consider when using this technology. Details regarding the principles and practices of ISCO for groundwater remediation are presented in the subsequent chapters of this volume, as highlighted in Table 1.3. While the focus of this volume is on use of ISCO for groundwater remediation, many of the principles and practices presented are applicable to the use of chemical oxidation for remediation of contaminated soils and other media.

    Table 1.3.

    Chapters in this Volume and Coverage in Each

    1.2 Isco as a Remediation Technology

    A wide range of organic chemicals are susceptible to oxidative degradation through chemical reactions with oxidants, such as catalyzed hydrogen peroxide (also known as CHP or modified Fenton’s reagent), potassium and sodium permanganate, sodium persulfate, and ozone (Table 1.4). There are also a few new oxidants or combination of oxidants, such as peroxone, percarbonate, and calcium peroxide. There is less information on these oxidants and they have been implemented less frequently so they are not discussed extensively in this volume.

    Table 1.4.

    Characteristics of Chemical Oxidants Used for Destruction of Organic Contaminants (adapted from Huling and Pivetz, 2006)

    aThose oxidants with an * have been most commonly used for ISCO applications (see Section 1.3).

    Under the right conditions, oxidants yield reactive species (Table 1.5) that can transform and often mineralize many COCs including chlorinated hydrocarbons (e.g., TCE, PCE), fuels (e.g., benzene, toluene, methyl tert-butyl ether), phenols (e.g., pentachlorophenol), polycyclic aromatic hydrocarbons (e.g., naphthalene, phenanthrene), polychlorinated biphenyls, explosives (e.g., trinitrotoluene), and pesticides (e.g., lindane). Through laboratory experimentation, reaction stoichiometries, pathways, and kinetics have been established for many common COCs. Degradation reactions tend to involve electron transfer or free radical processes, involve simple to complex pathways with various intermediates, and follow second-order kinetics. The need for activation to generate free radicals and the sensitivity to matrix conditions, such as temperature, pH, and salinity, vary with the different oxidants and specific contaminants.

    Table 1.5.

    Reactive Species Associated with Oxidant Chemicals (Huling and Pivetz, 2006)

    The ISCO systems that can be (and have been) applied in the field vary significantly in their features based on the nature and extent of contamination, the site conditions present, and the remediation goals and objectives (Siegrist et al., 2001; ITRC, 2005; Huling and Pivetz, 2006; Krembs, 2008). Remediation objectives for ISCO are established within the context of the overall remediation goals and cleanup levels established for a particular site. The objectives generally fall into one of the following three categories:

    Reduce the contaminant concentration or mass in the ISCO-treated zone by some fraction (e.g., >90%).

    Achieve a specified post-ISCO contaminant concentration in an ISCO-treated zone (e.g., ≤1 mg/kg in subsurface solids or ≤100 μg/L in groundwater).

    Achieve a specified concentration in a groundwater plume at compliance points downgradient from an ISCO-treated source zone.

    In some cases, these objectives might be met by combining ISCO with another remedial technology or approach, as described in Chapter 7.

    Oxidant can be delivered into the subsurface at varied concentrations and mass loading rates in liquid, gas, or solid phases through different subsurface delivery methods. Oxidant delivery has most commonly been accomplished through permeation by vertical direct-push injection probes or flushing by vertical groundwater wells (Figure 1.4). Other delivery methods have also been employed using horizontal wells, infiltration galleries, soil mixing, and hydraulic or pneumatic fracturing.

    A978-1-4419-7826-4_1_Fig4_HTML.gif

    Figure 1.4.

    In situ chemical oxidation using (a) direct-push injection probes or (b) well-to-well flushing to deliver oxidants (shown in blue) into a target treatment zone of groundwater contaminated by DNAPL compounds (shown in red).

    Research completed by Siegrist et al. (2006, 2008a), along with that of other investigators funded under the DoD ISCO initiative (http://www.serdp-estcp.org/ISCO.cfm), has demonstrated that the successful use of ISCO for remediation of contaminated groundwater fundamentally depends on the reaction chemistry of the oxidant used and its ability to degrade the COCs at a site. Success also depends on effective delivery of oxidants into the subsurface, which is determined by the oxidant reactive transport under the hydrogeologic and geochemical conditions present at a site. Oxidant reaction chemistry (both with the COCs and with natural organic matter and minerals) and subsurface transport affect ISCO application, COC destruction, and cost effectiveness.

    The application of ISCO can be affected by, and also effect changes in, the pre-ISCO subsurface conditions (e.g., groundwater flow direction and velocity, oxidant concentrations, Eh, pH, temperature, dissolved organic carbon). These effects should be considered and accounted for during ISCO system design and implementation (e.g., Figures 1.5 and 1.6).

    A978-1-4419-7826-4_1_Fig5_HTML.gif

    Figure 1.5.

    Macro- and micro-scale features of contaminated groundwater zones where organic contaminants of concern may exist in the aqueous, sorbed, and nonaqueous phases (adapted from Siegrist et al., 2006). Note: DOC - dissolved organic carbon, NOM - natural organic matter.

    A978-1-4419-7826-4_1_Fig6_HTML.gif

    Figure 1.6.

    ISCO delivery methods can affect groundwater velocity, oxidant concentrations, and NAPL depletion within a target treatment zone (Siegrist et al., 2006; Petri et al., 2008a). Oxidant injection out of a well yields a forced gradient such that Condition 1 has a high flow velocity and high oxidant concentration; Condition 2 has a lower velocity and moderate oxidant concentration; and Condition 3 has near-ambient groundwater velocity and near-zero oxidant level.

    1.3 Evolution of Isco

    1.3.1 Research and Development Activities

    ISCO has evolved out of a long history of using chemical oxidation to destroy organic contaminants in water within the municipal and industrial water and waste treatment industry. The first step in the evolution of ISCO involved research and development (R&D) to adapt use of chemical oxidants like hydrogen peroxide and ozone to treat organic COCs in groundwater that was pumped to the surface and containerized in tank-based reactors (i.e., ex situ treatment) (Barbeni et al., 1987; Glaze and Kang, 1988; Bowers et al., 1989; Watts and Smith, 1991; Venkatadri and Peters, 1993). While a concept underlying in situ chemical oxidation was patented in 1986 (Brown and Norris, 1986), the first commercial in situ application of hydrogen peroxide occurred in 1984 to treat groundwater contaminated with formaldehyde (Brown et al., 1986).

    Starting in 1990, researchers began to explore hydrogen peroxide and modified Fenton’s reagent oxidation as applied in soil and groundwater environments (Watts et al., 1990, 1991, 1997; Watts and Smith, 1991; Tyre et al., 1991; Ravikumar and Gurol, 1994; Gates and Siegrist, 1993, 1995). Research was also initiated with alternative oxidants such as ozone (Bellamy et al., 1991; Nelson and Brown, 1994; Marvin et al., 1998) and potassium permanganate (Vella et al., 1990; Vella and Veronda, 1994; Gates et al., 1995; Schnarr et al., 1998; West et al., 1997; Siegrist et al., 1998a, b, 1999; Yan and Schwartz, 1998, 1999; Tratnyek et al., 1998; Urynowicz and Siegrist, 2000). The development of ISCO continues to expand as evidenced by more recent research with newer oxidants like sodium persulfate and sodium carbonate peroxide (percarbonate) (e.g., Brown et al., 2001; Block et al., 2004; Crimi and Taylor, 2007).

    In a parallel development, alternative methods of delivering oxidants into the subsurface were explored. ISCO was deployed by a range of methods including permeation through vertical probes (e.g., Jerome et al., 1997; Siegrist et al., 1998c; Moes et al., 2000), flushing by vertical and horizontal groundwater wells (e.g., Schnarr et al., 1998; West et al., 1997, 1998; Lowe et al., 2002), and reactive zone emplacement by either soil mixing (e.g., Gates and Siegrist, 1993) or pneumatic and hydraulic fracturing (e.g., Murdoch et al., 1997a, b; Siegrist et al., 1999).

    In the late 1990s and early 2000s, case study reports became available (e.g., USEPA, 1998; ESTCP, 1999), followed by the publication of the first reference book (Siegrist et al., 2001) and the first technical and regulatory reference manual (ITRC, 2001). These documents provided valuable insight into principles and practices, field experiences, and regulatory requirements. However, they did not provide the state-of-the-science knowledge and engineering know-how needed for a standard of practice to ensure effective, timely, and cost effective site-specific application of ISCO alone or in combination with other remedial options. As a result, the implementation of ISCO was plagued by uncertain and variable design and application practices. This caused ISCO performance to be unpredictable for some site conditions and remediation applications.

    To advance the science and engineering of ISCO and resolve questions regarding ISCO design and performance capabilities, ISCO research and development efforts escalated during the early 2000s. This research has been catalyzed in large part by the promising potential of ISCO and growing interest in its use, notably at DoD sites across the country. Around 2002, a major ISCO research initiative was launched within DoD’s Strategic Environmental Research and Development Program (SERDP) and Environmental Security Technology Certification Program (ESTCP) (http://www.serdp.org/Research/ISCO.cfm).

    A portfolio of ISCO projects and activities, both within the DoD ISCO initiative and through other sponsored research programs, has increased the understanding of ISCO. Advancements have included improved fundamental understanding of the following:

    COC oxidation chemistry and treatment (e.g., Gates-Anderson et al., 2001; Jung et al., 2004; Qiu et al., 2004; Smith et al., 2004; Watts et al., 2005a).

    Oxidant interactions with subsurface media (e.g., Siegrist et al., 2002; Crimi and Siegrist, 2003, 2004a, b; Anipsitakis and Dionysiou, 2004; Shin et al., 2004; Jung et al., 2005; Monahan et al., 2005; Mumford et al., 2005; Bissey et al., 2006; Jones, 2007; Teel et al., 2007; Sun and Yan, 2007; Sirguey et al., 2008; Urynowicz et al., 2008; Woods, 2008).

    Oxidative destruction of DNAPLs (e.g., Crimi and Siegrist, 2005; Heiderscheidt, 2005; Kim and Gurol, 2005; Urynowicz and Siegrist, 2005; Watts et al., 2005b; Siegrist et al., 2006; Smith et al., 2006; Heiderscheidt et al., 2008a; Petri et al., 2008a).

    Examination of newer oxidants (e.g., Liang et al., 2004a, b; Crimi and Taylor, 2007; Waldemer et al., 2007; Liang and Lee, 2008).

    Oxidant transport processes and deliverability (e.g., Choi et al., 2002; Lowe et al., 2002; Struse et al., 2002; Lee et al., 2003; Tunnicliffe and Thomson, 2004; Heiderscheidt, 2005; Ross et al., 2005; Zhang et al., 2005; Petri, 2006; Heiderscheidt et al., 2008a; Petri et al., 2008a).

    Combining ISCO with other remedies (e.g., Sahl, 2005; Dugan, 2006; Sahl and Munakata-Marr, 2006; Sahl et al., 2007).

    Treatability test methods (e.g., Haselow et al., 2003; Mumford et al., 2004; ASTM, 2007).

    Formulation of mathematical models and decision support tools (e.g., Kim and Choi, 2002; Heiderscheidt, 2005; Heiderscheidt et al., 2008b).

    The escalation in interest and R&D activities in ISCO during the early 2000s led to a dramatic increase in the published literature concerning the fundamental and applied understanding regarding chemical oxidants and their application for in situ remediation. In 2008, a critical review of the published literature relevant to ISCO science and technology was completed by Petri et al. (2008b) as a component of a DoD ESTCP project (Siegrist et al., 2010). Petri et al. identified nearly 600 ISCO-relevant publications encompassing four oxidants: permanganate, persulfate, hydrogen peroxide, and ozone. The publications were initially characterized and classified to provide insight into the evolution and current state of ISCO science (Figures 1.7 and 1.8). Then, a detailed review and analysis was made to elucidate key findings regarding ISCO and the implications for site-specific engineering and performance.

    A978-1-4419-7826-4_1_Fig7_HTML.gif

    Figure 1.7.

    Evolution of the (a) published literature base concerning ISCO and (b) the types of publications within a database of about 600 articles, theses and reports (Petri et al., 2008b).

    A978-1-4419-7826-4_1_Fig8_HTML.gif

    Figure 1.8.

    Classification of ISCO literature in Figure 1.7 based on the scope of the work (Petri et al., 2008b).

    The Petri et al. (2008b) review of the published literature concerning ISCO revealed the dramatic increase in published information beginning in the late 1990s (Figure 1.7). It also revealed marked differences in the scope of work carried out and the breadth and depth of understanding for each of the four oxidants and their application in situ (Figure 1.8). For example, the published literature concerning ISCO using hydrogen peroxide or permanganate oxidants markedly exceeds the literature base available for ozone and persulfate (Figure 1.8). The published information on ISCO using hydrogen peroxide is primarily focused on reaction chemistry with relatively less information concerning oxidant transport and modeling. In contrast, the published information on ISCO using permanganate spans reaction chemistry as well as transport and modeling (Figure 1.8).

    The review by Petri et al. (2008b) also included several findings that apply to all four oxidants and ISCO in general. For example, regardless of oxidant type, ISCO is challenged by hydrogeologic conditions, as are many non-ISCO remediation technologies. Also, ISCO can enhance or interfere with the use of other remediation technologies. The impacts of ISCO on biological processes and bioremediation were the most extensively investigated, yielding the following conclusions:

    ISCO does not sterilize the subsurface.

    Bioprocesses are often disrupted, but activity rebounds with time.

    Some ISCO effects may be beneficial to bioprocesses.

    Further detailed information concerning the fundamental understanding that has evolved out of the R&D completed concerning the use of hydrogen peroxide, permanganate, persulfate, and ozone for ISCO is provided in Chapters 2–5, respectively. Each of these chapters describes the respective oxidant properties, reaction chemistry, interactions in the subsurface, and treatability of different COCs. Information regarding the reactive transport of oxidants in a groundwater environment, including mathematical modeling approaches, is given in Chapter 6. The potential to combine ISCO with bioremediation, MNA and other technologies and approaches is described in Chapter 7.

    1.3.2 Field Applications

    The number of field applications of ISCO in the United States and abroad has grown rapidly during the past decade. Around 2000, the applications of ISCO using hydrogen peroxide (or modified Fenton’s reagent), ozone, and permanganate were highlighted in several published reports and texts (ESTCP, 1999; USEPA, 1998; Yin and Allen, 1999; Siegrist et al., 2001). More recently, Krembs (2008) completed a review of prior case history studies and carried out a critical review of 242 projects involving ISCO field applications to examine the site-specific design and performance of ISCO projects. Based on the projects reviewed by Krembs (2008), the number of applications of ISCO has grown rapidly since the late 1990s (Figure 1.9).

    A978-1-4419-7826-4_1_Fig9_HTML.gif

    Figure 1.9.

    Cumulative number of ISCO projects from 1995 to 2006 (Krembs, 2008).

    Based on the review by Krembs (2008), ISCO has been applied to sites with different geologic conditions (Figure 1.10a) and used to achieve different remediation goals and objectives (Figure 1.10b). Figure 1.11 shows several photographs illustrating the conditions and infrastructure present at sites where ISCO has been deployed. Further details concerning ISCO field applications and experiences are presented in Chapter 8.

    A978-1-4419-7826-4_1_Fig10_HTML.gif

    Figure 1.10.

    Relative number of ISCO applications (a) to sites with different geologic conditions (n = 149) and (b) for full-scale projects with different remediation objectives (n = 99) (Krembs, 2008). In (b) the length of the bar corresponds to the number of sites with the goal specified and the % label indicates the percentage of those sites meeting the goal set. Note: ACL - alternative (risk-based) cleanup levels, cm/s - centimeters per second, ft/d - feet per day, K - saturated hydraulic conductivity.

    A978-1-4419-7826-4_1_Fig11a_HTML.gifA978-1-4419-7826-4_1_Fig11b_HTML.gif

    Figure 1.11.

    Photographs illustrating some of the conditions and infrastructure at sites where ISCO has been deployed for groundwater remediation. Image captions give oxidant used, delivery method, and target contaminants. Note: CB - chlorobenzene, MGP - manufactured gas plant.

    1.4 System Selection, Design, and Implementation

    ISCO system selection, design, and implementation practices should rely on a clear understanding of ISCO and its applicability to a given set of contaminant and site conditions to achieve site-specific remediation objectives. A number of key issues may be relevant and need to be addressed during the selection, design, and implementation of ISCO regardless of the oxidant and delivery system being employed. These issues are listed below:

    Amenability of the target COCs to oxidative degradation.

    Effectiveness of the oxidant for NAPL destruction.

    Optimal oxidant loading (dose concentration and delivery) for a given target treatment zone in a given subsurface setting.

    Nonproductive oxidant consumption due to natural oxidant demand (NOD) exerted by natural organic matter, reduced inorganic species, and some mineral phases in the subsurface.

    Nonproductive oxidant consumption by autodecomposition reactions and free radical scavenging reactions.

    Potential adverse effects (e.g., mobilizing metals such as chromium, forming toxic byproducts, reducing formation permeability, generating off-gases and heat).

    Potential to combine ISCO with other remediation technologies and approaches.

    Matching the oxidant and delivery system to the COCs and site conditions is important for achieving site-specific remediation objectives. Moreover, the objectives need to be realistic and achievable, for ISCO as a stand-alone technology or as part of a combined remedy, based on the resources allocated for the ISCO project.

    The selection, design, and implementation of a remedial action are generally accomplished within a phased project approach, as depicted in Figure 1.12 (GAO, 2005) and described for ISCO in Chapter 9. During the feasibility study for a given site, consideration of ISCO as a viable remedial option is often based on the general benefits that ISCO can offer. These include rapid and extensive reactions with various COCs, applicability to many subsurface environments, ability to tailor ISCO to a site, and rapid implementation that can support property transfers and site redevelopment projects. Potential limitations for ISCO are also considered during decision-making, including the resistance of some COCs to complete chemical oxidation, the level of NOD exerted by the subsurface, the stability of the oxidant in the subsurface, constraints on effective oxidant distribution, possible fugitive gas emissions, the potential for contaminant rebound, and the effects of chemical addition on water quality.

    A978-1-4419-7826-4_1_Fig12_HTML.gif

    Figure 1.12.

    Selected phases and milestones in DoD’s environmental cleanup process (GAO, 2005).

    If ISCO is selected as a viable alternative for a particular site and a site-specific design must be accomplished, many choices and decisions have to be made. For example, choices must be made between oxidant type (e.g., hydrogen peroxide, persulfate, permanganate, ozone), delivery method (e.g., direct push probes, injection wells, air sparging wells), process control and performance monitoring, and so forth. As described in Chapter 9, these choices should be carefully made to improve the chances that the ISCO system will yield a sufficient concentration of a suitable oxidant in contact with the target COCs under amenable conditions over a sufficient period of time for the COCs to be destroyed.

    As discussed earlier, the ISCO systems that can be (and have been) applied in the field are highly varied in their features. Different oxidants and additives (such as stabilizers or activators) have been used. Concentrations and injection flow rates can vary widely, and a variety of subsurface delivery methods can be employed. In addition, overall site remediation goals and regulatory constraints may influence the remediation objectives established for the ISCO technology. To improve the confidence level in the choices and decisions made, treatability studies and field-scale pilot tests are frequently necessary. If ISCO is selected for a site, remedial design and system construction must be accomplished. Lastly, ISCO operation and performance monitoring ensues. Eventually site closure should be achieved through ISCO alone or in conjunction with another remediation technology or approach.

    Given the properties of chemical oxidants, the use of ISCO requires diligent attention to safety and waste management issues. ISCO involves the use of hazardous chemicals under field conditions and as such, worker and environmental safety must be carefully considered and effectively managed. Potential worker safety risks include those typically associated with standard construction operations and work at a contaminated site, as well as risks specific to ISCO, which are primarily associated with the aboveground handling of the oxidants. The nature and seriousness of the risks will depend on the oxidant form and concentrations used and the methods employed for general handling and subsurface delivery. Once oxidants are delivered into the subsurface, worker risk should be low because the ground surface is normally not disturbed by well-designed in situ delivery methods. Environmental risk includes daylighting of oxidants into streams or water bodies where there can be concern over water quality degradation and adverse effects on aquatic life. This risk can be managed through adequate engineering controls to prevent oxidant or contaminant movement away from the site to a sensitive environmental receptor.

    Waste management issues are related to type and volume of ISCO-generated waste. Possibly the most difficult wastes to manage are the containers, materials, and personal protective equipment that have oxidant residual in them. Flushing such materials with water can reduce the residual oxidant concentration and mass to acceptable levels. A reducing agent (e.g., sodium thiosulfate) can also be used to neutralize oxidants such as permanganate, however this will generate manganese (IV) oxide (MnO2) solids that need to be dewatered and disposed of appropriately. Care must be taken to avoid disposing of combustible materials that have contacted oxidants in bulk waste containers because fires are possible in such situations. Additional generated wastes expected during ISCO are drill cuttings produced during installation of oxidant delivery access points and other monitoring locations. Driven casing techniques are available to minimize this waste when it is a concern.

    1.5 Project Performance and Costs

    The effectiveness of ISCO varies: at some sites, ISCO has been applied and the destruction of the target COCs has occurred such that cleanup goals have been met in a cost-effective and timely manner, whereas at other sites, ISCO applications have had uncertain or poor in situ treatment performance. Poor performance has often been attributed to poor uniformity of oxidant delivery caused by low permeability zones and formation heterogeneity, excessive oxidant consumption by natural subsurface materials, or presence of large masses of DNAPLs (Siegrist et al., 2001, 2006, 2008a). In some applications, there have been concerns over secondary effects, such as mobilization of metals, loss of well screen and formation permeability, and gas evolution and fugitive emissions, as well as health and safety practices (Siegrist et al., 2001; Crimi and Siegrist, 2003; Krembs, 2008).

    Of the 242 ISCO projects examined by Krembs (2008) (the results of the study are included in Chapter 8), PCE or TCE were the targeted COCs at 70% of the sites, the subsurface conditions were characterized as permeable at 75% of the sites, and oxidants were delivered using permanent or temporary injection wells at 70% of the sites. Krembs (2008) also examined the reported performance and cost of ISCO projects. The percentage of full-scale ISCO projects that attempted to meet a specific goal and reported that they had met it were as follows:

    21% of 28 Projects attempting to achieve MCLs met this goal.

    44% of 25 Projects attempting to achieve ACLs met this goal.

    33% of Six projects attempting to reduce the COC mass by a certain percentage met this goal.

    82% of 34 Projects attempting to reduce the COC mass and/or time to cleanup met this goal.

    100% of Six projects attempting to evaluate effectiveness and optimize future injections met this goal.

    Krembs (2008) reported the median total cost for 55 ISCO projects to be $220,000; the median unit cost was $94 per cubic yard (cy) treated based on 33 projects with unit cost data. McDade et al. (2005) reported median and unit costs of $230,000 and $125/cy, respectively, for 13 ISCO projects. It is important to recognize that the cost of an ISCO project can vary by an order of magnitude or more depending on various factors. For example, sites with fuel hydrocarbons and permeable subsurface conditions typically cost less than those with DNAPLs or complex subsurface conditions. High unit costs can also result where ISCO has been used to treat relatively smaller source zones.

    To gain an understanding of the current status of technology practices and the performance associated with ISCO, an ESTCP workshop was convened at the Colorado School of Mines in 2007 (Siegrist et al., 2008b). More than 40 experts attended, representing chemical companies, technology vendors, environmental consultants, academics, and remedial project managers. The workshop included a series of presentations, panel sessions, and breakout sessions, along with an exercise where six site scenarios were used to obtain views regarding ISCO design and performance. The following views expressed and consensus developed at the workshop provide insights into ISCO applications and performance expectations:

    ISCO may be considered for a wide range of situations, but site-specific conditions interacting with ISCO technology attributes determine the performance that could be considered reliably achievable.

    As an illustration of this view, a strong majority of the workshop participants indicated that they would consider ISCO for a range of contaminated site scenarios. However, the degree of anticipated ISCO performance, the timeframe necessary, and the costs varied between different scenarios.

    The success or failure of an ISCO application is strongly dependent not only on the site-specific conditions but also on the remediation objectives and the resources (e.g., time and money) made available to implement the ISCO system.

    Multiple injection events are commonly required and should be anticipated.

    Where ISCO has not achieved treatment goals, the two fundamental reasons are:

    Oxidant was not distributed throughout the target treatment zone (TTZ).

    An insufficient amount of oxidant was delivered to the TTZ.

    Performance deficiencies based on these two reasons were viewed as more likely to occur under the following circumstances:

    Site characterization is inadequate and the contaminant mass is poorly understood.

    The subsurface is highly heterogeneous.

    The design neglects the mass of contaminants that are sorbed.

    Presence of DNAPLs is unknown or not accounted for.

    Co-contaminants are present that also consume oxidants.

    Oxidants migrate out of the target treatment zone.

    The oxidant does not persist as long as expected.

    Rebound, where post-ISCO concentrations of target COCs in groundwater within a TTZ return to levels near or even above those present prior to ISCO, is a relatively common occurrence that may or may not be a negative condition or reflect an inherent shortcoming of ISCO or a site-specific performance deficiency.

    The rebound observed at an ISCO treated site can be beneficial if it is used in an observational approach to refine the conceptual site model and refocus subsequent treatment.

    The use of ISCO can be viewed as an ongoing, iterative process that will take advantage of contaminant rebound rather than view it as an indication that the technology was inappropriate for a site or was applied improperly.

    1.6 Summary

    Groundwater contamination by organic chemicals is a widespread problem throughout the United States and industrialized world. At sites with contaminant source zones remaining in the subsurface (e.g., NAPLs), groundwater contamination can persist for decades, requiring large expenditures of public and private funds to mitigate public health and environmental risks. ISCO has proven to be a useful remediation technology for the most prevalent organic contaminants in groundwater. Its advantages include its applicability to several of the most prevalent organic COCs, its adaptability to a wide range of subsurface conditions, its potential to treat a site rapidly, and its ability to be combined with other in situ technologies and remediation approaches. Its limitations include the resistance of some COCs to complete chemical oxidation, the NOD exerted by the subsurface, the short longevity of many oxidants in the subsurface, constraints on effective oxidant distribution, potential for COC rebound, and detrimental side effects, such as fugitive gas emissions or mobilization of some metals.

    The evolution of a successful remediation technology typically involves a series of stages, from the initial idea generation through research and development, demonstration and testing, increasing field applications by practitioners, and eventual establishment of widely accepted standards of practice. The evolution of ISCO has followed this path. During the 1990s, fundamental and applied laboratory research elucidated many aspects of the chemical oxidation reactions for common organic chemicals in aqueous systems. Laboratory research also explored the transport processes affecting oxidant delivery and dispersal in soil or groundwater systems. Pilot-scale and full-scale applications demonstrated in situ treatment of low levels of solvents and petrochemicals and, to a lesser degree, NAPLs. Results of this work were disseminated in journal articles and technical reports. In the late 1990s and early 2000s, case study reports became available (e.g., USEPA, 1998; ESTCP, 1999) and the first reference book (Siegrist et al., 2001) and reference manual (ITRC, 2001) were published. Subsequently, research and development efforts were catalyzed by a major initiative within SERDP and ESTCP. Fundamental understanding of ISCO evolved, and the number of publications in the open literature grew exponentially. Simultaneously, the number and diversity of field applications expanded substantially. A standard of practice for ISCO began to emerge and be documented and disseminated for widespread use by new and experienced practitioners and regulators (e.g., ITRC, 2005; Huling and Pivetz, 2006; Siegrist et al., 2008b, 2010).

    The site-specific selection, design, and implementation of ISCO require consideration of a variety of questions (Table 1.6) and careful attention to key points to enable success and avoid problems (Table 1.7). The chapters that follow provide details regarding the principles and practices of ISCO, providing a sound basis for site-specific engineering to achieve successful remediation of contaminated groundwater.

    Table 1.6.

    Frequently Asked Questions on ISCO Selection, Design, and Implementation (Crimi et al., 2008)

    Table 1.7.

    Key Points to Consider Which Can Enable Success with ISCO and Help Avoid Problems (Siegrist et al., 2010)

    References

    Anipsitakis GP, Dionysiou DD. 2004. Radical generation by the interaction of transition metals with common oxidants. Environ Sci Technol 38:3705–3712.CrossRef

    ASTM (American Society for Testing and Materials). 2007. ASTM D7262-07 Standard Test Method for Estimating the Permanganate Natural Oxidant Demand of Soil and Aquifer Solids. ASTM International, West Conshohocken, PA, USA, 5 p.

    ATSDR (Agency for Toxic Substances and Disease Registry). 2009. ATSDR 2007 CERCLA Priority List of Hazardous Substances. http://www.atsdr.cdc.gov/cercla/. Accessed June 23, 2010.

    Barbeni M, Nfinero C, Pelizzetti E, Borgarello E, Serpon N. 1987. Chemical degradation of chlorophenols with Fenton’s reagent. Chemosphere 16:2225–2237.CrossRef

    Bellamy WD, Hickman PA, Ziemba N. 1991. Treatment of VOC-contaminated groundwater by hydrogen peroxide and ozone oxidation. J Water Pollut Control Fed 63:120–128.

    Bissey LL, Smith JL, Watts RJ. 2006. Soil organic matter-hydrogen peroxide dynamics in the treatment of contaminated soils and groundwater using catalyzed H2O2 propagations (modified Fenton’s reagent). Water Res 40:2477–2484.CrossRef

    Block PA, Brown RA, Robinson D. 2004. Novel Activation Technologies for Sodium Persulfate In Situ Chemical Oxidation. Proceedings, Fourth International Conference on the Remediation of Chlorinated and Recalcitrant Compounds, Monterey, CA, USA, May 24–27, Paper 2A-05.

    Bowers AR, Gaddipati P, Eckenfelder WW, Monsen RM. 1989. Treatment of toxic or refractory wastewaters with hydrogen peroxide. Water Sci Technol 21:477–486.

    Brown RA, Norris RD. 1986. Method for Decontaminating a Permeable Subterranean Formation. U.S. Patent 4,591,443.

    Brown RA, Norris RD, Westray M. 1986. In Situ Treatment of Groundwater. Presented at Haz Pro ’86, Baltimore, MD, USA, April 1–3.

    Brown RA, Skaladany G, Robinson D, Fiacco RJ. 2001. Comparing Permanganate and Persulfate Treatment Effectiveness for Various Organic Contaminants. Proceedings, First International Conference on Oxidation and Reduction Technologies for In-Situ Treatment of Soil and Groundwater, Niagara Falls, Ontario, Canada, June 25–29.

    Choi H, Lim H-N, Hwang T-M, Kang J-W. 2002. Transport characteristics of gas phase ozone in unsaturated porous media for in-situ chemical oxidation. J Contam Hydrol 57:81–98.CrossRef

    Crimi ML, Siegrist RL. 2003. Geochemical effects associated with permanganate oxidation of DNAPLs. Ground Water 41:458–469.CrossRef

    Crimi ML, Siegrist RL. 2004a. Association of cadmium with MnO2 particles generated during permanganate oxidation. Water Res 38:887–894.CrossRef

    Crimi ML, Siegrist RL. 2004b. Impact of reaction conditions on MnO2 genesis during permanganate oxidation. J Environ Eng 130:562–572.CrossRef

    Crimi ML, Siegrist RL. 2005. Factors affecting effectiveness and efficiency of DNAPL destruction using potassium permanganate and catalyzed hydrogen peroxide. J Environ Eng 131:1716–1723.CrossRef

    Crimi ML, Taylor J. 2007. Experimental evaluation of catalyzed hydrogen peroxide and sodium persulfate for destruction of BTEX contaminants. Soil Sediment Contam 16:29–45.CrossRef

    Crimi ML, Siegrist RL, Petri B, Krembs F, Simpkin T, Palaia T. 2008. In Situ Chemical Oxidation for Remediation of Contaminated Groundwater: Frequently Asked Questions. Prepared for the DoD Environmental Security Technology Certification Program (ESTCP), Arlington, VA, USA, 25 p. http://www.estcp.org/.

    Dugan P. 2006. Coupling In Situ Technologies for DNAPL Remediation and Viability of the PITT for Post-Remediation Performance Assessment. PhD Dissertation, Environmental Science and Engineering Division, Colorado School of Mines, Golden, CO, USA, August.

    ESTCP (Environmental Security Technology Certification Program). 1999. Technology Status Review: In Situ Oxidation. ESTCP, Arlington, VA, USA, 50 p. http://www.estcp.org/documents/techdocs/ISO_Report.pdf. Accessed June 23, 2010.

    GAO (U.S. Government Accountability Office). 2005. Report to Congressional Committees. Groundwater Contamination: DOD Uses and Develops a Range of Remediation Technologies to Clean up Military Sites. GAO-55-666. GAO, Washington, DC, USA, 46 p. http://www.gao.gov/products/GAO-05-666. Accessed June 23, 2010.

    Gates DD, Siegrist RL. 1993. Laboratory Evaluation of Chemical Oxidation Using Hydrogen Peroxide. Report from The X-231B Project for In Situ Treatment by Physicochemical Processes Coupled with Soil Mixing, ORNL/TM-12259. Oak Ridge National Laboratory, Oak Ridge, TN, USA.

    Gates DD, Siegrist RL. 1995. In situ chemical oxidation of trichloroethylene using hydrogen peroxide. J Environ Eng 121:639–644.CrossRef

    Gates DD, Siegrist RL, Cline SR. 1995. Chemical Oxidation of Contaminants in Clay or Sandy Soil. Proceedings, American Society of Civil Engineering (ASCE) National Conference on Environmental Engineering, Pittsburgh, PA, USA, July.

    Gates-Anderson DD, Siegrist RL, Cline SR. 2001. Comparison of potassium permanganate and hydrogen peroxide as chemical oxidants for organically contaminated soils. J Environ Eng 127:337–347.CrossRef

    Glaze WH, Kang JW. 1988. Advanced oxidation processes for treating groundwater contaminated with TCE and PCE: Laboratory studies. J Am Water Works Assoc 5:57–63.

    Haselow JS, Siegrist RL, Crimi ML, Jarosch T. 2003. Estimating the total oxidant demand for in situ chemical oxidation design. Remediation 13:5–15.CrossRef

    Heiderscheidt JL. 2005. DNAPL Source Zone Depletion During In Situ Chemical Oxidation (ISCO): Experimental and Modeling Studies. PhD Dissertation, Environmental Science and Engineering Division, Colorado School of Mines, Golden, CO, USA, August.

    Heiderscheidt JL, Siegrist RL, Illangasekare TH. 2008a. Intermediate-scale 2-D experimental investigation of in situ chemical oxidation using potassium permanganate for remediation of complex DNAPL source zones. J Contam Hydrol 102:3–16.CrossRef

    Heiderscheidt JL, Crimi ML, Siegrist RL, Singletary M. 2008b. Optimization of full-scale permanganate ISCO system operation: Laboratory and numerical studies. Ground Water Monit Remediat 28:72–84.CrossRef

    Huling SG, Pivetz BE. 2006. Engineering Issue Paper: In-Situ Chemical Oxidation. EPA 600-R-06-072. U.S. Environmental Protection Agency (USEPA) Office of Research and Development. National Risk Management Research Laboratory, Cincinnati, OH, USA, 60 p. http://www.epa.gov/tio/tsp/issue.htm#EF. Accessed June 23, 2010.

    ITRC (Interstate Technology & Regulatory Council). 2001. Technical and Regulatory Guidance for In Situ Chemical Oxidation of Contaminated Soil and Groundwater (ISCO-1). Prepared by the Interstate Technology & Regulatory Cooperation Work Group In Situ Chemical Oxidation Work Team. http://www.itrcweb.org/guidancedocument.asp?TID=13. Accessed June 23, 2010.

    ITRC. 2005. Technical and Regulatory Guidance for In Situ Chemical Oxidation of Contaminated Soil and Groundwater, 2nd ed (ISCO-2). Prepared by the ITRC In Situ Chemical Oxidation Team. http://www.itrcweb.org/guidancedocument.asp?TID=13. Accessed June 23, 2010.

    Jerome KM, Riha B, Looney BB. 1997. Demonstration of In Situ Oxidation of DNAPL Using the Geo-Cleanse Technology. WSRC-TR-97-00283. Westinghouse Savannah River Company, Aiken, SC, USA.CrossRef

    Jones LJ. 2007. The Impact of NOD Reaction Kinetics on Treatment Efficiency. MS Thesis, University of Waterloo, Waterloo, ON, Canada.

    Jung H, Kim J, Choi H. 2004. Reaction kinetics of ozone in variably saturated porous media. J Environ Eng 130:432–441.CrossRef

    Jung H, Ahn Y, Choi H, Kim IS. 2005. Effects of in-situ ozonation on indigenous microorganisms in diesel contaminated soil: Survival and regrowth. Chemosphere 61:923–932.CrossRef

    Kavanaugh MC, Rao PSC, Abriola L, Cherry J, Destouni G, Falta R, Major D, Mercer J, Newell C, Sale T, Shoemaker S, Siegrist RL, Teutsch G, Udell K. 2003. The DNAPL Cleanup Challenge: Is There a Case for Source Depletion? EPA/600/R-03/143. USEPA National Risk Management Research Laboratory, Cincinnati, OH, USA, 129 p. http://www.epa.gov/nrmrl/pubs/600R03143/600r03143.htm. Accessed June 23, 2010.

    Kim J, Choi H. 2002. Modeling in situ ozonation for the remediation of nonvolatile PAH-contaminated unsaturated soils. J Contam Hydrol 55:261–285.CrossRef

    Kim K, Gurol MD. 2005. Reaction of nonaqueous phase TCE with permanganate. Environ Sci Technol 39:9303–9308.CrossRef

    Krembs FJ. 2008. Critical Analysis of the Field Scale Application of In Situ Chemical Oxidation for the Remediation of Contaminated Groundwater. MS Thesis, Environmental Science and Engineering Division, Colorado School of Mines, Golden, CO, USA, April.

    Lee ES, Seol Y, Fang YC, Schwartz FW. 2003. Destruction efficiencies and dynamics of reaction fronts associated with the permanganate oxidation of trichloroethylene. Environ Sci Technol 37:2540–2546.CrossRef

    Liang C, Lee IL. 2008. In situ iron activated persulfate oxidative fluid sparging treatment of TCE contamination: A proof of concept study. J Contam Hydrol 100:91–100.CrossRef

    Liang C, Bruell CJ, Marley MC,

    Enjoying the preview?
    Page 1 of 1