You are on page 1of 63

23.

HYDRAULIC DESIGN FOR GROUNDWATER


CONTAMINATION
Paul C. Johnson

Department of Civil and Environmental Engineering

Arizona State University

Tempe, Arizona

23.1. INTRODUCTION
This chapter focuses on in situ treatment systems designed to eliminate or minimize the impacts of hazardous chemicals on
groundwater quality. This chapter builds on the principles introduced in Chapter 4. First, typical soil and groundwater
contamination scenarios of interest are introduced. Then a discussion of general remediation design principles and strategies
follows. Finally, the design, monitoring, and refinement of selected in situ remediation technologies is presented. These include
both conventional and developing technologies, and include containment, reaction barrier, active remediation, and passive
remediation systems.

23.1.1. Unique Features of In Situ Treatment Technology Design


The design of in situ treatment systems for groundwater contamination is uniquely different from the design of above-ground
treatment and hydraulic conveyance systems. The subsurface is an environment that is difficult to describe precisely; it is
naturally heterogeneous and, due to practical constraints, characterization data are generally limited. Thus, treatment systems
must be designed under conditions of great uncertainty–usually the amount of chemical spilled is unknown, the timing of the
release is unknown, and the system into which the spill occurred is poorly characterized. For this reason, the design of in situ
treatment systems relies heavily on

conceptual models,

screening-level calculations,

empiricism, heuristics, and experience, and

monitoring and refinement of the design.

A conceptual model is a realization of how the subsurface might look and how contaminants might move, based on the
available data. Often, there are several plausible conceptual models, and the treatment system designer must allow for a range
of possibilities in the design. It is also recognized that the conceptual model is continuously refined as new data become
available.

Screening level calculations are used to estimate treatment performance limits. They are generally biased toward estimating
best-case performance, and are useful for making rough estimates of treatment times and treatment limits. They are primarily
used for feasibility screening and creating the initial treatment system design. More sophisticated models are available and can
be used; however, given the typical level of site characterization data and gross approximations that must be made, their use is
rarely warranted, except at those sites where treatment costs are projected to be very high.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Empiricism and experience play a key role in the design as well. In many ways, each site is unique; yet, the collective experience
from many sites is invaluable. This design field is still relatively young and most experiences are not well-documented. As is
often the case, successful applications are more likely to be reported than failures, although the knowledge from both would be
equally valuable. In practice, there are many “rules of thumb” that designers use and some are well founded in theory, while
others are indefensible (yet they still persist in practice). In any case, it is useful to be aware of these rules-of-thumb even if
their bases are not well documented in the literature.

Finally, given the inherent large initial uncertainty, it is important to recognize that the design phase continues well past the
construction, installation, and operation of the initial system. Following sound design practices does not guarantee success; it
does, however, provide a higher probability of success at most sites. Appropriate monitoring is essential to verify assumptions
built into the conceptual model and initial design, and system design refinement should be anticipated in the initial design. For
this reason, it is important to build robust systems that are both flexible and expandable.

23.1.2. Overview
It has been estimated that costs associated with the remediation of known hazardous waste sites in the United States will
exceed $700 billion (all costs are quoted in U.S. dollars) and that the current average rate of expenditure is about $20 billion
(Bredehoeft, 1994). Despite the fact that society has been tackling this difficult problem since the 1970s, the environmental
profession still struggles to find cost-effective remediation processes. Remediation costs for service station sites now range
from $100,000 to $500,000 per site, while costs for the larger landfill and manufacturing sites often exceed $10,000,000 per
site. Costs associated with sites where groundwater is contaminated generally exceed those where only soil is impacted by at
least an order of magnitude. Clearly, these are not trivial expenses; thus, appropriate technology selection, design, and
optimization are important activities.

23.1.3. Groundwater Contamination Scenarios–Point Versus Area


Sources
Groundwater contamination is prevalent in most industrial societies. Leaking fuel tanks, petroleum storage and refining, solvent
degreasing activities, chemical manufacturing and storage, landfills, and mining activities have all contributed to groundwater
contamination problems. Groundwater contamination is also prevalent in most modern agricultural areas, where the
combination of irrigation and fertilizer, pesticide, and herbicide use have contributed to many large-scale groundwater
contamination problems. Fig. 23.1 [US Enviromental Protection Agency (USEPA), 1990] summarizes the more common sources
of groundwater contamination and the relative frequency at which they are expected to cause groundwater impacts. Septic
tanks and leaking underground fuel storage tanks are considered to be the largest single contributor to the number of
groundwa ter contamination sources. Agricultural activities are likely the largest contributor to groundwater contamination in
terms of the actual extent of impact.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 23.1 Common sources of groundwater contamination and the number of states and
territories considering each to be major threat to gro (USEP A, 1990, after Fetter, 1993).

For the purposes of this chapter, we first segregate the universe of groundwater contamination scenarios into two main
categories:

Point sources,

Area, or distributed sources

Underground storage tanks, above-ground storage tanks, landfills, pipeline releases, chemical manufacturing and petroleum
refining locations, wood treating facilities, and so forth are all considered to be point sources. Agricultural activities are often
considered to be area, or distributed sources.

By virtue of their large areal extent, the environmental engineering profession has yet to develop effective solutions for the
remediation of area source contamination problems. These often involve widespread region- or basin-scale nitrate or pesticide
contamination. Currently, the only practicable solution for these problems appears to be above-ground water blending and
direct well-head treatment. In other words, contaminated groundwater is either diluted with water of higher quality as it is
pumped from the aquifer, or else it is treated as necessary above-ground as it is removed from the aquifer. As stated, area
source problems are rarely addressed with in situ treatment systems, and therefore, these will not be considered further in this
chapter.

23.1.4. Groundwater Contamination Scenarios–Segregation by


Contaminant Type
Point source groundwater contamination problems are divided here into three main categories according to the contaminant
properties. The three categories are

Light nonaqueous phase liquids (LNAPLs)

Dense nonaqueous phase liquids (DNAPLs)

Inorganics and other dissolved constituents

LNAPLs, in their pure liquid form, are less dense than water (ρ < 1 g/cm3 ). Most LNAPL sites involve the release of petroleum
products or crude oil (e.g., service stations, refineries, pipeline spills). DNAPLs, in their pure liquid form, are more dense than
water (ρ > 1 g/cm3 ). DNAPL contamination is generally found near sites where dry cleaning and aviation, automobile, and
electronic circuit board degreasing operations took place. Historically these used chlorinated solvents, such as
trichloroethylene (TCE) and perchloroethylene (PCE). Metals and salts fall into the category of inorganics and other dissolved
constituents. Mining operations, electroplating operations, leaking wastewater treatment facilities, and landfills are examples of
sources that historically have added contaminants to groundwater in dissolved form.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
23.1.5. Groundwater Contamination Scenarios–Subsurface
Contaminant Distributions
Figures 23.2a, 23.2b, and 23.2c schematically illustrates release scenarios involving LNAPLs, DNAPLs, and inorganics. While
very simplistic, these figures highlight key characteristics that impact subsurface characterization activities, remediation
design, and monitoring.

Figure 23.2 Example contaminant release scenarios: (a) LNAPL release.(b) DNAPL release.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 23.2 (Continued) (c) Example contaminant release scenarios: (c) dissolved contaminant
release.

As their density is less than that of water, one can expect to find LNAPLs in the general vicinity of the water table. Provided that
a given spill or leak is of sufficient size and that there are not any impermeable vertical flow barriers, LNAPLs travel downward
and pool on top of the water table. Then over time, water table fluctuations cause a vertical smearing so that immiscible-phase
LNAPL can be found trapped by capillary forces in the soil pores above and below the current groundwater table level. LNAPL
solubilities are generally so low that dissolution (and other natural loss processes) occurs very slowly and pools of LNAPL will
persist for decades.

Dissolution of LNAPLs into groundwater leads to the formation of dissolved contaminant plumes. Aquifer characteristics and
chemical properties control the growth and extent of the dissolved contaminant plume. Fortunately, most LNAPL sites involve
the release of petroleum hydrocarbons and some of these are aerobically degradable by indigenous microorganisms. The
result is that natural chemical degradation often limits the growth of the dissolved LNAPL plume. For example, two survey
studies have shown that most benzene, toluene, ethylbenzene, and xylenes (BTEX) dissolved plumes at leaking underground
storage tank (LUST) sites do not extend more than 300 ft (100 m) from the downgradient edge of the source zone (Mace et al.,
1997; Rice et al., 1995). It is important to note, however, that not all fuel-related contaminants degrade at an appreciable rate.
For example, some oxygenates (e.g., MTBE), are highly soluble and resistant to degradation. Consequently, these compounds
generally migrate much further distances than the more degradable monoaromatic chemicals (e.g., benzene, toluene, xylenes).
The behavior of these compounds is currently the focus of intense research.

DNAPL sites are generally more complex and challenging. Provided that a given spill or leak is of sufficient size and that there
are not any impermeable vertical flow barriers, DNAPLs travel downward to the water table. Initially, they too will pool on top of
the water table; but, provided that the spill or leak is large enough, the weight of the DNAPL pool will exceed the entry pressure
of the medium and at that point the DNAPL will penetrate through the groundwater table. It will continue to migrate downward
until a vertical flow barrier is encountered, or until all of the DNAPL liquid is bound by capillary forces. DNAPL vertical migration
in aquifers appears to be very sensitive to small changes in soil structure, and as a DNAPL migrates downward, it may also
spread horizontally along thin, more permeable, soil layers. DNAPL solubilities are also generally low so that dissolution (and
other natural loss processes) occurs very slowly and DNAPL found below the water table will persist for decades. The
assessment and remediation of DNAPL sites is generally much more challenging than that of LNAPL sites.

Most DNAPLs of interest are chlorinated solvents (e.g., TCE, PCE, and so on) and these degrade slowly, if at all, under natural
conditions. This results in dissolved plumes that are much longer than the dissolved LNAPL plumes that might be found in
similar settings, and plumes > 1500 ft (500 m) in length are not uncommon; in fact, plumes several miles (kilometers) in length
can be found in regions of concentrated historical aviation or electronics manufacturing.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
In the case of inorganics, such as those originating from the landfill inFig. 23.2, the contaminants are already dissolved in
solution when they reach the water table. Thus, these chemicals follow groundwater flow once they reach the aquifer. Dissolved
inorganic chemical migration is governed by complex geochemical interactions; however, in some cases of interest (e.g.,
nitrate contamination), the contaminants typically are not transformed or retarded, and may also form long dissolved plumes
that persist for extended periods of time. These solutions generally behave as neutrally buoyant liquids (ρ ≈ 1 g/cm3 ); however,
if the total concentration of inorganics exceeds about 10,000 mg/L H2 O, then density gradients will cause these solutions to
migrate vertically downward within aquifers.

23.2. REMEDIATION GOALS


When subsurface soil and/or groundwater contamination have been identified, decisions must be made as to whether or not
treatment is necessary, and if so, to what levels. Often regulatory guidelines prescribe the actions that must be taken. In most
cases, contaminant concentrations in soil and groundwater are compared with levels that have been deemed acceptable.
These levels are sometimes referred to as target levels. If exceedences are noted, then treatment or further study is required;
otherwise, no action other than monitoring is typically required.

While the development and enforcement of target levels is not the focus of this chapter, it is important to recognize that target
levels play a significant role in the selection and design of in situ treatment systems. There is no universal set of target levels,
and they vary from state to state and country to country. Thus, it is worthwhile to briefly discuss the range of situations that
might be encountered in practice.

Target levels may be prescribed in terms of acceptable groundwater concentrations, or in terms of soil concentrations that are
expected to be sufficiently protective of groundwater quality. In the United States, target levels are often linked to

maximum contaminant levels (MCLs) or maximum contaminant level goals (MCLGs)

risk-based considerations, or

resource protection goals

23.2.1. Maximun Contaminant Levels (MCL)


In the United States, maximum contaminant levels (MCLs) are concentrations that must not be exceeded at any drinking water
supply point (e.g., the home faucet). Water suppliers are responsible for meeting these criteria. MCLs are promulgated by
USEPA, and are based on considerations of health effects, aesthetic impacts, available treatment technologies, analytical
limitations, and cost. Table 23.1 lists MCLs and MCLGs for some chemicals of interest (USEPA 1996a). Historically, it was
common for regulatory agencies to uniformly adopt MCLs as target levels for groundwater quality.

Table 23.1 Regulatory and Human Health Benchmarks Used for Soil and Groundwater Target Cleanup Levels

Chemical Maximum Contaminant Level Goal Maximum Contaminant Level (MCL) Water Health-Based Limit*
(MCLG) (μg/L) (μg/L) Mq/L

Acenapthene 2,000

Acetone 4,000

Aldrin 0.005

Anthracene 10,000

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Chemical Maximum Contaminant Level Goal Maximum Contaminant Level (MCL) Water Health-Based Limit
(MCLG) (μg/L) (μg/L) Mq/L

Antimony 6 6

Arsenic 50

Barium 2000 2,000

Benz(a)anthracene 0.1

Benzene 0.005

Benzo(b)fluoranthene 0.1

Benzo(k)fluoranthene 1.0

Benzoic acid 100,000

Benzo(a)pyrene 0.2

Beryllium 4 4

Bis (2-chloroethyl) ether 0.08

Bis (2-ethylhexyl) phthalate 6

Bromodichloromethane 100*

Bromoform 100*
(tribromomethane)

Butanol 4,000

Butyl benzyl phthalate 7,000

Cadmium 5 5

Carbazole 4

Carbon disulfide 4,000

Carbon tetrachloride 5

Chlordane 2

p-Chloroaniline 100

Chlorobenzene 100 100

Chlorodibromomethane 60 100*

Chloroform 100*

2-Chlorophenol 200

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Chemical Maximum Contaminant Level Goal Maximum Contaminant Level (MCL) Water Health-Based Limit
(MCLG) (μg/L) (μg/L) Mq/L

Chromium 100 100

Chromium (III) 40,000

Chromium (VI) 100

Chrysene 10

Cyanide (amenable) 200 200

DDD 0.4

DDE 0.3

DDT 0.3

Dibenze(a,h)anthracene 0.01

Di-n-butyl phthalate 4,000

1,2-Dichlorobenzene 600 600

1,4-Dichlorobenzene 75 75

3,3-Dichlorbenzidene 0.2

1,1-Dichloroethane 4,000

1,2-Dichloroethane 5

1,1-Dichloroethylene 7 7

cis–1,2-Dichloroethylene 70 70

trans–1,2-Dichloroethylene 100 100

2,4-Dichlorophenol 100

1,2-Dichloropropane 5

1,3-Dichloropropene 0.5

Dieldrin 0.005

Diethylphthalate 30,000

2,4-Dimethylphenol 700

2,4-Dinitrophenol 40

2,4-Dinitrotoluene 0.1

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Chemical Maximum Contaminant Level Goal Maximum Contaminant Level (MCL) Water Health-Based Limit
(MCLG) (μg/L) (μg/L) Mq/L

2,6-Dinitrotoluene 0.1

Di-n-octyl phthalate 700

Endosulfan 200

Endrin 2 2

Ethylbenzene 700 700

Fluoranthene 1,000

Fluorene 1,000

Heptachlor 0.4

Heptachlor epoxide 0.2

Hexachlorobenzene 1

Hexachloro-1,3-butadiene 1 1

α-HCH (α-BCH) 0.01

β-HCH (β-BCH) 0.05

HCH (Lindane) 0.2 0.2

Hexachlorocyclopentadiene 50 50

Hexachloroethane 6

Indeno(1,2,3-cd)pyrene 0.1

Isophorone 90

Mercury 2 2

Methoxychlor 40 40

Methylbromide 50

Mthylene Chloride 5

2-Methylphenol (o-cresol) 2,000

Napthalene 1,000

Nickel 100

Nitrobenzene 20

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Chemical Maximum Contaminant Level Goal Maximum Contaminant Level (MCL) Water Health-Based Limit
(MCLG) (μg/L) (μg/L) Mq/L

N-Nitrosodiphenylamine 20

N-Nitrosodi-n-propylamine 0.01

Pentachlorophenol 1

Phenol 20

Pyrene 1,000

Selenium 50 50

Silver 200

Styrene 100 100

1,1,2,2-Tetrachloroethane 0.4

Tetrachloroethylene 5

Thallium 0.5 2

Toluene 1,000 1,000

Toxaphene 3

1,2,4-Trichlorobenzene 70 70

1,1,1-Trichloroethane 200 200

1,1,2-Trichloroethane 3 5

Trichloroethylene 0 5

2,4,5-Trichlorophenol 4,000

2,4,6-Trichlorophenol 8

Vandium 300

Vinyl acetate 40,000

Vinyl chloride 2
(chloroethene)

m-, o-, and p-Xylene 10,000 10,000

Zinc 10,000

23.2.2. Risk-Based Target Levels

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
In recent years some regulatory agencies have moved away from using MCLs as fixed cleanup levels, and have begun to
consider risk-based target levels [e.g., American Society of Testing and Materials (ASTM), 1995]. Like MCLs, risk-based target
levels are concentrations also deemed to be protective; however, they differ from MCLs in that they take into account site-
specific considerations, the beneficial uses of the aquifer (or soil), the distance from the contaminant source zone and actual
water supply wells (or other potential receptors), and any dilution that might occur during groundwater pumping. Risk-based
target levels may be larger, or smaller, than MCLs, depending on the site-specific conditions.

23.2.3. Resource Protection Goals


In some areas aquifers are considered to be valued resources, and therefore must be protected from any impacts. In such
areas, no level of impact is acceptable. These anti degradation policies require that all contamination must be remediated and
the resource must be restored to pristine, or background, conditions. Of the three approaches, this is typically the most stringent
as the cleanup goal is either zero, or the background concentration. It is also typically the most difficult goal to achieve.

23.2.4. Application of the Target Levels—Remediation, Points of


Compliance, and Containment
Once target levels are selected, one must choose how and where to apply them. Two obvious choices are

apply target levels everywhere in the soil and groundwater, or

apply the target levels to the boundary of a compliance zone.

While total cleanup of soils and groundwater is always preferred, it is not always practicable. For this reason, a compliance zone
is sometimes negotiated. Contaminant concentrations must not exceed the target levels outside of this compliance zone, but
may exceed them within the compliance zone. This approach is similar to that taken when permitting mixing, or dilution zones
for surface water discharges.

The allowance of a compliance zone greatly impacts the range of technologies that can be selected at a site. In this case
complete remediation of the source zone and groundwater is not always necessary and this opens the door for consideration of
contaminant containment systems. In summary, prior to selecting and designing in situ treatment technologies, the user must

define the target treatment levels (both in soil and groundwater),

define the points of compliance where these goals are to be applied, and

define the time frame within which compliance must be achieved.

23.3. INSITU TREATMENT TECHNOLOGIES–GENERAL


CLASSIFICATIONS
In situ treatment technologies are generally designed to perform one, or more of the following functions:

Contaminant source zone removal

Aquifer restoration

Prevent, or minimize continued contaminant migration

Rarely does a single technology accomplish all three of these goals, and so combinations and sequences of complementary
technologies are often used.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
23.3.1. Source Zone Treatment Technologies
Source zone treatment technologies target removal and destruction of the residual contaminants in soil that serve as the source
for groundwater contamination. The goal here is to treat the cause of the groundwater contamination, rather than the symptom.
Example source zone treatment technologies include

Free-product recovery

Excavation and disposal or above-ground treatment

In situ soil venting

Bioventing

In situ air sparging

Enhanced in situ soil venting with soil heating and/or soil fracturing

In situ vitrification

Phytoremediation

Groundwater pump and treat systems

Fig 23.3 depicts simple process schematics of the most common of these source zone treatment technologies.

Figure 23.3 Example source zone treatment technology processes schematic: (a) free-product
recovery, (b) in situ air sparging, (c) soil vapor extraction, and (d) bio venting.

23.3.2. Aquifer Restoration Technologies


Aquifer restoration technologies target treatment of dissolved contaminant plumes. These technologies may be employed
before, during, or after source zone treatment. Examples of aquifer restoration technologies include

Groundwater pump and treat systems

Natural attenuation

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
In situ air sparging

Enhanced bioremediation

Fig. 23.4 displays simple process schematics of the first three of these aquifer restoration technologies. Restoration of
groundwater quality and use of these technologies is only feasible if a complementary source zone treatment technology is
also utilized.

Figure 23.4 Example aquifer restoration process schematics: (a) groundwater pump and treat, (b) in
situ air sparging, (c) natural attenuation.

23.3.3. Contaminant Migration Prevention


Contaminant migration prevention technologies are designed to minimize future impacts of contaminants on groundwater.
These technologies are often employed at sites where: a) the source zone location is not known and/or b) there are no
practicable source zone and aquifer restoration options. Examples of these technologies, or strategies, include

Natural attenuation

Groundwater pump and treat systems

In situ reaction walls

In situ air sparging curtains

Infiltration barriers

In situ containment options (grout walls, sheet piling walls, and so on)

Figure 23.5 displays simple process schematics of common contaminant migration prevention technologies.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 23.5 Example contaminant migration technology processes schematics: reaction barrier.

23.4. GENERIC TECHNOLOGY SELECTION AND DESIGN


PROCESS
Fig. 23.6 presents a generalized sequence of actions to be followed when selecting and designing any in situ treatment
technology. Each key step is discussed briefly below in this section, and then discussions specific to common source zone
treatment, aquifer restoration, and contaminant migration prevention technologies follow in Secs. 23.5, 23.6, and 23.7.

Figure 23.6 Generalized technology selection and design flowchart.

23.4.1. Site Assessment and Conceptual Model Development


As stated above in Sec. 23.1.1, the design of in situ treatment systems for groundwater contamination is challenging and
unique because the subsurface is an environment that is difficult to describe precisely; it is naturally heterogeneous and, due to
practical constraints, characterization data is generally limited.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Thus, the first step of any design process involves characterizing the site and, from the limited data available, proposing a
conceptual model (or models) of how the subsurface might look and how contaminants might move. At a minimum, the
following data are essential to the design process:

A historic assessment

A geologic/hydrogeologic assessment

A contaminant distribution assessment

A receptor assessment

A political assessment

The historic assessment reviews the history of activities at the site, including chemical inventory records, plot plans (locations
of structures, pipelines, rail roads, storage areas, maintenance areas, and so on); records of known spills, leaks, and other
releases, and any existing site assessment data and reports for this site and any other nearby sites.

The geologic/hydrogeologic assessment often involves collecting soil cores, installing groundwater wells, performing aquifer
characterization tests, monitoring groundwater elevations, and using other geotechnical tools (e.g., cone penetrometer) to
characterize the subsurface.

The contaminant distribution assessment involves conducting chemical analyses of soil, water, groundwater, and soil gas
samples, and measuring free-product levels in monitoring wells.

The receptor assessment involves visiting the site and nearby area and reviewing water well inventory records for the immediate
vicinity. The goal is to identify persons or resources that have been, or could be impacted by the contamination.

Finally, the political assessment involves identifying the appropriate regulatory authorities, the potentially responsible parties,
and any other persons or groups that may be involved in the decision-making process for the site.

The collective goal of these various assessment activities is to identify the key components of the conceptual model(s):

Primary sources (storage tanks, pipelines, degreasing operations, etc.)

Time of release and amount of contaminant released

Depth to groundwater

Direction of groundwater flow and groundwater velocity

Subsurface geology (soil type, distinct layers)

Chemicals present

Contaminant distribution in the subsurface

Existence and location of any mobile free-product liquid

Potential receptors and any existing adverse impacts

Persons that will be involved in the decision making—process

Often the time of the release and the quantity released are unknown, unless a single catastrophic well-documented event took
place (many slow leaks from underground storage tanks go undetected for years). However, if they are known, then this
information can be used to better choose the number and locations of soil and groundwater samples. For example, for a spill
size of volume Vspill (m3 or ft3 ), the maximum depth dspill (m or ft) penetrated by the spill can be estimated by

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(23.1)

dspill = maximum depth penetrated by the spill (m or ft)

V spill = volume of liquid contaminant spilled (m3 or ft3 )

Aspill = cross-sectional area (plan view) impacted by spill (m2 or ft2 )

SR = residual saturation of contaminant liquid in soil (m3 fluid/m3 pores or ft3

fluid/ft 3 pores). stet

T = total soil porosity ( ≈ 0.40 m3 pores/m3 soil or ft3 pores/ft3 soil)

Where the residual saturation S R is specific to the contaminant fluid and porous medium. S R increases with increasing
surface/interfacial tension and decreasing pore size (finer soils). Table 23.2 [American Petroleum Institute (API), 1989]
provides reasonable values for S R T for petroleum fuels as a function of soil type. Also given are equivalent soil
concentrations Csoil (mgcontaminant/kg soil). While these values are specific to the petroleum mixtures and soils given in the table,
they also provide reasonable estimates for many other liquid contaminants of concern.

As an example, Fig. 23.7 presents depths of spill penetration as a function of spill size and infiltration area, for a range of soil
types.

In addition to soil concentration measurements, groundwater and soil gas samples can also be used to help define
contaminant source zone locations. A general rule of thumb is to look for areas where groundwater or soil gas concentrations
exceed 1 percent of the maximum chemical-specific concentration expected for the contaminants released.

Table 23.2 Ranges of Residual Liquid Hydrocabon Concentrations in the Unsaturated Zone.

Residual Liquid Hydrocarbon Concentrations in Unsaturated Zone Soils (= SR T)

[Gasolines [Gasolines [Gasolines [Gasolines [Middle [Middle [Middle [Fuel Oils] [Fuel Oils] [Fuel Oils]
] Soil ] (gal/ft3) ] (L/m3) ] (mg/kg) Distillates] Distillates] Distillates] (gal/ft3) (L/m3) (mg/kg)
(gal/ft3) (L/m3) (mg/kg)

Abreviations: (gal/ft3) = gallons of liquid per ft 3 of soil; L/m3] b = L of liquid/m3 of soil, (mg/kg) = mg of liquid/kg of soil; (mg/kg)
concentrations based on ρ soil = 1.85 g soil/cm 3 soil, ρliquid = 0.7, 0.8, and 0.9 g,-liquid/cm 3 liquid for gasoline, middle distillates, and
fuel oils, respectively.

Source: API, 1989.

Coarse 0.02 2.5 950 0.04 5 2200 0.07 10 4,800


gravel

Coarse 0.06 7.5 2800 0.1 15 6500 0.22 30 15,000


sand

Fine sand/ 0.15 20.0 7500 0.3 40 17000 0.6 80 39,000


silts

In the case of single-component spills, the maximum dissolved concentration CW max [mg/LH2O] is the pure component solubility
S (mg/L H2 O), while the maximum vapor concentration CVmax (mg/Lvapor) is derived from tabulated values of vapor pressure PV
(torr), the molecular weight MW (g/mol), and the Ideal Gas Law. In summary, for single-component spills:

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(23.2)

(23.3)

Figure 23.7 Depth of spill penetration vs. spill size and infiltration area.

Table 23.3 contains and for a range of chemicals of interest. For chemical-specific properties of other chemicals,
the reader is referred to the USEPA Superfund Chemical Data Matrix databases (USEPA, 1996b).

For the case of mixtures, the appropriate equations for estimating the maximum dissolved and vapor concentrations of each
compound are

(23.4)

(23.5)

Table 23.3 Chemical Properties of Selected Soil and Groundwater Contaminants of Interest (derived from USEPA SCDM
1996b)

Chemical Molecular Weight Pure Component Solubility Vapor Pressure Maximun Vapor Conc.
(g/mole) (mg/LH2O) (torr) (mg/Lvapor)

Source: Johnson, et al., (1990a, b).

Acetone 58 106 23 730

Aldrin 365 0.017 6 × 10-6 0.0001

Anthracene 178 0.043 3 × 10-6 0.00003

Atrazine 216 70 3 × 10-7 0.000004

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Chemical Molecular Weight Pure Component Solubility Vapor Pressure Maximun Vapor Conc.
(g/mole) (mg/L ) (torr) (mg/L )

Benz(a)anthracene 238 0.009 1 × 10-7 0.000001

Benzene 78 1800 9 405

Benzo(a)pyrene 252 0.002 6 × 10-9 0.00000008

Benzo(b)fluoranthene 252 0.002 5 × 10-7 0.000007

1,3-Butadiene 54 740 2100 6200

Butanol 74 106 74,000 300,000

Carbon tetrachloride 154 790 120 1010

Chlordane 410 0.056 5 × 10-5 0.001

Chlorobenzene 113 470 12 14

Chloroform 119 7900 200 1300

Chloromethane 51 5300 4300 12,000

Chrysene 228 0.006 9.5 × 10-5 0.001

m-Cresol 108 23,000 0.14 0.83

Cyclohexane 84 55 97 450

DDD 320 0.09 7 × 10-7 0.00001

DDE 318 0.12 6 × 10-6 0.0001

DDT 354 0.025 2 × 10-7 0.000004

1,3 Dichlorobenzene 147 130 2.2 18

1,2-Dichloroethane 99 8500 79 430

trans 1,2- 97 6300 330 1800


Dichloroethylene

1,2-Dichloropropane 113 2800 52 320

Dieldrin 381 0.2 5 × 10-6 0.0001

Endrin 381 0.2 3 × 10-6 0.00006

Ethylbenzene 106 170 9.6 0.056

Ethylene glycol 62 106 0.092 0.31

Formaldehyde 30 550,000 5200 8500

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Chemical Molecular Weight Pure Component Solubility Vapor Pressure Maximun Vapor Conc.
(g/mole) (mg/L ) (torr) (mg/L )

Heptachlor 373 0.18 0.0004 0.008

Hexane 86 12 152 720

Isobutanol 74 85,000 10 40

Lindane 291 7.3 0.0004 0.006

Methanol 32 106 130 230

Methylethylketone 72 220,000 95 370

Napthalene 128 31 0.085 0.60

PCBs 356 0.07 7.7 × 10-5 0.0015

Phenol 94 83,000 0.28 1.4

Styrene 104 310 6.1 35

Tetrachloroethylene 166 200 19 170

Toluene 92 530 28 141

Toxaphene 413 0.74 9.8 × 10-7 0.00002

Trichlorbenzene 181 49 0.43 4.3

1,1,2-Trichlorethane 133 4400 23 170

Trichloroethylene 131 1500 73 523

Vinyl chloride 63 8800 3000 10,000

m-xylene 106 160 8.5 49

Gasoline (fresh)* 95 1300

Gasoline (weathered)* 111 220

where Xi(moles–i/total moles in mixture) is the mole fraction of the chemical of interest in the mixture. In many cases, when the
complete composition of the mixture is unknown, the mole fraction can be approximated by the mass fraction Wi (mass–i/mass
of mixture):

(23.6)

For example, if a spill of benzene liquid occurred, we would expect to see dissolved and soil gas vapor benzene concentrations
approaching 1800 mg/LH2 0 and 400 mg/Lvapor, respectively, near the source zone. If, on the other hand, a gasoline spill
occurred, then we would expect to see dissolved and soil gas vapor benzene concentrations approaching 18 mg/LH2O and 4
mg/Lvapor, respectively, for a gasoline containing ≈ 1 percent by weight benzene.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
For the site assessment, how much data collection is enough? Clearly, it is desirable to be able to completely characterize a
site, but this option is cost-prohibitive. There are no universal guidelines that define minimum numbers of soil borings, samples,
and so on; however, some useful questions to consider are

Have you adequately defined primary sources, depth to groundwater, direction of groundwater flow, groundwater velocity,
subsurface geology (soil type, distinct layers), chemicals present, contaminant distribution in the subsurface, existence and
location of any mobile free-product liquid, potential receptors and any existing adverse impacts?

Is it likely that the additional data will change the proposed site conceptual model?

Are additional data critical to the prescreening of treatment technologies?

What are the risks of designing a system based on the current data?

One other important factor to consider is that some treatment technologies are more robust than others. Here “robustness” is a
measure of how easily an initial system design can be refined once operation begins. It is also a measure of how sensitive
treatment effectiveness is to small variations in subsurface conditions and contaminant distributions. In other words, a very
robust technology is one that can be designed with little site-specific information, and that can be made to perform as desired
through minor operating condition changes rather than major system equipment changes. Table 23.4 provides a qualitative
indication of the robustness of many treatment technologies in use today.

23.4.2. Select Target Treatment Levels


At this step, appropriate target treatment levels are identified for the chemicals that have been identified at the site. The
discussion in Sec. 23.2 outlines the range of target treatment goals that might be encountered.

Briefly, at this stage the user must

define the target treatment levels (both in soil and groundwater),

define the points of compliance where these goals are to be applied, and

define the time frame within which compliance must be achieved.

After this step is completed, site concentrations are compared with the target treatment values. Any exceedences will trigger
further assessment, and treatment may be required.

23.4.3. Identify Potential Technologies


If treatment is necessary, a gross pre-screening of technologies takes place. This prescreening considers the site conditions,
documented performance of the available technologies at similar sites, the state of understanding of the technology, economic
factors, and the treatment goals (target levels and target time frame). Table 23.4 contains information that can be used to
provide a rough screen of technologies.

At the end of this step the user is generally considering one to three treatment options.

23.4.4. Screening Level Calculations


At this stage a more refined screening of technologies occurs based on basic fundamental screening calculations. These
calculations are generally biased toward predicting optimal system performance. Typical calculations involve estimating
maximum treatment rates. Examples of these screening level calculations are presented in Secs. 23.5 and 23.7.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Table 23.4 Example Technology Prescreening Matrix.

Technology Application Relative Time Robustness Typical Settings


Frame

Excavation and above- Source treatment Short •••• Shallow cotamination (<20 ft BGS)
ground treatment

Groundwater pump and Containment Long •••• All, except very low permeability settings
treat

Bioventing Source treatment Medium ••• Petroleum release sites and permeable soils

Soil vapor extraction Source treatment Medium ••• Petroleum fuel and solvent release sites and
permeable soils

Physical barriers Containment Long ••• Mixed waste sites involving very hazardous
chemicals, low permeability settings

Natural attenuation Dissolved plume Long ••• Petroleum release sites


management

Free-product recovery Source zone Medium •• Sites with mobile and pumpable free-product liquid
systems treatment

In situ air sparging Source treatment Medium •• Petroleum fuels and solvents trapped below the water
table in permeable soils

Chemical migration Dissolved plume Long ••• Chlorinated solvent spill sites
/reaction barriers containment

23.4.5. Decision Point—Is the Technology Appropriate?


The user then compares the output of the screening level calculations with treatment expectations. If, even under optimal
conditions, the performance predicted by the screening-level calculations does not meet treatment expectations, then the user
has two options:

eliminate that technology as an option, or

revise treatment expectations (target levels or treatment time) and retain the technology.

23.4.6. Pilot Testing


In situ pilot tests are usually conducted for the most promising treatment technologies. Often these are designed to test critical
conceptual model assumptions, such as achievable flow rates, maximum effluent concentrations, and so on. Convention and
economic constraints dictate that most pilot tests will be less than 1 week in duration.

Key guidelines for planning and conducting pilot tests include

Design the pilot test so that the equipment used is easily integrated into any full-scale system that might be installed

As it is very unlikely that the short-term performance of a pilot test can be accurately extrapolated to long-term full-scale
system performance, pilot tests should be designed to look for infeasibility. In other words, the pilot test should be designed
to identify behavior or fatal flaws that suggest that the technology would not work at the site (e.g., maximum achievable
flow is too low, geology prohibits access to the contaminants, and so on).

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
23.4.7. Decision Point—Is the Technology Appropriate?
The user then compares the pilot test results with treatment expectations. If the performance does not meet treatment
expectations, then the user has two options:

eliminate that technology as an option or,

revise treatment expectations (target levels or treatment time) and retain the technology

23.4.8. Initial Design


Based on all available data, the site conceptual model is revised, and an initial system design is created and installed. The
design process often relies on a combination of screening level models, heuristics, and sometimes more sophisticated
computational design tools. It is important at this stage to

Design a robust system that can succeed under a range of actual site conditions

Design a system that can easily be modified and expanded

Design the system to be easily monitored

23.4.9. Operation and Monitoring


Once the system is installed and operated, adequate monitoring should be conducted to assess if the system is meeting the
treatment goals and regulatory objectives, and if there are modifications that can be made to improve system performance.

23.4.10. Design Refinement


As stated above, given the inherent large initial uncertainty, the design phase continues well past the construction, installation,
and operation of the initial system. It should be recognized that following sound design practices does not guarantee success;
it does, however, provide a higher probability of success. For this reason, it is likely that refinements will be made to the original
system design based on the performance monitoring data. These refinements may include changes in operating conditions
(e.g., flow rates) or the addition of new hardware (e.g., additional wells).

23.4.11. Decision Point–Have the Treatment Goals Been Met?


To determine when it is appropriate to terminate the treatment process, the following questions must be addressed:

Has the treatment system achieved the treatment goals?

If the system is turned off, will treatment goals continue to be met?

If the treatment goals have not yet been met, is there another technology that can more effectively be used to progress
toward the treatment goals?

If the answer to any of these questions is yes, then operation of the current treatment system should be terminated. If treatment
goals have yet to be met, it is time to transition to the next technology in the treatment train. For example, at petroleum fuel
release sites, a typical sequence would involve free-product liquid recovery → soil vapor extraction/bioventing → natural
attenuation.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
23.5. SOURCE ZONE TREATMENT
In this section, general design principles are presented for common source zone treatment technologies, including free-product
recovery, pump and treat, soil vapor extraction, bioventing and in situ air sparging.

23.5.1. Free-Product Recovery


Most regulatory programs require removal of any mobile and pumpable immiscible free-product liquid that can be collected in
monitoring wells. In the case of LNAPLs, free product is found floating on top of groundwater in a monitoring well; in the case
of DNAPLs, free=product liquid is found at the bottom of the well. This of course assumes that the well is screened across the
interval containing the free=product liquid. The following discussion is specific to LNAPL sites, as our understanding of these
sites greatly exceeds our knowledge of free-product recovery at DNAPL sites.

Fig. 23.8 presents a simplistic drawing of a general LNAPL site where free-product is observed in the wells. In this figure free-
product is present on top of the water table, and LNAPL liquid has been trapped below the water table by rising groundwater
levels.

23.5.1.1. Free-Product Liquid Monitoring.


First, it is important to note that free-product levels measured in groundwater monitoring and recovery wells Hwell may not
reflect actual product thicknesses in the formation Hformation . Assuming vertical equilibrium, Kemblowski and Chiang (1990)
show that the two may be related by

(23.7)

where hcapaw denotes the capillary rise height for water in the soil matrix.Table 23.5 (Bear ,1979) contains hcapaw values for a
range of soil types. As can be seen, hcapaw increases as the soil becomes finer-grained. Thus, the measured product thickness
is expected to be roughly equal to the actual formation thickness in coarse soils. In finer-grained soils, thicknesses measured in
monitoring wells can be very misleading, and several feet of liquid product may be detected in wells where there are only a few
inches of actual formation product thickness.

Figure 23.8 Equilibrium LNAPL fluid distribution prior to free-product recovery.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Table 23.5 Water Capillary Rise Heights in Soils

Soil Type Capillary Rise Height (cm) Capillary Rise Height (in)

Source: Bear (1979).

Coarse Sand 2–5 1–2

Sand 12–35 5–14

Fine Sand 35–70 14–28

Silt 70–150 28–60

Clay 200–400 80–160

In general, the use of product thickness measurements for estimating actual spill volumes and recovery performance is
discouraged. In addition to the reasons discussed above, product thickness measurements are also affected by water table
elevation changes. In general

measured and actual product thicknesses always decrease as the water table rises;

measured and actual product thicknesses decrease as the water table lowers, if there is no trapped residual below the
current water table elevation; and

measured and actual product thicknesses increase as the water table lowers, if there is trapped residual below the current
water table elevation.

The last two items reflect the fact that trapped liquid concentrations below the water table are always greater than drained
residual liquid concentrations above the water table.

23.5.1.2. Maximum Achievable Free-Product Liquid Recovery.


While the use of product thickness measurements for estimating spill volumes and recovery performance is discouraged, the
information from such calculations is often useful as a design consideration. The maximum recoverable volume of LNAPL
Vmaxproduct (m3 or ft3 ) can be estimated by

(23.8)

where

A spill = cross-sectional area (plan view) impacted by spill (m2 or ft2 )

Htrapped = thickness of trapped liquid below water table (m or ft)

Hformation = thickness of free-product liquid in formation (m or ft)

T = total soil porosity ( ≈ 0.40 m3 pores/m3 soil or ft3 pores/ft3 soil)

ST = trapped saturation of contaminant liquid in soil (m3 fluid/m3 pores or

ft3 fluid/ft3 pores)

SR = residual saturation of contaminant liquid in soil (m3 fluid/m3 pores or ft3

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
fluid/ft3 pores)

Smax = maximum saturation of contaminant liquid in free-product layer in the

formation (m3 -fluid/m 3 pores or ft3 fluid/ft3 pores)

Table 23.2 provides values for TS R as a function of soil type and petroleum product type. In the absence of site-specific
values, reasonable estimates for S T and S max are

(23.9)

(23.10)

Estimated residual soil contaminant concentrations that would remain after complete mobile liquid product recovery are given
in Table 23.2. On a mass of contaminant per mass of soil basis, the total soil concentrationCT is given by

(23.11)

where ρ and ρb are bulk density of the contaminant liquid and the soil, respectively. As can be seen fromTable 23.2, free-
product recovery is not likely to reduce soil concentrations to values less than 1000 mg total contaminant/kg soil. For this
reason, it is often the first treatment technology applied at a site, but is rarely the final treatment technology applied to a site.

23.5.1.3. Free-Product Liquid Recovery System Designs.


Liquid contaminant recovery systems generally incorporate

interceptor trenches or drains, and/or

vertical recovery pump wells.

These components are depicted schematically in Fig. 23.9. Trench systems also rely on pumping wells to collect the free-
product that flows to the interception trench.

23.5.1.4. Trench Systems.


Trench systems are preferred for settings where

the water table is shallow, or some shallow confining layer limits the maximum depth of liquid penetration to about 20 ft (6
m) below ground surface (BGS), and

soils are heterogeneous with thin layers of highly permeable strata.

Trench systems are constructed as shown in Fig. 23.9. Soils are excavated to a depth below where liquid product is expected,
and then the very permeable gravel is back-filled into the trench. One or more vertical wells for free-product and (possibly)
groundwater collection are placed in the trench prior to backfilling. Often a surface seal is placed on top to minimize vapor
losses. The dimensions of the trench are usually dictated by soil stability, but in general, smaller trenches are preferred (1–3 m
[2–6 ft] wide). Modern trenching equipment can install drainage pipe and gravel in a continuous trenching operation. Liquid
product recovery rates rarely exceed 5 gal/min (20 L/min). The selection of vertical recovery well pumping schemes is
discussed below in Sec. 23.5.1.5.

23.5.1.5. Vertical Recovery Well Schemes.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
In situations where interceptor trenches are not feasible, product recovery systems use vertical recovery wells. As stated above,
interceptor trench designs also incorporate vertical recovery wells in their design. Thus, it is important to discuss the different
vertical pumping well recovery schemes used in practice. These include

Skimming systems for LNAPLs only

Single, total fluids pumps that pump water and free-product mixtures

Dual recovery pump systems, where a groundwater pump is used to depress the water table, while the second pump skims
the LNAPL free-product that flows into the well

Figures. 23.10, 23.11, and 23.12 (API, 1989) present design schematics of these recovery configurations.

Skimming systems are designed to remove free-product liquid with little, or no groundwater production. They are normally used
when

A significant quantity of liquid is not trapped below the water table

Water table depression is not needed to accelerate flow to the well

The formation is composed of very low permeability materials (clays, silty clays)

It is cost-prohibitive, or not possible to treat produced groundwater above ground

Figure 23.9 Free-product recovery systems: (a) interceptor trench, (b) vertical recovery well

Figure 23.10 Free-product skimming system (after API 1989).

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 23.11 Free-product total fluids recovery system. ( From API 1989).

Figure 23.12 Free-product dual pump recovery system. (After API 1989).

Single- and dual-pump recovery systems are used when

A significant quantity of liquid is trapped below the water table

Water table depression is needed to accelerate flow to the well

The formation is composed of relatively permeable materials (sands, silty sands, and so)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
The dual=pump system is more complex than the single pump system, but is attractive because it eliminates the need for an
above-ground oil-water separator, and minimizes the load to the above-ground water treatment system.

When selecting the operating conditions for a single-, or dual-pump recovery system, the practitioner must balance two factors:

In general, increasing the drawdown in the well will increase the free-product flow rate to the well

Increased drawdown below the lowest elevation of trapped free-product liquid will decrease the cumulative liquid recovery.

Given these factors, it can be expected that there will be an optimum drawdown that maximizes the volume recovered for a
given setting. It can also be expected that this drawdown might cause the contamination of soils not previously impacted. This
behavior is illustrated schematically in the sequence of drawings shown in Fig. 23.13.

Figure 23.13 Schematic progression of free-product recovery operation with groundwater


drawdown.

In general, for the case of groundwater table depression, the maximum cumulative free-product recovery as a function of the
groundwater table depression is given by

(23.12)

where ΔVGW is the volume of new soil below the lower elevation of trapped residual LNAPL that becomes exposed by
groundwater drawdown (m or ft), and all other quantities are as defined for Eq. (23.8).

For example, Fig. 23.14 presents results for the case of a uniform groundwater table depression, where

Htrapped =1m

Hformation = 0.5 m

fT = 0.40 m3 pores/m3 soil

ST = 0.20 m3 fluid/m3 pores

SR = 0.10 m3 fluid/m3 pores

Smax = 0.95 m3 fluid/m3 pores

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
For other geometries and nonuniform groundwater table depression, designers should consider using more sophisticated
numerical tools, such as the USEPA Hydrocarbon Spill Screening Model (Weaver et al., 1994). In addition to estimating the
cumulative recovery of free-product, these tools can also be used to estimate the liquid recovery rates as a function of time.

When choosing a skimmer or fluids pump for a site, it is best to consult with the manufacturers. The information that is needed
to choose an appropriate pump includes:

Target pumping rate

Depth to groundwater

Figure 23.14 Effect of groundwater drawdown on maximum cumulative free-product recovery for
example problem.

Depth below ground surface (height fluid must be pumped)

Type of fluid to be pumped and its properties (density, reactivity, and so on)

Availability of electricity and compressed air supply

23.5.2. Soil Vapor Extraction


23.5.2.1. Soil Vapor Extraction Overview.
Soil vapor extraction systems are designed to maximize contaminant removal by volatilization. A basic in situ soil vapor
extraction system, such as the one depicted in Fig. 23.15, couples vapor extraction wells with blowers or vacuum pumps to
remove contaminant vapors from the subsurface. The vacuum applied to the vapor extraction well induces air flow from the
atmosphere, through the soil, and to the well. Contaminants volatilize from the soil to replace those swept away by the induced
air flow. Above-ground equipment often consists of control valves to adjust air flow to individual wells, pressure gauges, and
flowmeters at each wellhead, an air-liquid separator (for moisture removal from the extracted gases), a pressure gauge and
flowmeter at the vacuum pump/blower, and a vapor treatment unit. Vapor treatment systems are included if air treatment is
required by local regulations. Also shown in Fig. 23.15 is a groundwater depression system, which may be a necessary addition
to expose trapped contaminants below the water table. More complex soil vapor extraction systems incorporate trenches,
horizontal wells, forced-air injection wells, passive air inlet wells, impermeable surface seals, multiple vapor extraction wells in
single boreholes, and sophisticated unsaturated zone moisture, contaminant, and temperature monitoring systems.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 23.15 Soil vapor extraction treatment system schematic.

23.5.2.2. Feasibility Assessment.


Soil vapor extraction systems are generally applicable to permeable soils and volatile contaminants. Screening level
calculations can be performed to decide if the technology may be feasible at a specific site. Below, we follow the approach
suggested by Johnson et al., (1990) to assess feasibility. As will be seen below, the similar calculations are performed again
when planning the initial design. Users may find the USEPA HyperVentilate (Johnson and Stabeneau, 1991) software useful in
performing these calculations.

The three main factors governing the behavior of any in situ soil venting operation are

Achievable vapor flow rates

Maximum contaminant vapor concentrations

The location of the vapor flowpath relative to the contaminant location

By multiplying estimates for the first two, one can estimate the maximum removal rate by volatilization. The third factor affects
the efficiency of the system and defines how significant mass transfer resistances will be. In general, it is these mass transfer
limitations that control the performance of most in situ remediation systems. Most often they result from diffusion-limited
chemical transport.

The following questions will be answered in making the initial feasibility assessment. These are questions that should be
addressed prior to performing any field-scale pilot tests.

What vapor flowrates are achievable?

What maximum vapor concentrations are achievable?

What maximum mass removal rates can be achieved?

Are there likely to be significant mass transfer resistances?

What changes in performance are to be expected with time?

Is this behavior/expected performance acceptable?

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Vapor extraction flow rates can be estimated with the following equation (Johnson et al., 1990a):

(23.13)

where Q = vapor flow rate (cm3 /s), H = thickness of permeable zone, or well screen thickness, whichever is less (cm),k = soil
permeability to air flow (cm2 or darcy), μ = viscosity of air = 1.8 × 10-4 (g · cm)/s or 0.018 cp., Pw = gauge pressure at extraction
well, with vacuums taken to be positive=quantities [(g · cm)/s2 or atm], Patm = absolute ambient pressure ≈ 1.01 × 106 (g ·
cm)/s2 or 1 atm Rw = radius of vapor extraction well (cm) and RI = radius of influence of vapor extraction well (cm)

This equation assumes simple steady-state, radial, compressible flow. While actual flow fields are not one-dimensional, Eq.
(23.13) provides reasonable flow rate estimates. As k is dictated by soil structure, and Pw and Rw are design parameters, then
the only unknown parameter is the empirical “radius of influence” RI. Values ranging from 9 m (30 ft) to 30 m (100 ft) are
reported in the literature for a variety of soil conditions. However, Eq. (23.13) is fairly insensitive to large changes in RI, and
therefore, for estimation purposes a value of RI = 15 m (≈50 ft) can be used without a significant loss of accuracy.Figure 23.16
presents single-well flow rate estimates per unit thickness for a range of soil types and applied wellhead vacuums. Most vapor
extraction well radii are either 1.3 cm (1 in) diameter, 2.54 cm (2 in) diameter, or 5 cm (4 in) diameter. The results shown in Fig.
23.16 are based on a well radius of 5 cm (2 in). It is advised that an order of magnitude range in soil permeability values be
used when making these flow rate estimates as this is likely the parameter with the largest uncertainty.

Maximum soil gas contaminant vapor concentration is easily calculated from knowledge of the compound’s (mixture's)
molecular weight, vapor pressure at the soil temperature, residual soil contaminant composition, and the Ideal Gas Law using
Eq. (23.5) (Sec. 23.4.1), with the total vapor concentration being the sum of individual component concentrations. If possible,
the maximum vapor concentration can also be measured by analyzing soil gas samples during the site assessment phase.

Table 23.3 presents Cvmax values for some chemicals and mixtures often accidently released to the environment. There are
more sophisticated equations for predicting vapor concentrations in soil systems based on equilibrium partitioning arguments
(Baehr and Corapcioglu, 1987; Corapcioglu and Baehr, 1987; Johnson et al., 1990a), but these require more detailed information
(organic carbon content, soil moisture) that is normally available. If a site is chosen for remediation, the residual total
hydrocarbons in the soil typically exceed 500 mg/kg. In this residual concentration range the majority of hydrocarbons will be
present as a separate or “free” phase, the contaminant vapor concentrations become independent of residual concentration
(but still depend on composition), and Eq. (23.5) is applicable (Johnson et al., 1990a). In any case, it should be noted that these
are estimates only for vapor concentrations at the start of venting, which is when the removal rates are generally greatest.
Contaminant concentrations in the extracted vapors will decline with time due to changes in composition, residual levels, or
increased diffusional resistances.

Maximum achievable mass removal rates Mmax (kg/day or lb/day) are obtained by multiplying the estimated flow rates, vapor
concentrations, and the number of vapor extraction wells Nwells:

(23.14)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 23.16 Soil vapor extraction single-well flow rate estimates for Rw = 5 cm (2 in).

Figure. 23.17 presents estimated maximum removal rates for a range of soil vapor concentrations and flow rates.

Once this estimate has been made, it should be compared with treatment expectations. Actual system performance will be less
than this prediction, so if the performance predicted by the screening-level calculation does not meet treatment expectations,
then the user has two options:

eliminate that technology as an option, or

revise treatment expectations (target levels or treatment time) and retain the technology.

As mentioned above, other factors ensure that actual performance is less than that predicted above inEq. (23.14). These
factors are

Composition changes with time, and

Mass transfer resistances

Even under optimum conditions (maximum vapor-contaminant contact and local equilibrium partitioning), removal rates will
decline over time as compositions and soil concentrations decrease. More volatile compounds are removed preferentially and
with time the remaining mixture is enriched in the less volatile compounds. This is reflected in declining removal rates with
time. A more detailed discussion of this topic can be found in Baehr and Corapcioglu 1987 (1989), Johnson et al. (1990a), and
Benson et al. (1993). For the purposes of this discussion, it suffices to say that there are computational tools available to
examine the impact that this has on removal (see USEPA, 1995 for a review of available models).

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 23.17 Estimated maximum mass removal rates Mmax as a function of maximum vapor
concentration and total vapor extraction flow rate Qtot = Nwells Q.

As a screening calculation, Johnson, et al. (1990b) suggest performing simulations involving a well-mixed system. Sample
results for a weathered gasoline are presented in Fig. 23.18. Rather than presenting mass removal rates or concentrations as
functions of time, the results have been normalized so that the fractional removal rate relative to the initial maximum removal
rate [(M(t)/M(t = 0))] is plotted as a function of the cumulative vapor volume removed per unit initial residual contaminant mass
Qt/Mspill.

To use this information to assess feasibility, one estimates the spill mass using site characterization data, or other information,
and selects a target treatment time ttarget and treatment level. Then, from a graph such as Fig. 23.18, the value of (Qt/Mspill)
obtained. For example, in this case, if a 90% removal efficiency were desired, then (Qt/Mspill) pred ≈ 100 L vapor/g initial residual.
This information is then used in the equation:

(23.15)

where Qmin is the minimum total vapor flow rate that would be required to meet the target treatment objectives. The user can
then assess whether that flow rate is achievable; if not, then the user can eliminate that technology as an option, or revise
treatment expectations (target levels or treatment time) and retain the technology.

Mass transfer limitations will also decrease system performance relative to the ideal case. Fig. 23.19 presents two common
macro-scale mass transfer limited situations. In the first, vapor flows parallel to, but not through, the zone of contamination.
This resembles the situation where a layer of free-product rests on top of an impermeable strata or the water table. For the case
of a single component, Johnson et al. (1990) present the solution:

(23.16)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 23.18 Effect of composition change with time on treatment performance as a function of the
normalized volume of vapor (Q x t) per initial mass of residual contamination M spill (Johnson, et al.
1990).

Figure 23.19 Macroscopic mass transfer limiting scenarios. (From Johnson, et al 1990a).

where η= efficiency relative to maximum removal rate, D = effective vapor-phase diffusion coefficient (cm2 /s), μ = viscosity of
air = 1.8 × 10-4 (cm · s), k = soil permeability to vapor flow (cm2 ), H = thickness of screened interval (cm), RI = radius of influence
of venting well (cm), Rw = venting well radius (cm), PAtm = absolute ambient pressure = 1.016 × 106 g/(cm · s2 ), Pw = absolute
pressure at the venting well g/(cm · s2 ) and R1 < r < R2 = defines region in which contamination is present

The efficiency η is inversely proportional to the screened interval thickness H because a larger interval will, in this geometry, pull
in unsaturated air that has passed above the liquid-phase contamination. D is calculated by the Millington-Quirk expression,
which utilizes the molecular diffusion coefficient in air Do , the vapor-filled soil porosity A, and the total soil porosity T:

(23.17)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
As an example, consider removing a layer of contamination bounded by sandy soil (k = 1 darcy). A 5.1 cm (4 = in) radius
extraction well is being operated at Pw = 0.90 atm [0.91 × 106 g(cm · s2 )], and the contamination extends from the region R1 =
Rw = 5.1 cm to R2 = 9 m (30 ft). The well is screened over a 3 m (10 ft) interval. Assuming that

ρb = 1.6 g/cm3

A = 0.23

Do = 0.087 cm2 /s

T = 0.30

RI = 12 m

then the venting efficiency relative to the maximum removal rate is η = 0.09 = 9 percent. In Fig. 23.19b, vapor again flows past,
rather than through the contaminated soil zone. In this case, however, the contaminant is incorporated into the soil matrix. This
would be the case for a contaminated silt lens surrounded by sandy soils. Again, diffusion limits the removal rate, but in this
case, the diffusion path length increases with time as contaminant is depleted from the lower permeability layer. Estimates of
the removal rate Rest and clean zone thickness d(t) from a contaminated zone extending from R1 to R2 is:

(23.18)

(23.19)

where D (cm2 /s) is the effective vapor diffusion coefficient, CVmax is the estimated maximum vapor concentration (see Table
23.3), t(s) is time, Csoil (mg/kg) is the initial residual contaminant concentration in soil, and ρb (kg/L) is the bulk soil density.

Figure. 23.20 presents estimated clean zone thicknesses δ (cm) as a function of time for the ratio (CVmax/ρbCsoil) for a value of
D = 0.01 cm2 /s.

Equations (23.16) through (23.19) are specific to single component systems. Presently there are no simple analytical solutions
for mixtures, but it is likely that the removal rates will be less than those predicted for the single component systems. More
sophisticated numerical simulations can be performed to study these effects (Baehr and Corapcioglu 1987 Benson, et al.,
1993).

23.5.2.3. Soil Vapor Extraction Pilot Tests.


Figure 23.21 displays the setup of a typical soil vapor extraction pilot test. A basic system consists of

Test vapor extraction well—constructed as one would expect to construct a vapor extraction well for the proposed full-scale
system

Soil gas monitoring probes

Vacuum pump and/or blower to induce air flow

Vapor treatment system (if required)

Groundwater depression system (if required)

Calibrated flowmeter(s)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Calibrated pressure/vacuum gauge(s)

Figure 23.20 Remedial progress as measured by the growth of the clean zone thickness with time.

Figure 23.21 Soil vapor extraction pilot test schematic.

Soil gas monitoring points are used to collect soil gas samples and to measure the subsurface vaporphase pressure
distribution. Those used primarily for soil gas sampling should be placed within the target treatment zone. Others used
primarily for pressure distribution measurements should be spaced distances of roughly H/2, H,3H, and 5H away from the
extraction well, where H is the depth below ground surface to the top of the extraction well screen. Probes should be placed
within each distinct soil strata, especially those that are expected to have significant impact on the vapor flow field. At a
minimum, pilot tests should be designed to measure

Extraction flowrate versus applied vacuum

Extracted vapor concentrations

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Soil gas concentrations

When multiplied together, the first two provide a measure of the mass removal rate to be expected at startup, and should be
used to confirm and refine key assumptions made when assessing feasibility with screening level Eq. (23.14) through (23.19).
The second and third items provide a measure of the efficiency of vapor removal.

In addition, pilot tests can be designed to measure.

soil gas pressure distributions with time and space

Other possible significant effects of applied vacuum (e.g., water table upwelling)

Soil gas pressure distributions can be used to assess the vapor flowfield and soil permeability to air flow as described in and
Baehr and Hult (1991), Beckett and Huntley (1994), and Johnson et al., (1990a).

Tests designed to measure the quantities listed above can generally be conducted within 1 day as vapor flow rates and soil gas
pressures reach near steady-values within minutes to hours at most sites.

Care should be taken in the manifolding of test wells, flowmeters, pressure gauges, and blowers or vacuum pumps. As most
blowers/vacuum pumps are driven by fixed-speed motors, extraction/injection flow rates are often controlled by including
gate/block/globe valves and an air inlet/outlet pipe on the manifold as shown in Fig. 23.21. While a single in-line valve is
sufficient to control the injection/extraction flow rates, the air inlet/outlet pipe is typically included to allow the same level of
control, while also preventing blower/pump overheating. It is very important to ensure that flowmeters and pressure/vacuum
gauges are placed between the wellhead and the first encountered valve or piping junction, otherwise the flow rate and applied
pressure/vacuum at the wellhead are not being measured. Unfortunately this is not always the case in practice, and these
measuring devices are often found incorrectly located between the blower/vacuum pump and an air inlet/outlet valve.

23.5.2.4. Soil Vapor Extraction System Design.


Our intent here is not to prescribe a standard system design process that must be followed; indeed, there are many approaches
applied today in practice. Instead, a range of possible design approaches are presented, and these encompass most current
practices. For each approach presented, data requirements are specified and the relative advantages and disadvantages of that
approach are discussed. The approaches discussed below are presented in the order of relative complexity and expertise
required.

At this point it is worth pointing out to the reader that, in principle, the system design process should not be initiated until one
has made a feasibility evaluation. Aside from very general tables, graphs, or guidelines, however, one of the best methods for
evaluating feasibility involves first designing a system and then assessing its implementability. This is the approach that was
taken above in discussing the use of Eq. (23.15).

Components that are typically specified in a soil vapor extraction-based system design include:

Number of vapor extraction wells

Number of air injection wells

Well location(s)

Well construction(s) (depth, screened interval, materials, and so forth)

Extraction blower(s) or vacuum pump(s)

Injection blower(s)

Vapor treatment unit(s)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Equipment manifolding and piping

Instrumentation (flowmeters, sampling ports, vapor concentration monitoring, and the like)

Operating conditions

Below we discuss how each is prescribed in the various design approaches. With the exception of the number of wells selected,
well locations, and the operating conditions, strategies for dictating the remaining design parameters are relatively similar for
all approaches. To avoid repetitious discussions, therefore, these common features are first discussed prior to specific
discussions of each approach.

Extraction/injection well construction. For the most part, vapor extraction-based processes have predominantly employed
vertical wells. However, given recent advances in horizontal drilling techniques, the use of horizontal wells is being explored, as
these offer some unique advantages (vapor extraction beneath buildings, greater effective air flow per well in a narrow depth
interval). In addition, vapor extraction trenches have been utilized for shallow groundwater sites. Despite the obvious physical
differences in each of these scenarios, the same questions must be addressed when specifying their construction. These are:

What size conduit (pipe) should be used?

What conduit material should be used?

What interval should be perforate or “screened” to vapor flow?

What are the packing and screened interval specifics?

How should the well be installed in the subsurface?

What will be done to minimize vapor flow “short circuiting”?

Well diameters ranging from 1 to 6 in are commonly employed in extraction/injection well construction. Smaller diameter wells
have the advantage that they may be pushed (rather than drilled) to depths of about 30 ft if the site geology is amenable;
however, the drawback is that there may be significant line pressure drops when high vapor flow rates are required. Larger
diameter wells are almost always installed by first drilling a borehole, and then placing the well casing within this borehole. A
permeable (less resistant to vapor flow than the surrounding formation) packing material is placed in the annulus around the
screened interval, and the remaining annular region is then grouted (using cement grout or bentonite pellets) to prohibit short-
circuiting of the air flow. An example of this type of construction is shown in Fig. 23.22. For most applications polyvinyl chloride
(PVC) piping is adequate; however, other materials should be used in high pressure, high temperature scenarios, or extreme
environmental conditions constructed by human beings. In the case of systems using air injection wells, it is important to
recognize that air passing through an injection blower may undergo a significant temperature rise (especially for high injection
pressures).

Screened interval locations should be chosen to maximize air flow through the desired zone. In the absence of accessible
predictive tools, the practitioner must develop a good intuitive feel for subsurface vapor flow paths and the influence of
geologic conditions and manmade obstructions and conduits.constructed by numan beingso In the case of vapor extraction
systems the screened interval is usually set across the hydrocarbon impacted zone, unless it extends to ground surface. There
is rarely any advantage to extending the well screen above the zone of contamination.

Screen size should be selected to maximize the area open for vapor flow, while maintaining an open conduit. Packing materials
should be at least as permeable to vapor flow as the formation itself.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 23.22 Soil vapor extraction well schematic.

Vapor treatment. Vapor treatment units are selected to allow the operator to meet regulatory requirements, which often take
the form of a prescribed percent reduction in hydrocarbon concentrations across the treatment unit (e.g., 95 percent reduction
in total hydrocarbon concentration), a required limit on emission rate of specific or total hydrocarbons (e.g., 1 lb/day total
hydrocarbon emission restriction), or some combination of the two. In some areas it is permissible to discharge untreated
vapors, if dispersion modeling results indicate adequate concentration reduction by within a given distance of the source stack.

A variety of units to meet most needs are currently offered by vendors for lease or purchase. Most fall under one of the
following categories:

Thermal oxidation units. Vapors are destroyed by combusting (burning) them in a chamber that is designed to maintain a
temperature of >1700° F and a gas residence time of about 1.

Internal combustion engines. Use the intake manifold of an internal combustion engine, which then serves as the extraction
pump as well as vapor treatment device.

Catalytic oxidation units, Vapors are passed through a catalyst bed maintained at an elevated temperature (often 500–700°
F), where they are combusted.

Absorption units. Hydrocarbon vapors are absorbed out of the flowing gas stream onto a support material, such as vapor-
phase activated carbon.

Biofilters. Biologically based units are in the development and evaluation phase. Here vapors are transferred to a bacteria-
containing media (e.g., liquid in a fluidized biological reactor, or soil/peat in a “biofilter”) where bacteria convert the
contaminant vapors to carbon dioxide and water.

Selecting a vapor treatment unit is relatively simple; one needs to provide a vendor with the range of influent vapor
concentrations, range of total vapor flowrates to be processed, and treatment requirements. Qualified vendors can then specify
which of their models are appropriate, and provide you with estimates of operating costs (supplemental fuel needs) and

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
maintenance requirements. Some care should be taken in selecting a vapor treatment strategy as this often limits the range of
potential operating conditions, and it is often the largest capital and operating expense. In many cases, the incremental cost for
larger–capacity units (e.g., cost differential between a 100=and 200 FSCFM unit) is not significant, and it is worthwhile
investing the money to allow greater flexibility.

Extraction pump/blower and injection blower. To select extraction blowers or vacuum pumps, one simply specifies the flow rate
necessary at a given design vacuum or pressure. This information is collected during the pilot test as discussed above. Based
on these data, pump/blower manufacturers can specify which of these units will meet your needs based on measured
performance, or “pump curves,” for their equipment. Given a range of manufacturers and blower/pump types, one generally
chooses a unit based on cost, maintenance requirements, and experience with similar units. Given that extraction
blowers/pumps are the “heart” of most vapor extraction processes, it is recommended that most systems be designed with
multiple, or redundant blowers/pumps to allow for periodic failure and maintenance.

While the lower cost blowers have historically been the predominant choice for vapor extraction, there is now a trend toward
installing more versatile vacuum pumps as the incremental cost for upgrading is usually not significant in comparison to total
remediation costs.

Instrumentation. At a minimum, flow rates, applied vacuums or pressures, and extracted hydrocarbon vapor concentrations for
each well and for the total system effluent should be measured. With the exception of the total system flow and concentration
measurements, sampling ports and measuring devices must be placed between extraction wellheads and the first downstream
flow restriction or manifold junction.

Total hydrocarbon measuring devices (e.g., continuous flame ionization detector or explosimeter) are often placed in-line to
assess changes in the total concentrations of hydrocarbons in the extracted vapors. In some cases these are integral
components of the vapor treatment unit.

In addition, there may be instrumentation associated with vapor treatment units, such as thermocouples to measure catalyst
bed temperatures. Many newer units are outfitted with a dial-in modem that alerts you whenever the system shuts itself down.

Manifolding. The manifolding of system components plays an important role in the monitoring and operation of vapor
extraction-based processes; yet traditionally little attention has been paid to this aspect of system design. As a first rule, the
designer must allow for flexibility and future expansion. This means that the system design must easily accommodate the
installation of additional wells, blowers, or vapor treatment units with minimal cost and disruption to an operating system.
Valves must be installed to allow independent control of flow to/from each well, blower, and vapor treatment unit in the
system. Furthermore, the piping should be done in a manner to facilitate replacement of any equipment, especially
blowers/pumps and instrumentation.

Care should be taken to locate instrumentation where it can be easily monitored, or replaced, and guidelines given above for
flowmeter, pressure gauge, and sampling port placement should be followed.

Number of wells, well locations, and operating conditions. Historically, the radius of influence based approach has been the
most widely practiced. The number of wells and their spacing is based solely on the subsurface pressure distribution measured
during a pilot-scale test. Data required for this approach include:

A summary of the geologic/hydrogeologic assessment

Steady-state soil gas pressure distributions measured during a pilot-scale field test

Measured flow rate as a function of applied pressure/vacuum

Extracted vapor hydrocarbon concentrations from the pilot test (to better understand vapor treatment needs)

Fig. 23.23 outlines the radius of influence design approach. First, steady-state subsurface pressure distribution data are plotted
as a function of distance from the pilot extraction well and then a best-fit line is drawn through the data. While there is no

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
established convention for plotting the data, it is most often done on log-log or semi log coordinates. The “radius of influence”
of the well is then graphically determined to be that distance away from the well where the pressure distribution has been
extrapolated to reach some specified value. Numerous criteria exist in practice, including 0.1 in H2 O, 1.0 in H2 O, and 10 percent
of the applied vacuum at the extraction wellhead.

Figure 23.23 Graphical illustration of the radius of influence–based design approach.

Once the radius of influence RI has been established, circles of this radius are drawn on a site map so that the number of wells
and their locations are assigned to ensure that all the radii encompass the hydrocarbon-impacted zone.

After well locations have been specified the designer selects extraction/injection pumps/blowers and a vapor treatment unit.

This approach is widely practiced because it is simplistic, graphical in nature, and mimics common groundwater recovery
design practices. Unfortunately, this approach at best ensures containment of hydrocarbon vapors (as is the intent of
groundwater recovery system design), and systems are installed without any estimate of their expected performance or long-
term costs (Johnson and Ettinger, 1994).

The major limitation of the radius-of-influence approach discussed above is that it does not consider expected system
performance (cleanup levels, duration of remediation, and so on) or constraints (costs, and so forth). To begin to incorporate
these considerations, the practitioner must have predictive models that facilitate estimation of system performance as a
function of a wide range of parameters, such as geologic conditions, hydrocarbon type, number of wells, and so on. In this
approach the screening-level models used above to assess feasibility are again used in the system design process. As an
example we adopt the approach presented by Johnson, et al 1990a, which is captured in the HyperVentilate (Johnson and
Stabenau, 1991) software guidance system distributed by USEPA. The data requirements include:

Extraction flow rate as a function of applied vacuum for conceptual well design

Estimate of average residual soil concentration and volume of impacted soil

Remedial goals—cleanup levels and remediation time

Minimum volume of vapor to achieve required cleanup (calculated)

Steady-state pressure distribution from well

Effluent vapor concentration data from pilot test

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Geological cross-sectional map with hydrocarbon distribution

Constraints—costs, regulatory requirements, and so forth

Utilizing Eq. (23.15), the cumulative minimum system flow rate Qmin and the number of wells Nwells required to achieve our
goals as a function of the single-well flowrate Q, target remediation time ttarget, the estimated mass of contamination in the
subsurface Mspill, and the minimum volume of vapor required to achieve cleanup per unit mass of hydrocarbon (Qt/Mspill) pred (L
vapor/g initial hydrocarbon)=

(23.20)

The value for (Qt/Mspill) pred can be determined by consideration of equilibrium modeling results and possible mass transfer
resistances.

The number of wells Nwells predicted by Eq. (23.20) should be regarded as a minimum estimate of wells actually required, as
any real system performs less efficiently than an ideal system. Once the number of wells and applied pressure are specified, we
utilize steady-state pressure distribution data and the radius of influence based approach described above to ensure that our
system achieves vapor containment. If not, additional wells are needed. Table 23.6 (Johnson et al., 1994) illustrates a sample
worksheet for this design approach.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Table 23.6 Example Design Summary Worksheet for a Soil Vapor Extraction System.

Description of Data or Calculation Value/Result Units

Site data

Attach geological cross section showing hydrocarbon sampling results, subsurface conduits, and groundwater
elevation

Attach hydrocarbon analyses results

Average hydrocarbon concentration in treatment zone < Cs> g/g soil

Estimated volume of soil in treatment zone Vsoil cm3

Approximate soil density ρ soil 1.7 g soil/cm3


soil

Design flowrate per well Q@ΔP (attach pilot test data) L/min @ in
H2O

Radius of influence (attach steady-st pressure dist. figure) m

Remedial goals

Cleanup target g/g soil

Desired duration of remediatio t target days

Cost goal $

Calculations

Minimum volume of vapor required (ideal case) ( Qt/Mspill)pred L/g


hydrocarbon

Efficiency factor (describe basis for choice) 0<e<1

Nwells = (Mspill (Qt/Mspill)pred )/(Q t target *1440 min/day) minimum # of


=

Total flow rate estimate = Q Nwells L/min

Additional wells required to achieve vapor containment

Attach system layout schematic

We can extend this approach with the use of more complex mathematical models. In most cases these models require more
input data than is likely to be available; however, there will be sites where the projected cost of remediation warrants a greater
investment in the design phase of the project. To be used most effectively, complex modeling must be performed in an iterative
mode such that the predictive model is continuously refined, calibrated, and updated based on performance data collected from
system operation.

For the purpose of illustration it is worthwhile to compare the radius-of-influence and screening-level modeling design
approaches for the site data presented in Figs. 23.24 and 23.25 (Johnson et al., 1991). While this site has a number of distinct

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
strata, for the sake of simplicity we will design vapor extraction systems for the fine sandy layer located between the clay layer
and water table at a depth interval of about 43–49 ft BGS.

Based on the steady-state pressure data given in Fig. 23.24, a radius of influence of approximately 60 ft would be selected. In
comparison, Fig. 23.25 indicates that most hydrocarbons are distributed within a radial distance of 30 ft, it would therefore be
concluded that a single well should be sufficient for remediation.

To use the screening-level model approach, the following parameter values are selected from the data:

Vsoil = π (30 ft)2 × 6 ft = 17,000 ft3 = 4.8 × 108 cm3

<Csoil> = 20,000 mg/kg soil = 0.020 g/g soil × 1.7 g/cm3 soil = 0.034

g/cm3 soil

(Qt/Mspill) pred = [100 L/g hydrocarbon (ideal case)]/[0.50 (efficiency)] = 200

L/g hydrocarbon

Q = 10 SCFM @ 100 in H2 O = 4.7 L/s

ttarget = 360 days = 3.1 × 107 s

The volume of soil Vsoil is obtained by approximating the hydrocarbon-containing zone as a cylinder of radius 30 ft and depth 6
ft, as estimated from Fig. 23.25. The average soil concentration <Csoil> in this zone is assumed to be 20,000
mg/hydrocarbon/kg soil, which is consistent with data given in Fig. 23.25. Flow rate data combined with our knowledge of
groundwater drawdown system limitations indicates that the maximum extraction flowrate is likely to be about 10 SCFM (4.7
L/s) per well in this situation, where air flow is confined between the clay layer and groundwater. In this example we assume
that our goal is a remediation duration of 1 year (360 days = 3.1 × 107 s). Equation (23.20) predicts that a minimum of 23 wells
is required to achieve remediation at this site. Consequently, the system must be sized to handle and treat 23 × 10 SCFM = 230
SCFM of vapor at 100 in H2 O gauge vacuum.

Figure 23.24 Example site data for design process (From Johnson et al.,1991).

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
It is interesting to note the differences of the system designs. In the first case, a simple single extraction well system is
prescribed. Screening level model predictions, however, yield a prohibitively high areal density of extraction wells results that
should lead the user to question the efficacy of vapor extraction at this site. Thus, this sample exercise has illustrated the
inherent drawbacks of simplistic design practices that are commonly used. Based on the screening-level model predictions,
systems installed by the first method is unlikely to achieve satisfactory remediation in a reasonable time frame, although there
is nothing in the design procedure to indicate this.

23.5.3. Groundwater Pump and Treat Systems for Source Zone


Treatment
Groundwater pump and treat systems are most often used for hydraulic containment purposes. Given the low dissolved
concentrations of most contaminants (< 1 mg/L) and typical groundwater recovery rates (1–23 gal/min), most mass removal
rates will be <<0.05 kg/day. Consequently, these systems are rarely considered to be a viable source zone treatment option.

However, there are some situations in which remediation of source zones by pump and treat could be attractive. Fuel
oxygenates and alcohols are very soluble in water, and could possibly be effectively removed through dissolution. Also,
solubility enhancing additives might be added to an aquifer (alcohols, surfactants) to increase solubilization rates [Advances
Applied Technology Demonstration Facility Program (AATDF), 1997].

Figure 23.25 Example pilot test data for design process. Valves to the right of boring indicators are
soil hydrocarbon concentration (mg/kg). (From Johnson et al 1991)

If this is the case, pump and treat systems can be selected and designed using the approach discussed above inSec. 23.5.2 for
soil vapor extraction systems. Indeed the same design equations and principles can be used. One needs only to substitute
maximum dissolved concentrations Cwmax for maximum vapor concentrations Cvmax, water flow rates for air flow rates, and
effective water phase diffusion coefficients for effective vapor diffusion coefficients. Rather than repeat the discussions here, it
is assumed that the reader can make the analogy between the two systems.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 23.26 Bioventing system schematic.

23.5.4. Bioventing
23.5.4.1. Background.
Fig. 23.26 presents the schematic for a simple bioventing system. Readers will note that the components are similar to those
used in a simple soil vapor extraction system, with the notable exception that there is no vapor treatment system and air is
being injected rather than being withdrawn. Bioventing and soil vapor extraction are complementary technologies and most soil
vapor extraction systems are easily converted to bioventing operation. The main advantage of bioventing relative to soil vapor
extraction is the elimination of the need for above-ground vapor treatment (often the greatest capital and operation expense). In
theory, contaminant vapors are degraded in situ with a well-designed bioventing system.

Bioventing differs from soil vapor extraction in that the design goal is to maximize in situ biodegradation of contaminants while
minimizing vapor emissions. In contrast, the design goal for a vapor extraction system is to optimize removal by volatilization.
By virtue of drawing air into the soil, most soil vapor extraction systems induce some aerobic biodegradation; however, in that
case it is regarded as a side benefit of the process and is not a design consideration.

Much of our knowledge of bioventing performance comes from the data published from the U.S. Air Force Bioventing Initiative
(Battelle Press 1996 a, b, c). In that work petroleum fuel spills (primarily jet fuel) were the target contaminants. In summary, it
has been observed that aerobically degradable fuel compounds will biodegrade at most sites when supplied sufficient oxygen.
It appears that >5 percent by volume O2 is sufficient to maintain aerobic degradation. Observed zero-order degradation rates
typically fall in the 1–23 (mghydrocarbono , kg soil)/day range.

23.5.4.2. General Bioventing Design Principles.


The bioventing design approach follows the sequence:

1. Assess whether or not the contaminants are aerobically degradable.

2. Assess whether or not reasonable air flow rates are achievable.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
3. Conduct an in situ respirometry test to determine the site-specific respiration rate, achievable vapor flow rates, and treatment
zone boundaries.

4. Use this information to select the optimal well spacing and air injection rates.

The first is accomplished by identifying the target contaminants. Most aliphatic and monoaromatic petroleum hydrocarbons are
aerobically degradable, with the ease of degradation decreasing with increasing molecular weight, number of aromatic rings,
and increasing branched structure. Many chlorinated solvents of interest (e.g., TCE, PCE) are not by themselves readily
aerobically degradable; however, researchers are investing the possibility of cometabolic systems that involve the injection of a
gaseous primary substrate (e.g., butane) with the air stream.

The second and third design steps involve a bioventing pilot test, which is discussed below, followed by the details of specific
design calculations.

23.5.4.3. In situ Respirometry Tests.


The setup for an in situ respirometry test is shown inFig. 23.27. It is very to the simple bioventing system in Fig. 20.26, with the
exception that a conservative tracer gas (usually helium) is added to the injection air. In addition to the air injection well, blower,
and flow and pressure measuring devices, soil gas monitoring probes are distributed throughout the target treatment zone.

An in situ respirometry test generally involves the following sequence (Hinchee and Ong, 1992):

Begin air injection at a steady air injection flow rate (usually 2–20 ft3 /min). Adjust the helium injection rate so that 2 to 5
percent helium by volume is maintained in the injection stream.

Record the air injection flow rate, injection pressure, and helium concentration.

Begin monitoring and recording the soil gas vapor probes for oxygen and helium, and soil gas pressure.

Continue steady air/helium injection until steady values of oxygen and helium are detected in the soil gas monitoring
probes.

Stop the air/helium injection and monitor and record the oxygen and helium concentrations in the soil gas monitoring
probes until oxygen concentrations decline to ≈ 5 percent by volume.

Repeat these steps in an uncontaminated (background) area of the site.

Reduce the data to determine the zero-order biodegradation rate and extent of the zone of treatment.

To obtain the zero-order biodegradation rate, the oxygen and helium soil gas concentrations are plotted versus time and the
slopes (%/day) are obtained from best-fit linear regressions. If the helium loss rate exceeds the oxygen loss rate, then the data
should be discarded. This is an indication that other mechanisms (diffusion losses, sampling errors, and so on) are significant.
Provided that the data is acceptable, the zero-order biodegradation rate kB [(mg · kg)/day] is derived from the oxygen loss rate
k o (%/day):

(23.21)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 23.27 Bioventing pilot test schematic.

where:

A = air-filled porosity ( ≈ 0.25 cm3 vapor/cm3 soil)

ρo = density of oxygen gas ( ≈ 1330 mg/L)

S = stoichiometric coefficient (≈ 1 g hydrocarbon/3.5 g O2 )

ρb = soil bulk density (≈ 1.4 g/cm3 )

For the reasonable parameter values listed above, Eq. (23.21) simplifies to

(23.22)

Rates obtained from the treatment area and background pilot test area should be compared to determine of other significant
oxygen uptake mechanisms are occurring.

The zone of treatment for a given flow rate is easily defined by the soil gas oxygen helium data. Provided that the zero-order
biodegradation rate is significant, then air injection is inducing treatment anywhere that oxygen concentrations are >5 percent
by volume and the helium concentration is roughly equal to the injection concentration.

23.5.4.4. Design calculations.


Given the data from the pilot test, the system specifications can be arrived at by the following approach:

The number and placement of wells is dictated by the single-well treatment zone determined as discussed above from the
soil gas measurements.

The minimum required air injection rate for a given treatment zone of volume V (ft3 ) can be determined from Eq. (23.23):

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(23.23)

The blower size is selected based on the design air injection flow rate and soil permeability to air flow determined from the
pilot test and Eq. (23.13).

With respect to system monitoring, this generally involves:

Periodic soil gas sampling to ensure that sufficient oxygen is found in the target treatment zone

Periodic in situ respirometry tests to assess changes in respirometry rate with time; usually declining oxygen utilization
rates with time are positive indications of remedial progress

23.5.4.5. Sample pilot test data and design calculations.


Table 23.7 presents sample in situ respirometry test data (Battelle Press, 1996 a, b, c). The slope in the linear portion of the
oxygen concentration versus time data (0–7.8 h) is 25 percent/day. This translates to an estimated biodegradation rate of 18
(mg · kg)/day, using Eq. (23.22).

If the target treatment volume was 50 ft × 50 ft × 20 ft (50,000 ft3 , 1400 m3 ), then the minimum air injection rate required
predicted by Eq. (23.23) is ≈ 10 ft3 /min (0.005 m3 /s).

23.5.5. In Situ Air Sparging


23.5.5.1. Background.
Unlike soil vapor extraction and bioventing treatment systems,insitu air sparging (IAS) targets contaminant source zones
beneath the water table. A schematic of a simple system is shown in Fig. 23.28. In this process, air is injected below the water
table, with the hope that it will volatilize contaminants and deliver oxygen to stimulate aerobic degradation.

Traditionally, in situ air sparging has been viewed as a vehicle for delivering oxygen to, and promoting volatilization from,
contaminated aquifers. Now the process is also being considered for other purposes. For example, air sparging might be used
to improve air flow distribution near the capillary zone for bioventing purposes. Air sparging may also be used to deliver other
gases to the subsurface; for example, one may use the process to deliver hydrocarbon vapors (e.g., methane, ethane, propane)
required to promote the cometabolic degradation of chlorinated compounds.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Table 23.7 Sample in Situ Respirometry Bioventing Test Data

Time (h) Oxygen (% vol/vol) Dioxide Carbon (% vol/vol) Helium (%)

Source: Battelle. (1996).

–24 0 9.6 –

0 20 0.5 1.5

1.9 13.1 0.5 1.7

3.8 12.5 0.9 1.3

7.8 10.5 1.1 1.5

20 7.5 1.5 1.5

32 6.0 4.5 1.7

Figure 23.28 In situ air sparging process schematic.

23.5.5.2. Design Principles.


At this time IAS design practices are largely empirical. Despite the wide range of geologic conditions encountered in the field,
there seems to be little variance in system designs and operating conditions. Table 23.8 summarizes data presented in the
state-of-the-practice review prepared by Marley and Bruell (1995).

The major components of an IAS system typically include: air injection wells, air compressor(s), monitoring components, and
vapor extraction system (if needed). Each of these are discussed briefly below.

Air injection wells are generally constructed from polyviny chloride (PVC) and installed according to typical monitoring well
installation practices. When using vertical wells, the well screen interval is often about 2 ft long and placed entirely below the
groundwater table. The specific depth at any given site is based on considerations of the contaminant depth and subsurface
stratigraphy. To date, most injection well screened intervals have been installed 5–20 ft below the groundwater table. This

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
observation, however, may be biased by the fact that most systems have been installed at petroleum fuel release sites. It is
also possible to use driven well points as air injection wells. At this time, sufficient performance data are not available in the
literature to judge the performance of driven points relative to typical well installations.

The number of air injection wells installed is often based on a “zone of influence” approach. Based on the results of a pilot test,
or experience, the practitioner estimates the areal extent of air flow in the aquifer. Injection wells are then placed to insure that
air flows through the target treatment zone. As reported by Marley and Bruell (1994), most practitioners tend to use injection
wells that are spaced roughly 20–60 ft apart. In general, it is felt that better performance will always be achieved with higher
densities of air sparging wells.

Table 23.8 Typical Design Parameters

Parameter and Range Most Often Used Value Second Most Often Used Value Third Most Often Used Value Total N0. of
(No. of Sites) (No. of Sites) (No. of Sites) Sites

Source: Marley and Bruell, (1995).

Abbreviations:IAS ROI, radi us of influence of air injection wells; SCFM, SVE ROI, radius of influence of vapor extraction wells.

Scree length 0.5–10 ft 2 ft 3 ft 5 ft 40

(16 sites) (8 sites) (7 sites)

Well diameter 1–4 in 2 in 4 in 1 in 37

(17 sites) (7 sites) (5 sites)

Overpressure 0.35–18.2 0.35–5 psig 5–10 psig 10–15 psig 31


psig
(14 sites) (9 sites) (5 sites)

Well screen depth below 5–10 ft 10–15 ft 2–5 ft 31


water table
(10 sites) (8 sites) (6 sites)

2–26.5 ft

Air injection flow rate 1.3–5 SCFM 5–10 SCFM 15–20 SCFM 39

(16 sites) (9 sites) (5 sites)

1.3–40 SCFM

Air injection pressure 3.5– 5–10 psig 10–15 psig 20–25 psig 40
25 psig
(17 sites) (8 sites) (6 sites)

SVE ROI/IAS ROI 0.16– 1–2 0.16–1 3–4 26


7.42
(12 sites) (6 sites) (3 sites)

Compressors, or blowers, are typically selected based on anticipated injection flow rates and pressure requirements. According
to current design practices, the practitioner should expect to inject air at flow rates ranging from 1 to 20 ft3 /min per injection
well. Practitioners have varied opinions on the effect of changing air injection flow rates. Some feel that lower air injection rates
will favor biological treatment relative to volatilization (“biosparging”), while others feel that changes in flowrate can affect air
distributions. For example, the USEPA (1992) suggests that as the injection flow rate and injection pressure are increased, there
is an increased likelihood that the injected air will bypass the target treatment zone. Other laboratory-scale studies (Ji et al.,

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
1993; Rutherford and Johnson, 1996) indicate that increasing injection flow rate results in positive effects, especially in
stratified geological settings.

Vapor extraction systems are installed in conjunction with IAS systems when one must recover contaminant vapors, or where
there is the concern that contaminant vapors could migrate to enclosed spaces (utility conduits, basements, buildings, and so
on). Practitioners seem to select vapor extraction well spacing based on subsurface soil gas pressure measurements during a
pilot test. For example, it is common for practitioners to define their vapor extraction radius of influence to be that distance
where, during a pilot test, vacuums in the subsurface fell to less than 0.1 in H 2 O gauge vacuum (Johnson and Ettinger, 1994).
Then, when operating the full-scale system, extraction well flow rates may be adjusted so that extraction flow rates are greater
than injection flow rates. Injection and extraction flow rates may also be adjusted so that a vacuum is measured in soil vapor
probes at desired locations. It has been shown, however, that this approach can lead to a false sense of security. Johnson et al.
(1997) conducted a continuous helium recovery tracer test at their field study site and were only able to recover about 50
percent of the injected helium at steadystate, despite the extraction flow rate being six times greater than the injection flow rate
and vacuums being measured in all soil vapor probes.

In summary, there seem to be no “proven” design practices, although most practitioners seem to follow the same approach.
Fortunately, it seems that this approach is generating positive results at a number of sites. Probably the best practical design
approach at this time is a phased one, where a preliminary design is created based on available data and experience, a system
is installed, and then that system is optimized and modified based on the information and insight gained from monitoring data.

23.5.5.3. Short-term Pilot Tests.


Short-term pilot tests play a key role in the selection and design of IAS systems. A single air injection well is used in a typical
pilot test, and the duration of air injection is often less than 24 h. Practitioners look for changes in easily measured aquifer and
vadose zone parameters, such as dissolved oxygen, water level changes, soil gas pressures, and soil gas contaminant
concentrations. More often than not, the practitioner monitors these changes within conventional monitoring wells that already
exist at the test site. Increases in dissolved oxygen and decreases in dissolved contaminant levels are taken to be favorable
indicators of feasibility, whereas changes in water levels within the wells and vadose zone air pressures are assumed to be
indicators of the area of treatment.

In a few cases investigators have supplemented conventional practices with other more innovative approaches. Most of these
have focused on better assessing air distributions. Lundegard (1994) reports on the use of electrical resistance tomography
(ERT) to measure air flow distributions in situ, Acomb et al. (1997) applied geophysical tools, Beckett (personal communication,
1995) has utilized a combination of tracer gases and conductance probe measurements, and Clayton, et al. (1995) used time-
domain reflectrometry (TDR) probes to assess air content changes in the saturated zone.

With the exception of the work by Lundegard (1994) and the discussion given in Boersma et al. (1994), few have questioned
conventional pilot test procedures and monitoring practices, as they relate to the design and assessment of IAS performance.
Lundegard, who compared ERT imaging results with traditional pilot test measurement and data reduction practices, concluded
that conventional monitoring approaches overestimate the zone of air flow at relatively homogeneous sites. Boersma et al.
(1994) suggested that there may be a large degree of uncertainty associated with system design and performance
assessment, when using conventional monitoring approaches. In both cases, these authors focused their discussions primarily
on the extent of air distribution.

Recently, Johnson et al. (1997) conducted a field study in order to compare short-term pilot test data with longer-term
performance data. At their site a number of discrete drivepoint sampling implants were installed in addition to conventional
monitoring wells. Groundwater quality indicators (dissolved hydrocarbons, dissolved oxygen, and so on) were measured with
time and tracer gas tests were performed. Based on their data it was concluded that

short-term pilot-test data collected from conventional monitoring wells yielded the most optimistic picture of long-term
system performance.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
short-term pilot-test discrete implant monitoring results were less promising as they showed little effect of in situ air
sparging during the 3 day pilot test and

there was no clear correlation between any of these short-term measurements and the longer-term system performance at
this site.

It was also concluded that, in the absence of validated predictive tools, it is very unlikely that short-term pilot-test data can be
extrapolated to long-term system performance. Therefore, the authors recommended that practitioners design their short-term
pilot tests to assess infeasibility rather than feasibility; that is to say that short-term pilot tests should be designed to identify
conditions indicative of poor system performance. These might include:

Inability to inject significant air flow (>1 ft3 /min) into the aquifer at pressures less than the combined hydrostatic head and
soil overburden

Observation of adverse impacts caused by air injection (e.g., uncontrollable migration of vapors)

Poor air distributions (e.g., highly stratified or unidirectional air flows)

The inability to detect any qualitative indicators of air distribution (e.g., no observed dissolved oxygen changes, little tracer
gas recovery, and so forth)

The interaction between air sparging behavior and subsurface geology was also highlighted in that study. Strata that were not
identified during the initial site assessment controlled the performance at the study site, and these caused the formation of a
large horizontal trapped air bubble beneath the water table. This illustrates the need for careful geologic assessment at
potential in situ air sparging sites.

23.6. DISSOLVED PLUME TREATMENT TECHNOLOGIES


The remediation of dissolved contaminant plumes is generally accomplished by groundwater pump and treat systems, in situ
air sparging, or natural attenuation. In recent years, natural attenuation has emerged as the preferred low-cost option, and
therefore it is the focus of this section.

23.6.1. Natural Attenuation


Natural attenuation refers to the collective effect that natural dispersion, degradation, volatilization, and other loss mechanisms
have on dissolved groundwater contaminant migration. In some cases, the combination of these mechanisms significantly
limits dissolved plume growth. When this is the case, and the source is removed at those sites, often the dissolved plume
dissipates within a few years.

Natural attenuation is a “passive” remediation process in that there is not any associated above-ground equipment to be
installed or operated. The only “design” activities involve identifying that natural attenuation is an acceptable option, and
designing an appropriate monitoring protocol. The latter is installed to monitor and track the disappearance of the dissolved
contaminant plume.

The primary questions that need to be addressed before selecting natural attenuation for treatment are the following:

Is the dissolved contaminant plume adversely impacting human health or a protected resource?

Is the dissolved plume stable (stationary), or receding, or growing?

At what rate could the plume migrate?

Does the dissolved plume concentration decrease with distance as one moves downgradient?

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Is there another more cost-effective treatment option?

Answers to these questions are used to determine the appropriateness of natural attenuation. This generally requires at least 3
years of quarterly groundwater quality monitoring data at 5–10 monitoring wells distributed along the dissolved plume. Multiple
years of historical data are needed to filter out seasonal effects and water table level fluctuations. As subsurface transport
processes are relatively slow and one would normally expect concentration changes to occur slowly over long periods of time,
the questions above often cannot be answered with great technical certainty, even if much site monitoring data is available.

If natural attenuation is the selected remedial option, then a monitoring plan is designed. These plans generally involve the
placement of sentinel wells between the dissolved contaminant plume and the nearest groundwater pumping well or valued
resource. The sentinel well is placed at least 1 year’s travel time upgradient from the water supply well or valued resource, and
along the projected plume centerline.

Secondary-level data collection activities are geared toward identifying the mechanisms that control natural attenuation. In
most cases these focus on identifying circumstantial evidence that contaminant biodegradation is taking place. Two recent
two survey studies have shown that most benzene, toluene, ethylbenzene, and xylenes (BTEX) dissolved plumes at leaking
underground storage tank (LUST) sites do not extend more than 300 ft (100 m) from the downgradient edge of the source zone
Rice et al., 1995). Mace et al., 1997). The only explanation that is consistent with all the data is that this behavior is attributed to
biodegradation as the plume lengths would have been expected to be much longer based solely on advection and dispersion.
Thus monitoring plans often focus on quantifying whether or not sufficient electron receptors are present outside the plume
(e.g., O2 , NO3 -, SO4 2-, etc.). McAllister and Chiang (1994) recommend looking for inverse relationships between contaminant
concentrations and electron receptor concentrations. For example, at fuel hydrocarbon spill sites, one often sees that dissolved
oxygen levels within the contaminant plume are depressed, indicating oxygen uptake, and suggesting that aerobic
biodegradation is taking place.

23.7. CONTAMINANT MIGRATION BARRIERS


23.7.1. Background
In some cases treatment of the source zone is not practicable, or the treatment is projected to take a long time. In such cases
there is often concern about continued spreading of the contamination through the aquifer.

Historically contaminant migration barriers have focused on preventing flow into and away from the source zone as shown in
Fig. 23.5. This can be accomplished through use of physical barriers, such as combinations of grout walls, sheet piling,
geosynthetic membranes, and surface infiltration caps. It can also be accomplished by controlled groundwater withdrawal,
treatment, and reinjection (groundwater pump and treat systems). Groundwater pump and treat is probably the most widely
applied technology of this group because of its maturity and robustness.

Recently, a third option for contaminant migration barriers has emerged. Rather than focusing on flow restrictions, this new
approach focuses on in situ treatment as the contaminants are brought by groundwater flow to the treatment zone. Reaction
barriers may utilize chemical or biological destruction, or other physical/chemical treatment approaches such as in situ air
stripping.

The design principles involved with hydraulic control systems and the reaction barriers concept are briefly discussed below.

23.7.2. Hydraulic Containment Systems


The design objective here is to hydraulically contain the dissolved groundwater contamination, while minimizing the pumping
and water treatment required to accomplish this. Groundwater extraction and injection wells are generally placed between the

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
downgradient plume edge and the point beyond which migration is not allowed.

The typical design process involves the following steps:

Characterize the dissolved groundwater contamination plume: determine the extent of contamination and the spatial
distribution of contaminant concentrations.

Collect aquifer characterization data: geologic cross section, aquifer transmissivity, storativity, hydraulic conductivity and
thickness, regional and local groundwater gradient.

Define the design objectives: containment at what point(s) and treatment to what level(s).

Propose initial system design(s): locations of wells, pumping rates, and so on.

Using a groundwater flow model, predict the resulting groundwater flow field; display the results as contours of hydraulic
head and as streamlines, and define the capture zone.

Determine if the proposed design meets the design objectives. If so, perform a sensitivity analyses to see if other well
configurations and operating conditions will reduce the amount of pumping necessary and total system costs. Repeat this
step until an acceptable hydraulic design results (optimization).

Select an appropriate groundwater treatment technology for the produced water (e.g., air stripping, activated carbon
adsorption, biological treatment).

Install the system and adjust pumping rates to maintain desired drawdowns at compliance points; revise the groundwater
flow simulator with new data and again perform a sensitivity analyses to see if other well configurations and operating
conditions will reduce the amount of pumping necessary and total system costs.

Characterization of the dissolved groundwater plume is accomplished through groundwater sampling and chemical analyses,
and has presumably already been done during the site assessment phase. Characterization of aquifer properties is generally
accomplished through traditional in situ aquifer testing employing slug tests and constant head pump tests. It is recommended
that these tests be conducted with more than one well (as many as is feasible), as the results often vary by orders of
magnitude. Sustained groundwater pump tests are preferred over the shorter and more economical slug tests as the pump test
provides a more direct measure of the expected groundwater drawdown and capture with distance.

Groundwater modeling relies heavily on software-based numerical codes. Analytical solutions are available for some
conditions (Javendel and Tsang, 1986); however, so many user friendly graphical-based software packages are available that
there is little reason not to use one as a design tool. Many are based on the U.S. Geological Surey (USGS) MODFLOW
groundwater flow code (McDonald and Harbaugh, 1984).

While most groundwater capture systems are designed by trial and error, using a software simulator, some are moving toward
more rigorous optimization approaches. Ahlfeld (1994); Ahlfeld et al, (1995), and Gailey and Gorelick (1993), present example
designs generated from a more formal optimization approach. In general, the major cost of such systems is the operation and
maintenance expense; thus, minimizing pumping rates is very desirable.

Conventional pump and treat system performance is easily monitored through groundwater elevations and groundwater quality
analyses. Compliance can be judged by measuring groundwater elevations, contouring the data, and then inspecting the results
to ensure gradients towards the capture wells at the compliance boundary. Conventional pump and treat systems are also
robust in the sense that adjustment in operating conditions (pumping rates) can make up for uncertainty inherent in the initial
design.

Recently, the use of groundwater recirculation wells has been proposed as an alternative to conventional pump and treat
designs (Herrling and Buermann, 1990; Herrling et al., 1994). As shown in Fig. 23.29, groundwater is extracted at one elevation
in the aquifer and reinjected at another, thereby establishing a vertical circulation flow cell. In addition, inwell groundwater
treatment occurs (usually through the use of in-well air strippers). In some cases, due to regulatory loopholes, this technology

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
has been selected because it eliminates the need for costly groundwater reinjection permits. Little performance data are
available in the literature. Fundamental considerations suggest that performance of such systems will be more sensitive to
aquifer stratigraphy than conventional systems, and that, due to the presence of significant vertical gradients, conventional
monitoring approaches will not be appropriate.

The design process for such systems, however, should follow the same design approach described above for conventional
pump and treat, although the numerical simulations will likely be more complex and difficult to interpret.

23.7.3. Reaction-Based Contaminant Migration Barriers


The use of in situ reaction barriers has gained interest because it eliminates the need for costly above-ground water treatment.
The treatment occurs as groundwater flows through an in situ treatment system. In some cases flow barriers and/or
groundwater extraction and injection may be used to direct the dissolved contaminant plume to the treatment zone. For
example, Fig. 23.30 shows a funnel and gate design proposed by Starr and Cherry (1994). Such hybrid systems can be
desirable when the cost of construction of the treatment zone dominates the economics, as proper control of the hydraulics can
be used to minimize the width of the treatment zone.

Chemical, physical/chemical, and biological treatment barriers have been proposed in the literature. Gillham and O’Hannesin
(1992) have showed that elemental iron will induce the dechlorination of some chlorinated contaminants of interest (e.g., TCE,
PCE, TCA, DCA). The reaction rates are slow by conventional above-ground treatment standards, but are sufficiently fast for
slow flow groundwater systems. Pankow et al., (1993) propose the use of in situ air sparging in trenches to prevent plume
migration.

Figure 23.29 Groundwater recirculation well schematic.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 23.30 Funnel and gate reaction wall chemical migration barrier schematic.

No matter how treatment is to be accomplished, the design approach will follow the same general sequence of activities:

1. Characterize the dissolved groundwater contamination plume: determine the extent of contamination (especially the width)
and the spatial distribution of contaminant concentrations.

2. Collect aquifer characterization data: geologic cross section, aquifer transmissivity, storativity, hydraulic conductivity and
thickness, and regional and local groundwater gradient.

3. Define the design objectives: containment at what point(s) and treatment to what level(s).

4. Propose initial hydraulic system design(s): locations of any barriers, wells, and so on.

5. Using a groundwater flow model, predict the resulting groundwater flow field display the results as contours of hydraulic head
and as streamlines, and determine the groundwater flow rate and contaminant concentrations entering the treatment zone.

6. Select the in situ groundwater treatment technology and design the treatment zone to obtain the desired contaminant
concentration reduction.

7. Perform a sensitivity analyses to see if other configurations and operating conditions will reduce the total system costs.
Repeat this step until an acceptable design results (optimization).

8. Install the system and adjust pumping rates to maintain desired drawdowns at compliance points; revise the groundwater
flow simulator with new data and again perform a sensitivity analyses to see if other well configurations and operating
conditions will reduce the amount of pumping necessary and total system costs.

Many of the steps are similar to those described above for the design of pump and treat systems. In this case one needs to
balance the capital costs of constructing the treatment trench with the operation and maintenance costs associated with any
groundwater extraction that might be necessary to focus the plume through a narrower treatment zone. Again, the hydraulics
and groundwater flow will be estimated with groundwater flow simulators.

The thickness of the treatment zone will be a function of the contaminant concentration, treatment technology, cleanup goal,
and groundwater flow rate. For the case of the iron treatment walls proposed by Gillham and O’Hannesin (1992), kinetic data
suggest that the reaction rates can be adequately described by a first-order kinetic expression. Therefore, the design equation
becomes

(23.24)

or rearranged:

(23.25)
o

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
where, C = concentration in groundwater leaving the treatment zone (mass/volume),Co = concentration in groundwater entering
the treatment zone (mass/volume), λ = first-order degradation rate (one/time), L = length of treatment zone in the direction of
flow (length), qtreatment = groundwater specific discharge through the treatment zone [volume/(area · time)],qupgradient = natural
groundwater specific discharge upgradient of the treatment zone [volume/(area · time)], Wcapture = upgradient width of capture
zone (length) and Wtreatment = width of treatment zone orthogonal to groundwater flow (length)

Occasionally, half-lifes t1/2 (time) are cited instead of first-order reaction rates. To convert between the two, the following
relation can be used:

(23.26)

Figure 23.31 displays results calculated from Eq. (20.24) for a range of first-order degradation rates and residence times
(=L/qtreatment ).

As an example, consider the construction of an iron treatment wall for TCE-contaminated groundwater. Suppose that the
groundwater concentration is 100 μg/L, the treatment goal is 5 μg/L, the groundwater velocity is 1 ft/day (0.3 m/day), and the
plume width is 300 ft (100 m). If the upgradient groundwater specific discharge is 0.3 ft/day (0.1 m/day), the half-life of TCE is
0.5 h, and the desired width of treatment zone is 100 ft (30 m), then the required thickness would be

(23.27)

23.7.4. Air Sparging Cut-Off Trenches


In situ groundwater treatment can also be accomplished by air stripping within a trench through which groundwater flows.
Pankow et al. (1993) assumed that the water within the trench was well mixed and in equilibrium with the sparge air leaving the
system. Under these conditions, they derived the design equations

(23.28)

and

(23.29)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 23.31 Effect of residence time and first–order degradation rate on treatment wall
performance

where

(23.30)

Here all variables are defined as in the discussion for chemical reaction barriers, with:

KH = dimensionless Henry’s law constant [(mg/L air)/(mg/L water)]

Atreatment = vertical cross sectional area of treatment zone (qtreatment Atreatment = total

groundwater flow rate into trench) (area)

Q air = air injection rate (volume/time)

Figure 23.32a and b displays treatment efficiencies for a range of air injection flow–rates, groundwater flow rates, Henry’s law
constants, and number of sequential trenches.

As an example, consider the scenario outlined above for the design of the iron treatment wall.KH for TCE is ≈ 0.35 [(mg/L
air)/(mg/L water)]. Thus, if the treatment zone were 10 ft thick, then for a single air sparging trench, the required air injection
rate would be

(23.31)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 23.32 Predicted performance for single and multiple well-mixed air sparging trench system
as a function of the dimensionless Henry’s law constant and the number of trenches in series.

with a required air injection rate this low there would appear to be little incentive to move toward a multiple trench design in this
case.

23.8. SUMMARY
This chapter has focused on in situ treatment systems designed to eliminate or minimize the impacts of hazardous chemicals
on groundwater quality. General and specific design approaches were introduced for a range of technologies; however, not all
technologies were discussed. Also, in the interest of providing breadth, not all aspects of the design process were addressed
for each technology. Readers are referred to the more detailed design-oriented documents for technologies of interest.

In closing, it is important to keep in mind that the design of in situ treatment systems for groundwater contamination is
uniquely different from the design of above-ground treatment and hydraulic conveyance systems, as the subsurface is difficult
to describe precisely, it is naturally heterogeneous and, due to practical constraints, characterization data are generally limited.
Thus, treatment systems must be designed under conditions of great uncertainty. As a result, the design of in situ treatment
systems will always rely heavily on conceptual models, screening-level calculations, empiricism, heuristics, and experience, and
monitoring and refinement.

Finally, given the inherent large initial uncertainty, it is important to recognize that the design phase continues well past the
construction, installation, and operation of the initial system. Following sound design practices does not guarantee success; it
does, however, provide a higher probability of success at most sites. Appropriate monitoring is essential to verify assumptions
built into the conceptual model and initial design, and system design refinement should be anticipated in the initial design. For
this reason, it is important to build robust systems that are both flexible and expandable.

23.9. REFERENCES
1. AATDF/AATDF Technology Practices Manual for Surfactants and Cosolvents: 2d ed. Advanced Applied Technology
Demonstration Facility Program. Rice University, The Energy and Environmental Systems Institute, 1997.
2. Acomb, L. J., D. McKay, P. Currier, S. T. Berglund, T. V. Sherhart, and C. V. Benedicktsson, “Neutron Probe Measurements
of Air Saturation Near an Air Sparging Well.” In In Situ Aeration: Air Sparging, Bioventing, and Related Remediation
Processes, Battelle Press, Columbus, OH, 1995.
3. Ahlfeld, D., “Applications of Optimal Hydraulic Control to Ground Water Systems,”Journal of Water Resources Planning
and Management, 12(3):351–365, 1994.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
4. Ahlfeld, D., R., H. Page, and G. F. Pinder, “Optimal Ground-Water Remediation Methods Applied to a Superfund Site: From
Formulation to Implementation” Ground Water, 33(1):58–70, 1995.
5. American Petroleum InstituteA Guide to the Assessment and Remediation of Underground
6. Petroleum Releases, API Publication 1628, American Petroleum Institute Marketing Department, Washington, DC, 1989.
7. American Society for Testing and Materials (ASTM), Standard Guide for Risk-Based Corrective Action (RBCA) at
Petroleum Release SitesE1739–95, ASTM, 1995.
8. Baehr, A. L., and M. Y. Corapcioglu, “A Compositional Multiphase Model for Groundwater Contamination by Petroleum
Products, 2. Numerical Solution” Water Resources Research, 23(1): 231–214, 1987.
9. Baehr, A. L. and M. H. Hult. “Evaluation of Unsaturated Zone Air Permeability Through Pneumatic Tests” Water Resources
Research, 27(10):2605–2617, 1991.
10. Battelle. Press, Bioventing Manual, Vol. I, Bioventing Principles, Battelle Press, 1996 a.
11. Battelle. Press, Bioventing Manual, Vol. II, Bioventing Design, Battelle Press, 1996 b.
12. Battelle. Press, Bioventing Manual, Vol. II, Additional Considerations for Bioventing Systems. ess, 1996.
13. Bear J., Hydraulics of Groundwate., McGraw-Hill, 1979.
14. Beckett, G., Personal communication, 1995.
15. Beckett, G., and D. Huntley, Characterization of flow parameters controlling soil vapor extraction.Groundwater32(2):239–
247, 1994.
16. Benson, D. A., D. Huntley, and P. C. Johnson., “Modeling Vapor Extraction and General Transport in the Presence of NAPL
Mixtures and Nonideal Conditions” Groundwater, 31(3):437–445, 1993.
17. Boersma, P., P. Newman, and K. Piontek, “The Role of Groundwater Sparging in Hydrocarbon Remediation,” American
Petroleum Institute Pipeline Conference, Houston, TX, April 25–28, 1994.
18. Clayton, W., R. A. Brown, and D. H. Bass, Presentation at the In Situ and On-Site Bioreclamation Conference, 3rd
International Symposium, San Diego,CA, April 24–27, 1995.
19. Corapcioglu, M. Y., and A. L. Baehr, “A Compositional Multiphase Model for Groundwater Contamination by Petroleum
Products, 1. Theoretical Considerations,” Water Resources Research, 23(1), 1987.
20. Fetter, C. W., Contaminant Hydrology, Macmillan, New York, 1993.
21. Gailey, R. M., and S. M. Gorelick, “Design of Optimal, Reliable Plume Capture Schemes: Application to the Gloucester
Landfill Ground-Water Contamination Problem” Ground Water, 31(1):107–114, 1993.
22. Gillham, R. W., and S. F. O’Hannesin, “Metal-Catalysed Abiotic Degradation of Halogenated Organic Compounds.”Modern
Trends in Hydrogeology, International Association of Hydrogeologists, Hamilton, Ontario, May 10–13, 1992.
23. Herrling, B,. and W. Buermann, “A New Method for in Situ Remediation of Volatile Contaminants in Groundwater—
Numerical Simulation of the Flow Regime,” in G. Gambolati, A. Rinaldo, C. A. Brebbia, W. G. Gray, and G. F. Pinder, eds.,
Computational Methods in Subsurface Hydrology, Springer, Berlin, 1990, pp. 299–304.
24. Herrling, B., J. Stamm, E. Alesi, G. Bott-Breuning, and S. Diekmann, “In Situ Bioremediation of Groundwater Containing
Hydrocarbons, Pesticides, or Nitrate Using Vertical Circulation Flows,” in R. E. Hinchee, ed., Air Sparging for Site
Remediation, Lewis Publishers, Boca Raton, FL. 1994, 56–80.
25. Hinchee, R. L., R. N. Miller, and P. C. Johnson, In Situ Aeration: Air Sparging, Bioventing, and Related Remediation
Processes, Battelle Press, 1995.
26. Hinchee, R. E., and S. K. Ong, “A Rapid In Situ Respiration Test for Measuring Aerobic Biodegradation Rates of
Hydrocarbons in Soil,” Journal of the Air & Waste Management Association, 1992,
27. Hutzler, N. J., B. E. Murphy, and J. S. Gierke, State of Technology Review Soil Vapor Extraction Systems, EPA/600/2-
89/024, U.S. Department of Commerce, NTIS Document PB89-195184, 1989.
28. Ji, W., A. Dahmani, D. Ahlfeld, J. D. Lin, and E. Hill, “Laboratory Study of Air Sparging: Air Flow
29. Visualization,” Ground Water Monitoring and Remediation, Fall1993.
30. Johnson, P. C., and R. A. Ettinger, ”Some Considerations for the Design of In Situ Vapor Extraction Systems: Radius of
Influence -vs- Radius of Remediation,” Ground Water Monitoring and Remediation, 14(3):123–128, 1994.
31. Johnson, P. C., G. E. Hoag, R. H. Hinchee, R. A. Brown, and A. L. Baehr,Innovative Site Remediation Technology: Vapor
Extraction-Based Technologies (Soil Vapor Extraction, Bioventing, Air Sparging, and Thermally-Enhanced Soil Vapor
Extraction), USEPA/AAEE WASTECH Monograph, 1994.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
32. Johnson, P. C., R. L. Johnson, C. Neaville, E. E. Hansen, S. M. Stearns, and I. J. Dortch, “An Assessment of Conventional In
Situ Air Sparging Tests,” Ground Water, 1997.
33. Johnson, P. C. and A. J. Stabenau, HyperVentilate—A Software Guidance System Created for Vapor Extraction
Applications—Users Manual, USEPA 600/R-93/028, 1991.
34. Johnson, P. C., C. C. Stanley, D. L. Byers, D. A. Benson, and M. A. Acton. “Soil Venting at a California Site: Field Data
Reconciled with Theory,” Hydrocarbon Contaminated Soils and Groundwater: Analysis, Fate, Environmental and Public
Health Effects, Remediation, Vol I.P. T. Kostecki and E. J. Calabrese, eds., Lewis Publishers, Boca Raton, FL, 1991, pp.
253–282.
35. Johnson, P. C., C. C. Stanley, M. W. Kemblowski, D. L. Byers, and J. D. Colthart, “A Practical Approach to the Design,
Operation, and Monitoring of In Situ Soil-Venting Systems,” Ground Water Monit. Rev., 10(2):159–178, 1990a
36. Johnson, P. C., M. W. Kemblowski, and J. D. Colthart, “Quantitative Analysis for the Cleanup of Hydrocarbon-
Contaminated Soils by In Situ Soil Venting,” Ground Water, 3 (28):413–429, 1990b
37. Johnson, R. L., P. C. Johnson, D. B. McWhorter, R. Hinchee, and I. Goodman, “An Overview of In Situ Air Sparging,”Ground
Water Monitoring and Remediation, 13(4):127–135, 1993.
38. Kemblowski, M. W., and C. Y. Chiang. “Hydrocarbon Thickness Fluctuations in Monitoring Wells,” Ground Water,
28(2):244–252, 1990.
39. Lundegard, P., “Actual Versus Apparent Radius of Influence—An Air Sparging Pilot Test in a Sandy Aquifer,” American
Petroleum Institute (API)/NGWA Conference, Petroleum Hydrocarbons and Organic Chemicals in Ground Water:
Prevention, Detection and Restoration, Houston, TX, 1994.
40. Mace, R. E., R. S. Fisher, D. M. Welch, and S. P. Parra, “Extent, Mass, and Duration of Hydrocarbon Plumes from Leaking
Underground Storage Tank Sites in Texas,” Geologic Circular 97-1, Bureau of Economic Geology, The University of Texas
at Austin, 1997.
41. Marley, M. C., and C. J. Bruell, In Situ Air Sparging: Evaluation of Petroleum Industry Sites and Considerations for
Applicability, Design, and Operation, API Publication 4609, American Petroleum Institute. Washington, DC, 1995.
42. Massmann, J. W., “Applying Groundwater Flow Models in Vapor Extraction System Design,”Journal of Environmental
Engineering, 115(1):129–149, 1989.
43. McAllister, P. M., and C. Y. Chiang, “A Practical Approach to Evaluating Natural Attenuation of Contaminants in
Groundwater,” Ground Water Monitoring and Remediation, 161–173, 1994.
44. McDonald, M. G., and A. W. Harbaugh, “A Modular Three-Dimensional Finite-Difference GroundWater Flow Model,” U.S.
Geological Survey Open-File Report83–875, 1984.
45. Pankow, J. F., R. L. Johnson, and J. A. Cherry, “Air Sparging in Gate Wells and Cutoff Walls and Trenches for Control of
Plumes of Volatile Organic Compounds,” Ground Water. 31(4):654–663, 1993.
46. Rice, D. W., R. D. Grose, J. C. Michaelsen, B. P. Dooher, D. H. MacQueen, S. J. Cullen, W. E. Kastenberg, L.G. Everett, and
M.A. Marino, California Leaking Underground Fuel Tank Historical Case Analyses, UCRL-AR-122207, Lawrence Berkeley
Laboratories, Environmental Restoration Division, Environmental Protection Department, 1995.
47. Rutherford, K., and P. C. Johnson. “Effects of Process Control Changes on Aquifer Oxygenation Rates During In Situ Air
Sparging in Homogeneous Aquifers,” Ground Water Monitoring and Remediation. Fall 1996.
48. Shan, C. S., R. W. Falta, and I. Javandel, “Analytical Solutions for Steady Gas Flow to a Soil Vapor
49. Extraction Well,” Water Res. Res., 28(4):1105–1120, 1992.
50. Starr, R. C., and J. A. Cherry. “In Situ Remediation of Contaminated Ground Water: The Funnel and Gate System,”Ground
Water, 32(3):465–476, 1994.
51. U.S. Environmental Protection Agency, National Water Quality Inventory—1988 Report to Congress, EPA-440-4-90-003,
Environmental Protection Agency, 1990.
52. U.S. Environmental Protection Agency, A Technology Assessment of Soil Vapor Extraction and Air Sparging, EPA/600/R-
192/173, Environmental Protection Agency, 1992.
53. U.S. Environmental Protection Agency. Review of Mathematical Modeling for Evaluating Soil Vapor Extraction Systems.
Office of Research and Development, 1995. EPA/540/R-95/513. July.
54. U.S. Environmental Protection Agency, Soil Screening Guidance: Technical Background Document, Office of Solid Waste
and Emergency Response, EPA/540/R-95/128, PB96-963502, Environmental Protection Agency, 1996a.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
55. U.S. Environmental Protection Agency, 1996b. Superfund Chemical Data Matrix (SCDM). Can be downlo—aded from the
USEPA web site www.epa.gov, U.S. Environmental Protection Agency, 1996b.
56. U.S. Geological Survey, Documentation of AIR3D, an Adaptation of the Ground-Water-Flow Code MODFLOW to Simulate
Three-Dimensional Air Flow in the Unsaturated Zone, Open File Report94–533, 1994.
57. Weaver, J. W., R. J. Charbeneau, B. K. Lien, and J. B. Provost, The Hydrocarbon Spill Screening Model (HSSM) Vol, 1:
User’s Guide, Office of Research and Development, U.S. Environmental Protection Agency, EPA/600/R-94/039a, 1994.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.

You might also like