Sei sulla pagina 1di 57

Journal of Contaminant Hydrology, 6 (1990) 10%163 107

Elsevier Science Publishers B.V., Amsterdam

Review Paper

A REVIEW OF IMMISCIBLE FLUIDS IN THE SUBSURFACE:


PROPERTIES, MODELS, CHARACTERIZATION AND
REMEDIATION

JAMES W. MERCER and ROBERT M. COHEN


GeoTrans, Inc., Herndon, VA 22070 (U.S.A.)
(Received February 5, 1990; revised and accepted May 8, 1990)

ABSTRACT

Mercer, J.W. and Cohen, R.M., 1990. A review of immiscible fluids in the subsurface: Properties,
models, characterization and remediation. J. Contain. Hydrol., 6: 107-163.

In the past few years~ as hazardous waste sites have been studied more often and in more detail,
immiscible fluids have been encountered in the subsurface with greater frequency. These
nonaqueous phase liquids (NAPL's) behave differently than dissolved solutes in the subsurface.
This behavior depends on fluid properties such as interfacial tension, viscosity and density. In
addition, mass transfer produces vapor transport in the vadose zone and solute transport in
groundwater. Mass transfer depends on properties associated with volatilization and aqueous
solubility. As a consequence, characterization techniques as well as remediation efforts must be
modified at sites where NAPL's are present. Although considerable research is necessary before
NAPL problems are well understood, sufficient work has been performed to permit a review of
NAPL properties and behavior in the subsurface.

INTRODUCTION

Nonaqueous phase liquids (NAPL's) have been discovered at numerous


hazardous waste sites (e.g., Faust, 1985; Mercer et al., 1985; Cohen et al., 1987).
In addition, NAPL often is identified with contamination problems associated
with underground storage tanks. According to Villaume (1984), typical
chemical and industrial processes that may involve NAPL include transformer
oil containing polychlorinated biphenyls (Roberts et al., 1982; Schwartz et al.,
1982), trichloroethene and related chlorinated hydrocarbons (Palombo and
Jacobs, 1982; Carpenter, 1984), coal tars from manufactured gas plants (D.C.
Wilson and Stevens, 1981; Yazicigil and Sendlein, 1981; Lafornara et al., 1982;
Unites and Houseman, 1982; Villaume, 1982, 1984; W.R. Adams and Atwell,
1983; Anastos et al., 1983; Thompson et al., 1983; Villaume et al., 1983a, b), steel
industry coking operations (Coates et al., 1982), wood treating operations (Hult
and Schoenberg, 1981; Ramsey et al., 1981; Ehrlich et al., 1982; Hickok et al.,
1982; Pereira et al., 1983), and petroleum products (Holzer, 1976; P.L. Hal! and
Quam, 1976; Pfannkuch, 1983). As an example of the size of the potential

0169-7722/90/$03.50 © ].990 - - Elsevier Science Publishers B.V.


108 J.W. M E R C E R A N D R.M. C O H E N

NOTATION

Definition nf terms and symbols

A area
c~ concentration of chemical in water (tool m-3)
g gravitational acceleration constant
x, non-advective flux of species i in the ~ phase
hc capillary head
K. Henry's law constant (atm. m a mo1-1)
partition coefficient of species i between the ~ and fl phases
Kow octanol/water partition coefficient
goo organic carbon/water partition coefficient (ml g-l)
k intrinsic permeability
kg gas phase permeability
km volumetric mass exchange coefficient
kra relative permeability of air
krn relative permeability of NAPL
k rna relative permeability of NAPL in an air-NAPL system
k~w relative permeability of NAPL in a water-NAPL system
k*w relative permeability of NAPL at the residual saturation of water
k~ relative permeability of water
m mass of NAPL source
mx mass exchanged
n porosity
P partial pressure of chemical in gas phase (atm.)
PA vapor pressure of a solution
vapor pressure of the pure solvent
Pc capillary pressure
PNAP~ NAPL pressure
Pw water pressure
Q strength of a hydrocarbon source
R volumetric retention capacity (liters NAPL/m a of medium)
external supply of species i to the ~ phase
r pore radius
mass exchange of species i due to interphase diffusion and/or phase change
s saturation
Sr residual saturation
t time
V volume of NAPL source
NAPL volume of NAPL
voids volume of pore space
V~ mass average velocity of the ~ phase
:t
Wi mass fraction of species i in the • phase
x~ mole fraction of the solvent
compressibility of ~ phase
fraction of volume occupied by the ~ phase
contact angle
II absolute viscosity
p~ intrinsic mass density of the ~ phase
PN NAPL density
Pw water density
O'Nw interfacial tension between NAPL and water
O'N.~ interfacial tension between NAPL and solid
O'ws interfacial tension between water and solid
A REVIEW OF IMMISCIBLE FLUIDS IN THE SUBSURFACE 109

problem, U.S.E.P.A. (1979) estimates that in 1974 ~ 310,200 t* of waste solvents


alone were produced by degreasing operations. In addition, there are an
estimated 796,000 individual motor fuel storage tanks in the U.S.A. (U.S.E.P.A,
1986a).
Although chemical properties and site conditions vary from site to site, the
basic principles governing the fate and transport of NAPL's are the same.
These principles may be used to understand the contamination problem and to
evaluate remediation. Unfortunately, development of state-of-the-art
technology for dealing with NAPL problems lags behind the technology
developed for many other groundwater contamination problems. For example,
several models are available to simulate the flow of NAPL; however, obtaining
chemical-specific and site-specific data is difficult. Consequently, for most sites,
models only may be used in a conceptualization mode.
NAPL migration in the subsurface is affected by (Feenstra and Cherry,
1988): (1) volume of NAPL released; (2) area of infiltration; (3) time duration of
release; (4) properties of the NAPL; (5) properties of the media; and (6)
subsurface flow conditions. The cross-sectional schematic in Fig. la depicts the
distribution of organic chemicals in multiple phases resulting from a release of
lighter-than-water nonaqueous phase liquid (LNAPL). When introduced into
the subsurface, gravity causes the NAPL to migrate downward through the
vadose zone as a distinct liquid. This vertical migration also is accompanied to
some extent by lateral spreading due to the effect of capillary forces (Schwille,
1988) and due to medium spatial variability (e.g., layering) which is not shown
in Fig. la. As the NAPL progresses downward through the vadose zone, it
leaves residual liquid (residual saturation) trapped in the pore spaces. This
entrapment is due to surface tension effects. In addition to migration of NAPL,
some of the immiscible fluid may volatilize and form a vapor extending beyond
the NAPL.
If the release is sufficiently large, some of the NAPL will eventually reach
the saturated zone. Here LNAPL will spread laterally along the capillary
fringe. It may also depress natural groundwater levels. The LNAPL distribu-
tion depends on LNAPL, water and air pressures and the pore size distribution.
Fig. la is more typical of a homogeneous, permeable medium. In heterogeneous
media, the LNAPL distribution will be more complex. As the NAPL encounters
flowing water, soluble components may dissolve to form a solute plume that can
migrate due to hydraulic gradients. Examples and descriptions of dissolved
chemical plumes are provided in Mackay and Cherry (1989).
Denser-than-water nonaqueous phase liquid (I)NAPL) will displace water
and continue its migration under pressure and gravity forces (Fig. lb).
Preferential spreading will occur where DNAPL encounters relatively
permeable layers, fractures, or other pathways that present less capillary
resistance to entry than underlying less permeable strata. Given sufficient
volume, DNAPL will continue its downward migration until it encounters a
*lt = 1 metric tonne = 103kg.
110 J.W. MERCER AND R.M. COHEN

GROUND SURFACE
"' 6
~.-:~..~~:::~-.~-

,-APL Z O N E - . . - ~ ~.~:i;:~;=.~ VADOS~ ZONE

~;'i=~i!~:;'t:~:~;!-::~=~-~ GAS ZONE (evaporation envelope)

CAPILLARY FRINGE ~ ~": ~:~',I=~


:~: N:.~/.~.~.:.~i~.
~-
~"

.~> . ~..; ..~ .:~ ,, ~ . - . . . - , ~ .~, . ;, '~.

TABLE ""'3 ~ , .il;:. i! ~ " ~ '~:' ',


WATE •, ~ ~ ~ . t . ~ ,,'::.

• N . . . co. / zo.
(soluble corrkoonent~)

(b) ~ . _~DNAPL

AIR OR WATER-
DNAPL RELEASE FILLED PORE SPACE

RESIDUAL DNAPL
, ..~\\\~...,,'~'-'-'-~," SINK,NG VAPORS TOP OF
/"
. . :.' ". .
• • ....
"o" "'.~.Q~l=w,~
i~-:.:'~
. . . . ; . ~ (~.,:=....,
~ CAPILLARY FRINGE
~ ' ~ . i ....... .. WATER TABLE

~i~!j::.:.";"'DI's'S O (.~ E'~)"'""::;!::~


"--'.
"" CHE M ICAL "- '""- ':"""
:~;:.':~'.;; PLUME ,.~;::!.;" =
, 4 , G R O U N I~WAT E R'." .¢.'": :'.~:'..!";;!.~:!!'; / LOW E R
FLOW .=,"::: •, -~""."j".:'::::i:~.:,i:~'!!:'.':::::::: 1.... / PERMEABILITY
STRATA

DNAPL
WATER-FILLE
PORE SPACE

Fig. 1. a. LNAPL infiltration schematic (modified from Pinder and Abriola, 1986).
b. DNAPL infiltration schematic (modified from Feenstra and Cherry, 1988).

barrier layer upon which it may continue to flow under pressure and gravity
forces. As in the vadose zone, some of the NAPL will be held in the pore space
within the saturated zone. This residual NAPL will serve as a chemical source
to the flowing groundwater depending on the aqueous solubility of the organic
compounds. In the vadose zone, infiltrating rainwater may dissolve organic
vapors or the residual NAPL and transport these organic components to the
saturated region.
As indicated, many of the processes associated with NAPL movement are
understood conceptually. In addition, considerable effort recently has been
devoted to NAPL studies. The purpose of this paper, therefore, is to review the
A REVIEW OF IMMISCIBLE FLUIDS IN THE SUBSURFACE ill

work performed on NAPL's. To establish a basic foundation, properties


generally associated with NAPL's are discussed and tabulated values provided.
A brief review of the mathematical equations that describe NAPL flow is
presented. Because of the importance of data acquisition, a discussion is given
concerning field techniques used to characterize NAPL movement. Finally,
technologies considered and used to remediate NAPL contamination are
presented. Throughout the discussion, the importance of NAPL properties is
emphasized.

PROPERTIES OF FLUID AND MEDIA

Organic contaminants can reach the water table either through vapor
transport, dissolved in water, or as NAPL. The subsurface transport of NAPL
is governed by various factors, some of which are different from those for
dissolved (or miscible) contaminants. In this section, properties associated
with NAPL flow are discussed. See the Notation for symbols used in this paper.

Saturation

The saturation, s, of a fluid is the volume fraction of the total void volume
occupied by t h a t fluid. Saturations vary from zero to one and the saturations
of all fluids sum to one. Saturation is important because other properties, such
as capillary pressure and relative permeability, are represented as functions of
saturation.
Measuring saturation is difficult. Ferrand et al. (1989) present a dual-gamma
(~37Cs-241Am) technique for laboratory determination of three-fluid saturation
profiles in porous media. They provide porosity and saturation profiles for
sands containing air, water, and trichlo~'oethene or tetrachloroethene. Drilling
and sampling, pumping tests, and borehole geophysical logging can be used to
facilitate field estimates of saturation. These estimates are qualitative add the
methods used are largely undocumented.

Interracial tension

Liquid interfacial tension is equal to the free st.rface energy at the interface
formed between two immiscible or nearly immiscible liquids (Villaume, 1985).
It results from the difference between the mutual attraction of like molecules
within each fluid and the attraction ef dissimilar molecules across the fluid
interface (Schowalter, 1979). Liquid interfacial tension is directly related to the
capillary pressure across an N A P L - w a t e r interface and is a factor controlling
wettability.
The interfacial tension between a liquid and its own vapor is called vapor
tension or surface tension. Surface tension is responsible for capillary effects.
The magnitude of the liquid interfacial tension is always less than the larger
of the surface tensions for the pure liquids. This is due to the mutual attraction
of unlike molecules at the immiscible liquid i~lterface.
112 J.W. MERCER AND R.M. COHEN

Measured in units of energy per unit area, interfacial tension decreases with
increasing temperature ( ~ 5.5"10- ~ dyn cm-~ °C-~ for crude oil-water systems)
and may be affected by pH, surface-active agents and gas in solution
(Schowalter, 1979). Interfacial tensions for N A P L - w a t e r range from zero, for
completely miscible liquids, to 72 dyn cm-1, the surface tension of water at
25 °C (Lyman et al., 1982). Values of interfacial and surface tensions for NAPL-
forming chemicals are listed in Appendix B and generally range between 15 and
50 dyn cm 1. In general, larger surface tensions result in higher capillary
pressure, which may produce larger residual saturations (see subsequent
section).

Wettability

Wettability describes the preferential spreading of one fluid over solid


surfaces in a two-fluid system; it depends on interfacial tension. Whereas the
wetting fluid will tend to coat the surface of grains and occupy smaller spaces
(i.e. pore throats) in porous media, the nonwetting fluid will tend to be con-
stricted to the largest openings. W.G. Anderson (1986a, b, c, 1987a, b, c) recently
prepared a comprehensive literature review on wettability, its measurement,
and effects on capillary pressure, relative permeability, residual NAPL
saturation and enhanced NAPL recovery.
A measure of wettability is the contact angle at the fluid-solid interface (Fig.
2). For two fluids, such as NAPL and water, in contact with a solid, Young's
equation describes the contact angle of the interface:
cos ~b = (a~s - aw~)/aNw (1)
where aNs is the interracial tension between NAPL and solid; aws is the in-
terracial tension between water and solid; aN~ is the interracial tension
between NAPL and water; and ~ is tLe contact angle measured into the water
(in degrees). The contact angle indicates whether the porous medium will be
preferentially wetted by NAPL or water and may vary between 0 and 180°. If
~b < 70° , the system is water-wet; if ~b > 110°, it is NAPL-wet; and if
= 70° - 110°, it is considered neutral (W.G. Anderson, 1986a). Methods for
measuring contact angles are described by Gould (1964), W.G. Anderson (1986b)
and Honarpour et al. (1986).
With the exception of mercury, liquids (NAPL or water), rather than air,
preferentially wet solid surfaces in the vadose zone. Wettability relations in
NAPL-water systems are affected by several factors including medium
mineralogy, NAPL chemistry, water chemistry, the presence of surfactants or
organic matter, and medium saturation history. With the exception of organic
matter (such as coal, peat and humus), graphite, sulfur, talc and talc-like
silicates, and many sulfides, most natural porous media are strongly water-wet
if not contaminated by NAPL (W.G. Anderson, 1986a). Although water is often
the wetting fluid in N A P L - w a t e r systems and has been considered a perfect
wetting agent in certain petroleum reservoirs (Smith, 1966; Berg, 1975;
A R E V I E W O F IMMISCIBLE FLUIDS IN T H E S U B S U R F A C E 113
WAT ER - W E T
q)<~70 °

r///~///A

NAPL-WET

(0>110 °

On
. ~ NA
uoY'/.,,'//////L//'/A

Fig. 2. Wettability configurations.

Schowalter, 1979; Corey, 1986), other researchers have documented that


petroleum reservoirs, particularly limestone and dolomite, may be partially or
preferentially wet by oil (Nutting, 1934; Benner and Bartell, 1941; Leach et al.,
1962; Craig, 1971; Treiber et al., 1972; Salathiel, 1973).
NAPL wetting usually increases due to adsorption and/or deposition on
mineral surfaces of organic matter and surfactants derived from NAPL or
w~,ter (Treiber et al., 1972; Schowalter, 1979; JBF Scientific Corp., 1981;
Thomas, 1982; Honarpour et al., 1986). NAPL wetting has been shown to
increase with aging during contact angle studies (Craig, 1971; JBF Scientific
Corp., 1981), presumably due to mineral surface chemistry modifications
induced by NAPL presence. Similarly, a hysteresis effect has been documented
in which the contact angle is less when NAPL advances over an initially
water-saturated medium than when NAPL is receding from an NAPL-contami-
nated medium (Villaume, 1985).
Given the heterogeneous nature of subsurface media and the factors that
influence wettability, some investigators have concluded that the wetting of
porous media by NAPL can be heterogeneous, or fractional, rather than
uniform (W.G. Anderson, 1986a; Honarpour et al., 1986). Unfortunately, few
wettabiliW studies have been conducted on non-petroleum NAPL's. Results of
contact angle experiments using several DNAPL's and various substrates are
provided in Table 1 (Arthur D. Little, Inc., 1981).

Capillary pressure
Capillary pressure is a property that causes porous media to draw in the
wetting fluid and repel the nonwetting fluid (Bear, 1972). If capillary pressure
114 J.W. MERCER AND R.M. COHEN

TABLE 1

Results of contact angle experiments conducted using DNAPL's by Arthur D. Little, Inc. (1981)

Immiscible Fluid Substrate Medium Contact


angle
(°)

Tetrachloroethene clay APL 23-48


Tetrachloroethene clay air 153--168
1,2,4-Trichlorobenzene clay APL 28--38
1,2,4 -Trichlorobenzene clay air 153
Hexachlorobutadiene clay water 32-48
Hexachlorocyclopentadiene clay water 32-41
2,6.Dichlorotoluene clay water 30-38
4-Chlorobenzotrifluoride clay water 30-52
Carbon tetrachloride clay water 27-31
Chlorobenzene clay water 27-34
Chloroform clay water 29-31
S-Area DNAPL clay APL 21-54
S-Area DNAPL clay water 20-37
S-Area DNAPL clay air 170-171
S-Area DNAPL fine sand and silt water 30-40
S-Area DNAPL clayey till (30-40 % clay) water 20-37
S-Area DNAPL Ottawa fine to coarse sand water 33-50
Tetrachloroethene Ottawa fine to coarse sand water 33-45
Tetrachloroethene Lockport Dolomite water 16-21
Tetrachloroethene Lockport Dolomite air 171
S-Area DNAPL Lockport Dolomite water 16-19
S-Area DNAPL Lockport Dolomite air 164-169
S-Area DNAPL NAPL-contaminated fine sand APL 45-105
S-Area DNAPL soils with vegetative matter water 50-122
S-Area DNAPL paper water 31
S-Area DNAPL wood water 34-37
S-Area DNAPL cotton clot~ water 31-33
S-Area DNAPL stainless s,~, : water 131-154
S-Area DNAPL clay water (SA) 25-54
S-Area DNAPL with solvents clay water 15-45

Adsorbed S-Area (New York, U.S.A.) chemicals were detected on some of the clay samples. APL
refers to aqueous phase liquids (water containing dissolved chemicals). S-Area DNAPL is com-
prised primarily of tetrachlorobenzene, trichlorobenzenes, tetrachloroethene, hexachlorocy-
clopentadiene, and octachlorocyclopentene. SA refers to surface-active agents (Tide ® and Alco-
nox ®) which were added to the water.

is assumed positive, it is defined as the difference between the nonwetting fluid


pressure and the wetting fluid pressure. For a water-NAPL system with water
being the wetting phase, capillary pressure, Pc, is defined as:
Pc - P N - Pw (2)
where PN is the NAPL pressure; and Pw is the water pressure.
Capillary pressure is related, to interfacial tension, contact angle, and pore
size by (Bear, 1979):
A REVIEW O F I M M I S C I B L E F L U I D S IN T H E S U B S U R F A C E 115

Liquid Interfacial Tension : 15 D y n e s / C r n Liquid Interfacial Tension : 3 0 D y n e s / C r n


H¢ (cm water) Hc (cm water)
10OO~ 1000 L
k Con[acl Angles
r...~'~_ - - CA - 0 ~.grees ~ - . . . : . -.~ , ~ I _ _ r ~ . 0 =,~,ees
,oo ~ --.. : : ~ .... cA. ,o ~.~,.,

'°Ff -

o.o01 O.Ol o.1 1 o.ool O.Ol o.1 1


Pore R a d i u s (turn) Pore R a d i u s (ram)

Liquid Interracial Tension : 4 5 D y n e s / C m Liquid Interfacial Tension : 6 0 D y n e s / C m

Hc (cm water) Hc (cm water)


IO~DO~ , . " 'L 1000 IE ~ •

I I E
L ".-:..~... I- ~°°'~"" I F-.. " - - : . - . " ~ i - - ~-o~,~ ....
,oo~ --.: ~ I .... cA.,0o,~,., ! ,oo i ~ --.::.-~..... ~ .... ~.,o0,0,~,

.... J I

o.ool o.ol o.1 1 o.ool o.ol o.1 1


Pore R a d i u s (mm) Pore R a d i u s (mrn)

Fig. 3. Capillary pressure as a function of liqmd interfacial tension and contact angle ( C A ) .

Pc = (2a cos ¢)/r (3)


where r is the radius of the water-filled pore that the NAPL must enter; and a
is the interfacial tension between NAPL and water with the subscripts
dropped. Eq. 3 is valid only for interfaces that form subsections of a sphere.
Capillary pressure increases as r and ¢ decrease and as a increases. Capillary
pressure affects the shape of a spill in the vadose zone. While the shape of a spill
is influenced by many factors, the horizontal component is due largely to
capillarity (de Pastrovich et al., 1979) and medium spatial variability.
The capillary pressure t h a t must be overcome for a nonwetting NAPL to
enter w a t e r - s a t u r a t e d media is known as the threshold or displacement entry
pressure. Because capillary forces can restrict the movement of NAPL into
w a t e r - s a t u rat ed media, layers with small r can serve as capillary barriers. That
i~, before an N A P L can p e n e t r a t e into a water-saturated porous medium, the
NAPL pressure head must exceed the resistance of the capillary forces (e.g.,
Schwille, 1988). The threshold entry pressure for NAPL migration can be
estimated as an equivalent head of wetting fluid (water) by the following
equation:
116 J.W. M E R C E R AND R.M. C O H E N

0 ~,,,I

,,-4

H
o
a
I
II ....., I ~

I 0 I
0
o
e~

.2 c~
II
H
H H
a ,o
,el

°~

, 0
"~
"~ "" ~ ~

Z~
"~I ~
o o ~,

,.., ~'

• ,.~

"'~ I~, ~,

,,'4

o o
o
<
Z
I I
$
r~

]P
A
~1~
I
I

~ ~
I I

~
+
I

o'~

~1~ ~ ~ ~
H

::k

,-.< ::L

II II II
t~ H
+
II

~.~ ~ -
= ~ ~
,.~o ~0,-~ =

t~
~o o ~ ~
~ . ~ o °

• .~ i~! ~ ~ ~ ~<~:. o=~ : ' ~ i~.~'~


118 J.W. MERCER AND R.M. COHEN

hc = (2a cos ¢)/(r p~ g) (4)


where hc is the capillary rise of the wetting fluid; Pw is the density of water; and
g is the gravitational constant. This equation is solved for a range of a-, ~b- and
r-values in Fig. 3. Several related useful threshold entry pressure relationships
are given in Table 2. Equations 3, 4, and those in Table 2 are approximations
for interfaces in porous media. To determine entry pressure more accurately,
it must be measured.
Laboratory experiments show that capillary pressure can be represented as
a function of saturation (e.g., Thomas, 1982). Changes in capillary pressure
with saturation depend on whether the medium is undergoing wetting
(imbibition) or draining of the wetting fluid. Capillary hysteresis is related to
trapped nonwetting fluid and to the differences in the contact angles with
wetting and draining, and causes different wetting and drying curves to be
followed depending on the previous wetting/drying history. During drainage,
the larger pores drain quickly while the smaller pores drain reluctantly, if at
all. This capillary retention explains why capillary pressure corresponds to
higher saturations on the dr~ inage curve. On wetting, the smaller pores fill first
and the larger ones are least likely to fill. This condition leads to a lower
capillary pressure curve with saturation (Thomas, 1982).
Lenhard and Parker (1987b) measured saturation-capillary pressure
relations for benzene, o-xylene, p-cymene and benzyl alcohol in a sandy porous
medium. As part of this work, they evaluated a scaling procedure (Parker et al.,
1987) applied to s-Pc relations of two-phase ~ir-water, air-NAPL and N A P L -
water porous media systems. Relatively good fits were obtained by fitting the
experimental data to multifluid versions of the Brooks and Corey (1964) and
van Genuchten (1980) retention functions. They concluded that using scaling
factors from interfacial tension data permits prediction of s-Pc relations for
any two-phase fluid system in a porous medium from measurement of a single
two-phase system and appropriate interfacial tension data. Based on the
assumptions of Leverett (1941)~ three-phase system behavior may also be
predicted.

Residual saturation

Residual saturation (Sr) of NAPL is the saturation (VNAPL/Vvoids)at which


NAPL becomes discontinuous and is immobilized by capillary forces under
ambient groundwater flow conditions. The physics of oil entrapment and de-
velopment of methods to reduce residual saturation by enhanced oil recovery
are of obvious importance to the petroleum industry (Melrose and Brander,
1974; Chatzis et al., 1983; McCaffery and Batycky, 1983; W.G. Anderson, 1987c;
Mohanty et al., 1987; Chatzis et al., 1988; Morrow et al., 1988; Wang, 1988).
Similarly, residual saturation of NAPL has important consequences in the
remediation of subsurface contamination. Residual NAPL's may provide a
lasting source of significant groundwater contamination because drinking
A R E V I E W O F IMMISCIBLE FLUIDS IN T H E S U B S U R F A C E 119

water standards for many NAPL's are orders of magnitude less than their
solubility limits. Combined with practical limitations on residual NAPL
recovery (J.L. Wilson and Conrad, 1984), NAPL dissolution necessitates
perpetual hydraulic containment at some contamination sites (Cohen et al.,
1987; Mackay and Cherry, 1989).
Residual saturation results from capillary forces and depends on several
factors, including: (1) the medium pore size distribution; (2) wettability; (3) fluid
viscosity ratio and density ratio; (4) interracial surface tension; (5) gravity]
buoyancy forces; and (6) hydraulic gradients. Residual saturation for the
wetting fluid is conceptually different from that for the nonwetting fluid. The
nonwetting fluid is discontinuous at residual saturation, whereas the wetting
fluid is not. Field-scale values of sr are difficult to measure or estimate
accurately and are subject to considerable error.
In the vadose zone, NAPL is retained as wetting pendular rings and as
nonwetting blobs in the presence of water (e.g., Cary et al., 1989). The ability
of the vadose zone to trap NAPL is sometimes measured and reported as the
volumetric retention capacity:
R = sr × (porosity) × 1000
where R is liters of residual NAPL per cubic meter of medium (de Pastrovich
et al., 1979; Schwille, 1984; J.L. Wilson and Conrad, 1984). Residual saturation
measurements involving various NAPL's and media compiled in Table 3
indicate that st-values typically range from 0.10 to 0.20 in the vadose zone. In
general, residual saturation and retention capacity values in the vadose zone
increase with decreasing intrinsic per_m_eability~ effective porosity and
moisture content (Fussell et al., 1981; Schwille, 1984; Hoag and Marley, 1986;
M.R. Anderson, 1958).
Usually more NAPL is immobilized in the saturated zone than in the vadose
zone because: (1) the fluid density ratio (NAPL/air vs. NAPL/water above and
below the water table, respectively) favors greater drainage in the vadose zone;
(2) as the nonwetting ~tuid in m~st saturated media, NAPL is trapped in the
larger pores; and (3) as the wetting fluid (with respect to air) in the vadose zone,
NAPL tends to spread into adjacent pores and leave a lower residual content
behind, a process that is inhibited in the saturated zone where NAPL is usually
the nonwetting fluid (M.R. Anderson, 1988). Values of sr in saturated media
generally range from 0.15 to 0.50 (Table 3).
On the pore scale, residual NAPL below the water table is immobilized as a
result of snap-off and bypassing mechanisms (Chatzis et al., 1983). Snap-off is
prevalent in high aspect ratio pores wherc the pore body is much larger than
the pore throat, resulting in single droplets or blobs of residual NAPL.
Bypassing occurs when wetting fluid flow disconnects the nonwetting fluid,
causing NAPL ganglia to be trapped in clusters of large pores surrounded by
smaller pores. Residual saturation is thought to depend on soil pore size
distribution in the vadose zone, but depends more on pore aspect ratio in the
saturated zone (J.L. Wilson and Conrad, 1984). Residual saturation tends to
120 J.W. M E R C E R A N D R.M. C O H E N

~
o

3
II II II II II II II II II II II II II II H H II II II II II II II II II II II II II II II II II

0 ~ "~ ~ 0~

~ ~ ~ ~ .~ ~o

"0 ~ 0 0

.o

"~ .~ ~
q~ rll ~ ~rj

~d
A REVIEW OF IMMISCIBLE F L U I D S IN T H E S U B S U R F A C E 121

~ ~ ~ ~ . ~ ~ .

II II II II II II II II II II II II II II II II II II A A A A II II II II II II II II II II II
~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ~ ' ~ ~ ~ ~ oo
co o~ o0

° ~

°~

° ~

II II , ~

m ;~ moo m m ~ ~" ~ ~ m

ii It

oo
~J

o o ~
.,-i

o o o o o o o o o o o o o , - - ,-~ ,-~

H
H
~ ~ ~ ~ ~ ~ ~ z Z
~NNNNN~!NNNa. a . ~ ~ ~ m ~
122 J.W. M E R C E R A N D R.M. C O H E N

increase with increasing pore aspect ratios and pore size heterogeneity
(Chatzis et al., 1983), and with decreasing porosity, probably due to reduced
pore connectivity and a ~lecrease in mobile nonwetting fluid in smaller pore
throats (J.L. Wilson and Conrad, 1984). Residual saturation is reduced in
near-neutral wettability media because the capillary forces that trap NAPL are
minimized (W.G. Anderson, 1987c).
Some residual NAPL at contamination sites can be mobilized by increasing
the prevailing hydraulic gradient or reducing interfacial tension. Dissolution
and volatilization cause the residual saturation to decrease with time. These
topics are discussed in more detail in subsequent sections.

Relative permeability

When more than one fluid exists in a porous medium, the flowing fluids
compete for pore space. The net result is that the mobility is reduced for each
fluid. The reduction can be quantified by multiplying the intrinsic permeability
by a dimensionless ratio, known as relative permeability. Relative permeability
is the ratio of the effective permeability of a fluid at a fixed saturation to the
intrinsic permeability. Relative permeability varies from zero to one. Like
capillary pressure, relative permeability can be represented as a function of
saturation and also can exhibit hysteresis. An example of relative permeabili-
ties in a water-oil system is shown in Fig. 4a. At residual saturation, the
respective relative permeability becomes zero; that is, flow ceases to occur.
Setting the relative permeability of the wetting fluid to zero at residual
saturation is an approximation, whereas the relative permeability of the
nonwetting fluid is zero. Similar curves are obtained for air-NAPL systems.
Extensive reviews of two- and three-phase relative permeability data, measure-
ment methods, and controlling factors are presented by Saraf and McCafferty
(1982) and Honarpour et al. (1986).
If water, NAPL and air are flowing simultaneously at a point, three-phase
relative permeabilities are required to describe the movement of each. The
functional dependence of relative permeabilities is based on experiments
(Corey et al., 1956; Snell, 1962). Unfortunately, given the expense and difficulty
of measurement, actual site-specific data and the functional form of three-phase
relative permeability are generally not available, particularly for NAPL's
other than petroleum. As a result, theoretical models have been developed to
characterize three-phase relative permeability (Stone, 1970, 1973; Dietrich and
Bonder, 1976; Fayers and Matthews, 1984; Parker et al., 1987; Delshad and
Pope, 1989). For example, Stone (1973) has proposed a model for estimating
three-fluid relative permeabilities based on data for two-fluid relative per-
meabilities. In this approach, relative permeability data for NAPL are obtained
in both water-NAPL and air-NAPL systems. The relative permeability of
NAPL in the three-fluid system is determined from the following equation:
krn = k*rnw [(krn./k*rnw ~- krw)(krna/k*rnw + kr a) - (krw + kra)] (5)
A REVIEW OF IMMISCIBLE FLUIDS IN THE SUBSURFACE 123

Sa=l
(a) 100 I I I I I I I I I (b)

8O

a.
70 n

>:
~- 60
NAPL
u,i 50-
,1-
u,i
o. 40
uJ
I,-
30-
Sw = 1 Sn = 1
20-

S a = AIR S A T U R A T I O N
10-

S w = WATER SATURA11ON
I I . . . . . . I 1
O0 10 20 30 40 50 60 70 80 90 100
S n = NONAQUEOUS P H A S E
WATER SATURATION, percent SATURATION

Fig. 4. a. Water-NAPL relative permeability (modified from Honarpour and Mahmood, 1988).
b. Ternary diagram showing the relative permeability of the nonaqueous phase as a function of
phase saturations (Faust, 1985).

where k * * is the relative permeability of NAPL at the residual saturation of


water in the water-NAPL system; kr,, is the relative permeability of NAPL in
the water--NAPL system (a function of water saturation); and k~a is the relative
permeability of NAPL in the air-NAPL system (a function of air saturation).
This equation may be used to construct a ternary diagram, as presented in Fig.
4b. Delshad and Pope (1989) recently evaluated seven different three-phase
relative permeability models by comparing predicted permeabilities with three
sets of experimental data.
Relative permeability data are generally not available for NAPL's found in
hazardous waste sites. However, Lin et al. (1982) made laboratory measure-
ments of pressure-saturation relations for water-air and trichloroethene
(TCE)-air systems in homogeneous sand columns. These data were converted
to two-fluid saturation-relative permeability data by Abriola (1983) using
Mualem's (1976) theory. Other pressure-saturati°n data for tetrachloroethene
are provided in Kueper et al. (1989).

Solubility
The aqueous solubility of a chemical is the maximum concentration of the
chemical that will dissolve in pure water at a particular temperature. Results
of laboratory dissolution experiments (M.R. Anderson, 1988; Schwille, 1988)
show that concentrations approximately equal to the aqueous solubility of the
compound are obtained in water flowing at 10-100cm d a y 1 through NAPL-
124 J.W. MERCER AND R.M. COHEN

contaminated sands. According to Mackay et al. (1985), however, organic


compounds are commonly found in groundwater at concentrations of < 10% of
NAPL solubility limits, even when NAPL is known or suspected to be present.
This discrepancy between laboratory and field measurements is probably
caused by diffusional limitations of dissolution in conjunction with heteroge-
neous field conditions, such as non-uniform groundwater flow, variable NAPL
distribution, and mixing of stratified groundwater in a well (Mackay et al.,
1985; Feenstra and Cherry, 1988). These processes favor the creation of large
plumes of groundwater with low chemical concentrations that may greatly
exceed acceptable levels. Another factor to consider is that dissolved chemical
concentrations will also be less than aqueous solubilities reported for pure
chemicals where the NAPL is composed of multiple liquids. For this case, the
actual aqueous solubility of a particular component of the multi-liquid NAPL
can be approximated by multiplying the mole fraction of the chemical in the
NAPL by its pure form aqueous solubility (Banerjee, 1984).
NAPL's vary widely in their aqueous solubility (Appendix A). Nonpolar
hydrophobic compounds are less soluble than polar hydrophilic compounds.
Solubilities may be measured experimentally or estimated based on empirical
relationships developed between solubility and other chemical properties such
as partition coefficients and molecular structure. For example, Kenaga and
Goring (1980) and Lyman et al. (1982) present numerous regression equations
that correlate aqueous solubility with Kow (octanol/water) and Koc (organic
carbon/water) partition coefficients for different chemical groups. Kow and Koc
data for NAPL's are given in Appendix A. Nirmalakhandan and Speece (1988)
developed a predictive equation for aqueous solubility based on correlations
between the solubilities and molecular structures of 200 environmentally
relevant chemicals. Organic concentrations in water can also be calculated
from an equilibrium relationship based on Raoult's law and Henry's law (Cor-
apcioglu and Baehr, 1987).
Several factors influence solubility, including temperature, cosolvents,
salinity and dissolved organic matter. Although the aqueous solubility of most
organic chemicals increases with temperature, the direction and magnitude of
the temperature-solubility relationship are variable (Lyman et al., 1982).
Similarly, the effect of cosolvents (multiple organic compounds) on chemical
solubility depends on the specific mix of compounds and concentrations.
Banerjee (1984) and Groves (1988) describe methods for predicting the solubili-
ties of organic chemical mixtures in water based on activity coefficient
equations. The aqueous solubility of organic chemicals generally declines with
increasing salinity (Eganhouse and Calder, 1976; Rossi and Thomas, 1981).
However, dissolved organic matter, such as naturally occurring fulvic and
humic acids, has been shown to enhance the solubility of hydrophobic organic
compounds in water (Lyman et al., 1982; Chiou et al., 1986).
If NAPL has a range of intermixed components of varying individual solu-
bilities, the more soluble components will dissolve more rapidly and leave
behind a less soluble residue (Senn and Johnson, 1987)~ For a gasoline, this
A REVIEW OF IMMISCIBLE FLUIDS IN THE S U B S U R F A C E 125
TABLE 4

Mass-exchange coefficients* for various hydrocarbons

Hydrocarbons Mass-exchange coefficients, km


(10-3mgm-2s-~)

Gasoline, tar oil 100


Fuel oil, diesel, kerosene 10
Lube oils, heavy fuel oil 1
*Working Group "Water and Petroleum" (1970).

process is referred to as weathering, where the light end compounds are lost to
dissolution and volatilization. Thus, the ratios of chemicals in the NAPL and
dissolved plume change with time.
Subsurface NAPL trapped as ganglia at residual saturation and NAPL
contained in pools, such as LNAPL floating above the water table or DNAPL
in depressions along the base of an aquifer, present long-term sources of
groundwater contamination. Factors controlling NAPL dissolution into
groundwater and its eventual depletion include the aqueous solubility of
NAPL chemicals, groundwater velocity, NAPL-water contact area and the
molecular diffusivity of the NAPL chemicals in water (Pfannkuch, 1984; M.R.
Anderson, 1988; Feenstra and Cherry, 1988; Hunt et al., 1988a; Schwille, 1988).
Various experimental, theoretical and conceptual analyses of these factors and
mass exchange across the NAPL-water interface are discussed below.
Pfannkuch (1984) reviewed literature related to the mass exchange of
petroleum hydrocarbons to groundwater. Laboratory studies of LNAPL
transfer to water were conducted by Hoffmann (1969, 1970), Working Group
"Water and Petroleum" (1970), van der Waarden et al. (1971), Zilliox et al.
(1973, 1974), and Fried et al. (1979). Schwille (1988) and M.R. Anderson (1988)
performed experiments to analyze the mass transfer of DNAPL chemicals into
groundwater. Estimates of mass-exchange coefficients (kin), which represent the
mass of NAPL dissolved into groundwater per unit contact area per unit time,
are given in Table 4. The NAPL-water contact area is difficult to determine
where residual NAPL is trapped as globules in the saturated zone. To describe
mass exchange in these cases, Ziiliox et al. (1973) utilize the volumetric
exchange coefficient (kmv)which equals (dm/dt)/V where m and V are the mass
and volume of the source, respectively.
Experimental data indicate that mass-exchange co,~fficients generally
increase with greundwater ve!nc~ty~ except at low velocities where the
exchange rate is controlled by molecular diffusion, and decrease with time as
the NAPL ages (Zilliox et al., 1973, 1978; Pfannkuch, 1984). However,
Pfannkuch (1984) notes that fluctuations of the water table rejuvenate the
dissolution process (presumably by changing the exposed NAPL surface area)
and effectively increase k~-values for LNAPL trapped below the water table.
The mass-exchange rate (m, T- ~), or strength of the dissolved contamination
source, can be expressed simply as the product of the mass-exchange coefficient
126 J.W. MERCER AND R.M. COHEN

(m~ L-2T - 1) and some measure of the contact area (L2). The contact area of a
given mass of residual NAPL ganglia is more difficult to estimate, but much
greater than that of an equivalent mass of pooled NAPL. Consequently, dis-
solution of residual NAPL produces higher concentrations of NAPL chemicals
in groundwater and depletes the NAPL source more quickly than dissolution
of an NAPL pool of equivalentmass.
Several relationships have been derived for residual and pooled NAPL
geometries to predict dissolved chemical concentrations in groundwater and
the time required for NAPL source depletion (Azbel, 1981; M.R. Anderson, 1988;
Hunt et al., 1988a). Given current limitations characterizing subsurface NAPL
distributions and mass-exchange processes, these models are primarily useful
as a conceptual tool to evaluate the long-term contamination potential
associated with subsurface NAPL. For example, the time required to
completely dissolve an NAPL source given an existing or induced interstitial
groundwater velocity, 0, can be estimated as:
t = m/(O n Cw A )
where A is the cross-sectional area containing NAPL through which ground-
water flow exits with a dissolved NAPL concentration, Cw. Considering limits
to solubility and groundwater velocities, it is apparent that dissolution is an
ineffective removal process for significant quantities of many NAPL's.

Volatilization

Volatilization refers to the transfer of matter from liquid and soil to the
gaseous phase. Thus, chemicals in the soil gas may indicate the presence of
NAPL or dissolved chemicals. Chemical properties affecting volatilization are
vapor pressure (Appendix A) and solubility in water. Other factors affecting
volatilization rate include: concentration in the soil, soil moisture content, soil
air movement, sorptive and diffusion characteristics of the soil, temperature,
and bulk properties of the soil such as organic carbon content, porosity, density
and clay content (Lyman et al., I982). For example, Zytner et al. (1989) observed
that the greater the organic carbon content, the higher the tetrachloroethene
(PCE) sorption on soil and, consequently, the slower the volatilization rate for
both aqueous and pure PCE. On the other hand, volatilization increases with
soil air movement. Volatilization losses in the subsurface from NAPL are
expected where NAPL exists close to the ground surface or in dry pervious
sandy soils, or where NAPL has a very high vapor pressure (Feenstra and
Cherry, 1988).
Once a chemical volatilizes and becomes part of the gas phase, it is trans-
ported and ultimately condenses, is sorbed onto soil particles, degraded, or is
released to the atmosphere. In the case of flammable organics, volatilization in
soil can result in a fire or explosion hazard (Fussell et al., 1981). The vapor can
also adsorb to the soil. Using kerosene experiments, Acher et al. (1989) found
that adsorption of vapor decreased with increasing soil moisture content.
A REVIEW OF IMMISCIBLE FLUIgS IN T H E S U B S U R F A C E 127

Estimating volatilization from soil involves two steps: (1) estimating the
organic partitioning between water and air, and NAPL and air, and (2)
estimating the vapor transport from the soil. Henry's law is commonly used to
determine the partitioning between water and air, whereas the partitioning
between an NAPL and air is described by Raoult's law. Vapor transport in the
soil is usually described by the diffusion equation and several approximate
methods have been developed where the main transport mechanism is macro-
scopic diffusion (Lyman et al., 1982). More complicated models are also
available (e.g., Falta et al., 1989; Sleep and Sykes, 1989; Jury et al., 1990).
Because a chemical can volatilize from either a dissolved state or from
NAPL, both conditions need to be considered to characterize the total amount
of the chemical that is volatilized. In general, local equilibrium is assumed
between the air and other fluids. Henry's law relates the concentration of a
dissolved chemical in water to the partial pressure of the chemical in the gas:
P = KH Cw (6)
where P is the partial pressure of the chemical in the gas phase (atm.); Cw is the
concentration of the chemical in water (molm-3); and KH is Henry's law
constant (atm. m 3mol-1). Henry's law is obeyed for sparingly soluble, non-elec-
trolytes where the gas phase is considered ideal (Noggle, 1985). Henry's law
constants for NAPL compounds are given in Appendix A. The tendency of a
chemical to volatilize increases with an increase in Henry's law constant.
Raoult's law has been used to quantify the ideal reference state for the

TABLE 5

Vapor concentration and total gas density data for selected NAPL's at 25° C (from Falta et al., 1989)

Chemical Molecular Vapor Saturated vapor T~tal gas


weight, M pressure concentration density
(gmo1-1 ) (kPa ~a 25°C) (kgm -~) (kgm -~)

Trichloroethene 131.4 9.9 0.52 1.58


Toluene 92.1 3.8 0.14 1.27
Benzene 78.1 12.7 0.40 ! .42
Chloroform 119.4 25.6 1.23 2.11
Tetrachloroethene 165.8 2.5 0.17 1.31
1,1,1 -Trichloroethene 133.4 16.5 0.89 1.87
Ethylbenzene 106.2 1.3 0.06 1.22
Xylene 106.2 1.2 0.05 1.21
Dichloromethane 84.9 58.4 2.00 2.50
1,2-Dichloroethene 96.9 43.5 1.70 2.37
1,2-Dichloroethane 99.0 10.9 0.44 1.48
Chlorobenzene 112.~ 1.6 0.07 1.23
1,1 -Dichloroeth ~me 99.0 30.1 1.20 2.03
Tetrachloromethane 153.8 15.1 0.94 1.93
Air at 1 atm., 25° C 28.6 101.3 n.a. 1.17

n.a. = not applicable.


128 J.W. MERCER AND R.M. COHEN

equilibrium between NAPL and air (Corapcioglu and Baehr, 1987). Raoult’s
law relates the ideal vapor pressure and the relative concentration of a
chemical in solution to its vapor pressure over the solution:
PA = XAPA
where PA is the vapor pressure of the solution; XA is the mole fraction of the
solvent; and Pi is the vapor pressure of the pure solvent.
Volatilization represents a source to vapor transport. Recent studies have
examined the gas-phase advection resulting from gas pressure and gas density
gradients (Falta et al., 1989; Sleep and Sykes, 1989; Mendoza and McAlary,
1990; Hughes et al., 1990; Mendoza and Frind, 1990). Density-driven gas flow
can be an important transport mechanism in the vadose zone that may result
in contamination of the underlying groundwater and significant dissipation of
residual NAPL.
Density-driven gas flow is a function of the gas-phase permeability, the
gas-phase retardation coefficient, and the total gas density which depends on
the NAPL molecular weight and saturated vapor pressure (Falta et al., 1989).
Saturated vapor concentrations and total gas densities calculated for some
common NAPL’s using the ideal gas law and Dalton’s law of partial pressures,
respectively, are given in Table 5. Density-driven gas flow will likely be signifi-
cant where the total gas density exceeds the ambient gas density by > 10 % and
the gas-phase permeability exceeds l*lO- ‘I m2in homogeneous media (Falta et
al., 1989). Dense gas emanating from NAPL in the vadose zone will generally
sink to the water table where it and gas that has volatilized from the saturated
zone will spread outward. Gas migration patterns will be strongly influenced
by subsurface heterogeneities.

Density is the mass per unit volume of a substance. It often is presented in


terms of specific gravity, which is the ratio of a substance’s density to that of
some standard substance, usually water. Density varies as a function of
different parameters, most notably temperature.
According to Mackay et al. (1985), density differences of * 1% influence
fluid movement in the subsurface. In many situations, NAPL densities differ
from water by 10-50 %. The specific gravities of gasoline and other petroleum
distillates may be as low as 0.7 (Appendix B). Halogenated hydrocarbons
generally are more dense than water. Chlorinated aliphatic compounds
containing one and two carbon atoms have specific gravities from 1.2 to 1.5
(Mackay et al., 1985). The densities of selected NAPL’s are given in Appendix
B.
LNAPL’s and DNAPL’s can be referred to as floaters and sinkers, respective-
ly. After a spill or release, LNAPL’s in sufficient quantity, migrate in response
to pressure and elevation head gradients. Thus, knowing the shape of the
water-table surface may help locate LNAPL, which will tend to move toward
A REVIEW OF IMMISCIBLE FLUIDS IN THE SUBSURFACE 129

water-table lows. Downward DNAPL movement below the water table will be
impeded by capillary and/or permeability barriers upon which DNAPL may
flow downdip counter to tae hydraulic gradient (Faust, 1985). Preferential
DNAPL spreading will occur along pathways of least capillary and permeabil-
ity resistance. Although site stratigraphy and structural geology can be helpful
in locating DNAPL, little success has been achieved at some sites in even
locating subsurface DNAPL sources (Mackay and Cherry, 1989), particularly
in fractured media.

Viscosity
Viscosity is the internal friction within a fluid that causes it to resist flow.
Following a spill, a product of low viscosity will penetrate more rapidly into
the soil than a product with higher viscosity. NAPL viscosity can change with
time. For example, fresh crude oils with volatile components become increas-
ingly viscous as they evaporate. Absolute viscosity data for selected NAPL's
are provided in Appendix B.
Also significant is the N A P L - w a t e r viscosity ratio, which is p a ~ of a teim
used in the petroleum industry known as the mobility ratio. In a water flood,
the mobility ratio is defined as the mobility of the displacing fluid (relative
permeability/viscosity for water) divided by the mobility of the displaced fluid
(relative permeability/viscosity for NAPL). Mobility ratios of > 1 favor the
flow of water whereas those < 1 favor the flow of NAPL. In the displacement
process, lower mobility ratios favor NAPL recovery.
When a fluid in a porous medium is displaced by another fluid, the interface
between them may become unstable. This phenomenon, known as fingering,
results, in part, because of the viscosity difference between the two fluids (e.g.,
Chouke et al., 1959; Homsy, 1987). Where viscous fingering starts is also
influenced by heterogeneities. These factors are discussed by Kueper and Frind
(1988). As a result of viscous fingering, NAPL may not occupy the complete
cross-sectional area through which it passes, thus allowing water to flow
through and increase dissolution. In addition, for a given volume of NAPL,
viscous fingering will allow the NAPL to penetrate deeper than it would
without viscous fingering.

M A T H E M A T I C A L DESCRIPTION O F N A F L F L O W

Petroleum reservoir codes for simulating the flow of immiscible fluids have
existed for more than 20 years (e.g., Crichlow, 1977, or Peaceman, 1977); but,
with few exceptions, it only has been in the last few years that these techniques
have been used to examine oil and chemical spill problems. Codes used to
examine NAPL flow are reviewed in Pinder and Abriola (1986) and Abriola
(1988).
Early recognition of NAPL movement in shallow groundwater as a two-fluid
flow phenomenon is attributed to van Dam (1967). Later, several models were
130 J.W. M E R C E R A N D R.M. C O H E N

developed to describe mathematically the flow of NAPL in the subsurface


(Mull, 1969, 1971, 1978; Holzer, 1976; Schiegg, 1977; Dracos, 1978; Hochmuth
and Sunada, 1985). Common to each of these is the assumption of negligible
capillarity (piston-like flow).
Brutsaert ._~973) presents an early code used to examine multifluid well flow
that account: .%r capillarity. The model is radial and based on a finite-differ-
ence approximation. Later, Guswa (1985) developed a one-dimensional
(vertical) finite-difference, two-fluid flow simulator. Faust (1985) extended this
work to accommodate two dimensions as well as a static air phase, a necessary
step to simulate NAPL flow in the vadose zone. A model similar to Faust's
(1985) model, which did not consider an air phase, was applied to the Hyde Park
Landfill, Niagara Falls, New York, U.S.A., by Osborne and Sykes (1986).
Abriola and Pinder (1985a, b) developed a two-dimensional model that also
considers volatilization and dissolution. A similar model is presented in Baehr
and Corapcioglu (1987) and Corapcioglu and Baehr (1987). Subsequently,
Lenhard and Parker (1987a) and Parker and Lenhard (1987) incorporated
hysteretic constitutive relations. More recently, a three-dimensional model
that extends Faust's (1985) model is described in Faust et al. (1989).
Because of previous reviews on NAPL models, a detailed review is not
provided. The basic governing equations are presented for completeness, along
with constitutive relationships that concern many of the properties discussed
in the preceding section. This discussion follows closely that presented by
Abriola (1988).

Mass-balance equations

The equation development begins with the mass-balance equation for


species i in phase ~, where ~ stands for soil, air, water and NAPL, or a subset
of these. A species is defined as a specific chemical that is present in one or more
phases or is considered as a group of chemicals with average characteristics.
The mass-balance equation is written as (Abriola, 1988):

co ) + V- v - V. = + (8)

where v~ is the mass average velocity of the ~ phase; o~ is the mass fraction of
species i in the ~ phase; ~ is the fraction of volume occupied by the • phase; p~
is the intrinsic mass density of the ~ phase; J~ is the non-advective flux of a
species i in the ~ phase; S~ represents the exchange of mass of species i due to
interphase diffusion and/or phase change; R~ represents an external supply of
species i to the ~ phase; and V is the differential operator.
The first term in eq. 8 accounts for the accumulation of the mass of species
i in phase ~. The second term accounts for the movement of mass due to
advection of the phase. Motion due to non-advective effects (such as diffusion
and dispersion) is accounted for by the third term. The first term on the
right-hand side of eq. 8 is a source/sink term due to phase changes. R7
A REVIEW OF IMMISCIBLE FLUIDS IN T H E SUBSURFACE 131

represents the destruction or creation of the species due to chemical or


biological transformations. Eq. 8 is subject to the following constraints:
to~ = 1 and ~e~ = 1 (9a), (9b)
i

which follow from the definition of mass and vohame fraction. Also, when mass
is lost by one phase due to interphase exchange, an equal a m o u n t of mass is
gained by a n o t h e r phase, or:
S~ = 0 (9c)

Based on eq. 8, a mass-balance relationship for a specific phase may be


developed by summing over all species present in the phase or alternatively
summing over all phases.

Immiscible flow equations

A set of equations can be developed from eq. 8 where it is assumed that there
is no mass exchange between phases and no chemical reaction or biological
transformation. Eq. 8 is summed over all species to yield (Abriola, 1988):

~(p~) + V - ( p ' e ~ v ~) = 0 (10)

where use has been made of constraint (9a). The non-advective flux terms,
which deal with relative motion of the species within a phase, also sum to zero.
Eq. 10 has been used to model the migration of pure NAPL's, t h a t is, NAPL's
with one chemical or physical properties t h a t are spatially invariant.
In general, one equation is written for each of the four phases: soil (s), air
(a), water (w) and NAPL (N). If the porous medium is rigid (porosity is invariant
in time), the soil equation is not needed. Similarly, if the gas phase remains at
atmospheric pressure, the gas equation aJso can be eliminated to yield:

n-~ (s~p~) + V " (p~ s~nv ~) = O, • - : (11)

Here n is porosity and s= is the saturation of the ~ phase (~ = ns~). If fluid


compressibilities are neglected, then:

n-~ (s~) + V " ( s ~ n v ~) = O, ~ = w, N (12)

Compositional equations

For the interphase transfer of mass (i.e. the formation of a dissolved plume
or the t r a n s p o r t of organic vapors), balance equations for each species are
written. The species-balance equations are obtained by summing eq. 8 over all
132 J.W. M E R C E R A N D R.M. C O H E N

phases to yield (Abricla, 1983):

~-a,s,w,N (13)
~Lt, l, _I

where constraint (9c) has been incorporated. For nonreactive components, the
right-band side of eq. 13 is zero. The number of equations that are required
depends upon the number of species. If the soil matrix is rigid, the soil species
equation may be eliminated. Solving the governing eqs. 13 yields the distribu-
tion of each fluid and the composition of the fluids in space and time.

Constitutive relations

Assuming it is valid, Darcy's law may be substituted into a system of mass-


balance equations (such as eq. 11 or eq. 13) to yield the equations governing the
multiphase flow of fluids in a porous medium. As an example, consider NAPL
flow in a rigid matrix (Faust et al., 1989):

n-~(s~o ~) - V- O5 (VP~ - o~g) = 0 (14)


P~
Eq. 14 is formulated in the unknowns P~, which are continuous in space.
Porosity n and intrinsic permeability k are assumed known properties of the
matrix. Generally, viscosity (p~), a weak function of pressure, is assumed
constant. Relative permeability is generally considered a function of
saturation. For the incompressible fluid case, p~ is a constant. In general,
however, density will depend on the fluid pressure, P~. Thus, fluid density may
be expanded in terms of fluid pressure by incorporating fl~, the compressibility
of the ~ phase. For slightly compressible fluids, fl~ is essentially constant.
Saturation is considered a known function of capillary pressure, which can
exhibit hysteretic behavior.
The compositional model equations (13) have additional considerations due
to the dependence of properties on composition. For example, viscosity and
density may be functions of composition. There are also two additional
equation terms, those accounting for non-advective flux and chemical reaction~
which must be evaluated. The non-advective flux term is commonly assumed to
have a Fickian form. This is the form used in most existing solute transport
models and accounts for both molecular diffusion and hydrodynamic
dispersion effects (Bear, 1979). There is no general functional form that may be
specified for the reaction terms that appear in eqs. 13. Simple decay rates can
be directly substituted in eqs. 13 or a system of chemical reaction equilibrium
or rate equations can be solved to determine these term~
Finally, for a compositional model, expressions m'e needed to relate mass
fractions of a given species within all the phases. Generally, the assumption of
local equilibrium is used to develop these i'z~lations (e.g., for sorption, see
Valocchi, 1985). Local equilibrium implies that witLin some relatively short
time scale, contiguous phases reach a thermodynamic equilibrium. Thus, the
A REVIEW OF IMMISCIBLE FLUIDS IN THE SUBSURFACE 133

mass fraction of a species in one phase can be related to the mass fractions of
the same species in other phases via partition expressions of the following form
(Abriola, 1988):
o~ = K~~oJ~ (15)
where K~~ is called the partition coefficient of species i between the ~ and fl
phases. In general, partition coefficients are functions of phase compositions
and pressures. These coefficients may be determined from Henry's law
constants (Appendix A) and solubility relations. See, for example, Baehr (1987),
Baehr and Corapcioglu (1987), Corapcioglu and Baehr (1987).
Although NAPL models can, in theory, simulate a variety of problems, the
data required for applications are generally lacking. Heterogeneity influences
NAPL flow and solute transport; however, spatial variability of pore size
(effecting displacement pressures) and intrinsic permeability are rarely suffi-
ciently defined to permit accurate prediction. As pointed out in the sections on
solubility and volatilization, the theoretical description of mass transfer in
porous media has not been adequately developed, and therefore adds additional
uncertainty to any modeling approach. Currently, existing NAPL models can
be used primarily in a conceptualization mode.

FIELD TECHNIQUESUSED TO CHARACTERIZENAPL FLOW


As important as the ability to theoretically describe NAPL flow is the ability
to characterize it in the field. NAPL source locations and volumes are often
puurly defined. Medium heterogeneities promote complex and unpredictable
spreading of NAPL. Preferential NAPL migration in the saturated zone occurs
through permeable pathways (soil and rock fractures, root holes, sandy layers,
etc.) ~ue to capillary pressure arid hydraulic effects. Guidance for conducting
field investigations at NAPL ~ites is given by the A.P.I. (1980, 1989) and
Villaume (1985).
Characterizing the presence, composition and properties of mobile and
residual NAPL is fundamental to determining the nature, extent and rate of
chemical migration during a site assessment. Data collection typically occurs
in phases using many of the same techniques employed at contamination sites
where NAPL is not present. However, special precautions and considerations
must be recognized where NAPL is present. Field techniques used to charac-
terize NAPL transport discussed here include soil gas analysis, soil and rock
sampling, well measurements, and fluid sampling.

Soil gas analysis

For volatile NAPL's, soil gas analysis may be a screening tool for locating
potential sources of contamination and for siting monitoring wells. Volatile
organic compounds (VOC's) include fuel chemicals such as benzene, toluene,
ethylbenzene and xylenes, and solvents such as trichloroethene, tetra-
134 J.W. M E R C E R A N D R.M. C O H E N

chloroethene, dichloroethene, trichloroethane, and Freon®'s. The transfer of


VOC's from groundwater or N A P L via volatilization has already been
discussed. Although this process may be enhanced by water-table fluctuations
(Lappala and Thompson, 1983), equilibrium vapor concentrations are probably
never attained in the field due to rapid vapor diffusion at the water table
(Marrin and Thompson, 1984).
Once VOC's enter the soil gas, they diffuse in response to the chemical
concentration gradient. Volatilization of chemicals with high molecular
weights and saturated vapor concentrations (Table 5) can induce density-
driven gas transport in media with high gas phase permeability (i.e.kg > 1-
10-n m 2 in uniform media) (Falta et al.,1989; Sleep and Sykes, 1989; Mendoza
and Frind, 1990; Mendoza and McAlary, 1990). Density-driven gas flow causes
VOC's to sink and move outward above the water table. Although advective
processes due to density effectsor high vapor pressure gradients may influence
V O C migration, gaseous diffusion is considered the predominant transport
mechanism at most sites (Marrin and Kerfoot, 1988). At steady state, vertical
solute flux is proportional to the air-filledporosity, the V O C diffusion coeffi-
cient, and the gas-phase concentration gradient.
Subsurface geologic heterogeneities, soil porosity, moisture conditions,
V O C concentrations in groundwater and sorption equilibria can significantly
affect V O C gradients in soil gas (Marrin and Thompson, 1987). For example,
false negative interpretations may result from the presence of vapor barriers
(perched groundwater, clay lenses, or irrigated soils) below the gas probe
intake. Thus, sample locations and depths influence the measured vapor con-
centrations.
Several chemical characteristics indicate whether a measurable vapor con-
centration can be detected (Devitt et al., 1987; Marrin, 1988). Ideally,
compounds such as VOC's monitored using soil gas analysis will: (1) be subject
to littleretardation in groundwater; (2) partition significantly from water to
soil gas (Henry's law c,)nstant > 0.0005atm. m3mol-1); (3) have sufficient
vapor pressure to diffuse ~ignificantlyupward in the vadose zone ( > 1.0m m H g
@ 20 °C); (4) be persistent; and (5) be susceptible to detection and quantitation
by affordable analytical techniques.
Grab and passive sampling techniques are used to collect soil gas. During
static grab sampling, samples are collected from a quiescent soil gas sample. In
dynamic grab sampling, samples are collected from moving soil gas as it is
pumped through a hollow probe. Grab samples can be analyzed on site during
the sampling episode or shipped to a laboratory. Passive sampling provides an
integrated measure of V O C concentrations over time. Charcoal sorbents to
trap solutes that diffuse through the soil gases are buried up to one month and
then retrieved for analysis. Soil gas sampling is discussed in more detail in
Devitt et al. (1987).
Although soil gas analysis has been used successfully at m a n y contamina-
tion sites, it can provide misleading results if subsurface conditions are not
understood adequately. Thus, information regarding the location, extent and
A REVIEW OF IMMISCIBLE FLUIDS IN THE SUBSURFACE 135

composition of VOC's in soil and groundwater generated by a soil gas survey


must be confirmed by analysis of soil and fluid samples taken at depth.

Drilling investigations

Borings and monitoring wells are the primary means for evaluating the
distribution of subsurface NAPL. Selection of drilling locations and depths
should be based on available information regarding site hydrogeologic and
contaminant conditions. Care must be exercised to hmit the potential for
fugitive vapor and particulate emissions, damage to subsurface utilities, small
flash fires or explosions, and cross-contamination. In particular, it is important
to prevent downward migration of mobile perched Db~:APL or NAPL-contami-
nated soils that may result from drilling through a ba~:rier layer. One approach
is to work in from the edge of the plume to characteri.~e the site prior to drilling
at the source of contaminants.
Drilling at NAPL-contaminated sites occurs mo~;t frequently using hollow-
stem augers with split-spoon samplers in soils and wireline-coring or rotary
drill rigs in bedrock. In addition to logging soil characteristics such as grain
size distribution, consistency and moisture state, it is important to note visual
observations and olfactory evidence (where consistent with site safety plans)
of NAPL or other chemicals. Inner sections of cores should be examined to
prevent misinterpreting core surfaces that may be contaminated during the
sampling process. The presence and density of NAPL relative to water can be
easily identified in the field by shaking free liquid or NAPL-contaminated soil
in a jar with water and watching for phase separation. Free liquids or NAPL
extracted using centrifugation or other methods can be tested for fluid
properties such as density, interfacial tension and viscosity. Physical or
chemical extraction processes can also be used to estimate NAPL saturation,
and analyses can be made to determine NAPL composition.
Monitoring well design considerations for DNAPL and LNAPL sites are
discussed by Villaume (1985} and the A.P.I. (1989), respectively. Well screens in
LNAPL-monitoring wells should be long enough to ensure that the entire free
LNAPL layer will be within the screened interval during seasonal water-table
fluctuations. Misleading LNAPL measurements will result from a well that is
screened through perched LNAPL in the vadose zone to a deeper water table.
At DNAPL sites, a well that is fully screened from above the DNAPL-water
inmrface to the cop of the underlying stratigraphic barrier is usually appro-
priate. Sandpucks should be coarser than surrounding media to ensure the
movement of NAPL to the well. NAPL can chemically attack well casing,
screen and pumps. The compatibility of NAPL with well materials including
grout and Bentonite @ seals should be considered.

NAPL thickness and elevation measurements

If NAPL is at residual saturation or under negative pressure, it will not flow


136 J.W. M E R C E R A N D R.M. C O H E N

into a monitoring well. Under such circumstances, soil or core samples


showing NAPL or high aqueous concentrations may be the only indication, of
NAPL presence. If the NAPL is mobile, then product thickness, elevation and
composition need to be determined. Therefore, fluid levels must be measured
and NAPL samples obtained.
Measurements of NAPL elevation and thickness in monitoring wells are
typically made using hydrocarbon-detection paste on a steel tape, an interface
probe, or a transparent, bottom-loading bailer. For all methods, care should be
exercised to minimize disturbance of the fluid column during static level
measurements. Hydrocarbon-detection paste applied to a measuring tape
reveals the top of floating hydrocarbons as a wet line and the top of water as
a distinct color change. This method is accurate to ~ ___0.01 ft. ( ~ ± 3 ram), but
cannot be used to measure DNAPL.
Interface probes employ optical and conductivity sensors to distinguish the
air-fluid and NAPL-water interfaces, respectively. These probes are expensive
(typically US $1500-3000) and can be used to measure both LNAPL and DNAPL
to within 0.01-0.10ft. (~ 0.3-3cm) but may produce spurious results in the
presence of conductive NAPL, emulsified NAPL, or viscous NAPL that coats
the sensors. Standard electric-line water-level probes can detect interfaces
between non-conductive NAPL and water, but may fail to identify the interface
between LNAPL and air.
NAPL can also be measured and sampled using a transparent bottom-load-
ing bailer. For LNAPL, the bailer should be long enough so that its top is in
air when its check valve is in water. NAPL rise due to displacement by the
bailer may result in slight overestimation of product thickness in the well. This
influence can be minimized by allowing time for well equilibration after
lowering the bailer.
Various other means are available to estimate NAPL thickness and
elevation in wells. For example, measurements can be made by taking small
fluid samples from specific levels using mechanical discrete-depth (Kemmerer
type) samplers or by measuring DNAPL that coats a weighted string following
its retrieval from a well. The cost of measuring device decontamination should
be cgnsidered when selecting a measurement method, particularly given uncer-
tainties involved in interpreting NAPL thickness and elevation data.

Interpreting well measurements

LNAPL elevations measured in wells that are screened across the air-fluid
interface are generally assumed to indicate the surface of the LNAPL "pool"
or "pancake" in the surrounding formation, although this may not always be
the case (Hampton, 1988). Water-table elevations can be calculated by adding
the product of the density and thickness of LNAPL measured in a well to the
LNAPL-water interface elevation.
Measured LNAPL thickness in wells (hw) typically exceeds corresponding
LNAPL-saturated formation thickness (h0 by a factor between 2 and 10. This
A REVIEW OF I M M I S C I B L E FLUIDS IN THE S U B S U R F A C E 137

occurs because LNAPL perched above the water-saturated zone will flow into
the well and depress the well water level. Generally, the hw/hf ratio increases
with decreasing formation grain size, increasing capillary fringe height, and
increasing LNAPL density.
Much theoretical and experimental research has been conducted to quan-
titatively relate hw to hf (van Dam, 1967; Zilliox and Muntzer, 1975; de
Pastrovich et al., 1979; Blake and Hall, 1984; R.A. Hall et al., 1984; Schiegg,
1985; Abdul et al., 1989; Fiedler, 1989; Farr et al., 1990; Kemblowski and Chiang,
1990; Lenhard and Parker, 1990). A reliable predictive relationship would
facilitate analysis of LNAPL plume geometry and remedial options. Hampton
(1988) and Hampton and Miller (1988) examined this problem by conducting
laboratory experiments and comparing results to various equations proposed
for estimating hf from hw. They conclude that all the equations and methods
investigated lack predictive capabilities, but suggest using a simple equation
proposed by de Pastrovich et al. (1979):
h~/hf ~ Pn/(Pw - Pn)
for making crude estimates of LNAPL-saturated formation thickness.
Many factors confound development of a reliable predictive equation
between hw and hr. Capillary fringe heights are soil-specific and hysteretic with
fluid-level fluctuations. Increasing thicknesses of LNAPL can collapse the
water capillary fringe, causing the hw/hf ratio to decrease unpredictably. The
thickness of LNAPL-saturated soil is ambiguous because this layer is not
completely saturated with product (Hampton and Miller, 1988). Entrapvaent
and mobilization of LNAPL as ,1. ~ water -ol..,1-.-
~,,= ~ , ~ rlse~" and falls, respectively~ and
transient preferential vertical flow of liquids through wells in tight formations
may help explain the inverse relationship between h~ and changes in the
LNAPL-water interface elevation (Kemblowski and Chiang, 1990). Well con-
struction details can also affect the hw/hf ratio, particularly during periods of
fluid level change. As a result, mobile LNAPL volumes cannot be reliably
estimated based on well thickness measurements or bailing tests (A.P.I., :[980;
Hampton and Miller, 1988; Abdul et al., 1989; Testa and Paczkowski, 1989).
Kemblowski and Chiang (1990) suggest that a borehole geophysical method,
such as a dielectric log, may be the most promising approach to determining the
distribution of LNAPL below ground.
Less research has focused on interpreting DNAPL thickness and elevation
measurements in wells. DNAPL thickness and elevation measurements can be
made using wells that are screened downward from above the water-DNAPL
interface and completed to the top of the stratigraphic barrier under the
DNAPL. The elevatiu~, of DNAPL in a well, however, may exceed that in the
formation by a length equivalent to the capillary pressure exerted at the top of
the DNAPL pool (T.V. Adams and Hampton, 1990). If the well screen or casing
extends into the barrier layer, the measured DNAPL thickness will exceed that
in the formation by the length of well below the barrier layer surface. Both the
surface and thickness of measured DNAPL are likely to be erroneous in a well
138 J.W. MERCER AND R.M. COHEN

that connects a DNAPL pool above a stratigraphic barrier to a deeper


permeable formation. As noted previously (p. 135), such a well will cause
DNAPL to short-circuit the barrier layer and contaminate the lower permeable
formation. As a result of these factors, DNAPL presence in wells should be
evaluated in conjunction with evidence of NAPL presence obtained during
drilling.

Sampling NAPL from wells

NAPL samples should be taken after measuring the static elevations and
thicknesses of immiscible fluids in a well. A bottom-loading bailer or
mechanical discrete-depth sampler is often adequate for collecting NAPL
samples. These and other methods can be used to selectively sample NAPL from
multiple specific depths within a well. Well purging is generally not recom-
mended because it will cause mixing of the multiphase fluids (Villaume, 1985).
However, pumping or bailing tests can be used to examinee t_be mobility of
NAPL in the formation and the feasibility of NAPL recovery by pumping.
Advantages and disadvantages of many common sampling methods are
summarized by the A.P.I. (1989). The cost to decontaminate or replace NAPL-
contaminated equipment is usually a major consideration in selecting a
sampling method.

R E M E D I A T I O N OF N A P L C O N T A M I N A T I O N

Before methods for the in situ removal of NAPL from the subsurface can be
successfully implemented, the location and nature of NAPL must be known. If
NAPL contamination is shallow, then excavation may be an economical alter-
native for remediation. If the NAPL is immobile, hydraulic containment is
possible while purging and treating contaminated groundwater from the
dissolved plume. Purge wells and collector trenches effectively remove mobile
LNAPL; recovery wells remove mobile DNAPL less effectively. Partial
residual NAPL recovery may be possible only by using methods such as mobili-
zation, enhanced oil recovery techniques, or vacuum extraction. These NAPL
remediation methods are discussed below.

Product recovery

Experience in the recovery of mobile NAPL is currently limited and pertains


almost exclusively to the recovery of lenses of LNAPL's, most notably
petroleum products floating on the surface of the water table. Blake and Lewis
(1983) provide a summary of the special considerations ar, d procedures
involved in the recovery of LNAPL's. Other references describing recovery
techniques include de Pastrovich et al. (1979), A.P.I. (1980, 1989), and Fussell
et al. (1981). Examples of DNAPL recovery are prc~cided by Villaume et al.
(1983a, b) for coal tar from gravel, and by Ferry et al. (1986) for DNAPL from
fractured rock.
A REVIEW OF IMMISCIBLE FLUIDS IN THE SUBSURFACE 139

Where the w a t e r table is shallow, open or permeable trenches equipped with


pumps can be used. The t r e n c h can be equipped with pumps designed only to
skim the L N A P L as it collects in the trench. An alternative approach is to
pump g r o u n d w a t e r from the t r e n c h to reduce w a t e r levels and create gradients
t h a t will bring the L N A PL into the trench at g r e a t e r rates and a g a i n s t n a t u r a l
gradients.
Wells also may be used to collect mobile LNAPL. The general approach is
to use the gradients established in the cone of depression surrounding a
pumping well to collect the LNAPL. As the L N A P L collects in the cone of
depression, it can be recovered with a pump. This may be accomplished using
one well to establish the cone of depression and a second, n e a r b y well to
recover the L N A P L as it collects. Alternatively, a single well can be used to
both establish the cone of depression and recover the LNAPL.
A single pump can be used to collect the g r o u n d w a t e r and LNAPL. However,
the pump must be positioned n e a r the w a t e r - L N A P L interface to ensure that
both g r o u n d w a t e r and L N A P L are drawn into the well. A float is used to
regulate and maintain the proper drawdown in the well. Alternatively,
separate pumps can be used to drawdown the water table and collect the
LNAPL. These can be located in the same or separate wells. These pumping
a r r a n g e m e n t s also require special automatic controls to regulate the pumps
and drawdowns.
Screens must be positioned to allow L N A P L intake. Consequently, the
screen must extend above the level to which g r o u n d w a t e r will be drawn down
and into the a r e a in which L N A P L will collect. Otherwise, L N A P L will collect
above the screen and not be able to enter the well. To e n h a n c e recovery,
injection wells may be used in combination with the recovery wells to flush
c o n t a m i n a n t s from the vadose zone and increase the hydraulic gradient toward
the recovery wells.
Large drawdowns will increase the flow rate of L N A P L toward the well, thus
increasing the speed of recovery. Unfortunately, they also will create larger
dewatered a r e a s into which the L N A P L will collect. These areas will have been
previously u n c o n t a m i n a t e d by LNAPL. Substantial amounts of L N A P L will be
retained in the aquifer matrix at residual saturation, even after water levels
are allowed to recover (Blake and Lewis, 1983). Thus, recovery systems should
maintain recovery while minimizing drawdown.
Procedures for mobile D N A P L cleanup are experimental and are
documented poorly compared to procedures for L N A P L cleanup. Removing
DNAPL from the subsurface is extremely difficult. Problems associated with
DNAPL recovery are described by Feenstra and Cherry (1988) and M a c k a y and
Cherry (1989).
The properties that control D N AP L movement are largely responsible for
the difficulties associated with their cleanup. D N A P L movement is influenced
not only by pressure gradients but also by gravity. Thus, to move DNAPL's
toward a well, gradients must be created t h a t will overcome gravity. In cases
where D N A P L ' s are migrating along an inclined confining bed, the required
140 J.W. MERCER AND R.M. COHEN

10 1 I I I

=13 x10.3

0"=10 (dyne cm-1)


\
J o.1

O.Ol - '\

N~= "10"5 ~ ~k
0.001 I I
lO-e 10-7 lO-S lO-S 10-4 10-3 10-2 I0-I
I GRAVEL
--CLEAN SAN[} ]
s,,.,Y s , , , o - - , k (cm21
Fig. 5. Hydraulic gradient, J, necessary to initiate blob mobilization (at N~~) in soils of various
permeabilities, for hydrocarbons of various interfacial tensions, a. The upper curve represents the
gradient necessary for complete removal of all hydrocarbon (/V~*), with a = 10 dyn cm ~. Note t h a t
10 -scm 2 = 1 darcy (from J.L. Wilson and Conrad, 1984).

gradients may be large. For this reason, it is best to collect DNAPL's from low
points along the bedrock or at some downslope position where the pool is
migrating.
Mobile DNAPL may be recovered using the same basic techniques used in
other recovery programs (e.g., Ferry et al., 1986). If DNAPL is located near the
surface, drains may be capable of recovering these contaminants. Otherwise,
wells are required. In both cases, recovery requires pump intakes placed far
into the DNAPL to ensure that as much DNAPL as possible is collected.
Pumping rates should be used that discourage mixing and the formation of
emulsions at the interface between the DNAPL and groundwater. In most
cases, the groundwater above the pool of DNAPL will be contaminated and
also will require cleanup. This can be accomplished using separate wells or a
single well screen over the entire aquifer equipped with two pumps.

Mobilization

During mobilization, increased hydraulic gradients, resulting in increased


groundwater velocity, cause residual blobs of NAPL to move (J.L. Wilson and
Conrad, 1984). The capillary number, No, the ratio of capillary to viscous forces,
provides a measure of the propensity for NAPL trapping and mobilization. It
is defined as the product of intrinsic permeability, water density, gravitational
A REVIEW OF IMMISCIBLE FLUIDS IN THE SUBSURFACE 141

acceleration constant and hydraulic gradient divided by the interfacial


tension. The critical value, N*, of the capillary number is defined as the value
at which motion of some of the NAPL blobs is initiated. Through experiments
(Chatzis et al., 1983, 1988; Morrow et al., 1988), J.L. Wilson and Conrad (1984)
observed a strong correlation between displacement of residual NAPL and the
capillary number, when the hydraulic gradient was greater than that
producing the critical value of the capillary number. The hydraulic gradient
necessary to initiate blob mobilization for various permeabilities and in-
terfacial tensions is shown in Fig. 5. As may be seen, in very permeable media
(e.g., gravel or coarse sand), it is theoretically possible to obtain sufficient
hydraulic gradients to remove all NAPL blobs. In soils of medium permeability
(e.g., fine to medium sand), some of the residual can be hydraulically removed.
In less permeable media, removal is not possible.

Enhanced oil recovery (EOR) methods

Once mobile NAPL is collected, immobile NAPL remains at residual


saturation. Residual saturation decreases with decreasing capillary pressure;
therefore, enhanced oil recovery (EOR) methods have been suggested for
residual NAPL removal because they reduce the interfaciaI tension and/or
NAPL viscosity b flooding with one of the following: (1) hot water or steam,
(2) carbon dioxide, (3) surfactant, (4) alcohol, (5) alkaline, or (6) polymers. For
further information on these topics, see Shah (1981). If EOR is used to mobilize
NAPL at residual saturation, NAPL flow must be controlled carefully.
Otherwise, previously clean portions of the subsurface may become contami-
nated during remediation.
Thermal methods include hot water flooding and steam flooding (Sale and
Piontek, 1988). These methods decrease residual NAPL by increasing contami-
nant solubility and achieving a more favorable mobility ratio. Contaminant
solubility is increased because water solubility of many organics increases
with increasing water temperature. NAPL viscosity decreases with increasing
temperature causing a corresponding decrease in the mobility ratio. Therefore,
NAPL recovery increascs with decreas'ng viscosity. Steam flooding also
results in vaporization of NAPL, which may allow recovery. A limitation in the
use of thermal methods is that DNAPL may be converted to LNAPL. Therefore,
NAPL initially limited in extent may move through previously uncontaminat-
ed portions of the subsurface. Costs may also be high due to heat loss and the
need to heat large volumes of subsurface materials. Hunt et al. (1988a, b)
examined steam floodinb theory and performed laboratory experiments to
evaluate its suitability for DNAPL recovery. IIo~ ~ver, the effectivenes~ of ~bis
technology in environmental applications is unknown.
Carbon dioxide flooding is an EOR technique that also produces a decreased
mobility ratio. Carbon dioxide is injected under pressure and the viscosity of
the NAPL decreases as carbon dioxide dissolves into the NAPL. Because of the
high pressures, this method is applicable only at relatively large depths in a
142 J.W. MERCER AND R.M. COHEN

confined layer. The effectiveness of this technique in environmental applica-


tions is not known (Sale and Piontek, 1988).
Soil flushing with surfactant solutions (surfactant flooding) to extract
hydrophobic organic contaminants appears promising (Sale and Piontek, 1988).
Aqueous surfactant solutions are superior to water alone in extracting hydro-
phobic contaminants (Ellis et al., 1985). Surfactant addition improves both the
detergency of water and the mobility of NAPL in water. Improving the
detergency of water causes preferential wetting, increased NAPL solubiliza-
tion, and enhanced NAPL emulsification. Adding surfactants also may
increase residual NAPL mobility in water by lowering the interfacial tension
between water and NAPL, which facilitates the distortion of spherical NAPL
droplets as they pass through the media (Salager et al., 1979). Examples of
DNAPL and LNAPL soil washing using surfactants are given in Sale et al.
(1988) and Tuck et al. (1988), respectively.
In alcohol flooding (Taber, 1981), alcohol and NAPL dissolve each other,
thereby mobilizing the residual. Problems with this method include cost, phase-
behavior difficulties, and lack of field experience even in EOR (J.L. Wilson and
Conrad, 1984).
When in contact with certain hydrocarbon mixtures, alkaline agents (e.g.,
sodium carbonate) can react to form surfactants via a saponification reaction
(Sale and Piontek, 1988). These surfactants are created at the water-NAPL
interface, effectively reducing the interfacial tension. Combining alkaline
agents and surfactants may be a cost-effective way of reducing interfacial
tension and enhancing NAPL recovery (Surkalo, 1990). As in surfactant
flooding, decreased interfacial tension resulting from alkaline agents is not
likely to be effective unless favorable mobility ratios also are achieved.
Potential problems associated with alkaline agents include precipitation and
resultant aquifer plugging, dispersal and expansion of clays, and leaching of
trace metals.
Another EOR method that may have environmental applications is polymer
flooding. Adding polymer increases the effectiveness of a waterflood by
increasing the viscosity of the flood, thus lowering the mobility ratio (Sale and
Piontek, 1988). This method is expensive and has not been evaluated extensive-
ly for use in environmental applications.
As indicated, EOR techniques are largely untested for environmental
problems. There are technical problems, especially when dealing with shallow
water-table aquifers where injection pressure is limited by aquifer thickness.
Laboratory studies and small-scale field experiments are likely to yield over-
optimistic expectations (Mackay and Cherry, 1989). In the field, unexpected
spatial variability in porous medium parameters can detrimentally affect the
outcome of an EOR application. In addition to the technical problems, high
costs and regulatory restrictions are associated with injection. Finally,
primary recovery typically removes ~ 30-40 % of the NAPL. Secondary and
tertiary recovery, such as EOR, if successful, may remove only an additional
30-50 %. Thus, as much as 10-40 % of the NAPL can remain in the subsurface
after successful application of EOR methods.
A REVIEW OF IMMISCIBLE FLUIDS IN THE SUBSURFACE 143

Vacuum extraction

If NAPL is volatile and near or above the water table, vacuum extraction is
a potential technique for removing residual saturation. During vacuum
extraction, air is forced through soils contaminated with volatile organic
compounds. The resulting vapors generally are collected and treated. Simple
techniques have been developed to control subsurface hydrocarbon vapors and
are discussed in Dunlap (1984), Marley and Hoag (1984), and O'Connor et al.
(1984). In general, two principal types of vapor management systems are
available: positive differential pressure systems and negative differential
pressure systems. Positive differential pressure systems induce vapor flow
away from the control points while negative differential pressure systems
induce vapor flow toward the control points. The vapor management methods
may be either passive or active. Passive methods use naturally occurring
differences in gas pressures to induce the required flow regime. Active methods
require the artificial generation of differential gas pressures to accomplish the
same flow pattern. Practical experience has demonstrated that active
generation of negative differential gas pressures typically provides the most
favorable field results.
The air flow generates advective vapor fluxes that change the vapor-liquid
equilibrium, inducing volatilization of contaminants. The advantage of this
method is that it is implemented in place, causing minimum disruption. This is
especially important at active facilities or sites hindered by physical obstacles.
Vacuum extraction laboratory studies are described in Thornton and Wootan
(1982), Marley and Hoag (1984), and T.R.I. (1984). A field-scale experiment is
discussed in Crow et al. (1985, 1987). Simulation techniques used to simulate
vacuum extraction are presented in D.E. Wilson et al. (1987), Johnson et al.
(1988), Krishnayya et al. (1988), Stephanatos (1988), Baehr et al. (1989), and
Massmann (1989). Applications to hazardous waste sites include Agrelot et al.
(1985), Conner (1988), Regalbuto et al. (1988), and Hutzler et al. (1989). Knowing
subsurface spatial variability and proper design control the effectiveness of a
vacuum extraction application.
Vacuum extraction can be effective to remove NAPL at residual saturation
from the vadose zone. According to Hutzler et al. (1989), most chemicals that
have been successfully extracted have a low molecular weight and high
volatility. Most of the compounds have values of Henry's law constants
of > 0.01. Vacuum extraction tends to be most effective in homogeneous,
permeable media. If the water table is lowered, it can also be used to remove
residual NAPL from below the original water-table elevation. Groundwater
pumping and vacuum extraction are being used together to recover DNAPL
contaminants at the Tyson's Superfund site, Pennsylvania, U.S.A. (Wassersug,
1990). Vacuum extraction also can increase natural biodegradation processes
by introducing additional oxygen into the subsurface. Finally, vacuum
extraction is generally used in conjunction with ather remedial methods.
144 J.W. M E R C E R A N D R.M. C O H E N

Bioreclamation

Many water-table aquifers support aerobic microorganisms that degrade a


variety of organic contaminants. According to J.R. Wilson et al. (1986),
examples include benzene, toluene, xylenes, and other alkylbenzenes from
gasoline or solvent spills (Lee and Ward, 1984); naphthalene, the methyl-
naphthalenes, fluorene, acenaphthene, dibenzofuran and a variety of other
polynuclear aromatic hydrocarbons released from spilled diesel oil or heating
oil (B.H. Wilson and Rees, 1985); acetone, isopropanol, methanol, ethanol, and
t-butanol from solvent and gasoline spills (Jhaveri and Mazzacca, 1983; Lokke,
1984; Novak et al., 1984); and many methylated phenols and heterocyclic
organic compounds found in certain industrial waste waters. Many synthetic
organic compounds can also be degraded, including dichlorobenzenes (Kuhn et
al., 1985), the mono-, di- and tri-chlorophenols (Suflita and Miller, 1985), the
detergent builder nitrilotriacetic acid {Ward, 1985), and some of the simpler
chlorinated compounds such as methylene chloride (dichloromethane)
(Jhaveri and Mazzacca, 1983). Many DNAPL chemicals are halogenated
compounds that degrade under anaerobic conditions or are resistant to deg-
gradation (Feenstra and Cherry, 1988). Also, when they do degrade, equally
undesirable chemicals may result (Vogel et al., 1987).
J.R. Wilson et al. (1986) present an overview of bioreclamation as a ground-
water remediation technique. In the saturated zone, oxygen needed for aerobic
biodegradation is available in a dissolved state at relatively low concentra-
tions. Many of the bioreclamation systems are derived from a patent process at
Suntech, Marcus Hook, Pennsylvania, U.S.A. (Raymond, 1974). The first step
in this process is to recover mobile NAPL. Next, laboratory studies are
conducted to determine if the native microbial population can degrade the
contaminants. Finally, a combination of oxygen and nutrients is injected into
the contar~inated zone to enhance natural biodegradation. Controlling
groundwater flow in the saturated zone is critical for transporting oxygen and
nutrients to the contaminated zone and enhancing the degradation process.
Therefore, the limiting process in bioreclamation is convective transport of
oxygen and nutrients, not rate limitations due to degradation. Hence, site
characterization providing a detailed definition of hydraulic conductivity is
required for a successful bioreclamation design. Bioreclamation has been
applied mainly to LNAPL's (e.g., gasoline) with varying success (e.g., Yaniga
and Mulry, 1984). According to Downey (1990), while microbiologists have
proven the principles of biodegradation in the laboratory, engineers are having
less success achieving a uniform reaction in heterogeneous aquifers.
There is also related work on bio-emulsification of oils for microbial en-
hancement of oil recovery from petroleum reservoirs (Cooper and Zajic, 1980)
that should be directly applicable to petroleum product spills. Ehrlich et al.
(1985) demonstrated that bacteria from a well contaminated by JP-5 ~ jet fuel
could emulsify the fuel if the well water was supplemented with phosphate and
nitrate. In favorable geologic situations, these emulsions should be mobile and
A REVIEW OF IMMISCIBLE FLUIDS IN THE SUBSURFACE 145

could be removed by pumping for treatment on the surface (J.R. Wilson et al.,
1986). Microorganisms can also mobilize hydrocarbons by transforming them
to polar compounds such as alcohols, ketones, phenols, or organic acids (Perry,
1979). Here again, the success of the mobilization will depend on the site
characterization.

DISCUSSION AND FUTURE RESEARCH

NAPL chemicals result from a variety of industrial processes and are found
at numerous contaminated sites. Where present, NAPL's are a long-term
source of miscible groundwater contamination. Site characterization and re-
mediation depend on understanding the NAPL behavior in the subsurface. This
understanding depends on knowledge of NAPL properties, which is the main
topic of this paper.
Density determines if the immiscible fluid is lighter than water (LNAPL) or
heavier than water (DNAPL). LNAPL's are often associated with petroleum
hydrocarbons. Because of work performed in the petroleum industry, consider-
able information has been developed on properties associated with petroleum
products. In addition, because LNAPL is at or near the water table, charac-
terization and remediation are generally more successful than at DNAPL sites.
Remediation of NAPL contamination often involves a combination of
techniques to form a treatment train. Mobile NAPL should be recovered using
a pump-and-treat technology. This will leave behind a residual saturation that
might be remediated by another technique such as vacuum extraction or EOR
methods. Residual NAPL recovery, especially for DNAPL, is difficult with no
routine proven technology available. At some sites, hydraulic containment
may be the only practical contaminant management option.
Although much work has been accomplished to better understand NAPL
behavior, additional effort should be focused on the following areas:

Properties

(1) Development and demonstration of improved methods to measure in situ


saturations.
(2) Measurement of relative permeability functions, especially to improve
understanding of the field behavior of solvents.
(3) Because of the importance of residual saturation as a continuing source
for vapor transport and/or dissolved chemical transport, an improved un-
derstanding of in situ volatilization and dissolution is needed.

Models

(4) Although much of the data required for NAPL simulation are unavail-
able, model improvements can be made by adding gas and dissolved chemical
146 J.W. MERCER AND R.M. COHEN

transport to existing NAPL codes. This would also require an improved un-
derstanding of mass-transfer mechanisms.
(5) Few DNAPL simulation studies have been published. Conceptual
simulation analyses of the effects of variable source and subsurface conditions
on NAPL transport, containment and recovery should be conducted to provide
insight to site characterization, chemical migration and remediation.

Characterization

(6) As with any contamination problem, the ability to define spatial variabil-
ity, including fracturing, is important to understanding NAPL migration and
recovery.
(7) Improved manuals that provide documentation and guidance on
procedures and priorities for investigating NAPL contamination sites are
needed by NAPL site investigators.

Remediation

(8) Many NAPL remediation technologies are largely untested; field-scale


demonstration studies are required to determine which technologies work
effectively under what conditions. These studies require careful control to
evaluate actual mass recovered and left in situ.

Ongoing research and experience from current field work at sites such as
S-Area and Hyde Park in New York (Cohen et al., 1987) and Tyson's site in
Pennsylvania (Wassersug, 1990) should significantly advance our understand-
ing of the feasibility of using different methods to investigate and remediate
subsurface NAPL contamination during the 1990's.
A REVIEW OF IMMISCIBLE FLUIDS IN T H E SUBSURFACE 147

. . . . . . . . . . . . .

e4 ~ , ~ e4 ~ e4 e~ ~ ~

T
. . . . . . . . . . . . . . . . . . . . . . . . .

I I I I I I II 7 T~77

7~ T
~ ~ ~ ~ " ~~ ~ ~ ~ ~ ~

4~

o ~ # ~ # ~ ~ ~ ~ i ~ ~ ~ ~

0
g:U • ,.~ ~ .,~ ,.~ ~ ~ ~ "-~ o

o o ~

Z
{ oNNNNNN~NN ~ ~ ~o~
o o o
148 J.W. MERCER AND R.M. COHEN

i=

. . . . . ~ . . ~ . . . . . . .

, . . . . . . . . . . . . . . ° , , , , , ° ° ° ,

I:I

7 7 7 7
. . . . . 0

~ ~ v

~ = , 7 %%

~ ~ , ~ ~ ~ , ~ ~

I= I=

,,.,,,.,
o=
..i,,

°,,~,

i !
Z ~-~

< ~g e~ °~ 0~
A REVIEW OF IMMISCIBLE FLUIDS IN T H E S U B S U R F A C E 149

. . . . . . . .

.. il

o -~.q

~, T ~ T T
. . o o . . °

I ~ I

~ =6 v M N

'T ,
, , ~ Z ~

~-~ , ...~

• ~c~'~ ~ ,

cn 8 " ~ . ~ , ~ .. .
.....,
~ _ ~

....,
~ ~ =~
~ ~ ~o
•~ - .~

~{~ ,, ,, ,,~
150 J.W. M E R C E R A N D R.M. C O H E N

°,.~ A

ml

~0 ~ ~ 0 ' ~

. . . ~ ~ .

° i,,,i

i,,=,1

~ A A ~
o

~~
o

°1-i

°,,,~

°1-i

ggg~g~
~o~ ~o ~ ~
o ~
~ o o o o o 0 0 0
°~. ~oo
~ o~°~
°°~°~°
A REVIEW OF IMMISCIBLE FLUIDS IN THE SUBSURFACE 151

m + +~ m+ ++++

~
+++°+++++++++ ' .
+ ~ r+++ + + =-.-+m~+++

• ,
+ ~mmm mmm

<

- . om++
m++ + + + M m m+
m++++m+m
+2,+++

C~

i~ tt~

. . . .
®mmm . . . . . . . .
++++~+moO . .
m
,,-+ ++,.~ ~ ~ ~ ~

~i C'+4 t r + ~'~ +,,~ ,..+ tr,+

,::++ ,+,
~. ,+..,?

<3,I
+
m,
o~
_m
,,...., ..~ ,~
~t. ~-+ - - . "r

~ +. +-,
..t=~ ++."-, +., ,~

° ~ ~
152 J.W. M E R C E R A N D R.M. C O H E N

~t

b~

E-,

LI

.o

,=,
r~

• q ~

~ ~ ~ ~ •

°.

,=.#

~=,

.<
A REVIEWOF IMMISCIBLEFLUIDSIN THE SUBSURFACE 153

REFERENCES

Abdul, A.S., Kia, S.F. and Gibson, T.L., 1989. Limitations of monitoring wells for the detection and
quantification of petroleum products in soils and aquifers. Ground Water Monit. Rev., 9(2):
90-99.
Abr~ola, L.M., 1983. Mathematical modeling of the multiphase migration of organic comlx,~ds in
porous medium. Ph.D. Dissertation, Department of Civil Engineering, Princeton Un~.,ersity,
Princeton, N.J., 187 pp.
Abriola, L.M., 1988. Multiphase flow and transport models for organic chemicals: A review and
assessment. Electr. Power Res. Inst., Palo Alto, Calif., EA-5976, 93 pp.
Abriola, L.M. and Pinder, G.F., 1985a. A multiphase approach to the modeling of porous media
contamination by organic compounds, 1. Equation development. Water Resour. Res., 21(1):
11-18.
Abriola, L.M. and Pinder, G.F., 1985b. A multiphase approach to the modeling of porous media
contamination by organic compounds, 2. Numerical simulation. Water Resour. Res., 21(1):
19-26.
Acher, A.J., Boderie, P. and Yaron, B., 1989. Soil pollution by petroleum products, I. Multiphase
migration of kerosene components in soil columns. J. Contain. Hydrol., 4: 333-345.
Adams, T.V. and Hampton, D.R., 1990. Effects of capillarity on DNAPL thickness in wells and
adjacent sands. I.A.H. (Int. Assoc. Hydrogeol.) Conf. on Subsurface Contamination by
Immiscible Fluids, April 18-20, 1990, Calgary, Alta.
Adams, Jr., W.R. and Atwell, J.S., 1983. A dual purpose cleanup at a superfund site. Proc. Natl.
Conf. on Management of Uncontrolled Hazardous Waste Sites. Hazard. Mater. Control Res.
Inst., Silver Springs, Md., pp. 352-353.
Agrelot, J.C., Malot, J.J. and Visser, M.J., 1985. Vacuum: Defense system for ground water VOC
contamination. Proceedings of the Fifth National Symposium and Exposition on Aquifer
Restoration and Ground Water Monitoring. Natl. Water Well Assoc., Columbus, Ohio, pp.
485--494.
Alford-Stevens, A.L., 1986. Analyzing PCBs. Environ. Sci. Technol., 20(12): 1194-1199.
Anastos, G.J., Dul, E.F., Anderson. J.E. and Sherman, M.N., 1983. Coal tar disposal: Case study.
Proc. 6th Annu. Elect. Eng. Inst. and Envirosphere Conf. on Environmental Licensing and
Regulations Required Affecting the Electric Utility Industry, Washington, D.C.
Anderson, M.R., 1988. The dissolution and transport of dense non-aqueous phase liquids in
saturated porous media• Ph.D. Dissertation, Oregon Graduate Center, Beaverton, Oreg., 260 pp.
Anderson, W.G., 1986a. ~iettability literature survey, Part 1. Rock/oil/brine interactions, and the
effects of core handling on wettability. J. Pet. Technol., Oct. 1986, pp. 1125-1149.
Anderson, W.G., 1986b. Wettability literature survey, Part 2. Wettability measurement. J. Pet.
Technol., Nov. 1986, pp. 1246-1262.
Anderson, W.G., 1986c. Wettability literature survey, Part 3. The effects of wettability on the
electrical properties of porous media. J. Pet. Technol., Dec. 1986, pp. 1371-1378.
Anderson, W.G., 1987a. Wettability literature survey, Part 4. The effects of wettability on capillary
pressure. J. Pet. Technol., Oct. 1987, pp. 1283-1300.
Anderson, W.G., 1987b. Wettability literature survey, Part 5. The effects of wettability an relative
permeability. J. Pet. Technol., Nov. 1987, pp. 1453-1468.
Anderson, W.G., 1987c. Wettability literature survey, Part 6. The effects of wettability on water-
flooding. J. Pet. Technol., Dec. 1987, pp. 1605-1622.
A.P.I. (American Petroleum Institute). 1980. Underground spill cleanup manual. Am. Pet. Inst.,
Washington, D.C., Publ. No. 1628., 34 pp.
A.P.I. (American Petroleum Institute), 1989. A guide to the assessment and remediation of under-
ground petroleum releases Am. Pet. Inst., Washington, D.C., Publ. No. 1628, 2nd ed., 81 pp.
Arthur D. Little, Inc., 1981. The role of capillary pressure in the S Area landfill. Arthur D. Little,
Inc., Boston, Mass., Rep. to Wald, Harkrader & Ross, Washington, D.C., prepared for Environ.
Prot. Agency/State/City S Area Settlement Discussions, May 1981.
Ashland ® Product Catalog, 1989. Ashland Chemical Company, Dublin, Ohio.
154 J.W. MERCER AND R.M. COHEN

A.S.T.M. (American Society for Testing and Materials), 1985. Annual Book of ASTM Standards,
Section 5 Petroleum Products, Lubricants, and Fossil Fuels. Am. Soc. Test. Mater., Philadel-
phia, Pa.
Azbel, D., 1981. Two.Phase Flows in Chemical Engineering. Cambridge University Press,
Cambridge, 311 pp.
Baehr, A.L., 1987. Selective transport of hydrocarbons in the unsaturated zone due to aqueous and
vapor phase partitioning. Water Resour. Res., 23(10): 1926-1938.
Baehr, A.L. and Corapcioglu, M.Y., 1987. A compositional multiphase model for groundwater
contamination by petroleum products, 2. Numerical solution. Water Resour. Res., 23(1): 201-
214.
Baehr, A.L., Hoag, G.E. and Marley, M.C., 1989. Removing volatile contaminants from the un-
saturated zone by inducing advective air-phase transport. J. Contam. Hydrol., 4(1): 2-6.
Banerjee, S., 1984. Solubility of organic mixtures in water. Environ. Sci. Technol., 18(8): 587-591.
Bear, J., 1972. Dynamics of Fluids in Porous Media. American Elsevier, New York, N.Y., 764 pp.
(see especially pp.444-449).
Bear, J., 1979. Hydraulics cf Groundwater. McGraw-Hill, New York, N.Y., 569 pp.
Benner, F.C. and Bartell, F.E., 1941. The effect of polar impurities upon capillary and surface
phenomena ir, petroleum production. Drill Prod. Pract., pp. 341-348.
Berg, R.R., 1975. Capillary pressures in stratigraphic traps. Am. Assoc. Pet. Geol. Bull., 59(6):
939-956.
Blake, S.B. and Hall, R.A., 1984. Monitoring petroleum spills with wells: Some problems and
solutions. Proc. 4th Natl. Syrup. on Aquifer Restoration and Groundwater Monitoring. Natl.
Water Works Assoc., Columbus, Ohio, pp. 305-310.
Blake, S.B. and Lewis, R.W., 1983. Underground oil recovery. Ground Water Monit. Rev., 3(2):
40-46.
Brooks, R.H. and Corey, A.T., 1964. Hydraulic properties of porous media. Colo. State Univ., Fort
Collins, Colo., Hydrol. Pap. No. 3, 27 pp.
Brut~aert, W., 1973. Numerical solution of multiphase well flow. Proc. Am. Soc. Cir. Eng., J.
Hydraul. Div., 99: 1981-2001.
Callahan, M.A., Slimak, M.W., Gabel, N.W., May, I.P., Fowler, C.F., Freed, J.R., Jennings, P.,
Durfee, R.L., Whitmore, F.C., Maestri, B, Mabey, W.R., Holt, B.R. and Gould, C., 1979. Water-
related environmental fate of 129 priority pollutants, Vols. I and II, U.S. Environ. Prot. Agency,
Off. Water Plann. Stand., USEPA-440/4-79-029a&b.
Carpenter, D.W., 1984. Geologic influences on trichloroethylene distribution. Lawrence Livermore
Natl. Lab., Site 300, Build. 834 Complex, Abstr. Prog. 27th Annu. Meet., Assoc. Eng. Geol.,
Boston, Mass., p. 38.
Carsel, R.F. and Parrish, R.S., 1988. Developing joint probability distributions of soil and water
retention characteristic. Water Resour. Res. 24(5): 755-769.
Cary, J.W., McBridge, J.F. and Simmons, C.S., 1989. Trichloroethylene residuals in the capillary
fringe as affected by air-entry pressure. J. Environ. Qual., 18: 72-77.
Chatzis, I., Morrow, N.R. and Lira, H.T., 1983. Magnitude and detailed structure of residual oil
saturation. Soc. Pet. Eng. J., Apr. 1983, pp. 311-326.
Chatzis, I., Kuntamukklua, M.S. and Morrow, N.R., 1988. Effect of capillary number on the
microstructure of residual oil in strongly water-wet sandstones. Soc. Pet. Eng., Reserv. Eng.,
Aug. 1988, pp. 902-912.
Chiou, C.T., Malcolm, R.L., Brinton, T.[. and Kile, D.E., 1986. Water solubility enhancement of
some organic pollutants and pesticides by dissolved humic and fulvic acids. Environ. Sci.
Technol., 20(5): 502-508.
Chcu, S.F.J. and Griffin, R.A., 1983. Soil, clay, and caustic soda: Effects on solubility, sorpti.on, and
mobility of hexachlorocyclopentadiene. Ill. State Geol. Surv., Environ. Geol. Notes No. 104,
54 pp.
Chouke, R.L., Van Meurs, P. and Van der Poel, C., 1959. The instability of slow, immiscible, viscous
liquid-liquid displacements in permeable media. Pet. Trans. Am. Inst. Min. Metall. Eng., 216:
188-194.
A REVIEW OF IMMISCIBLE FLUIDS IN THE SUBSURFACE " 155

Coates, V.T., Fabian, T. and McDonald, M., 1982. Nineteenth-centu~ technology, twentieth-
century problems. Mech. Eng., 104(2):42-51.
Cohen, R.M., Rabold, R.R., Faust, C.R., Rumbaugh, HI, J.O. and Brid[,e,J.R., 1987. Investigation
and hydraulic containment of chemical migration: Four landfills in Niagara Falls. In: Civil
Engineering Practice. J. Boston Soc. Civ. Eng., Sect./Am. Soc. Cir. Eng., 2(1):33-58.
Cole-Parmer®Catalog, 1989/1990.Cole-Parmer Instrument Co., Chicago, Ill.,copyright 1988, 725 pp.
Conner, J.R., 1988. Case study of soil venting. Pollut. Eng., 20(7): 74-."-8.
Convery, M.P., 1979.The behavior and movement of petroleum products in unconsolidated surficial
deposits. M.S. Thesis, University of Minnesota, Minneapolis, Minn., 175pp.
Cooper, D.G. and Zajic, J.T., 1980. Surface-active compounds from microorganisms. Adv. Appl.
Microbiol., 26: 229-253.
Corapcioglu, M.Y. and Baehr, A.L., 1987. A compositional multiphase model for groundwater
contamination by petroleum products, 1. Theoretical considerations. Water Resour. Res., 23(1):
191-200.
Corey, A.T., 1986. Mechanics of Immiscible Fluids in Porous Media. Water Resources Publications,
Littleton, Colo., 259 pp.
Corey, A.T., Rathjens, C.H., Henderson, J.H. and Wyllie, M.R.J., 1956. Three-phase relative per-
meability. Soc. Pet. Eng. J., 207: 349-351.
Craig, Jr.,F.F., 1971. The Reservoir Engineering Aspects of Waterflooding Monograph, Vol. 3. Soc.
Pet. Eng., Am. Inst. Min. Metall., Henry L. Doherty Ser., Dallas, Texas.
Crichlow, H.B., 1977. Modern Reservoir Engineering--A Simulation Approach. Prentice-Hall,
Englewood Cliffs,N.J., 354 pp.
Crow, W.L., Anderson, E.R. and Minugh, E., 1985. Subsurface venting of hydrocarbon vapors from
an underground aquifer. Am. Pet. Inst.,Washington, D.C., Publ. No. 4410.
Crow, W.L., Anderson, E.R. and Minugh, E., 1987. Subsurface venting of hydrocarbons emanating
from hydrocarbon product on groundwater. Ground Water Monit. Rev., 7(1):51-57.
Dean, J.A. (Editor), 1979. Lange's Handbook of Chemistry. McGraw-Hill, N e w York, N.Y.
Delshad, M. and Pope, G.A., 1989. Comparison of the three-phase oil relativepermeability models.
Transp. Porous Media, 4(1):59-83.
de Pastrovich, T.L., Baradat, Y., Barthel, R., Chiarelli, A. and F,.Lsseil,D.R., 1979. Protection of
groundwater from oil pollution.C O N C A W E (Conservation of Clean Air and Water -- Europe),
The Hague, 61 pp.
Devitt, D.A., Evans, R.B., Jury, W.A., Starks, T.H., Eklund, B. and Ghalsan, A., 1987. Soil gas
sensing for detection and mapping of volatile organics. U.S. Environ. Prot. Agency--Environ.
Monit. Systems Lab., Las Vegas, Nev., EPA/600/8-87/036, 265 pp.
Dietrich, J.K. and Bonder, P.B., 1976. Three-phase oil relative permeability problem in reservoir
simulation. S P E 6044 presented at 51st Annu. Meet., Soc. Pet. Eng., N e w Orleans, La., Oct. 3-6,
1976.
Downey, D.C., 1909. Applying new technologies: A scientificperspective. In: Ground Water and
Soil Contamination Remediation: Toward Compatible Science, Policy, and Public Perception.
Colloq. sponsored by Water Sci. Technol. Board, National Academy Press, Washington, D.C.,
pp. 183-194.
Dracos, T., 1978. Theoretical considerations and practical implications on the infiltrationof
hydrocarbons in aquifers. In: Proceedings of the IAH International Symposium on Ground-
water Pollution by Oil Hydrocarbons. Int. Assoc. Hydrogeol., Prague, pp. 127-137.
Dunlap, L.E., 1984. Abatement of hydrocarbon vapors in buildings. In: Petroleum Hydrocarbons
and Organic Chemicals in Ground Water. Natl. Water Well Assoc., Worthington, Ohio, pp.
504-518.
Eganhouse, R.P. and Calder, J.A., 1976. The solubility of medium molecular weight aromatic
hydrocarbons and the effects of hydrocarbon co-solutes and salinity. Geochim. Cosmochim.
Acta, 40: 555-561.
Ehrlich, G.G., Goerlitz, D,F., Godsy, E.M. and Hult, M.F., 1982. Degradation of phenolic contami-
nants in ground water by anaerobic bacteria: St. Louis Park, Minnesota. Ground Water, 20(6):
703-715.
156 J.W. MERCER AND R.M. COHEN

Ehrlich, G.G., Schroeder, R.A. and Martin, P., 1985. Microbial populations in a jet-fuel contaminat-
ed shallow aquifer at Tustin, California. U.S. Geol. Surv., Open-File Rep., pp. 85-335.
Ellis, W.D., Payne, J.R. and McNabb, G.D., 1985. Treatment of contaminated solid with aqueous
surfactants. U.S. Environ. Prot. Agency, Cincinnati, Ohio, EPA/600]2-85/129.
Falta, R.W., Javandel, I., Pruess, K. and Witherspoon, P.A., 1989. Density-driven flow of gas in the
unsaturated zone due to evaporation of volatile organic chemicals. Water Resour. Res., 25(10):
2159-2169.
Farr, A.M., Houghtalen, R.J. and McWhorter, D.B., 1990. Volume estimates of light nonaqueous
phase liquids in porous media. Ground Water, 28(1): 48-56.
Faust, C.R., 1885. Transport of immiscible fluids within and below the unsaturated zone: A
numerical model. Water Resour. Res., 21(4): 587-596.
Faust, C.R., Guswa, J.H. and Mercer, J.W., 1989. Simulation of three-dimensional flow of
immiscible fluids within and below the unsaturated zone. Water Resour. Res., 25(12): 2449-2464.
Fayers, F.J. and Matthews, J.P., 1984. Evaluation of normalized Stone's methods for estimating
three-phase relative permeabilities. Soc. Pet. Eng. J., 24: 224-232.
Feenstra, S. and Cherry, J.A., 1988. Subsurface contamination by dense non-aqueous phase liquids
(DNAPL) chemicals. Pap. presented at Int. Groundwater Syrup., Int. Assoc. Hydrogeol., Halifax,
N.S., May 1-4, 1988.
Ferrand, L.A., Milly, P.C.D. and Pinder, G.F., 1989. Experimental determination of three-fluid
saturation profiles in porous media. J. Contam. Hydrol., 4(4): 373-395.
Ferry, J.P., Dougherty, P.J., Moser, J.B. and Schuller, R.M., 1986. Occurrence and recovery of a
DNAPL in a low-yielding bedrock aquifer. In: Proceedings of Petroleum Hydrocarbons and
Organic Chemicals in Ground Water: Prevention, Detection and Restoration. Natl. Water Well
Assoc., Houston, Texas, Nov. 12-14, 1986, pp. 722-733.
Fiedler, F.R., 1989. An Investigation of the Relationship between Actual and Apparent Gasoline
Thickness in a Uniform Sand Aquifer. M.S. Thesis, University of New Hampshire, Durham,
N.H. (unpublished).
Fried, J.J., Muntzer, P. and Zilliox, L., 1979. Ground-water pollution by transfer of oil hydrocar-
bons. Ground Water, 17(6): 586-594.
Fussell, D.R., Godjen, H., Hayward, P., Lilie, R.H., Marco, A. and Panisi, C., 1981. Revised Inland
Oil Spill Clean-Up Manual. CONCAWE (Conservation of Clean Air and Water - - Europe), The
Hague, Rep. No. 7/81, 150pp.
Gould, R.F. (Editor), 1964. Contact Angle Wettability and Adhesion. Adv. Chem. Ser., Am. Chem.
Soc., Washi-tgton, D.C., No. 43, 389pp.
Groves, Jr., F.R., 1988. Effect of cosolvents on the solubility of hydrocarbons in water. Environ. Sci.
Technol., 22(3): 282-286.
Guswa, J.H., 1985. Application of multi-phase flow theory at a chemical waste landfill, Niagara
Falls, New York. In: Proceedings of the Second International Conference on Groundwater
Quality Research. Natl. Cent. Ground Water Res., Stillwater, Okla., pp. 108-111.
Hall, R.A., Blake, S.B. and Champlin, Jr., S.C. 1984. Determination of hydrocarbon thickness in
sediments using borehole data. In: Proceedings of the 4th National Symposium on Aquifer
Restoration and Groundwater Monitoring. Natl. Water Well Assoc., Columbus, Ohio, pp.
300 304.
Hall, P.L. and Quam, H., 1976. Countermeasures to control oil spills in Western Canada. Ground
Water, 14(3): 163-169.
Hampton, D.R., 1988. Laboratory investigation of the relationship between actual and apparent
product thickness in sands. Am. Assoc. Pet. Geol., Symp. on Environmental Concerns in the
Petroleum Industry, Palm Springs, Call|\, May 10, 1988.
Hampton, D.R. and Miller, P.D.G., 1988. Lab,,,L-a~ory investigation of the relationship between
actual and apparent product thickne.~s in sands: In: Proceedings of the Conference on
Petroleum Hydrocarbons and Organic Chemicals in Ground Water: F'cevention, Detection and
Restoration. Natl. Water Well Assoc., Dublin, Ohio, pp. 157-181.
Hickok, E.A., Erdman, J.B., Simmonett, M.J., Boyer, G.W. and Johnson, L.L., 1982. Ground water
contamination with creosote wastes. In: Proceedings of the American Society Civil Engineering
A REVIEWOF IMMISCIBLEFLUIDSIN THE SUBSURFACE 157

National Conference on Environmental Engineering, Am. Soc. Civ. Eng., Minneapolis, Minn.,
pp. 430-437.
Hoag, G.E. and Marley, M.C., 1986. Gasoline residual saturation in unsaturated uniform aquifer
materials. J. Environ. Eng., Proc. Am. Soc. Civ. Eng., 112(3):586-604.
Hochmuth, D.P. and Sunada, D.K., 1985. Groundwater model of two-phase immiscible flow in
course material. Ground Water, 23(5):617-626.
Hoffmann, B., 1969. Urber die Ausbreitung gel6ster Kohlenwasserstoffe im Grundwasserleiter.
Mitt. Inst. Wasserwirtsch. Landwirtsch. Wasserbau Tech. Hochschule, Hannover, No. 16,
197 pp.
Hoffmann, B., 1971. Dispersion of soluble hydrocarbons in groundwater stream: Adv. Water Pollut.
Res., Pergamon, Oxford, 2HA.7b, pp. 1-8.
Holzer, T.L., 1976. Application of groundwater flow theory to a subsurface oil spill.Ground Water,
14(3): 138-145.
Homsy, G.M., 1987. Viscous fingering in porous media. Ann. Rev. Fluid Mech., 19: 271-311.
Honarpour, M. and Mahmood, S.M., 1988. Relative-permeability measurements: A n overview. J.
Pet. Technol., 40(8):963-966.
Honarpour, M., Koederitz, L. and Harvey, A.H., 1986. Relative Permeability of Petrolemn
Reservoirs. C R C Press, Boca Raton, Fla., 43 pp.
Hughes, B.M., Gillham, R.W. and Mendoza, C.A., 1990. Transport of trichloroethylene vapours in
the unsaturated zone: A field experiment. Int. Assoc. Hydrogeol. Conf. on Subsurface Con-
tamination by Ivo~miscible Fluids, Calgary, Alta., April 18-20, 1990.
Hult, M.F. and Schoenberg, M.E., 1981. Preliminary evaluation of ground water contRmination by
coal tar derivatives, St. Louis Park Area, Minnesota. U.S. Geol. Surv., Open-File Rep. 81-72,
57 pp.
Hunt, J.R., Sitar, N. and Udell, K.S., 1988a. Nonaqueous phase liquid transport and cleanup, 1.
Analysis of mechanisms. Water Resour. Res., 24(8):1247-1258.
Hunt, J.R., Sitar, N. and Udell, K.S., 1988b. Nonaqueous phase liquid transport and cleanup, 2.
Experimental studies. Water Resour. Res., 24(8):1259-1269.
Hutzler, N.F., Murphy, B.E. and Gierke, J.S.,1989.State of technology review soilvapor extraction
systems. Cooperative Agreement CR-814319-01-1,Hazard. Waste Eng. Res. Lab., Environ. Prot.
Agency, Cincinnati, Ohio, 87 pp.
JBF ScientificCorp., 1981. The interaction of S-Area soilsand liquids:Review and supplementary
laboratory studies. JBF Sci. Corp., Boston, Mass., Rep. submitted to U.S. Environ. Prot.
Agency, Washington, D.C., 26pp.
Jhaveri, V. and Mazzacca, A.J.,1983.Bio-reclamation of ground and groundwater. Presented at 4th
Natl. Conf. on Management of Uncontrolled Hazardous Waste Sites, Washington, D.C., Oct.
31-Nov. 2, 1983, pp. 242-247.
Johnson, P.C., Kemblowski, M.W. and Colthart, J.D., 1988. Practical screening models for soil
venting applications. In: Proceedings of Petroleum Hydrocarbons and Organic Chemicals in
Ground Water--Prevention, Detection and Restoration. Natl. Water Well Assoc., Houston,
Texas, pp. 521-546.
Jury, W.A., Spencer, W.F. and Farmer, W.J., 1984. Behavior assessment model for trace organics
in soil, IH. Application and screening model. J. Environ. Qual., 13(4): 573-579.
Jury, W.A., Russo, D., Streile, G. and E1 Abd, H., 1990. Evaluation of volatilization by organic
chemicals residing below the soil surface. Water Resour. Res., 26(1): 13-20.
Kemblowski, M.W. and Chiang, C.Y., 1990. Hydrocarbon thickness fluctuations in monitoring
wells. Ground Water, 28(2): 244-252.
Kenaga, E.E. and Goring, C.A.I., 1980. Relationship between water solubility, soil sorption, oc-
tanol-water partitioning, and concentration of chemicals in biota. In: J.G. Eaton, P.R. Parrish,
and A.C. Hendricks (Editors),Aquatic Toxicology. Am. Soc. Test. Mater., Washington, D.C.,
A S T M STP 707, pp. 78-115.
Krishnayya, A.V., O'Connor, M.J., Agar, J.G. and King, R.D., 1988. Vapour extraction system
factorsaffectingtheirdesign and performance. In: Proceedings of Petroleum Hydrocarbons and
Organic Chemicals in Groundwater Prevention, Detection and Restoration. Natl. Water Well
Assoc., Houston, Texas, pp. 547-567.
158 J.W. MERCER AND R.M. COHEN

Kueper, B.H. and Frind, E.O., 1988. An overview of immiscible fingering in porous media. J.
Contam. Hydrol., 2: 95-110.
Kueper, B.H., Abbott, W. and Farquhar, G., 1989. Experimental observations of multiphase flow in
heterogeneous porous media. J. Contam. Hydrol., 5(1): 83-95.
Kuhn, E.P., Colberg, P.J., Schnoor, J.L., Wanner, O., Zehnder, A.J.B. and Schwarzenbach, R.P.,
1985. Microbial transformation of substituted benzenes during infiltration of river water to
groundwater: Laboratory column studies. Environ. Sci. Technol., 19: 961-968.
Lafornara, J.P., Nadeau, R.J., Allen, H.L. and Massey, T.I., 1982. Coal tar: Pollutants of the past
threaten the future. In: Proceedings of the Hazardous Material Spill Conference. Bur. Explor.,
Washington, D.C., pp. 37-42.
Lappala, E.G. and Thompson, G.M., 1983. Detection of groundwater contamination by shallow soil
gas sampling in the vadose zone. In: Proceedings of the Characterization and Monitoring of the
Vadose Zone Conference. Natl. Water Well Assoc., Las Vegas, Nev., Dec. 8-9, 1983, pp. 659-679.
Leach, R.O., Wagner, O.R., Wood, H.W. and Harpke, C.F., 1962. A laboratory and field study of
wettability adjustment in waterflooding. J. Pet. Technol., 14: 206-214.
Lee, M.D. and Ward, C.H., 1984. Reclamation of contaminated aquifers: Biological techniques.
Proc. of Hazardous Material Spills Conference. Nashville, Tenn., Apr. 9-12, 1984, pp. 98-103.
Lenhard, R.J. and Parker, J.C., 1987a. A model for hysteretic constitutive relations governing
multiphase flow, 2. Permeability-saturation relations. Water Resour. Res., 23(12): 2197-2206.
Lenhard, R.J. and Parker, J.C., 1987b. Measurement and prediction of saturation-pressure rela-
tionships in three phase porous media systems. J. Contain. Hydrol., 1: 407-424.
Lenhard, R.J. and Parker, J.C., 1988. Experimental validation of the theory of extending two-phase
saturation-pressure relations to three-fluid phase systems for monotonic drainage paths. Water
Resour. Res. 24(3): 373-380.
Lenhard, R.J. and Parker, J.C., 1990. Estimation of free hydrocarbon volume from fluid levels in
monitoring wells. Ground Water, 28(1): 57-67.
Leo, A., Hansch, C. and Elkins, D., 1971. Partition coefficients and their uses. Chem. Rev., 71(6):
525-616.
Leverett, M.C., 1941. Capillary behavior in porous solids. Trans. Am. Inst. Min. Metall. Eng., Pet.
Eng. Div., 142: 152-169.
Lin, C., Pinder, G.F. and Wood, E.F., 1982. Water resources program report 83-WR-2. Water Resour.
Prog., Princeton Univ., Princeton, N.J., Oct. 1982, 33 pp.
Lokke, H., 1984. Leaching of ethylene glycol and ethanol in subsoils. Water, Air Soil Pollut., 22:
373-387.
Lyman, W.J., Reehl, W.F. and Rosenblatt, D.H., 1982. Handbook of Chemical Property Estimation
Methods--Environmental Behavior of Organic Compounds. McGraw-Hill, N e w York, N.Y.
Mackay, D.M. and Cherry, J.A., 1989. Groundwater contamination: Pump-and-treat remediation.
Environ. Sci. Technol., 23(6):620-636.
Mackay, D.M., Bobra, A., Chan, D.W. and Shiu, W.Y., 1982. Vapor pressure correlations for
low-volatility environments. Environ. Sci. Technol., 16(10):645-649.
Mackay, D.M., Roberts, P.V. and Cherry, J.A., 1985. Transport of organic contaminants in ground-
water. Environ. Sci. Technol., 19(5):384-392.
Marley, M.C. and Hoag, G.E., 1984. Induced soil venting for recovery/restoration of gasoline
hydrocarbons in the vadose zone. In: Proceedings of Petroleum Hydrocarbons and Organic
Chemicals in Ground Water. Natl. Water Well Assoc., Worthington, Ohio, pp. 473-503.
Marrin, D.L., 1988. Soil-gas sampling and misinterpretation, Ground Water Monit. Rev., 8(2):
51-54.
Marrin, D.L. and Kerfoot, H.B., 1988. Soil-gas surveying techniques. Environ. Sci. Technol., 22(7):
740-745.
Marrin, D.L. and Thompson, G.M., 1984. Remote detection of volatile organic contaminants in
ground water via shallow soil gas sampling, in: Proceedings of Petroleum Hydrocarbons and
Organic Chemicals in Ground Water. Natl. Water Well Assoc., Worthington, Ohio, pp. 172-187.
Marrin, D.L. and Thompson, G.M., 1987. Gaseous behavior of T C E overlying a contaminated
aquifer. Ground Water, 25(1):21-27.
Massmann, J.W., 1989. Applying groundwater flow models in vapor extraction system design. J.
Environ. Eng., 115(1):129-149.
A REVIEW OF IMMISCIBLE FLUIDS IN T H E S U B S U R F A C E 159

McCaffery, F.G. and Batycky, J.P., 1983. Flow of immiscible liquids through porous media. In: N.P.
Cheremisinoff and R. Gupta (Editors), Handbook of Fluids in Motion. Ann Arbor Science
Publishers, Ann Arbor, Mich., pp. 1027-1048.
McDevit, W.F. and Long, F.A., 1952. The activity coet~icient of benzene in aqueous salt solutions.
J. Am. Chem. Soc., 74: 1773-1777.
Melrose, J.C. and Brandner, C.F., 1974. Role of capillary forces in determination of microscopic
displacement efficiency for oil recovery by water flooding. J. Can. Pet. Technol., 13(4): 54-62.
Mendoza, C.A. and Frind, E.O., 1990. Advective-dispersive transport of der~.se organic vapours in
the unsaturated zone, 2. Sensitivity analysis. Water Resour. Res., 26(3): 388-398.
Mendoza, C.A. and McAlary, T.A., 1990. Modeling of groundwater contamination caused by
organic solvent vapors. Ground Water, 28(2): 199-206.
Mercer, J.W., Faust, C.R., Cohen, R.M., Andersen, P.F. and Huyakorn, P.S., 1985. Remedial action
assessment for hazardous waste sites via numerical simulation. Water Manage. Res., 3: 377-387.
Mohanty, K.K., Davis, H.T. and Scriven, L.E., 1987. Physics of oil entrapment in water-wet rock.
Soc. Pet. Eng., Reserv. Eng., Feb. 1987, pp. 113-128.
Montgomery, J.H. and Welkom, L.M., 1990. Groundwater Chemicals Desk Reference. Lewis,
Chelsea, Mich., 640 pp.
Morrow, N.R., Chatzis, I. and Taber, J.J., 1988. Entrapment and mobilization of residual oil in bead
packs. Soc. Pet. Eng., Reserv. Eng., Aug. 1988, pp. 927-934.
Mualem, Y., 1976. A new model for predicting the hydraulic conductivity of unsaturated porous
media. Water Resour. Res., 12(3): 513-522.
Mull, R., 1969. Modellm~issige Beschreibung der Ausbreitung von Mineral61-Produkten in Boden.
Mitt. Inst. Wasserwirtsch., Landwirtsch., Wasserbau, Tech. Univ. Hannover, Hannover, Tech.
Rep.
Mull, R., 1971. Migration ofoil products in the subsoil with regard to groundwater pollution by oil
In: Advances in Water Pollution Research. Pergamon, Oxford, pp. 1-8.
Mull, R., 1978. Calculations and experimental investigations of the migration of hydrocarbons in
natural soils. In: Proceedings of the L~H International Symposium on Groundwater Pollution
by Oil Hydrocarbons. Int. Assoc. Hydrogeol., Prague, pp. 167-181.
NIOSH (National Institute for Occupational Safety and Health), 1987. Pocket Guide to Chemical
Hazards, U.S. Government Printing Office, Washington, D.C.
Nirmalakhandan, N.N. and Speece, R.E., 1988. Prediction of aqueous solubility of organic
chemicals based on molecular structure. Environ. Sci. Technol., 22(3): 328-338.
Noggle, J., 1985. Physical Chemistry. Little, Brown and Co., Boston, Mass.
Novak, J.T., Goldsmith, C.D., Benoit, R.E. and O'Brien, J.H., 1984. Biodegradation of alcohols in
subsurface systems. Spec. Semin. on Degradation, Retention and Dispersion of Pollutants in
Groundwater. Copenhagen, Sept. 12-14, 1984, pp. 61-75.
Nutting, P.G., 1934. Some physical and chemical properties of reservoir rocks bearing on the
accumulation and discharge of oil. In: W.E. Wrather and F.H. Lahee (Editors), Problems of
Petroleum Geology. Am. Assoc. Pet. Geol., Tulsa, Okla., pp. 825-832.
O'Connor, M.J., Agar, J.G. and King, R.D., 1984. Practical experience in the management of
hydrocarbon vapors in the subsurface. In: Petroleum Hydrocarbons and Organic Chemicals in
Ground Water. Natl. Water Well Assoc., Worthington, Ohio, pp. 519-533.
Osborne, M. and Sykes, J., 1986. Numerical modeling of immiscible organic transport at the Hyde
Park Landfill. Water Resour. Res., 22(1): 25-33.
Palumbo, D.A. and Jacobs, J.H., 1982. Monitoring chlorinated hydrocarbons in groundwater. In:
Proceedings of the National Conference on Management of Uncontrolled Hazardous Waste
Sites. Hazard. Mater. Control Res. Inst., Silver Spring, Md., pp. 165-168.
Parker, J.C. and Lenhard, R.J., 1987. A model for hysteretic constitutive relations governing
multiphase flow, 1. Saturation-pressure relations. Water Resour. Res., 23(12): 2187-2196.
Parker, J.C., Lenhard, R.J. and Kuppusamy, T., 1987. A parametric model for constitutive
properties governing multiphase fluid flow in porous media. Water Resour. Res., 23(4): 618-6o,,4.
Peaceman, D.W., 1977. Fundamentals of Numerical Reservoir Simulation. Elsevier, Amsterdam,
176 pp.
160 J.W. MERCER AND R.M. COHEN

Pereira, W.E., Rostad, C.E., Garbarino, J.R. and Hult, M.F., 1983. Groundwater contamination by
organic bases derived from coal tar wastes. Environ. Toxicol. Chem., 2: 283-294.
Perry, J.J., 1979. Microbial cooxidations involving hydrocarbons. Microbiol. Rev., 43: 59-72.
Pfannkuch, H., 1983. Hydrocarbon spills, their retention in the subsurface and propagation into
shallow aquifers. Off. Water Res. Technol., Washington, D.C., Rep. W83-02895, 51 pp.
Pfannkuch, H., 1984. Mass-exchange processes at the petroleum-water interface. Pap. presented at
Toxic-Waste Tech. Meet., Tucson, Ariz., March 20-22, 1984. In: M.F. Hult (Editor), U.S. Geol.
Surv., Water-Resour. Invest. Rep. 84-4188, pp. 23-48.
Pinder, G.F. and Abriola, L.M., 1986. On the simulation of nonaqueous phase organic compounds
in the subsurface. Water Resour. Res., 22(9): 109S-119S.
Ramsey, W.L., Steimle, R.R. and Chaconas, J.T., 1981. Renovation of a wood treating facility. In:
Proceedings of the National Conference on Management of Uncontrolled Hazardous Waste
Sites. Hazard. Mater. Control Res. Inst., Silver Spring, Md., pp. 212-214.
Rathmell, J.J., Braun, P.H. and Perkins, T.K., 1973. Reservoir waterflood residual oil saturation
from laboratory tests. J. Pet. Technol., pp. 175-185.
Raymond, R.L., 1974. Reclamation of hydrocarbon contaminated groundwaters. U.S. Patent Off.,
Washington, D.C., 3,846,290, patented Nov. 5, 1974.
Regalbuto, D.P., Barrera, J.A. and Lisiecki, J.B., 1988. In-situ removal of VOCs by means of
enhanced volatilization. In: Proceedings of Petroleum Hydrocarbons and Organic Chemicals in
Groundwater: Prevention, Detection and Restoration. Natl. Water Well Assoc., Houston,
Texas, pp. 571-590.
Riddick, J.A. and Burger, W.B., 1970. Organic Solvents--Physical Properties and Methods of
Purification. Wiley-Interscience, New York, N.Y., 1041pp.
Roberts, J.R., Cherry, J.A. and Schwartz, F.W., 1982. A case study of a chemical spill: Poly-
chlorinated biphenyls (PCBs), 1. History, distribution, and surface translocation. Water
Resour. Res., 18(3): 525-534.
Rossi, S.S. and Thomas, W.it., 1981. Solubility behavior of three aromatic hydrocarbons in distilled
water and natural seawater. Environ. Sci. Technol., 15: 715-716.
Salager, J.L., Morgan, J.C., Schecter, R.S., Wade, W.H. and Vasquex, E., 1979. Optimum formula-
tion of surfactant]water/oil systems for minimum interracial tension or phase behavior. Soc.
Pet. Eng. J., pp. 107-115.
Salathiel, R.A., 1973. Oil recovery by surface film drainage in mixed-wettability rocks. J. Pet.
Technol., 25: 1216-1224.
Sale, T.C. and Piontek, K., 1988. In situ removal of waste wood-treating oils from subsurface
materials. Presented at U.S. Environ. Prot. Agency, Forum on Remediation of Wood-Preserving
Sites. San Francisco, Calif., Oct. 1988, 24 pp.
Sale, T.C., Stieb, D., Piontek, K.R. and Kuhn, B.C., 1988. Recovery of wood-treating oil from an
alluvial aquifer using dual drainlines. In: Proceedings of Petroleum Hydrocarbons and Organic
Chemicals in Ground Water. Natl. Water Well Assoc., Worthington, Ohio, pp. 419-422.
Saraf, D.N. and McCaffery, F.G., 1982. Two- and three-phase relative permeabilities: A review. Pet.
Recov. Inst., Calgary, Alta., Rep. No. 81-8.
Schiegg, H.O., 1977. Methode zur Absch~itzung der Ausbreitung von ErdSlderivaten in mit Wasser
und Luft erfiillten Boden. Mitt. Versuchsanst. Wasserbau, Hydrol. Glaziol., EidgenSss. Tech.
Hochschule, Zfirich, 256 pp.
Schiegg, H.O., 1985. Considerations on water, oil, and air in porous media. Water Science Technol.,
23(4/5): 467-476.
Schowalter, T.T., 1979. Mechanics of secondary hydrocarbon migration and entrapment. Am.
Assoc. Pet. Geol. Bull., 63(5): 723-760.
Schwartz, F.W., Cherry, J.A. and Roberts, J.R., 1982. A case study of a chemical spill: Poly-
chlorinated biphenyls (PCBs), 2. Hydrogeological conditions and contaminant migration.
Water Resour. Res., 18(3): 535-545.
Schwille, F., 1984. Migration of organic fluids immiscible with water in the unsaturated zone. In:
B. Yaron, G. Dagan and J. Goldschmid (Editors), Pollutants in Porous Media: The Unsaturated
Zone between Soil Surface and Groundwater. Springer, New York, N.Y., pp. 27--48.
A REVIEWOF IMMISCIBLEFLUIDSIN THE SUBSURFACE 161

Schwille, F., 1988. Dense Chlorinated Solvents in Porous and Fractured Media. Lewis, Chelsea,
Mich., 146 pp.
Senn, R.B. and Johnson, M.S., 1987. Interpretation of gas chromatographic data in subsurface
hydrocarbon investigations. Ground Water Monit. Rev., 7(1): 58-63.
Shah, D.O. (Editor), 1981. Surface Phenomena in Enhanced Oil Recovery. Plenum, New York, N.Y.
Shifrin, N.S., 1986. Affidavit re United States District Court (Buffalo, N.Y.) for the Western District
of New York, United States of America et al., Plaintiffs, v. Hooker Chemicals and Plastics
Corp., Niagara Falls, N.Y. (Hyde Park Landfill defendants).
Sleep, B.E. and Sykes, J.F., 1989. Modeling the transport of volatile organics in variably saturated
media. Water Resour. Res., 25(1): 81-92.
Smith, D.A., 1966. Theoretical considerations of sealing and non-sealing faults. Am. Assoc. Pet.
Geol. Bull., 50: 363-374.
Snell, R.W., 1962. Three-phase relative permeability in unconstituted sand. J. Inst. Pet., 84: 80-88.
Stephanatos, B.N., 1988. Modeling the transport of gasoline vapors by an advective-diffusion
unsaturated zone model. In: Proceedings of Petroleum Hydrocarbons and Organic Chemicals in
Ground Water--Prevention, Detection and Restoration. Natl. Water Well Assoc., Houston,
Texas, pp. 591-611.
Stone, H.L., 1970. Probability model for estimating three-phase relative permeability. J. Pet.
Technol., 20: 214-218.
Stone, H.L., 1973. Estimation of three-phase relative permeability and residual oil data. J. Can. Pet.
Technol., 12(4): 53-61.
Suflita, J.M. and Miller, G.D., 1985. Microbial metabolism of chlorc, phenolic compounds in ground
water aquifers. En,:ir6r,. Toxicol. Chem., 4: 751-758.
Surkalo, H., 1990. Enhanced alkaline flooding. J. Pet. Technol., 42(1): 6-7.
Taber, J.J., 1981. Research on enhanced oil recovery, past, present, and future. In: D.O. Shah
(Editor), Surface Phenomena in Enhanced Oil Recovery. Plenum, New York, N.Y., 750 pp.
Testa, S.M. and Paczkowski, M.T., 1989. Volume determination and recoverability of free hydrocar-
bon. Ground Water Monit. Rev., 9(1): 120-128.
Thomas, G.W., 1982. Principles of hydrocarbon reservoir simulation. Int. Human Resour. Dev.
Corp., Boston, Mass., 207 pp.
Thompson, S.N., Burgess, A.S. and O'Dea, D., 1983. Coal tar containment and cleanup: Plattsburgh,
New York. In: Proceedings of the National Conference on Management of Uncontrolled
Hazardous Waste Sites. Hazard. Mater. Control Res. Inst., Silver Spring, Md., pp. 331-337.
Thornton, J.S. and Wootan, W.L., 1982. Venting for the removal of hydrocarbon vapors from
gasoline contaminated soil. J. Environ. Sci. Health, A17(1): 31-44.
Treiber, L.E., Archer, D.L. and Owens, W.W., 1972. A laboratory evaluation of the wettability of
fifty oil producing reservoirs. J. Soc. Pet. Eng., 12(6): 531-540.
T.R.I. (Texas Research Institute), 1984. Forced venting to remove gasoline vapors from a large-scale
model aquifer. Am. Pet. Inst., Washington, D.C., 60 pp.
Tuck, D.M., Jaffe, P.R., Crerar, D.A. and Mueller, R.T., 1988. Enhanced recovery of immobile
residual non-wetting hydrocarbons from the unsaturated zone using surfactant solutions.
Proceedings of Petroleum Hydrocarbons and Organic Chemicals in Ground Water. Natl. Water
Well Assoc., Worthington, Ohio, pp. 457-479.
Unites, D.F. and Houseman, Jr., J.J., 1982. Field investigation and remedial action at sites contami-
nated with coal tars. In: Proceedings of the 5th Annual Madison Conference of Applied
Research and Practice on Municipal and Industrial Waste. Dep. Eng. Appl. Sci., Univ. of
Wisconsin Ext., Madison, Wise., pp. 344-355.
U.S.C.G. (U.S. Coast Guard), 1978. CHRIS Hazardous Chemical Data. U.S. Gov. Print. Off.,
Washington, D.C.
U.S.E.P.A. (U.S. Environmental Protection Agency), 1979. Organic Solvent Cleaners: Background
Information fGr Proposed Standards. U.S. Gov. Print. Off., Washington, D.C., NTIS No. PB80-
137912.
U.S.E.P.A. (U.S. Environmental Protection Agency), 1986a. Underground Motor Fuel Storage
162 J.W. M E R C E R A N D R.M. C O H E N

Tanks: A National Survey, Vol. 1. U.S. Gov. Print. Off.,Washington, D.C., Tech. Rep. EPA/560/
5-86-013.
U.S.E.P.A. (U.S. Environmental Protection Agency), 1986b. Superfund Public Health Evaluation
Manual. U.S. Gov. Print. Off., Washington, D.C., EPA/540/1-86-060, Appendix A-1.
Valocchi, A., 1985. Validity of the local equilibrium assumption for modeling sorbing solute
transport through homogeneous soils.Water Resour. Res., 21(6): 808-820.
van Dam, J., 1967. The migration of hydrocarbons in a water bearing stratum. In: P. Hepple
(Editor), The Joint Problems of the Oil and Water Industries. Elsevier, Amsterdam, pp. 55-96.
van der Waarden, M., Bridie, A.L.A.M. and Groenewoud, W.M., 1971. Transport of mineral oil
components to ground water, I. Water Res., 5: 213-226.
van Genuchten, M.'I'h.,1980. A closed-form equation for predicting the hydraulic conductivity of
unsaturated soils. Soil Sci. Soc. Am. J., 44: 892-898.
Villaume, J.F., 1982. The U.S.A.'s first emergency superfund site. In: J.E. Alleman and
J. Kavanaugh (Editors), Proceedings of the 14th Mid-Atlantic Industrial Waste Conference.
Ann Arbor Science Publishers, Ann Arbor, Mich., pp. 311-321.
Villaume, J.F., 1984. Coal tar wastes: Their environmental fate and effects, hazardous and toxic
wastes. In: S.K. Majumdar and F.W. Miller (Editors), Technology Management and Health
Effects. Pa. Acad. Sci., Easton, Pa., pp. 362-375.
Villaume, J.F., 1985. Investigations at sites contaminated with dense, non-aqueous phase liquids
(NAPLs). Ground Water Monit. Rev., 5(2):60-75.
Villaume, J.F., Lowe, P.C. and Lennon, G.P., 1983a, Coal tar recovery from a gravel aquifer:
Stroudsburg, Pennsylvania. In: Proceedings of the Conference on the Disposal of Solid, Liquid
and Hazardous Wastes, Bethlehem, Pa., Am. Soc. Civ. Eng.-Lehigh Univ., pp. 12-1-12-18.
Villaume, J.F., Lowe, P.C. and Unites, D.F., 1983b. Recovery of coal gasification wastes: A n
innovative approach. In: Proceedings of the Third National Symposium on Aquifer Restoration
and Ground Water Monitoring. Natl. Water Well Assoc., Worthington, Ohio, pp. 434-445.
Vogel, T.M., Criddle, C.S. and McCarty, P.L., 1987. Transformations of halogenated aliphatic
compounds. Environ. Sci. Technol., 21(8): 722-736.
Wang, F.H.L., 1988. Effect of wettability alteration on water/oil relative permeability, dispersion,
and flowable saturation in porous media. Soc. Pet. Eng. Reserv. Eng., M a y 1988, pp. 617-628.
Ward, T., 1985. Characterizing the aerobic and anaerobic microbial activities in surface and
subsurface soils. Environ. Toxicol. Chem., 4: 727-737.
Wassersug, S.R., 1990. Policy aspects of current practices and applications. In: Remediating
Groundwater and Soil Contamination. Rep. on a Colloq., Water Sci. Technol. Board, National
Academy Press, Washington, D.C., pp. 133-150.
Wilson, B.H. and Rees, J.F., 1985. Biotransformation of gasoline hydrocarbons in methanogenic
aquifer material. Proceedings of the N W W A / A P I Conference on Petroleum Hydrocarbons and
Organic Chemicals in Ground Water. Natl. Water Well Assoc.-Am. Pet. Inst., Houston, Texas,
Nov. 13-15, 1985, pp. 128-139.
Wilson, D.C. and Stevens, C., 1981. Problems arising from the redevelopment of gas works and
similar sites. Prepared for U.K. Gov., Dep. Environ., London, 175 pp.
Wilson, D.E., Montgomery, R.E. and Sheller, M.R., 1987. A mathematical model for removing
volatile subsurface hydrocarbons by miscible displacement. Water, Air, Soil Pollut., 33(3-4):
231-255.
Wilson, J.L. and Conrad, S.H., 1984. Is physical displacement of residual hydrocarbons a realistic
possibility in aquifer restoration? In: Proceedings of the NWWA/API Conference on Petroleum
Hydrocarbons and Organic Chemicals in Ground Water. Natl. Water Well Assoc., Worthington,
Ohio, pp. 274-298.
Wilson, J.R., Leach, L.E., Henson, M. and Jones, J.N., 1986. In situ biorestoration as a groundwater
remediation technique. Ground Water Monit. Rev., 6(4): 56-64.
Windholz, M., Budavar, S., Blumetti, R.F. and Otterbein, E.S. (Editors), 1983. The Merck Index.
Merck & Co., Rahway, N.J., 1463 pp.
Working Group "Water and Petroleum," 1970. Evaluation and treatment of oil spill accidents on
land with a view to the protection of water resources. Fed. Min. Inter.. F.R. Germany, Bonn,
2nd ed., p. 138.
A REVIEW O F IMMISCIBLE FLUIDS IN T H E S U B S U R F A C E 163

Yaniga, P.M. and Mulry, J., 1984. Accelerated aquifer restoration: In situ applied techniques for
enhanced free product recovery/adsorbed hydrocarbon reduction via bioreclamation. In:
Proceedings of Petroleum Hydrocarbons and Organic Chemicals in Ground Water. Natl. Water
Well Assoc., Worthington, Ohio, pp. 421--440.
Yazacigil, H. and Sendlein, L.V.A., 1981. Management of groundwater contamination by aromatic
hydrocarbons in the aquifer supplying Ames, Iowa. Ground Water, 19(8): 648-665.
Zilliox, L. and Muntzer, P., 1975. Effects of hydrodynamic processes on the development of ground-
water pollution: Study of,physical models in a saturated porous medium. Prog. Water Technol.,
7(3/4): 561-568.
Zilliox, L., Muntzer, P. and Menanteau, J.J., 1973. Probl~me de l'~change entre un produit p~trolier
immobile et reau en mouvement dans un milieu poreux. Rev. Inst. Fr. P~t., 28(2): 185-200.
Zilliox, L., Muntzer, P. and Schwille, F., 1974. Untersuchungen fiber den Stoffaustausch zwischen
Mineral61 und Wasser in porSsen Medien. Dtsch. Gewiisserkd. Mitt., 18(H2): 35-37.
Zilliox, L., Muntzer, P. and Fried, J.J., 1978. An estimate of the source of a phreatic aquifer
pollution by hydrocarbons, oil-water contact and transfer of soluble substances in ground
water. Proceedings of the International Symposium on Groundwater Pollution by Oil-Hydro-
carbons. Int. Assoc. Hydrogeol., Prague, pp. 209-227.
Zytner, R.G., Biswas, N. and Bewtra, J.K., 1989. PCE volatilized from stagnant water and soil. Proc.
Am. Soc. Civ. Eng., J. Environ. Eng., 115(6): 1199-1212.

Potrebbero piacerti anche