You are on page 1of 53

Progress Report for the Project Entitled:

Laboratory Studies to Characterize the Performance of a Sand Cap


over Tar-Contaminated Sediment: Grand Calumet River, Hammond, IN

Investigators:

Chad Jafvert1, Douglas Lane1, Nandita Basu1,


Suresh Rao 1,2, and Linda Lee2

School of Civil Engineering1


and Dept. of Agronomy2, Purdue University, West Lafayette, IN

Prepared for:

NiSource, 801 E. 86th Avenue, Merrillville, IN 46410

Date: August 29, 2003


This progress report documents the equilibrium partitioning experiments
for the 10 sediment core samples collected at the Hammond IN former
MGP site on October 23-24, 2002. Also included are results of sand
column experiments. The primary author of this report is Douglas Lane,
who graduated from Purdue University, May 2003, with an M.S.C.E.
degree. This report is a summary of the experimenta l work reported by
Doug in his M.S.C.E. thesis and column studies performed by Nandita
Basu.

1
Table of Contents

Table of Contents....………………………… ………………………………………… 2

1) Summary……………………………………………………………………………..3

2) Background....…………………………………………………………………….… 4
2.1. Polycyclic Aromatic Hydrocarbons…………………………………………... 4
2.2. PAH Toxicity…………………………………………………………………. 6
2.3. Aqueous PAH Photodegradation……………………………………………... 7
2.4. Equilibrium Partitioning……………………………………………………… 8
2.5. Sorption……………………………………………………………………….. 11
2.6. Groundwater Flow……………………………………………………………. 12

3) Methods and Materials……………………………………………………….…….. 20


3.1. Materials……………………………………………………………………… 20
3.2. Sample Collection…………………………………………………………….. 20
3.3. Water, Soil and Carbon Content Analysis……………………………………. 21
3.4. Equilibrium Partitioning of PAHs between Tar and Water……………..….… 21
3.5. Bench-Scale Sediment Flow-Through Experiments…………………………..22
3.6. Advection-Dispersion Experiments……………………..……………………. 24

4) Results and Analysis…………………………………...…………………………….26


4.1. Sediment Physical Properties and Appearance………………………………. 26
4.2. Water, Soil and Carbon Content Analysis…………………………………… 26
4.3. Total Sediment PAH Content Analysis……………………………………… 27
4.4. Equilibrium PAH Partitioning Between Coal Tar and Water………………...29
4.5. Bench-Scale Sediment Flow-Through Experiments………………………….32
4.6. Groundwater Flux Through Sediment……………………………………….. 34
4.7. Advection-Dispersion Experiments ………………………….……………… 38

5) Preliminary Conclusions ………..…………………….……………………………. 44


5.1. Additional Considerations………………………………………………….... 44

6) References ……………………………………………………………………….. 46

7) Appendix 1 ……………………………………………………………………….. 48

2
1) Summary

Equilibrium partition coefficients of coal tar components were measured on whole sediment
collected from the Grand Calumet River at the former manufactured gas plant in Hammond,
Indiana. Ten sediment samples were analyzed for water, volatile organic and total organic
content, yielding similar results with depth for all samples. Samples were then analyzed for 16
aromatic hydrocarbons common to coal tar. Total estimated coal tar varied among samples, but
coal tar composition was found to be almost constant. Estimated mole fractions of all 16
compounds were very similar within the coal tar of each sample. These mole fractions and
Raoult's law were used to predict aqueous PAH concentrations for water in equilibrium with the
coal tar. The ten samples were also equilibrated with water, and the resulting aqueous samples
were analyzed for the same PAHs. Measured and predicted equilibrium aqueous PAH
concentrations compared very favorably, implying that Raoult's law is accurate for
approximation of PAH partitioning between coal tar and water. In addition, sediment flow
through experiments were conducted to determine the validity of assuming equilibrium
partitioning between coal tar and pore water. Concentrations of five PAHs in effluent pore water
were measured, and were found to be at 20-56% of measured equilibrium values. Given that the
experiment used only 5 cm thickness of sediment, and an advective flow rate (≈10 cm/ day)
approximately 1000 times that of the in-situ flow rate calculated using hydrogeological
assessments, it is most likely valid to assume that equilibrium conditions are reached between
pore water and pure-phase coal tar in the sediment.

3
2.0 Background

From approximately the 1940's until the 1950's, gas generated at manufactured gas plants
(MGPs) served as the primary energy source for heating, lighting, and cooking in the
Northeastern United States. These plants eventually became obsolete because of the availability
of an extensive network of natural gas pipelines constructed in the mid 20th century, and
subsequently abandoned (Gratz et al., 1996). There are now over 3,000 former MGP sites in the
United States (Golchin et al., 1997), many of which were built on riverfronts, where raw
materials such as coal could be easily and cost-effectively shipped during that time period.

Gas was manufactured at MGPs through extraction from coal. A major by-product of the coal
pyrolysis process is coal tar, a dense non-aqueous phase liquid (DNAPL) containing a variety of
polycyclic aromatic hydrocarbons (PAHs), volatile organic compounds (VOCs), phenolic,
sulfuric and nitrogenous compounds and some metals (Golchin et al., 1997). Stored coal tar was
spilled at many MGP sites, either voluntarily, unknowingly or accidentally, resulting in extensive
soil and groundwater contamination at most of those sites (Mahjoub et al., 2000). Furthermore,
low water solubility and slow rates of biodegradation of coal tar constituents cause these
contaminants to persist in soil and become a long-term source of potential groundwater
contamination (Khodadoust et al., 2000).

One such forme r MGP site is located in Hammond, IN, adjacent to the Grand Calumet River.
This plant was used for the production and/or storage of manufactured gas from 1901 to 1950.
Residuals from the MGP operation, in particular pure-phase coal tar (DNAPL), have been found
in the soil and groundwater at the site, as well as in the river sediment (ThermoRetec, 2001).
This DNAPL plume is currently a constant source zone of contamination to the river, as NAPL
contaminants are in direct contact with the river water. This continual source of contamination
to the river will persist until it is depleted, removed, sequestered or degraded. Due to the size of
the plume, low groundwater velocity, slow subsurface degradation rates and the low aqueous
solubilities of the constituents, natural depletion of the plume will take an inordinately long time.
Thus, some sort of remediation effort is necessary at the site to reduce potential risks. If
sediment remediation in Hammond, IN proves to be successful, the site could serve as a model
for similar efforts at other MGP sites.

2.1 Polycyclic Aromatic Hydrocarbons . The term polycyclic aromatic hydrocarbon (PAH) is
used to describe any compound containing multiple aromatic carbon rings fused together.
Compounds containing only one aromatic ring are known as mono-aromatic hydrocarbons
(MAHs). Table 2.1 shows the molecular structure and some physical and chemical properties of
those aromatic hydrocarbons found in coal tar that have been routinely qua ntified in this project.

PAHs are extremely persistent in the environment, and have been detected in water, soil,
sediments, food, and the atmosphere (Ferreira, 2001). These compounds are often introduced
into the environment through anthropogenic activities, and are most commonly found as the by-
products and derivatives of coal. Industrial sources such as steel mills and manufactured gas
plants that include coal processing as a major part of their operatio n, are therefore some of the
heaviest contributors to PAH contamination.

4
Table 2.1 Properties of Aromatic Hydrocarbons Commonly Found in Coal Tar1
CAS Chemical Chemical MW, -log Sw , -log Sw L ,
Compound log Kow
Number Formula Structure (g / mol) (log M) (log M)
Ethylbenzene 100-41-4 C8 H10 106 2.8 3.15

m-Xylene 108-38-3 C8 H10 106 2.732 3.202

o-Xylene 95-47-6 C8 H10 106 2.76 3.12

p-Xylene 106-42-3 C8 H10 106 2.77 3.18

1,3,5 Trimethyl-
108-67-8 C9 H12 120 3.4 3.42
benzene
1,2,4 Trimethyl-
95-63-6 C9 H12 120 3.33 3.65
benzene
Naphthalene 91-20-3 C10 H8 128 3.06 3.36
2-Methyl-
91-57-6 C11 H10 142 3.623 4.112
naphthalene
Acenaphthylene 208-96-8 C12 H8 152 4.023 4.004

Acenaphthene 83-32-9 C12 H10 154 3.983 4.204

Fluorene 86-73-7 C13 H10 166 4.08 4.18

Phenanthrene 85-01-8 C14 H10 178 4.46 4.57

Anthracene 120-12-7 C14 H10 178 4.48 4.54

Fluoranthene 206-44-0 C16 H10 202 5.08 5.22

Pyrene 129-00-0 C16 H10 202 5.35 5.13

Benzo(a)-
56-55-3 C18 H12 228 5.96 5.91
anthracene

Chrysene 218-01-9 C18 H12 228 5.293 5.814


1
All data are from Schwarzenbach et al., 1993 at 25o C unless otherwise noted; 2 Chiou and
Schmedding, 1982; 3 Lee et al., 1992; 4 Schwarzenbach et al., 2003.

In Table 2.1, molecular weight is abbreviated as MW, S w is the pure-phase aqueous solubility of
those compounds that are liquid at room temperature, S wL is the hypothetical subcooled liquid
solubility for those compounds that solids at room temperature, and Kow is the octanol-water
partitioning coefficient. For compounds that are solids at room temperature (25°C), the
subcooled liquid solubility (L) is an important parameter as it basically defines what the
‘hypothetical’ water solubility of the compound would be at room temperature if the chemical
were a liquid at this temperature; hence, an adjustment to the true water solubility is made due to
the crystal energy of the solid phase. This is important because subcooled liquid solubilities of
solids, rather than there true solubilities, correlate to liquid- liquid distribution coefficients, and

5
each constituent in the coal tar exists in a liquid phase in contact with water. Subcooled liquid
solubilities are found by extrapolating the vapor pressure vs. temperature relationship from the
liquid phase into the solid phase, and then multiplying the solid phase aqueous solubility by the
ratio of Po (25° liquid)/ Po (25° solid) (see Schwarzenbach et al., 2003).

2.2 PAH Toxicity. The toxicity of PAH compounds is a very complex topic, as toxicity is
specific to each: (i) compound, (ii) affected organism, (iii) exposure pathway, (iv) exposure time,
and (v) set of surrounding environmental conditions. As a result of this complexity, only limited
specific data are available regarding PAH potency and toxicity in the environment, and even less
is known about the effects of PAH mixtures, or PAH compounds in the presence of other
contaminants (Eljarrat et al., 2001).

In general, polycyclic aromatic hydrocarbons are categorized as being nonpolar narcotic (i.e.,
anesthetic) chemicals. As such, in theory, they have a constant, toxic threshold for exposure
concentration and time for a given organism. PAHs tend to bioaccumulate in organisms,
particularly in blood, liver, muscle, gills, heart and fat tissues. Furthermore, a maximum steady-
state residual accumulated concentration exists for PAHs in organism tissue, dependent upon the
specific compound, the type of biomass (tissue) and the rate of PAH metabolism in the organism
(Lee et al., 2002a).

The effects of PAHs and the mechanisms of PAH toxicity depend upon the severity and duration
of exposure. Acute lethality from PAHs is usually caused by non-polar organic narcosis, the
partitioning of contaminants into cell membranes, which changes the properties of the
membrane. Sub- lethal effects are varied and occur through different processes (Lotufo, 1998).
In addition to narcosis, many PAH compounds are either known or suspected carcinogens. Sub-
lethal concentrations damage reproduction capability by impairing embryonic development in
some organisms (Lotufo, 1998). Some PAHs have been shown to exhibit toxicity effects similar
to that of tetrachlorodibenzo-p-dioxin (TCDD), which induces liver microsomal aryl
hydrocarbon hydroxylase in organisms. The potency and effects of PAH toxicity on hepatoma
cells has been compared to that of TCDD (Eljarrat et al., 2001).

Metabolites of PAHs, the by-products of biodegradation, have typically been shown to be less
toxic than their parent compounds. In this way, PAH metabolization can act as a sort of
detoxification process (Lee et al., 2002a). It is therefore possible for certain PAH metabolizing
organisms to thrive in a highly contaminated environment, provided that the PAH degradation
rate in the organism is greater than the PAH flux into the organisms, maintaining a PAH
concentration within the organism below a certain toxic threshold. In such organisms, PAHs are
degraded resulting in a body burden less than that estimated via thermodynamic bioaccumulation
models.

Toxicity to small animals and organisms is typically measured in time-dependent LC 50 values.


LC50 values represent the concentration that causes 50% lethality in a set of organisms (the
median lethal concentration) over a given time span (Lee et al., 2002a). Long-term (t=∞)
aqueous LC 50 values to the benthic organism Hyalella azteca have been measured at 4.19 µM
(0.696 mg/ L) for fluorene, 3.21 µM (0.571 mg/ L) for phenanthrene, and 0.51 µM (0.103 mg/ L)
for pyrene (Lee et al., 2002b). Toxicity becomes much more difficult to predict in the presence

6
of multiple contaminants, as in the case of coal tar, because each compound contributes to the
overall toxicity. Under such conditions, each individual compound will display a lower LC 50
value.

Bioavailability refers to the extent to which a contaminant is available for uptake by an


organism. When an orga nism is exposed to a contaminant solely through water, it is simple to
estimate the concentration of contaminant in that organism, because there is only one pathway
through which it can accumulate. Complications arise, however, when estimating bioavailability
to benthic organisms, due to the complexity of the sediment matrix. Possible exposure pathways
in sediment include pore water, pure phase contaminant and sediment particle ingestion, and
exposure through grazing on other contaminated organisms. Processes such as aging, phase
partitioning and sorption further complicate any approximation (Lotufo, 1998). Because of these
processes, a fraction of the contaminant may not be bioavailable, and thus benthic organisms
may not be able to ingest and accumulate that fraction.

Bioavailability of PAHs can potentially be significantly less than total concentration, leading to
the overestimation of sediment toxicity. Failed attempts at bioremediation with dredged PAH
contaminated sediments have shown that often much of the PAH contamination cannot be
removed in this way. Bioremediation is thus under scrutiny as a way to remediate such
sediments because it so often fails to meet regulatory cleanup objectives that only consider total
concentration. It is believed that the cause for the poor performance of bioremediation is low
bioavailability of PAHs in the sediments (Talley et al., 2002).

2.3 Aqueous PAH Photodegradation. Polycyclic aromatic hydrocarbons have been shown to
degrade in the presence of ultraviolet light, when in the pure phase or in another phase, such as
water. Research conducted on the photodegradation of PAHs dissolved in natural waters
compared to pure water has produced mixed results, although it has definitely been shown to be
significant in both media. Typically, PAH photodegradation is assumed to occur following a
psuedo- first order reaction. The table below lists several PAHs and first-order decay constants in
natural waters, measured near the water surface on July 21 at noon (Fasnacht and Blough, 2002).

Table 2.2 Measured Aqueous Photodegradation Rate Constants for PAHs


PAH Compound k p x 104 (s -1 ) k p (h-1 ) C/C0 after 5 hours
Acenaphthalene 0.30 0.108 0.58
Anthracene 4.7 1.692 0.00021
Benzo(a)anthracene 5.0 1.8 0.00012
Benzo (a)pyrene 21 7.56 0.00000
Benzo(b)fluoranthene 0.30 0.108 0.58
Benzo(k)fluoranthene 0.30 0.108 0.58
Chrysene 0.90 0.324 0.20
Fluoranthene 0.050 0.018 0.91
Fluorene 0.0090 0.00324 0.98
Perylene 4.4 1.584 0.00036
Phenanthrene 0.090 0.0324 0.85
Pyrene 3.9 1.404 0.00089
Reference: (Fasnacht and Blough, 2002)

7
This data is significant in assessing PAH contamination in surface waters, because it implies
rapid degradation of many PAH compounds in the water column.

Photo-enhanced toxicity refers to an increase in the toxicity of a chemical to an organism under


ultraviolet (UV) light conditions (Barron and Ka’aihue, 2001). Natural UV-B radiation alone
has been shown to be somewhat toxic at molecular, cellular, and organism levels, particularly in
the UVB range (280-320 nm), but causes far more severe damage in the presence of certain
chemical contaminants (Cleveland et al., 2000). Consequently, many compounds that are
relatively non-toxic under low-UV conditions can become acutely toxic in the presence of UV
light. 96 hour LC50 values of anthracene for juvenile shellfish were shown to be 190-1,800
times smaller (190-1,800 times the toxicity) in the presence of simulated sunlight than without.
Prior to this study, anthracene had been considered almost non-toxic to juvenile shellfish (Oris
and Giesy, 1985).

Toxicity photo-enhancement occurs through two mechanisms. The first is known as in- vitro
photomodification, or direct photo-oxidation of a contaminant, where the photoproduct is more
toxic than the original contaminant itself (Cleveland et al., 2000). These oxygenated products
(oxy-PAHs) are seldom regulated, and have been found in sediments at concentrations similar to
those of their parent PAH compounds (Lampi et al., 2001). The second mechanism is known as
in- vivo photosensitization, or indirect photosensitized oxidation, where bioaccumulated
photoactive chemicals become toxic with the exposure of the organism to solar radiation
(Cleveland et al., 2002). In this process, upon exposure to UV radiation, bioaccumulated
contaminants release light energy to other molecules in the form of free radicals, which cause
toxicity through tissue damage, rather than a narcosis mechanism (Barron and Ka’aihue, 2001).
In marine organisms, toxicity photoenhancement of oil products is believed to be primarily
caused by this photosensitization oxidation process, with photomodification contributing less to
photo-enhanced toxicity of the chemical (Barron and Ka’aihue, 2001).

Constituents in weathered oil products, including many PAHs, are particularly prone to have
photo-enhanced toxicity characteristics. Such compounds can be 2 to more than 1000 times
more toxic in the presence of UV than in conditions of very low or negligible UV intensity.
Some studies suggest that 3-5 ring compounds are the most phototoxic PAHs (Barron and
Ka’aihue, 2001). It is therefore important to consider solar radiation and photodegradation as
contributing parameters when assessing the toxicity of PAHs in the aquatic environment (Oris
and Giesy, 1985). Neglecting toxicity photoenhancement effects may thus lead to the
underestimation of toxicity (Cleveland et al., 2000).

2.4 Equilibrium Partitioning. In order to effectively understand, model, and predict the
behavior of contaminants in the natural environment, it is necessary to first have a good
understanding of how the specific contaminants of interest partition between phases. It is
therefore of interest to understand the equilibrium partitioning between water, air, sorbed (soil)
and pure contaminant phases.

In real environmental situations, equilibrium conditions may or may not actually be reached, but
can often be assumed or be used as a limiting condition when considering phase distribution.
When considering equilibrium conditions, it is convenient to use equilibrium partition

8
coefficients. The equilibrium partition coefficient for component i, Ka ,i, between the α phase
and the water (w) phase is defined by,

C
α,i
K
α,i C
w, i (2.1)

where Cα,i and C w,i are the concentrations (moles/ L) of constituent i in the α phase and the water
phase, respectively. Some commonly used equilibrium partition coefficients are Kd, the soil-
water partition coefficient (L water/ kg soil), KH, the air-water partition coefficient or the
dimensionless Henry's constant (L water/ L air), and Kow, the octanol- water partition coefficient
(L water/ L octanol),

Cs Cg Co
Kd KH Kow
Cw Cw Cw
(2.2 - 2.4)

where Cs (mg i/ kg soil), Cg (mg i/ L air) and Co (mg i/ L octanol) are the concentrations of
constituent i in the sorbed (soil), gaseous (air) and octanol phases in equilibrium with water,
respectively. In the present case, the non-aqueous (organic) phase is coal tar,

C
ct , i
K
ct , i C
w, i (2.5)

where Cct,i and Cw,i represent the concentrations of component i in the coal tar phase (moles i / L
coal tar) and water phase (moles i / L water), respectively. The following derivation indicates
that values of K have a strong thermodynamic basic. To stress the general applicability of the
final equation, it is derived using the subscript a (rather than ct) for the non-aqueous phase.
Recall that for convenience, the pure liquid or sub cooled liquid of each component may be
defined as that component ’s reference state (i.e., activity (a) = 1.0). By definition, the activity of
component i will be equal in both the water phase and the non-aqueous mixed solvent phase,

aa ,i = aw,i (2.6)

where aa ,i and aw,i are the activities of component i in the non-aqueous and aqueous phases,
respectively. The activity within a phase is defined by,

ai = γi xi (2.7)

where xi and γi are the mole fraction (moles of i/ total moles of α) and activity coefficient,
respectively, of constituent i in any phase. Mole fraction, xi, is related to molar concentration,
Ci, (moles / L) through the molar volume of the phase, V (L of solute / mol of phase),

xi = Ci ?V (2.8)

9
where V is further defined by MW/ ρ, where MW and ρ are the molecular weight (g / mol) and
density (g / L) of the phase. Although eq 2.8, per se, introduces no deviation from ideality, its
application to the present case is an approximation, since ideal solutions assume all molecules
are of identical size. Hence, even with V and MW of coal tar defined as the geometric average
value of all contributing components, molecular size differences among components within the
non-aqueous (coal tar) phase will lead to some non- ideal behavior.

For the pure liquid or subcooled liquid organic phase component i in equilibrium with water, eqs
2.6 and 2.7 combine to produce,

γpure,i xpure,i = γw,i xw,i (2.9)

where by definition, and assuming water solubility in the organic phase is negligible, xpure,i = 1.0
and γpure,i = 1.0 (recall the liquid is defined as the reference state). Following these assumptions,
γw,i xw,i = 1.0, and,

(xsat )w , i 1
γw , i
(2.10)

where (xsat )w,i is the mole fraction solubility of i in the aqueous phase. Substituting eq 2.10 into
eq 2.8,

xsat w,i = Sw,iVw (2.11)

where S w,i ( = Cw,i at saturation) is the pure-phase aqueous solubility of component i (moles i/ L
water). Combining eqs 2.10 and 2.11 gives,

1
S ⋅V
w, i w γw , i
(2.12)

If coal tar is the organic phase in equilibrium with water, combining equations 2.6 and 2.7 gives,

γct,i xct,i = γw,i xw,i (2.13)

Combining eqs 2.8 and 2.13 yields,

γct,i Cct,i Vct = γw,i Cw,iVw (2.14)

Combining equations 2.5, 2.12, and 2.14 defines the coal tar-water partitioning coefficie nt,

C
ct , i 1
K
p C ⋅γ ⋅V
w , i ct , i ct
w, i S
(2.15)

10
Because coal tar is composed of significant mass fractions of chemically- similar PAH
compounds, a good approximation is to assume that each individual organic constituent behaves
ideally within the complex organic mixture. Recall that this is the basic assumption of Raoult's
law. This assumption (i.e., γct,i = 1.0) further simplifies eq 2.15, defining values of Kp for coal tar
components as being proportional to the inverse liquid or subcooled liquid solubility. It also
indicates that aqueous phase concentrations of coal tar components are proportional to the
product of their aqueous solubility and their mole fraction concentration within the coal tar phase
(eqs 2.8 and 2.15),

Cw,i = xct,i Sw,i (2.16)

Lee et al. (1992) applied eq 2.16 to estimate aqueous phase concentrations of PAHs in
equilibrium with coal tar samples collected from several locations. These predictions were
compared with measured equilibrium aqueous concentrations and compared very favorably. It
was shown that γi-ct is approximately equal to 1.0, and that equilibrium aqueous concentrations
can be accurately predicted if the PAH mole fractions in coal tar are known. Making this
assumption and taking the base-10 log of equation 2.15,

log Kp = - log S w,i - log Vct (2.17)

Further substitution provided,

log Kp = - log S w,i - log (MWct / ρct ) (2.18)

This formula implies a linear inverse relationship between log Kp and log Sw,i, where the slope is
-1 and the intercept is a function of MWct /ρct . Experimental results published by Lee et al.
(1992) , and Lane et al. (1992) show a good correlation between observed and theoretical
behavior based upon equation 2.18.

2.5 Sorption. Sorption is defined as the process by which chemicals interact with and "dissolve "
in or adhere to a solid phase, such as soil. Two basic mechanisms of sorption exist, referred to
generally as adsorption and absorption. Adsorption refers to the sorption of a chemical onto a
two-dimensional surface, while absorption describes the dissolution of a chemical inside of a 3-
dimensional matrix (Schwarzenbach et al., 1993). Sorption mechanisms are important to
consider when assessing subsurface contaminant flux, as they will significantly affect the rate at
which contaminants move through soil.

After undergoing sorption to a solid surface or matrix, the chemical is then considered to have
transferred into the solid or "sorbed" phase. Phase transfer into the sorbed phase can take place
from aqueous, gaseous or organic phases, and can be shown to be governed by the same type of
equilibrium partition coefficient (K d) as transfer into other phases,

C
s, i
K
d C
w, i (2.19)

11
moles
i
where, C
s, i kg
solids

 moles i 
 kg solids  L
The units on Kd are defined as, K
  water
d
 moles i  kg
solids
 Lwater 
 

In addition, at equilibrium, the sorbed concentratio n of constituent i reaches a limit when the
water solubility of i is reached. Once such conditions exist, any amount of additional chemical
present in the system will exist as an organic liquid phase (or precipitated solid). River sediment
saturated with coal tar is an example of this. Once the solubility of coal tar constituents are
reached in the sediment pore water, no more can be sorbed, and coal tar components will be
present as an organic liquid phase within the pores. At this saturated condition, transfer from the
organic liquid phase to the aqueous phase will constitute a long-term source of water and soil
contamination until the organic liquid phase has been depleted.

2.6 Groundwater Flow. The fundamental theories of groundwater flow are mainly based upon
Bernoulli's equation and Darcy’s Law. Bernoulli's equation relates fluid pressure, velocity and
elevation between two points within a continuous fluid. The equation is written as follows
(Charbeneau, 2000),

p v 
2 p v 
2
h  + z+   + z+ 
γ 2g 
1 γ 2g 
2 (2.20)

where h = hydraulic head (length), p = fluid pressure (force/ area), γ = specific weight of fluid
(force/ volume), z = elevation with respect to an arbitrary, constant datum (length), v = fluid
velocity (length/time) and g = acceleration due to gravity (length/time2 ). In pressurized flow
situations such as water distribution pipes, where both pressure and velocity are significant, all
three terms of this equation (pressure, elevation and velocity) must be considered. However, in
non-pressurized flows, such as in open channels, the pressure term can be ignored in the
equation. Similarly, in extremely low velocity situations the velocity term can be neglected. The
velocity term is thus typically ignored in groundwater applications, where velocities rarely
exceed a few centimeters per day. In this way, after removing the ve locity term from Bernoulli's
equation, we are left with what is referred to as piezometric head, and commonly represented by
the Greek letter φ (Charbeneau, 2000),

 p + z  p + z
φ    
 γ 1  γ 2 (2.21)

12
Piezometric head is measured by a piezometer, which consists of a partially screened well
inserted into the ground to a desired depth. Simple piezometric measurement is shown in Figure
2.1,

Figure 2.1 A Simple Piezometer

The water surface in each well is known as the piezometric surface. The distance from the
reference datum to the piezometric surface (p/γ + z) is the piezometric head measured at the
screen of each well.

Henry Philibert Gaspard Darcy (1803-1858) developed a relationship for flow through porous
media after a series of experiments conducted in Dijon, France in 1855-1856. His findings,
published in the 1856 “Les Fontaines Publiques de la Ville de Dijon”, stated that the loss of
hydraulic head through porous media was proportional to the rate of fluid flow (Charbeneau,
2000). This leads to the following relationship, for a medium of finite volume,

φ1 − φ2
Q K⋅ A ⋅
L (2.22)

Where Q = volumetric flow rate (volume/time), K = saturated hydraulic conductivity


(length/time, empirically derived), A = cross-sectional flow area, φ = hydraulic head and L =
distance along the flow path. Simplifying, we find,


q K⋅
dL (2.23)

or q K⋅ I (2.24)

where q = Q/A = Darcy flux or specific discharge (length/time) and dφ/ dL = I = hydraulic
gradient (length/ length). Darcy flux represents the overall movement (speed) of the solvent
front. Because of the tortuous path that fluid must follow in a solid matrix, however, Darcy flux
does not represent the true velocity of the individual pore fluid molecules in such a matrix. The

13
true velocity of the fluid moving through the matrix is thus larger, and is expressed as
(Domenico and Schwartz, 1998),

q
v
ne (2.25)

where v = average pore-scale linear velocity and ne = effective porosity of the soil matrix. The
effective porosity of the soil matrix represents the fraction of total matrix volume that is
available for the fluid to flow through, not including “dead end” or non-interconnected pores
(Fetter, 1999).

Aqueous phase contaminants are transported through water via three processes, which are
advection, diffusion and mechanical dispersion (Fetter, 1999). For simplification, it will be
assumed that the sediment and the sand cap are each homogeneous media, and that groundwater
flow is steady and one-dimensional. The possibility for error in these assumptions must be
considered in making a final evaluation of results. Advection refers to the movement of flowing
water. Advective flux is thus the process of dissolved contaminants being carried along by
moving water, and is totally dependent upon the rate and direction of water flow. One
dimensional mass flux of contaminant due to advection is found as the product of the aqueous
concentration of the contaminant and the Darcy flux of the carrier water. Thus (Fetter, 1999),

Fa q ⋅C i , w
(2.26)
or
Fa v ⋅ne ⋅Ci , w (2.27)

where Fa = advective flux of constituent i (mass/ area · time) and Ci,w = aqueous concentration of
constituent i (mass/ volume). If advective flux is considered on its own, without the effects of
diffusion or dispersion, a behavior known as a plug flow is the result. Plug flow moves along as a
constant volume, constant concentration “plug” that displaces all of the water in front of it
without any mixing. The ideal condition of plug flow is conceptually important for
understanding contaminant flux in an aqueous medium; however it does not occur in real world
applications (Fetter, 1999).

Diffusion refers to the natural, molecular movement of a contaminant from an area of greater
concentration to an area of lesser concentration, until equilibrium conditions are reached.
Diffusive flux occurs whenever a concentration gradient exists, regardless water flow conditions.
Fick’s first law attempts to explain and model this flux, by assuming that the rate of diffusive
flux is proportional to the concentration gradient. Fick’s first law, for one dimension, is
expressed as (Fetter, 1999),

dC i , w
Fd − Dd ⋅
dx (2.28)

14
where Fd = diffusive flux of constituent i (mass/ area · time) in the x direction, Dd = the diffusion
coefficient (length2 /time; empirically derived), Ci,w = aqueous concentration of constituent i
(mass/ volume) and x = distance in the x direction (length). The quantity dC/ dx is the
concentration gradient (mass/ length4 ). Dd is typically considered to be constant for all directions
in a water medium at a given temperature for a given compound. In porous flow, effective
porosity must be taken into account, and thus diffusive flux becomes,

dCi, w
Fd −n e Dd⋅
dx (2.29)

Fick’s second law takes into account the time-dependency of systems where Ci,w changes with
time. Fick’s second law is written as (Fetter, 1999),

2
δC δ ⋅C
Dd ⋅
δt dx
2
(2.30)

where δC/ δt = change in concentration with time (mass/ volume/time). Due to the tortuous
pathways within porous media, the diffusion coefficient of a contaminant in water alone is not
accurate for use in a soil matrix, and an effective diffusion coefficient must be used. This
effective diffusion coefficient must be empirically estimated. One such estimation is (Fetter,
1999): De = ωDd, where ω is a dimensionless, empirically derived constant related to the matrix
tortuosity, Dd is the diffusion coefficient of the constituent in water, and De is the effective
diffusion coefficient of the constituent in matrix pore water. The parameter ω has been
estimated to be approximately 0.7 in laboratory sand column studies (Fetter, 1999). Another
estimation of effective porosity assumes a normal, Gaussian distribution. This estimate is
written as (Fetter, 1999),

(σc)2
De
2t (2.31)

where σc2 = normal Gaussian variance (length2 ) and t = time. The following relationship can
then be derived (Fetter, 1999),

 x 
Ci( x , t) C0⋅erfc  
 2 De ⋅t  (2.32)

where Ci(x,t) = concentration of constituent i at a distance x from the source at time t, C0 = initial
concentration of constituent i in the source zone and erfc( ) = the complimentary error function,
which is related to the Gaussian distribution function. The complimentary error function is
related to the error function, erf( ), according to the following equation (Fetter, 1999),

erfc(B) 1 − erf (B) (2.33)

15
where (Fetter, 1999),

B
2 ⌠ −t
2
erf ( B ) ⋅ e dt
π ⌡0
(2.34)

Equation 2.34 cannot be solved analytically. An analytical approximation of the error function is
(Fetter, 1999),

 −4B2 
erf ( B) 1 − exp  
 π  (2.35)

Mechanical dispersion refers to the mixing of aqueous phase contaminant within pore water, due
to turbulent and non- uniform flow conditions. This happens via three processes in a soil matrix.
The first mechanism is the mixing that occurs from water flowing faster through the center of
pores than along the sides. The second method of mechanical dispersion results from the
tortuosity of the medium, which causes contaminant particles to go through different flow paths
of varying length. Thus two particles, moving the same linear distance, will always follow
routes of different length. The third pathway for mechanical dispersion occurs because of the
great variety in pore sizes. Water flowing through larger pores will flow faster than water
flowing through small pores, causing mixing and turbulence due to different pore velocities
(Fetter, 1999).

If we assume that mechanical dispersion can be modeled in the same way as diffusion, using
Fick’s law, then a coefficient of mechanical dispersion can be empirically derived. This
coefficient is assumed to be proportional to the average linear velocity of flow in the soil matrix,
so that (Fetter, 1999):

Dlmd = α ivi and Dtmd = α jvi (2.36-2.37)

where Dlmd = coefficient of longitudinal mechanical dispersion, Dtmd = coefficient of transverse


mechanical dispersion, vi = average linear velocity in the i (longitudinal) direction (length/time),
α i = dynamic dispersivity in the i direction (length) and α j = dynamic dispersivity in the j
(transverse) direction (length).

Mechanical dispersion in the direction of flow will result in a dilution of the solute at the
advancing edge of solvent flow. This is referred to as longitudinal dispersion. Mechanical
dispersion in the direction normal to flow occurs because flow paths tend to diverge at the pore
scale, and is known as transverse dispersion (Fetter, 1999).

The effects of diffusion and of mechanical dispersion cannot, however, be separately detected in
groundwater applications. It is thus necessary to add the effects of both, resulting in what is
known as hydrodynamic dispersion. Because Fick’s law is assumed to apply to mechanical
dispersion as well as to diffusion, we can combine the diffusion coefficient and coefficients of

16
mechanical dispersion into a singular hydrodynamic dispersion coefficient, D. Longitudinal and
transverse hydrodynamic dispersion coefficients are defined as,

DL α L⋅ vi + De DT α T⋅ vi + De
and (2.38-2.39)

where DL = longitudinal hydrodynamic dispersion coefficient and DT = transverse hydrodynamic


dispersion coefficient. Hydrodynamic dispersive flux is thus found to follow the same
relationship as diffusive flux. For one dimension,

dCi, w
FDisp −ne DL ⋅
dx (2.40)

Furthermore, in the same way that diffusion is modeled using a Gaussian assumption,
hydrodynamic dispersion can be modeled using the same assumption, producing the following
relationship,

(σ L)2 (σ T )2
DL DT
2t and 2t (2.41-2.42)

where σL2 and σT2 = normal longitudinal and transverse Gaussian variance (length2 ), and t =
time. By considering advective contaminant flux, hydrodynamic dispersion and the conservation
of mass, a total relationship for advective and dispersive contaminant transport has been derived.
For one dimensional flow in a homogeneous, isotropic porous medium,

2
δ ⋅C δC δC
DL⋅ − vx ⋅
δx
2 δx δt
(2.43)

The analytical solution to this partial differential equation, for a continuous contaminant source,
is (Fetter, 1999):

C ( t) 1   L − vx⋅ t   vx⋅ L   L + vx ⋅t  
⋅ erfc  − exp  ⋅ erfc 
C0 2
  2 ⋅ DL⋅t   DL   2 ⋅ DL⋅ t   (2.44)

where C0 = constant concentration leaving the source (mass/ volume), and C(t) = time-dependent
concentration at a point that is L distance from the source (mass/ volume).

Retardation refers to the slowing of contaminant flux through a porous medium due to sorption
of contaminant onto the medium. Under these conditions, the flux of dissolved contaminant is
slower than the flux of water. In the presence of retardation, the advective-dispersive equation
(equation 2.43) becomes (Domenico and Schwartz, 1998),

17
DL δ 2 ⋅C vx δC δC
⋅ − ⋅
Rf 2 Rf δx δt
δx (2.45)

where Rf is the retardation factor, defined as (Domenico and Schwartz, 1998),

 1 − ne 
Rf 1+  ⋅ ρs ⋅ Kd
 ne  (2.46)

where ne = effective porosity of the soil matrix, ρs = density or specific gravity of the soil
particles (usually assumed to be 2.65 g / cm3 ) and Kd = soil / water partition coefficient (volume /
mass). With retardation, the analytical solution to the advective-dispersive equation then
becomes (Domenico and Schwartz, 1998),

C ( t) 1   R f ⋅L − v x ⋅t   v x ⋅L   Rf ⋅ L + vx⋅t  
⋅ erfc  − exp   ⋅ erfc 
C0 2   2 ⋅ R f ⋅DL⋅ t   DL   2 ⋅ Rf ⋅ DL⋅t   (2.47)

The velocity of the contaminant front will be related to the water velocity by the following
relationship (Domenico and Schwartz, 1998),

vw
vc
Rf
(2.48)

Figures 2.2 and 2.3 illustrates the advancement of a contaminant front, at a given time, both with
and without retardation, and the breakthrough curve of a contaminant, 20 cm from the source for
the flow and media conditions reported in the figure captions.

18
1
0.9
Rf
0.8
1
0.7
3
C/Co 0.6
0.5
0.4
0.3
0.2
0.1
0
0 50 100 150 200
Distance from Source (cm)

Figure 2.2 Concentration Profile at 1 Year with q = 0.1 cm / day,


D = 1.0 cm2 / day, ne = 0.4, and R = 1 and 3.

Conversely, at a fixed distance, the solvent front advances as in the chart below.

1
0.9
0.8
0.7
0.6
C/Co

0.5
0.4
Rf
0.3 1
0.2 3

0.1
0
0 50 100 150 200 250 300 350
Time (days)

Figure 2.3 Concentration at 20 cm from source versus time with


q = 0.1 cm / day, D = 1.0 cm2 / day, ne = 0.4, and R = 1 and 3.

19
3.0 Methods and Materials

While a great deal of research has been conducted to study and characterize coal tar
contamination of the subsurface at former manufactured gas plant sites, less has been done to
study and characterize coal tar contamination in the sediments of adjacent rivers. The purpose of
this research is to characterize the extent and behavior of PAHs in the sediment at the Hammond,
IN, MGP site. The specific tasks described in this report are:

1) Collection of sediment samples for analysis.


2) Determination of the water, soil, volatile organic content and total organic contents of the
sediment samples.
3) Measurement of the concentrations of 16 PAH and mono-aromatic hydrocarbon (MAH)
compounds in the sediment, and estimate the total PAH and MAH content of each sample.
4) Determination of the equilibrium aqueous concentrations and partition coefficients of the 16
compounds of interest, in water equilibrated with the sediment.
5) Evaluation of bench-scale flow-through studies to confirm the extent to which equilibrium
may occur.

3.1 Materials. A 16 compound PAH standard mixture was purchased from Ultra Scientific, and
individual standards of ethylbenzene, 1,3,5-trimethylbenzene, 1,2,4-rimethylbenzene, m-xylene,
p-xylene and o-xylene were obtained from Aldrich Chemical Company, to be used for compound
identification. Methanol (Methyl Alcohol), dichloromethane (methylene chloride) and hexane
were all obtained from Fisher Scientific International, Inc. Methanol was used to prevent
sediment samples from clumping during equilibration (thus maximizing exposed sediment
surface area and augmenting phase partitioning). Dichloromethane was used to extract PAHs
from the sediment for total sediment concentration analysis. Hexane was used to extract PAHs
from water equilibrated with the sediment, for equilibrium aqueous concentration analysis.
Acetonitrile was purchased from Mallinckrodt, through the Purdue University Chemistry Stores,
and used with water as the eluent for liquid chromatography analysis.

3.2 Sample Collection. In preparation for laboratory studies, Dr. Suresh Rao, Dr. Chad Jafvert,
Dr. Changhe Xiao and Mr. Douglas Lane of Purdue University participated in the collection of
river sediment samples at the Hammond, IN, MGP site on October 23-24, 2001. Figure A1 of
the Appendix 1 shows a map of the sampling locations along the river. Elliot Smith of AscI
(Detroit) and Damon Morris of Retec Corp (Seattle), collected five 4-inch by 8- foot long cores
from a pontoon boat, using a Vibrocorer. After Retec on-site personnel collected small samples
from the cores, the cores were divided into two 4-foot long sections. These 4- foot long sections
were placed into 10 individual stainless steel containers, each representing a depth of 0-4 foot
depth or 4-8 foot depth from one core. The containers were sealed for transport and stored in a
refrigerated room for preservation. Each sample was later homogenized with a paint mixer
attached to a power drill and a subsample from each bucket was immediately placed into a 1
quart Ball® jar. The 10 jars were immediately sealed and stored in a refrigeration room for
preservation. The sediment in each of these jars was homogenized with a spatula prior to
removing material for study.

20
3.3 Water, Soil and Carbon Content Analysis. Water content, ash content, total organic
content (including both “heavy” and volatile organic content), and organic carbon content were
measured to characterize each composite sample. Total volatiles (Water + volatile organic)
content was measured by weighing sediment samples before and after drying at 105° C for 24
hrs. Mass lost during drying includes the pore water in the sediment and any volatile organic
compounds that evaporate at 105° C. The mass remaining after drying at 105° C was taken as
the dry weight of the sediment. The true dry weight includes the volatile organic compounds lost
during drying, however this amount was not considered great enough to warrant correction.
These same dried samples were heated to 500° C for 3 hrs in a muffle furnace, after which they
were again weighed. The mass remaining after heating at 500° C was considered to be the ash
content of the sediment, and the mass lost during heating at 500° C was taken as the “ignitable”
mass. Fractional solids content is equal to the dry mass divided by the wet mass of the sediment.
Fractional ignitable content of a sample is equal to the ignitable organic mass divided by the dry
mass.

The total organic content of each composite sample was estimated from measured values of total
carbon, measured with a Carlo-Erva 1108 elemental analyzer. This instrument is designed to
analyze extremely small samp le sizes (1-10 mg), and therefore pulverization of the sediments
was necessary in order to provide homogenized samples. Each sediment was pulverized using a
Spex CertiPrep 6750 Freezer/Mill pulverizer, prior to sample analysis. Samples of approximately
4 mg were prepared in silver boats, and loaded onto the elemental analyzer. Total organic carbon
values measured by the instrument were multiplied by 1.7 to obtain an estimate of total organic
mass, as described in Schwarzenbach et al. (1993). The 1.7 multiplier may be somewhat less
accurate for this sediment than for most other sources of organic matter because so much of the
organic mass is carbon-rich coal tar (as opposed to natural organic matter). The multiplier 1.7
was used because of the difficulty in estimating a better correction factor. The total wet
fractional organic content was obtained by dividing the total organic mass by the wet (original)
mass of the sample. Total dry fractional organic content was found by dividing the wet
fractional orga nic content by the solids content of the sediment.

Volatile organic content was calculated as the difference between total and “heavy” organic
contents. Water content was calculated by subtracting volatile organic content from the total
volatile content (water + vo latile organic) found after heating at 105°C. Results for all of these
calculations can be found in Table A4 of Appendix 1.

3.4 Equilibrium Partitioning of PAHs between Tar and Water. There is no prescribed EPA
method for the extraction of PAHs from coal tar, and thus a method sufficient for the extraction
of the 16 compounds of interest was developed. For each of the 10 sediments, approximately 2 g
of wet sediment was placed with 4 mL methanol and 11 mL dichloromethane in a 35 mL Kimax
heavy duty centrifuge tube. Duplicate tubes were prepared for each sample. Dichloromethane
was used as the solvent because of its strong affinity for the compounds of interest. Methanol
was added to prevent the sediment from clumping, and to thus force maximum sur face area of
the sediment particles in the solvent and augment the extraction process. The tubes were sealed
with Teflon- lined caps and briefly vortexed, before being placed on a rotary shaker at
approximately 60 rpm for 24 hours in a dark, windowless room. Shaking the samples in a dark
environment was preferable to avoid photodegradation of the light-sensitive PAHs. After 24

21
hours of shaking, the tubes were centrifuged at 3,000 rpm for 20 minutes in a Jouan CR422
tabletop centrifuge. 5 mL of supernatant was then removed from atop each tube, and transferred
to another 35 mL Kimax tube with approximately 2g Na2 SO4 (previously heated to 350°C
overnight to dry). These tubes were then shaken for 3 hrs on a wrist action shaker in a dark room.
The Na2 SO4 was added as a drying agent to absorb any pore water from the sediment that had
partitioned into the dichloromethane during equilibration. After 3 hrs on the wrist action shaker,
each tube was centrifuged at 2,000 rpm to settle any particulate matter. Duplicate 1 mL-1.5 mL
samples of supernatant were then removed from each tube and transferred to autosampler vials
for GC-FID analysis. Thus, 4 vials were prepared for each sediment.

Concurrent to sample preparation for total PAH analysis, sample preparation was also performed
for water/coal-tar equilibration experiments. For each of the 10 sediments, approximately 20-25g
wet sediment was mixed with approximately 130 mL distilled water (enough to allow for no air
space) containing 2.5mM CaCl2 in a 150 mL Kimax® glass bottle. Duplicate bottles were
prepared for each sediment. CaCl2 was added to better simulate the natural waters in the river.
Each bottle was sealed with a Teflon-lined cap, briefly vortexed and left on an enclosed, light-
sealed end-over-end shaker for 48 hours. After 30 minutes settling time, the supernatant from
each bottle was transferred via pipette into 4 stainless steel centrifuge tubes each (30 mL per
tube). These tubes were immediately sealed with stainless steel and rubber caps, which were
wrapped in foil to cover the rubber o-rings. This precaution was taken because rubber will absorb
PAH from the dead space in the tubes. The tubes were centrifuged at 15,000 rpm in a Sorvall
ultracentrifuge to settle the particulate matter left behind by the sediment. 25 mL of supernatant
was then transferred from each of these 4 tubes to a 150 mL separatory funnel, to which 2.0 mL
of hexane was also added. Each funnel was shaken vigorously by hand for 5 minutes, to allow
for equilibration and for partitioning of the PAHs into the hexane, and then left to settle for a
minimum of 45 minutes. The aqueous phase was then discarded, leaving behind the hexane and a
small emulsified phase in the separatory funnel, both of which were transferred by pasture
pipette into two autosampler vials. Because hexane has a very high log Kow value of 4.11 (K ow =
12,900), while PAHs have similarly high Kow values, it is a valid assumption to neglect the PAH
compounds left behind in the aqueous phase. The two autosampler vials were centrifuged at
3,000 rpm to break the emulsion, and 200 µL of hexane was then transferred from each vial into
clean autosampler vials for GC-FID analysis. Thus, 4 vials were prepared for each sediment.

3.5 Bench-Scale Sediment Flow-Through Experiments. One-dimensional bench-scale


sediment flow-through experiments were conducted in order to yield an approximate
representation of the aqueous PAH loading that could be expected to come directly out of the
sediment under up- flow conditions.

A 30 cm long, 1 inch diameter threaded glass column with Teflon end-caps (Ace Glass, Inc.,
Vineland, NJ) was used for the experiments. The column was loaded from bottom to top with 23
cm of sand, 5 cm of sediment 4N-top (upper 4' of core), and 2 cm of sand. Glass frits were
placed inside the column on both ends to evenly distribute the inflow and collect the outflow.
The top 2 cm sand layer was placed above the sediment to keep the sediment from contaminating
the top frit through direct contact. The column was wrapped with aluminum foil to prevent
photodegradation.

22
Figure 3.1. Column Water flow from bottom to top was established through the column, to
Schematic imitate upward flux through sediment in a river. A FMI model QC
216 lab pump was connected to the column via 1/8” O.D. Teflon
tubing and was controlled via a ChronTrol XT timer. The timer and
pump were programmed to continuously dispense de- ionized, carbon
filtered water to the column, at the pump’s lowest incremental
dispense volume per cycle, for one second every 100 seconds. This
resulted in an average flow of approximately Q=0.015 mL/ min, or an
average Darcy flux in the column of approximately 10.7 cm/ day
(assuming ne=0.4 and q=Q/neπr2 ), which is about three orders of
magnitude faster than the expected flux through the sediment in the
river. Depletion of contaminants from the sediment in the column was
neglected, because of the extremely high sediment concentrations (≈17
g / kg naphthalene), the low volumetric flow rate and the relatively
low measured equilibrium aqueous concentrations (≈13 mg / L
naphthalene). Even at measured equilibrium, at this flow rate
naphthalene would be depleted at a rate of approximately 0.3 mg / day.

Effluent was directed from the top of the column through 1/16" O.D.
stainless steel LC tubing. During sampling, this tubing was pierced
directly through the septum of a sealed autosampler vial prepared with
200 µl acetonitrile. A stainless steel needle was also pierced through
the vial's septum to release air pressure from the vial. The acetonitrile
was used to prevent adsorption of the PAH compounds to the walls of the autosampler vial and
their volatilization. Between samples, waste flow was directed from the stainless steel tubing,
through polyethylene tubing and into a volumetric flask.

The column effluent was analyzed by high-pressure liquid chromatography (HPLC). Liquid
chromatography was used rather than gas chromatography due to the low sample volumes
generated and associated GC and HPLC detection limits. The HPLC consisted of a Waters 600
controller and LC pump, a Shimadzu SIL-9A auto injector, a Knauer Variable Wavelength
Monitor UV detector and a PC running ChromPerfect Spirit® 32-bit Chromatography Data
System software. Solvent flow was directed through the LC pump, into the auto sample injector,
through a Supelco Discovery C18® liquid chromatography column, through the UV detector,
and into a waste receptacle. The software simultaneously controlled both the pump and the auto
injector, while recording UV absorbance readings at a wavelength of 254 nm.

The disadvantage of HPLC analysis is that it is more difficult to obtain good peak separation. A
long solvent-gradient program (55 minutes) was necessary to obtain adequate peak separation,
and yet only five compound peaks were adequately separated and quantified at a wavelength of
254 nm. Those five compounds were anthracene, fluorene, naphthalene, phenanthrene and
pyrene. These five compounds cover a range of PAH properties and should provide an accurate
representation of the extent to which all of the PAHs reached equilibrium in water flowing
through the sediment.

23
De-ionized, carbon-filtered water and acetonitrile were used as solvents for LC analysis. The
solvent gradient program is listed below,

Table 3.1- LC Gradient Program


Time (min) % organic % water 100

0 20 80 80

% solvent
2 20 80 60 acetonitrile
52 100 0 water
40
55 100 0
20
56 20 80
75 20 80 0
0 15 30 45 60 75
Time (min)

The program varied solvent composition with a straight- line slope (% vs. time). The final 20
minutes of the program was meant to re-equilibrate the column and tubing to initial conditions.
Samples were first analyzed with the UV detector at a wavelength of 254 nm. This wavelength
was selected because it has a high absorbance value for a wide variety of compounds. During the
course of analysis, however, it was found that a large ethylbenzene peak was interfering with the
naphthalene peak at 254 nm. To allow fo r analysis of naphthalene, samples were re-analyzed at
276 nm, because at this wavelength naphthalene has a much higher absorbance than
ethylbenzene.

3.6 Advection-Dispersion Experiments. In order to investigate the effect of placing a sand cap
over contaminated sediments, additional column experiments were performed in which aqueous
extracts of coal tar were passed through sand columns to characterize the breakthrough curves.
The retardation coefficients were obtained by performing moment analysis on the breakthrough
curves,

C
∫C pdp
R= 0
(3.1)
C
∫ C0 dp
where, R is the retardation coefficient ; C = effluent concentration ; C0 = influent concentration ;
p = pore volumes = v t / L, where t = time ; v = velocity ; and L = length of the column. The
retardation coefficient, R, is a function of the sorption coefficient Kd of the compound,

ρbK d
R =1+ (3.2)
θ

where, ? b = bulk density of the sand; and ? = porosity of the media. Because the applied
advective flow rate was quite large, local equilibrium can not be assumed. Rather, a simple 2-
box sorption kinetic model was applied to the data. The model assumes instantaneous
equilibrium occurs to a fraction of the sorption sites (F), and that sorption to the remaining sites
(1 - F) is kinetically controlled. The kinetic (mass transfer) coefficient, k2 , and the fraction of

24
sorption sites, F, are the only new additional parameters. Sorption sites that are ‘instantaneous’
and ‘kinetically controlled’ are assumed to have the same partition coefficient. The equations
involving the two new parameters can be rearranged and combined to replace them with
dimensionless variables that have some immediate phenomenological meaning. These new
parameters, β and ω, are,

θ + ρ b FK d
β= (3.3)
θ + ρbK d

k 2 (1 − β )LR
ω= (3.4)
v

where all coefficients are as previously defined. At ? >>1, the system is near local equilibrium
conditions and mass transfer to and from sorption sites is not rate limiting; at ? <<1 the mass
transfer to and from sorption sites is rate limiting and severe local non-equilibrium conditions
exist. Additionally, as β becomes larger (i.e., F ? 1), the system approaches local equilibrium.

For these experiments, one-dimensional flow-through experiments were conducted with 25 mm


diameter, 120 mm long Omnifit® glass columns. Columns were packed with clean Quakrite®
sand. Deionized water was used as the eluent. In single compound experiments, a 2 mL pulse of
naphthalene, acenaphthene or phenanthrene was injected via an HPLC Rheodyne® injector. The
breakthrough curve was obtained by monitoring the fluorescence intensity of the effluent. The
experiments were performed at flow rates of 0.2 and 1.2 mL / min using a Waters HPLC pump.
The effect of a mixture on sorption was studied by injecting an aqueous extract of the coal tar-
contaminated sediment into the column. For this purpose, eluent and sample were introduced via
a syringe pump. A 60 mL pulse of the coal tar aqueous extract was injected at a flow velocity of
0.5 mL / min. Effluent samples were analyzed subsequently by HPLC. Due to dispersion within
the column and HPLC peak overlap, not all 16 peaks were quantified. The compounds that
showed good HPLC separation and were quantified in the mixture were naphthalene, fluorene,
phenanthrene and anthracene. Because the experimental methods for single and multi-
component studies were different, two more column experiments were performed with 60 mL
step inputs of single component aqueous solutions of naphthalene and phenanthrene. The
experimental method was the same as that of the aqueous coal tar extract experiment. To avoid
confusion, the two methods are referred to as Method 1 (loop injection and inline fluorescence
detection) and Method 2 (step input via syringe pump with subsequent HPLC detection) in the
results section.

25
4.0 Results and Analysis

4.1 Sediment Physical Properties and Appearance. All sediment samples were similar in
appearance, odor and physical form. The sediment samples were charcoal black in color.
Visible multicolored oil sheens could be observed in pools of fluid that formed on top of the
sediment, and were also visible as a thin coating over the solid material. Some natural organic
materials, such as leaves and twigs, also were observed in sediment samples. No grain size
analysis was performed; however the solid matrix of material appeared to consist of silt sized
materials.

The physical properties of the sample cores were generally consistent, differing somewhat with
depth. Generally speaking, the lower 4' of the cores, despite being saturated, were quite
cohesive. The top 4' of the cores were more fluid in their consistency, and could almost be
poured as an extremely viscous liquid. This trend can possibly be explained by simply
considering the forces of compaction on the sediment. The bottom 4' has more overburden
pressure put upon it, and will likely undergo more compaction, and thus be more dense and solid.
The more recently deposited top 4' has no overburden pressure put upon it. One very significant
thing to consider with a sand cap is how the sediment would be compacted by the weight of the
applied sand. For example, when loading soil columns with sediment 4N-top for the preliminary
flow-through experiments, very significant compaction of the sediment was observed, even with
only 2 cm of sand on top. Furthermore, the weight of the sand not only compacts the thickness of
the sediment layer, it also forces pore fluid from the sediment upward through the sand.

The sediment sample numbering scheme was simplified according to Table 4.1 Sediment
the scheme shown in Table 4.1. Sample Notation
Sediment Notation
4.2 Water, Soil and Carbon Content Analysis. Figure 4.1 displays 99-01 #1
the results of sediment content analysis for the ten sediment samples. 6N #2
Detailed calculations and analysis of these data can be found in 98-21 #3
Table A2 of Appendix 1. On Figure 4.1, the data for the top 4 ft of 4N #4
the sample cores are reported on the left, and the data for the bottom 98-22 #5
4 ft of the sample cores are reported on the right. One conclusion
that can be drawn from this figure is that although coal tar
concentrations vary with depth, the distribution of volatile and heavy organic content is more or
less constant along the length of the river. If barriers are installed along the banks of the river to
prevent horizontal flux of coal tar into/out of the river, the river can be modeled in one
dimension (vertical). One dimensional analysis greatly simplifies contaminant transport analysis,
and will likewise simplify the prediction of sand cap performance.

The top 4 ft of sediment (0-4 ft) at all sampling locations has noticeably more volatile organic
content than at the 4-8 ft depth. Given the high amount of volatile organic compounds, including
PAHs, in coal tar, this implies a significantly higher concentration of PAHs in the top 4 ft. Total
PAH content analysis confirmed a general trend to that effect, as only sample SD-99-01 was
found to have comparable concentrations of PAHs between the 0-4 and 4-8 ft depth ranges.

26
100%

80%

% volatile o
60% % heavy o
% ash
40% % water

20%

0%
1-T 2-T 3-T 4-T 5-T 1-B 2-B 3-B 4-B 5-B

Figure 4.1 Water, Ash, and Volatile Content of Sediment

These data are significant when considering dredging the river, since the most highly
contaminated sediment would be that which was removed from the top. It also helps to give
some evidence of how long the contamination has been present, since the age of the sediment
increases with depth.

Heavy organic content is nearly constant both along the river and with depth. This includes
leaves, twigs and other such naturally present organic matter. This regular consistency of heavy
organic matter would be expected in the river, but the data is important in that it implies no
major disturbances have occurred within the sediment bed. Possible disturbances, for example,
could be made by dumping in or dredging the river. Intuitively, one would think of a constant
progression in deposition time with sediment depth, however, drastic changes in flow patterns,
due to canals and other diversions in the river, were made in the 1900's, that could have altered
sediment transport in the river. These flow changes could have affected sediment deposition
rates, but the consistency of the data seems to indicate that the sediment on site was not
significantly disturbed.

4.3 Total Sediment PAH Content Analysis. Total concentrations of 16 MAHs and PAHs
found in each sediment sample are shown in Table 4.2. These data can also be found in Table
A4 of Appendix 1. The values in Table 4.2 were measured by GC-FID analysis of the
supernatant dichloromethane /methanol phase, after drying with Na2 SO4 . No internal standard
for analysis was used due to interference caused by the number and density of other peaks on the
chromatograms. Additionally, measured concentrations (µg/ mL) were multiplied by the volume
of the organic phase (15 ml = 11 ml dichloromethane and 4 ml methanol) to obtain a total mass
of the contaminant per sample. For each sample, the total mass of each contaminant was divided

27
Table 4.2 Total Sediment Concentrations (mg PAH/ kg wet sediment)
1-T 1-B 2-T 2-B 3-T 3-B 4-T 4-B 5-T 5-B

SD-99- SD-99- SD-00- SD-00- SD-98- SD-98- SD-00- SD-00- SD-98- SD-98-
PAH 01-Top 01-Bot 6N-Top 6N-Bot 21-Top 21-Bot 4N-Top 4N-Bot 22-Top 22-Bot
Ethylbenzene 274.9 219.3 749.8 397.0 744.2 147.9 1031 235.6 524.5 88.91

m,p-Xylenes 74.51 98.93 271.3 187.4 322.0 85.30 467.7 126.4 112.6 84.32
o-Xylene 105.1 103.3 273.9 173.5 324.6 59.00 333.3 129.3 154.8 100.4
1,3,5-Trimethyl-
benzene 77.91 29.05 91.66 82.96 114.1 29.23 290.6 67.96 122.2 38.62
1,2,4-Trimethyl-
benzene 181.2 110.3 419.8 252.0 584.8 110.6 591.6 159.6 296.1 56.23
Naphthalene 3573 4016 12430 7246 6161 2537 16740 5324 7755 1890
2-Methyl-
naphthalene 2195 2066 6010 3433 4283 1228 8873 2613 3709 822.2

Acenaphthalene 268.1 160.9 632.1 408.7 559.9 173.8 905.6 316.9 410.8 89.54

Acenaphthene 1301 1417 4113 2517 1991 803.7 6854 1988 2709 654.0
Fluorene 599.0 684.4 1999 1291 834.6 423.9 3144 1032 1217 327.2
Phenanthrene 1681 1932 5973 3827 2248 1167 8833 3102 3482 941.8
Anthracene 469.2 477.3 1409 987.2 679.6 569.4 7528 1174 1737 239.4
Fluoranthene 570.2 659.4 2121 1396 642.3 409.6 4800 1161 1284 385.3
Pyrene 765.3 879.2 2891 1897 1002 572.7 6407 1515 1733 476.0
Benzo(a)-
anthracene 323.5 380.9 1187 847.9 418.0 250.6 2563 618.3 673.3 222.1
Chrysene 299.2 331.3 928.0 656.3 448.8 269.5 2150 522.1 592.9 198.4

by the original (wet) mass of the sample prior to extraction, yielding the values reported in Table
4.2. It was assumed that all of the aromatic hydrocarbons of interest partitioned completely into
the dichloromethane/ methanol extracting fluid, and that none remained sorbed to the sediment
or in the non-extractable pitch fraction of coal tar.

The mass fractions of several PAHs within the coal tar in each sample were also estimated, and
are shown in Figure 4.2. Mass fractions represent a ratio of mass of constituent per total mass of
coal tar. Because coal tar is composed of a large number of different compounds, including
many that were not quantified, the total mass of coal tar for each sediment sample (mg/ g) was
estimated. This estimation was made by assuming that all unidentified peaks on the gas
chromatogram that came out between ethylbenzene and chrysene were organic coal tar
constituents. Ethylbenzene and chr ysene were the identified contaminants with the smallest and
largest retention times (8.48 and 54.53 minutes, respectively). All peaks between these were
assumed to have the same flame ionization detector (FID) response equal to that of phenanthrene
on a mass basis. A similar calculation using pyrene’s standard curve resulted in values generally
within 10% of those calculated based on phenanthrene’s standard curve. The sum concentration
was multiplied by 1.4 to give an estimated total mass of coal tar. The 1.4 multiplier was used as
an estimate to account for the amount of “pitch” within the coal tar for which there would be no
GC response. The value of this multiplier was based on a previous estimate (Lee et al., 1992).
This graph clearly shows that the composition of the coal tar is generally uniform throughout the
river, and with depth, even while the total amount of coal tar present may vary. This information
simplifies modeling of the river, while also implying a single source of the coal tar. The data

28
also seems to suggest either no degradation of PAHs within the sediment, or uniform degradation
or PAHs with depth and river station.

Mass Fraction (mg/mg NAPL) 1


Naphthalene Phenanthrene
Pyrene o-Xylene
Chrysene

0.1

0.01

0.001

1-T 1-B 2-T 2-B 3-T 3-B 4-T 4-B 5-T 5-B
Sample

Figure 4.2 Measured Mass Fractions in Coal Tar Samples

The total moles of coal tar constituents were estimated, assuming a geometric average molecular
weight of 250 g / mol for all coal tar constituents. The rationale for this number is based on the
average molecular weight of coal tar constituents measured by GC (i.e., 3 ring PAHs) and the
fact that the pitch fraction will contribute higher molecular weight compounds to the tar,
increasing the average molecular weight to a 4-5 ring average. Methyl, sulfate, hydroxyl and
other functional groups on non- volatile components will also contribute to a higher average
molecular weight. Once an estimate of total moles of coal tar was obtained for each sample,
mole fractions for each compound were calculated and used to predict equilibrium aqueous
concentrations assuming ideal mixing of components within the coal tar, as described previously.

4.4. Equilibrium PAH Partitioning Between Coal Tar and Water. Measured aqueous
equilibrium concentrations of PAHs found after equilibration with sediment samples are shown
in Table 4.3. These data can also be found in Table A4 of Appendix 1. For three of the samples,
the organic phase emulsion was problematic, and hence there was not adequate volume of
organic phase to analyze the samples. Those samples are omitted from the table.

29
Table 4.3 Measured Aqueous Equilibrium Concentrations (mg/ L)
2-Top 2-Bot 3-Top 3-Bot 4-Top 4-Bot 5-Top
SD-00-6N- SD-00-6N- SD-98-21- SD-98-21- SD-00-4N- SD-00-4N- SD-98-22-
Average
PAH Top Bot Top Bot Top Bot Top

Ethylbenzene 0.71 0.737 1.29 0.849 1.27 0.949 1.52 1.05

m,p-Xylenes 0.26 0.330 0.523 0.442 0.554 0.465 0.300 0.411

o-Xylene 0.275 0.262 0.589 0.351 0.444 0.355 0.457 0.390

1,3,5-Trimethylbenzene 0.0491 0.0356 0.0640 0.0484 0.0565 0.0493 0.0723 0.0536

1,2,4-Trimethylbenzene 0.149 0.170 0.368 0.264 0.282 0.229 0.204 0.238


Naphthalene 11.0 10.6 8.24 11.5 13.3 16.1 16.2 12.4

2-Methylnaphthalene 1.49 1.62 1.90 1.68 2.24 2.04 2.46 1.92

Acenaphthalene 0.096 0.0672 0.116 0.110 0.0739 0.119 0.114 0.0995

Acenaphthene 0.827 0.724 0.603 0.682 0.891 1.02 1.14 0.840

Fluorene 0.256 0.231 0.168 0.218 0.260 0.311 0.332 0.254

Phenanthrene 0.364 0.252 0.194 0.236 0.291 0.317 0.423 0.297

Anthracene 0.101 0.0661 0.0545 0.0752 0.0830 0.0865 0.111 0.0826

Fluoranthene 0.113 0.0352 0.0277 0.0383 0.0438 0.0454 0.0809 0.0549

Pyrene 0.112 0.0368 0.0169 0.0396 0.0414 0.0568 0.0776 0.0544

Benzo(a)anthracene 0.0157 0 0.00415 0 0 0 0.0253 0.00645

Chrysene 0.0111 0 0 0 0 0 0 0.00159

Figure 4.3 regresses the coal tar-water partitioning coefficient values, Kp ([mass of i per volume
coal tar] / [mass of i per volume water]) versus aqueous solubility or subcooled liquid solubility
of the pure liquid constituent for the PAHs listed in Table 4.3 for which data are reported. In
theory, if ideal mixing occurs in the NAPL phase, the data should have a slope of
-1, with a y- intercept that is a function of the molecular weight and density of the coal tar NAPL.
A solid line with a slope of -1 is shown on the graph, with an intercept of 0.6. The dotted lines
indicate a factor of 2 from this line. The data show a clear correlation and similar trend as data
reported by Lee et al. (1992) on the solubility of coal tar constituents measured in the absence of
sediment or other solid phase. It should be noted that the heavier compounds seem to tail off of
the line, exhibiting a trend of enhanced solubility with increased PAH size. The most likely
explanation for this solubility enhancement is that the presence of aqueous humic materials
enhances the solubility of these more hydrophobic compounds (Van Stempvoort et al., 2002; Lo
and Lee, 1995). Alternatively, size and molecular interaction differences of these larger
molecules may lead to non-ideal mixing in the NAPL phase, effectively decreasing the ir partition
coefficients to (i.e., solubility in) the NAPL.

30
6.5

log K p , (log (mL/gm))


5.5

2 Top
4.5
2 Bottom
3 Top
3 Bottom
3.5 4 Top
4 Bottom
5 Top
2.5
-6.5 -5.5 -4.5 -3.5 -2.5
log (S) , (log (M))

Figure 4.3 Meas ured log(Kp) versus log S

The solid line assumes ideal mixing in the NAPL (i.e., ?ct,i = 1). The intercept of this line (= 0.6)
theoretically equals - log (MWct / ρct ), as shown earlier. If the average density of the coal tar is
assumed to be ρct = 1,000 g/ L (Lee et al., 1992), the corresponding average molecular weight of
the coal tar is 252 g / mol. This average agrees with that assumed early (i.e., 250 g / mol) in
calculating the mole fraction composition of the tar. Assuming that the coal tar components
measured by GC have an average molecular weight equal to that of phenanthrene (178.23 g /
mol), that the pitch fraction is 40% of the coal tar mass, and that the total average molecular
weight in the coal tar is 252 g / mol as estimated from Figure 4.3, an average molecular weight
of 363 g / mol is calculated for the non-extractable/non- volatile pitch in the coal tar.

Equilibrium aqueous concentrations also were calculated for each sediment sample via eq 2.16
(Raoult's law). Figure 4.4 compares these values to measured equilibrium aqueous phase
concentrations. On this figure, each datum point represents the aqueous phase concentration of
one compound in equilibrium with the sediment coal tar at one location. Again, these data
indicate that a clear correlation exists between the theoretical and measured values. In a practical
sense, this implies that aqueous concentrations of PAHs in coal tar contaminated sediment pore
water can be accurately predicted if the chemical composition of the NAPL is known. Further,
the flux of PAH compounds from the sediment to the river can then be predicted if the
groundwater advective flow is known. Figure 4.4 also shows, as described above for Figure 4.3,
an apparent solubility enhancement for the larger PAH compounds.

31
100

2 Top

Caq (calculated), (mg / L)


10 2 Bottom
3 Top
3 Bottom
1 4 Top
4 Bottom
5 Top
0.1

0.01

0.001
0.001 0.01 0.1 1 10 100
Caq (measured) , (mg / L)
Figure 4.4 Predicted Caq versus Measured Caq

4.5 Bench-Scale Sediment Flow-Through Experiments. In the 1-D bench-scale sediment


flow-through experiments, column effluent samples were taken over a nine-day period, to
provide a good average of effluent concentrations from the pore water in the sediment. Recall
that in these experiments, a 5-cm layer of sediment was placed in the column and overlaid with 2
cm sand with an imposed advective up- flow though the sediment and sand layers. Effluent
exiting the sand layer was analyzed for anthracene, fluorene, phena nthrene and pyrene over time.
Results are shown in Figure 4.6 and chemical structures are shown in Figure 4.7.

0.14
Daily Average Effluent Concentration

anthracene
0.12
fluorene
0.1 phenanthrene

0.08 pyrene
(mg/L)

0.06

0.04

0.02

0
0 2 4 6 8 10
Time (days)

Figure 4.6 Measured Effluent Concentration with Time

32
Figure 4.7 PAH Compounds of Interest

After 17 days of continuous pumping, naphthalene was analyzed by HPLC at ? = 276 nm. This
wavelength was selected to minimize interference from the ethylbenzene peak. The overall
average effluent concentrations of the five compounds for all samples measured are compared to
the measured batch-equilibrium concentration for each compound in the same sediment in Table
4.4.

Table 4.4 Preliminary Column Data

Darcy Flux = 10.7 cm/ day


4N-Top sediment; 5 cm thickness
Standard
Average Standard
Deviation Measured
measured in Avg % of Deviation,
Compound measured in equilibrium
effluent, mg/ equilibrium % of
effluent, mg/ conc, mg/ L
L equilibrium
L
Anthracene 0.0196 0.0079 0.0830 23.56% 9.47%
Fluorene 0.0702 0.0322 0.2597 27.03% 12.42%
Naphthalene 7.475 0.658 13.320 56.12% 4.94%
Phenanthrene 0.0618 0.0377 0.2912 21.21% 12.94%
Pyrene 0.0182 0.0122 0.0414 44.08% 29.53%
file: /preliminary column data2.xls

Average effluent concentrations were 21-56% of measured equilibrium concentrations. This is


significant because it shows that substantial partitioning takes place between coal tar and
sediment pore water for a broad range of PAHs (i.e., 2 to 4 rings), even during the short time
scale under conditions of an advective flow rate of 10.7 cm / day passing through a 5 cm
sediment layer. Considering that this experiment was conducted under very high- velocity flow
conditions (compared to the field site) and through only 5 cm of sediment, it is logical to assume

33
that in situ pore water concentrations at the site are in equilibrium with the NAPL coal tar. In the
experiments described here, it is likely that the pore water may follow preferential flow pathways
that would further decrease exposure time and surface area of the flowing water to the stationary
coal tar phase, accounting for the flow through experiment values to be consistently less than
those measured in batch. At greater sediment depths, concentrations approaching the
equilibrium values are expected. If advective transport through the sediment is assumed to occur
in an upward direction, advective flow at the site is occurring through a minimum of 2.5-3
meters of contaminated sediment prior to reaching the river (8 ft cores contained NAPL at all
depths).
% Equilibrium Concentration

100.00%

80.00%

60.00%

40.00%

20.00%

0.00%

e
e

ne

ren
e
en

en
ore

len

Py
ac

thr
tha
Flu
thr

an
ph
An

en
Na

Ph

PAH Compound

Figure 4.8 Measured Percent of Equilibrium Value Achieved by


Effluent Pore Water

4.6 Groundwater Flux through Sediment. A Baseline hydrogeological assessment was


performed at the Hammond, IN former manufactured gas plant site to obtain estimates of the
properties of the subsurface on-site and through the river sediment (ThermoRetec, 2001). The
results of two slug tests conducted in the river sediment estimated the hydraulic conductivity (K)
of the sediment to be 9.2x10-7 cm / s. Piezometric measurements in the sediment produced
estimates between 0.07 - 0.15 ft / ft for the vertical hydraulic gradient (dφ / dx, or I), in the
upward direction (ThermoRetec, 2001). From Section 2,

q K⋅ I (2.24)

From the estimates on K and eq 2.24, an estimated range for Darcy flux through the sediment is
between q = 6.4 ×10-8 and 1.4 × 10-7 cm / s (between 0.0056 and 0.012 cm / day).

34
From the estimated upward groundwater flux and the measured equilibrium concentrations of
PAHs in the sediment pore water, the estimated advective contaminant flux from the sediment
entering the river can be calculated. Recall from Section 2,

Fa q ⋅ Ci , w
(2.26)

Table 4.5 reports estimated values of the advective contaminant flux leaving the sediment,
calculated with eq 2.26, the average aqueous concentration values reported in Table 4.3 for the
column experiments, and an upward groundwater Darcy flux range of 0.0056-0.012 cm / day
assuming steady one-dimensional flow through a constant source. Table 4.5 also lists calculated
steady-state values for the estimated daily and overall contaminant loads into the river via the
following methodology.

Table 4.5 Estimated Advective Contaminant Flux (mg/ m2 day) and


Estimated Daily (mg / day) and Overall (mg / L) Load to River
2 Top Q=0.0056 cm / d Q=0.012 cm / d Q=0.0088 cm / d

Average Steady-
State daily load to Average Steady-State
Average Ci,w Advective Flux, Fa Advective Flux, Fa Average Fa river† overall load to river ‡
Name mg/ L Mg/ m2 day mg/ m2 day mg/ m2 day mg/ day mg/ L

Ethylbenzene 1.047 0.06486 0.1137 0.08929 330.4 3.304E-06

m, p-Xylenes 0.4107 0.02905 0.04606 0.03755 138.9 1.389E-06


o-Xylene 0.3903 0.02307 0.05179 0.03743 138.5 1.385E-06

1,3,5-Trimethylbenzene 0.05358 0.003133 0.005628 0.004380 16.21 1.621E-07

1,2,4-Trimethylbenzene 0.2378 0.01492 0.03234 0.02363 87.42 8.742E-07


Naphthalene 12.43 0.9356 0.7253 0.8304 3073 3.073E-05

2-Methylnaphthalene 1.918 0.1425 0.1668 0.1547 572.2 5.722E-06

Acenaphthalene 0.09946 0.005914 0.01017 0.008041 29.75 2.975E-07

Acenaphthene 0.8397 0.06371 0.05308 0.05840 216.1 2.161E-06


Fluorene 0.2536 0.02029 0.01479 0.01754 64.90 6.490E-07
Phenanthrene 0.2967 0.02217 0.01710 0.01964 72.65 7.266E-07
Anthracene 0.08257 0.005820 0.004792 0.005306 19.63 1.963E-07
Fluoranthene 0.05490 0.003098 0.002433 0.002765 10.23 1.023E-07
Pyrene 0.05442 0.003238 0.001483 0.002361 8.734 8.734E-08
Benzo(a)anthracene 0.006445 0 0.0003650 0.0001830 0.6756 6.756E-09
Chrysene 0.001590 0 0 0 0 0
†- Using an estimated sediment-river exposure area of 3700 m 2
‡- Assuming no degradation and an average daily river flow of 1 × 10 8 L
file: /Batch Equilibrium Data2.xls

A sediment to river-water exposure area of 3,700 m2 was estimated from the approximate river
width of 50 feet (16 meters), and site length of 750 feet (230 meters), derived from site
schematic (ThermoRetec, 2001). The average daily stream flow data for the Grand Calumet

35
River was calculated using the daily mean stream flow measured in Hammond, IN between the
years 1992 and 2000 (USGS, 2003). Stream flow data are reported in Table A5 of Appendix 1.
Before taking into account the steady state loads from Table 4.5, it is important to consider how
long it will take for the sand cap to reach steady-state conditions, such that contaminant
concentrations in the cap do not change with time. Before these conditions exist, actual
concentratio ns will be less. Still, even under steady state conditions, the predicted load to the
river will be orders of magnitude less than the EPA toxicity reference values for the 16
compounds (Table A1 of Appendix 1). This also does not take into account any biodegradation
or photodegradation of PAHs in the river. Therefore, contributions of PAHs to the river are not
of concern at this site. The health of the benthic community and organisms that feed on this
community are of primary concern and the rationale for remediation.

The estimated contaminant flux through a sand cap, calculated with eq 2.44, assuming no
retardation, no degradation, and DL = 1 cm2 / day, q = 0.012 cm / day, and ne = 0.4 (i.e., v = 0.03
cm / day) is shown in Figure 4.9.

1
0.9
t=1 year
0.8
0.7 t=10 years
0.6 t=100 years
C/Co

0.5
0.4
0.3
0.2
0.1
0
0 200 400 600 800 1000
Distance from Source (cm)

Figure 4.9 Estimated Contaminant Flux at t = 1, 10, and 100 years (x=0 to 10 m)

36
Results scaled to the likely approximate dimensions of a sand cap (e.g., 2 ft thickness) are shown
in Figure 4.10.

1
0.9
0.8
0.7
0.6
C/Co

0.5
0.4
0.3 t=1 year
0.2 t=10 years
0.1 t=100 years
0
0 10 20 30 40 50 60
Distance from Source (cm)

Figure 4.10 Estimated Contaminant Flux at t = 1, 10, and 100 years (x = 0 to 0.6 m)

The longitudinal hydrodynamic dispersion coefficient, DL, was estimated by neglecting


mechanical dispersion due to extremely low advective flux, and by assuming a large diffusion
coefficient of 1 cm2 / day. The diffusion coefficient of H+, by far the largest Dd of any ion, has
been measured at approximately 8 cm2 / day. Most ions have Dd ≈ 1 cm2 / day (Domenico and
Schwartz, 1998). Since aqueous diffusion data for organic chemicals is generally much more
difficult to find, 1 cm2 / day was used as a conservative estimate. As will be shown
subsequently, it is reasonable to assume that retardation in a clean sand cap is minimal if the sand
contains negligible organic carbon. Neglecting retardation also results in the conservative
estimate (i.e., erring on the side of the environment).

The figures clearly indicate that under these conditions, prior to the development of steady state
flux from the sediment, the relatively large hydrodynamic dispersion (diffusion and mechanical
dispersion) results in ‘contaminant spread’ within the sand cap.

The contaminant breakthrough curve at the sand-water interface of a 60 cm deep sand cap placed
over coal tar NAPL containing sediment was calculated with eq 2.44, applying the same
assumptions as those made in developing Figures 4-9 and 4-10. The breakthrough curve is
shown in Figure 4-11 and indicates that it would take approximately 20-30 years for the
contaminant flux at the top of the sand cap to reach steady-state. At this point, pore water PAH
concentrations throughout the sand cap would approach the measured equilibrium values.

37
1
0.9
0.8
0.7
0.6
C/Co

0.5
0.4
0.3
0.2
0.1
0
0 10 20 30 40 50
time (years)

Figure 4.11 Estimated Contaminant Flux at 0.6 m with Rf = 1.0

Again, this conclusion assumes no retardation and no biodegradation occurs within the sand
layer. The extent to which pure-phase coal tar may move due to the advective flow of pore water
is unknown. The organic coal tar should move at a rate that is orders of magnitude slower than
the pore water, if at all, yet it cannot be considered immobile. The migration of coal tar upwards
through the sand cap could eventually compromise the intended benefits of the sand cap. A sand
cap could therefore, in theory, have a finite service life before it must be replaced with clean
sand. Initial compaction of the sediment due to the weight of the sand certainly would cause
contamination of the lower portion of the sand layer with the NAPL, effectively decreasing the
sand cap thickness.

4.7 Advection-Dispersion Experiments. The results of the advective-dispersion experiments,


where PAH compounds where introduced to the columns via pulse or step inputs, indicated that
the retardation coefficients measured or estimated in single component systems are good
predictors of retardation coefficients of components within the complex PAH mixture that results
from dissolution of coal-tar NAPL. Recall that 1-D flow-through experiments were conducted
with 25 mm diameter, 120 mm long Omnifit® glass columns packed with clean Quakrite® sand.
Pulse input experiments where performed with 2- mL loop injections and inline fluorescence
detection (method 1) and step input experiments where performed by introducing the PAH
compounds with a syringe pump with subsequent HPLC detection (method 2).
The breakthrough curves obtained with method 1 are shown in Figure 4.12. Retardation
coefficients (Table 4.6) were obtained by performing moment analysis on the breakthrough
curves. Retardation coefficients did not vary significantly with velocity. Kd (eqs 2.5 & 3.2) was
obtained using a porosity value of 0.35 and a bulk density of 1.74 gm / mL. The porosity and
bulk density were determined gravimetrically.

38
Table 4.6 Calculated Parameters for Experiments Conducted by Method 1
R ? β Kd (mL / gm) k 2 (hr -1) F
Compound
v=0.7 V=0.1 v=0.7 V=0.1 v=0.7 V=0.1 v=0.7 V=0.1 v=0.7 V=0.1 v=0.7 V=0.1
cm/min cm/min cm/min cm/min cm/min cm/min cm/min cm/min cm/min cm/min cm/min cm/min
Naphthalene 1.2130 1.3190 0.1298 0.1566 0.8498 0.8015 0.0426 0.0638 2.6704 0.3616 0.1446 0.1792
Acenaphthene 1.3250 1.4840 0.1550 0.1824 0.8444 0.7733 0.0650 0.0968 2.7762 0.3191 0.4098 0.3049
Phenanthrene 1.8330 2.0370 1.1120 0.7959 0.7716 0.7948 0.1665 0.2073 9.7956 1.1680 0.4974 0.5969

0.4

0.35

Order of Elution:
0.3 naphthalene at v = 0.7 then 0.1 cm/min,
acenaphthene at v = 0.7 then 0.1 cm/min,
0.25 phenanthrene at v = 0.7 then 0.1 cm/min
C/Co

0.2

0.15

0.1

0.05

0
0.5 1 1.5 2 2.5 3 3.5
Pore Volumes

Figure 4.12 Breakthrough Curves of 2 mL Pulse Inputs at 2 Velocities (Method 1)

Figure 4.13 shows the relation between Kd and Kow for the test compounds at the two flow rates.
The strong relationship confirms that hydrophobic forces are primarily responsible for sorption
to the sand. With the values of R obtained from moment analysis and a dispersion coefficient of
0.043 cm / min at the higher flow rate and 0.012 cm / min at the lower flow rate, the data were fit
to predictive curves by adjusting the parameters ? and ß. The chemical non-equilibrium model
in CXTFIT was used to perform the fit. The best- fit values of ? and ß are reported in Table 4.6.
The values of k2 and F were calculated with eqs 3.3 and 3.4 and are also presented in Table 4.6.
The value for k2 also can be estimated from the relationship proposed by Brusseau and Rao,

log k 2 = 0.301 − 0.668 × log K d (4.1)

This equation and the values of k2 obtained from the column experiments are shown in Figure
4.14 and indicate that the values calculated in this study are reasonable. The calculated values of
F are regressed against Kd in Figure 4.15.

39
-0.6

-0.8

log kd -1

-1.2

-1.4
v = 0.7 cm / min
v = 0.1 cm / min
-1.6
3 3.5 4 4.5 5
log Kow

Figure 4.13 log Kd vs log Kow at the Two Velocities (Method 1)

3
v = 0.3 cm / min, Method 2
2
v = 0.7 cm / min, Method 1
1
v = 0.1 cm / min, Method 1
0

-1
log k2

-2

-3

-4

-5

-6
-2 -1 0 1 2 3 4 5 6
log Kd

Figure 4.14 Values of log k2 versus log Kd (Parameters from Methods 1 & 2)

40
0.7

0.6 v = 0.7 cm / min


v = 0.1 cm / min
0.5

F 0.4

0.3 y = 2.5007x + 0.0852

0.2

0.1

0
0 0.05 0.1 0.15 0.2 0.25
Kd

Figure 4.15 Calculated Values of F versus Kd (Method 1 Parameters)

The breakthrough curves obtained with method 2 are shown in Figure 4.16. The three
experiments performed with this method were 60 mL step inputs of aqueous (1) naphthalene, (2)
phenanthrene and (3) coal tar extract. The retardation coefficients reported in Table 4.7 were
obtained by performing moment analysis on the breakthrough curves. It was observed that
retardation coefficients did not vary significantly between the single component PAH
experiments and the multi-component coal tar extract experiment indicating that sorption of each
compound is not significantly affected by the presence of other compounds in the mixture. The
slight difference in R values observed can be attributed to experimental variability. The sorption
coefficient Kd was obtained with eq 3.2. The relation between log Kd and log Kow (Figure 4.17)
again indicates correlation. From the values of R calculated from moment analysis and assuming
a dispersion coefficient of 0.012 cm / min, the values of ? and ß, again were calculated with the
chemical nonequilibrium model, CXTFIT. The ? and ß values obtained are presented in Table
4.7. The values of k2 and F were calculated with eqs 3.3 and 3.4 and are also presented in Table
4.7. Values of k2 are shown in Figure 4.14 (along with those of method 1), and F is regressed
against Kd in Figure 4.18.

Table 4.7 Model Parameters Calculated from Experiments Conducted by Method 2


(v = 0.3 cm / min)
R ? β Kd (mL / gm) k 2 (hr -1) F
Compound coal coal coal coal coal
tar tar tar tar tar coal tar
extract single extract single extract single extract Single extract single extract single
-
Naphthalene 1.16 1.11 0.1223 0.0594 0.86 0.90 0.0353 0.0212 1.0195 0.8454 0.000225 0.000009
fluorene 1.544 - 0.5055 - 0.72 0.1199 1.5498 0.200469
Phenanathrene 2.352 2.314 0.8126 0.4688 0.59 0.65 0.2980 0.2533 1.1320 0.8890 0.291964 0.381701
anthracene 3.364 - 0.6850 - 0.52 0.5210 0.5622 0.312685

41
1.2

naphthalene (single)
1
napthalene (mix)
fluorene (mix)
0.8
phenanthrene (single)
C / Co phenanthrene (mix)
0.6 anthracene (mix)

0.4

0.2

0
0 2 4 6 8 10 12
Pore Volume

Figure 4.16 Breakthrough curves for single compounds and coal tar extract at v = 0.3 cm /
min (Method 2)
-0.2
y = 0.8449x - 4.4011
-0.4 R2 = 0.9441
-0.6

-0.8
log Kd

-1

-1.2

-1.4

-1.6

-1.8
3 3.5 4 4.5 5
log Kow

Figure 4.17 Values of log Kd versus log Kow for single compounds and coal tar extract at v =
0.3cm / min (Method 2)

42
0.40

0.30
F

0.20
y = 0.6844x + 0.0554
R2 = 0.6311
0.10

0.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6
Kd

Figure 4.18 Values of F versus Kd for single solutes and solutes within the coal tar extract
at v = 0.3cm / min (Method 2).

43
5.0 Preliminary Conclusions

1. Coal tar is present as a NAPL in the sediment, and equilibrium conditions likely exist
between the MAH and PAH components in the coal tar NAPL and the pore water
throughout most of the sediment.

2. The equilibrium water phase concentrations of each PAH in contact with sediment
samples was measured and agreed within a factor of 2-3 to values calculated from
mole fraction composition of the sediment coal tar and literature pure phase solubility
data.

3. From measured physical, chemical and hydrogeological properties of the underlying


sediment, an estimate of contaminant flux through a sand cap was made.

4. Remedial action is required at these sites in order to protect the benthic community of
organisms.

5.1 Additional Considerations : The results and associated models presented in this study on
measured and calculated pore water concentrations and contaminant flux are valuable in
evaluating the potential performance of a sand cap. However, many other factors must also be
considered within an overall complete evaluation. These include (but are not limited to) the
following:

1) The rate of benthic biodegradation within a sand cap may significantly reduce exposure
concentrations in the benthic zone, enhancing the performance of a sand cap. At sub- lethal
concentrations, biodegradatio n by benthic organisms will attenuate the flux of aromatic
hydrocarbons through the cap. Facile biodegradation of MAH and PAH requires aerobic
conditions. This implies that any process that induces aeration of the top sand layer may
significantly reduce the concentration of MAH and PAH compounds within this layer, reducing
exposure and flux. It would therefore be beneficial to examine ways of inducing oxygen
transport to the sand layer and to measure how effective oxygen transfer is in inducing
biodegradation. Potential ways of inducing oxygen transport include: (1) active recirculation of
oxygenated water through the sand cap, (2) passive (via hydraulic head differences) circulation
of water through the sand cap, and (3) transport via rooted plants within the sand cap.

2) The weight of a sand cap will result in an overburden pressure upon the non-compacted
sedime nt, resulting in sediment compaction and forcing pore fluid (both water and NAPL)
upwards into the cap. In the actual event of a sand cap being applied, certain elevation
irregularities and non-uniformities in sand application will be inevitable. These irregularities
will cause pressure irregularities on the sediment, squeezing sediment around the sand during
application and causing mixing between the sand and sediment. Only a relatively very thick
sand cap would avoid these problems via minimizing the fraction of the cap for which this
mixing has occurred. In loading soil columns for experiments, some compression of the
sediment by the overlying sand was always observed.

44
3) Rapid photo-degradation, through the process of photo-oxidation, is likely to occur in the
river water above the sand cap, yet it is unlikely to occur in the cap itself due to the relative
scarcity of dissolved oxygen and light in the cap. This does not, however, necessarily mean
there will be no photodegradation products present in the benthic zone. The benthic zone could
be considered to be well- mixed, provided that enough bioturbation occurs within this zone. A
greater understanding of organic compound transport and dynamics within the benthic zone
would be helpful in explaining or predicting the presence of any photo-oxidation products in a
sand cap.

4) The sand cap itself could be physically or chemically modified to improve performance. A
layer of peat, applied below the sand cap would temporarily increase the time to breakthrough
due to the large partition coefficients that MAH and PAH compounds have for peat. One
possible problem with peat, however, is that preferential flow paths may be created reducing the
effectiveness of the permeable barrier. The thicker the peak layer, the less likely preferential
flow will occur. Geo-textile membranes may create similar problems, with flow around the
edges being significant, and settling inducing liquidity of the sediment to flow around the edges.

45
6.0 References

1. Barron, Mace G. and Lisa Ka’aihue. Potential for Photoenhanced Toxicity of Spilled Oil in
Prince William Sound and Gulf of Alaska Waters. Marine Pollution Bulletin, 2001. Vol 43,
Issues 1-6, January-June 2001. pp. 86-92.
2. Charbeneau, Randall J. Groundwater Hydraulics and Pollutant Transport. Prentice-Hall,
Inc.: Upper Saddle River, NJ, 2000.
3. Chiou, Cary T. and David W. Schmedding. Partitioning of Organic Compounds in Octanol-
Water Systems. Environmental Science & Technology, 1982. Volume 16, pp. 4-10.
4. Cleveland, Laverne, Edward E. Little, Robin D. Calfee, and Mace G. Barron. Photoenhanced
toxicity of weathered oil to Mysidopsis bahia. Aquatic Toxicology, 2000. 49, pp. 63-76.
5. Domenico, Patrick A. and Franklin W. Schwartz. Physical and Chemical Hydrogeology. 2nd
Ed. John Wiley & Sons, Inc., New York, 1998.
6. Eljarrat, Ethel, Josep Caixach, Josep Rivera, Mariona De Torres, and Antoni Ginebreda.
Toxic Potency Assessment of Non- and Mono-ortho PCBs, PCDDs, PCDFs, and PAHs in
Northwest Mediterranean Sediments (Catalonia, Spain). Environmental Science &
Technology, 2001. Volume 35, No. 18, pp 3589-3594.
7. Fasnacht, Matthew P. and Neil V. Blough. Aqueous Photodegradation of Polycyclic
Aromatic Hydrocarbons. Environmental Science & Technology, 2002. Volume 36, pp. 4364-
4369.
8. Ferreira, Marcia M.C. Polycyclic aromatic hydrocarbons: a QSPR study. Chemosphere,
2001. Volume 44, pp. 125-146.
9. Fetter, C.W. Contaminant Hydrogeology. Prentice-Hall, Inc.: Upper Saddle River, NJ, 1999.
10. Golchin, Johanshir, Greg A. Stenback, Bruce H. Knartanson, and Sam Nelson. Development
of a manufactured gas plant data management system. American Laboratory, 1997. Volume
29, No. 8, pp. 39-40.
11. Gratz, John H., Gary M. Grey, and O. Karl Scheible. Treatment of contaminated ground
water from manufactured gas plant sites. Hazardous and industrial wastes: proceedings of the
twenty-eighth Mid-Atlantic Industrial and Hazardous Waste Conference. Technomic
Publishing Co., Inc., Lancaster, PA, 1996.
12. Khodadoust, Amid P., Rajesh Bagchi, Makram T. Suidan, Richard C. Brenner, and Neal G.
Sellers. Removal of PAHs from highly contaminated soils found at prior manufactured gas
operations. Journal of Hazardous Materials, 2000. B80, pp. 159-174.
13. Lampi, Mark A., Xiao-Dong Huang, Yousef S. El-Alawi, Brendan J McConkey, D. George
Dixon, and Bruce M. Greenberg. Occurrence and Toxicity of Photomodified Polycyclic
Aromatic Hydrocarbon Mixtures Present in Contaminated Sediments. Environmental
Toxicology and Risk Assessment: Science, Policy, and Standardization- Implications for
Environmental Decisions: Tenth Volume. ASTM STP 1403. B. M. Greenberg, R. N. Hull, M.
H. Roberts, Jr., and R. W. Gensemer, Eds., American Society for Testing and Materials,
West Conshohocken, PA, 2001.
14. Lane, William F. and Raymond C. Loehr. Estimating the Equilibrium Aq ueous
Concentrations of Polynuclear Aromatic Hydrocarbons in Complex Mixtures. Environmental
Science & Technology, 1992. Volume 26, pp. 983-990.
15. Lee, Jong-Hyeon, Peter F. Landrum and Chul-Hwan Koh. Toxicokinetics and Time
Dependent PAH Toxicity in the Amphipod Hyalella azteca. Environmental Science &
Technology, 2002a. Volume 36, No. 14, pp. 3124-3130.

46
16. Lee, Jong-Hyeon, Peter F. Landrum and Chul-Hwan Koh. Prediction of Time-Dependent
PAH Toxicity in Hyalella azteca Using a Damage Assessment Model. Environmental
Science & Technology, 2002b. Volume 36, No. 14, pp. 3131-3138.
17. Lee, Linda S., P. Suresh C. Rao, and Itaru Okuda. Equilibrium Partitioning of Polycyclic
Aromatic Hydrocarbons from Coal Tar into Water. Environmental Science & Technology,
1992. Volume 26, pp. 2110-2115.
18. Lo, Jin-Hsiang Andy and Whei-May Grace Lee. Influence of Surfactant upon
Adsolubilization of Hydrophobic Organic Compounds in Fog Droplets. Journal of Aerosol
Science, 1995. Volume 26, pp. S199-S200.
19. Lotufo, Guilherme R. Lethal and subletha l toxicity of sediment-associated fluoranthene to
benthic copepods: application of the critical-body-residue approach. Aquatic Toxicology,
1998. Volume 44, pp. 17-30.
20. Mahjoub, Borhane, Emmanuel Jayr, Rémy Bayard and Rémy Gourdon. Phase partition of
organic pollutants between coal tar and water under variable experimental conditions. Water
Research, 2000. Volume 34, No. 14, pp. 3551-3560.
21. Oris, James T. and John P. Giesy, Jr. The photoenhanced toxicity of anthracene to juvenile
sunfish. Aquatic Toxicology, 1985. Volume 6, Issue 2, pp. 133-146.
22. Schwarzenbach, René P., Philip M. Gschwend, and Dieter M. Imboden. Environmental
Organic Chemistry. John Wiley & Sons, Inc.: New York, 1993.
23. Schwarzenbach, René P., Philip M. Gschwend, and Dieter M. Imboden. Environmental
Organic Chemistry, 2nd edition. John Wiley & Sons, Inc.: New York, 2003.
24. Van Stempvoort, D.R., S. Lesage, K.S. Novakowski, K. Millar, S. Brown, and J.R.
Lawrence. Humic acid enhanced remediation of an emplaced diesel source in groundwater.
1. Laboratory-based pilot scale test. Journal of Contaminant Hydrology, 2002. Volume 54,
pp. 249-276.
25. Talley, Jeffrey W., Upal Ghosh, Samuel G. Tucker, John S. Furey, and Richard G. Luthy.
Particle-Scale Understanding of the Bioavailability of PAHs in Sediment. Environmental
Science & Technology, 2002. Volume 36, pp. 477-483.
26. ThermoRetec. DRAFT Remediation Work Plan for Sediment Remediation in the Grand
Calumet River, Hammond Former Manufactured Gas Plant, Hammond, IN. ThermoRetec
Consulting Corporation, Concord Massachusetts, 2001.
27. USGS. Surface Water data for Indiana: Calendar Year Streamflow Statistics. United States
Department of the Interior, U.S. Geological Survey, 2003.
http://waterdata.usgs.gov/in/nwis/annual/calendar_year/?site_no=05536357

47
Appendix 1.

Figure A1. Plan View of Hammond MGP Site, Grand Calumet River Sampling Locations
(For Purdue U.), October 22-23, 2001. The site is at the corner of Hohman Avenue and Wilcox
Street, Hammond, IN. Samples were collected by Elliot Smith (AScI, Detriot) and Damon
Morris (Retec, Seattle). Jennifer Miles and Amy Wehnert (both of Retec, Indianopolis) cores
from these 5 sites plus 5 additional sites for chemical analysis.
Chad Jafvert, Changhe Xiao, Suresh Rao, and Doug Lane participated from Purdue.

Sample Locations and collection times (Purdue Samples) (From left to right on the map)

Sample ID ThermoRetec ID Collection Time


1 SD-99-01 Oct. 22 (Tues) afternoon
2 SD-00-6N Oct. 23 (Wed) morning
3 SD-98-21 Oct. 23 (Wed) morning
4 SD-99-4N Oct. 23 (Wed) afternoon
5 SD-98-22 Oct 22 (Tues) morning

48
Table A1. Sediment and Surface Water Toxicity
Reference Values for Selected PAH Compounds
Surface/ Pore
Sediment TRV Water TRV (mg/L)
Compound (mg/kg) † ‡
Acenaphthene 0.089 0.05585
Acenaphthylene 0.13 0.3069
Anthracene 0.25 0.02073
Benzo(a)anthracene 0.69 0.002227
Chrysene 0.85 0.002042
Fluoranthene 0.834 0.007109
Fluorene 0.14 0.0393
2-Methylnaphthalene 0.201 0.07216
Naphthalene 0.39 0.1935
Phenanthrene 0.54 0.01913
Pyrene 1.4 0.01011
†- Sediment TRVs are reported from research (ThermoRetec, 2001)
‡- Surface/ Pore Water TRVs are reported by the USEPA (ThermoRetec, 2001)

49
Table A2. Sediment Composition (all values are % wet weight)

Organic Matter Non-Volatile


Water and Total (OM) (Heavy) Volatile
Sediment Volatiles1 Ash2 Carbon3 (Carbon × 1.7) 4 OM5 OM6 Water7

1T 68.8 ± 0.3 24.7 ± 0.2 8.1 ± 2.6 13.8 ± 4.4 6.5 ± 0.7 7.3 ± 4.5 61.5 ± 4.7

1B 66.5 ± 0.06 27.6 ± 0.1 5.4 ± 0.4 9.3 ± 0.7 5.9 ± 0.05 3.3 ± 0.7 63.1 ± 0.7

2T 67.9 ± 0.03 24.3 ± 0.03 10.3 ± 0.5 17.4 ± 0.9 7.8 ± 0.03 9.7 ± 0.9 58.2 ± 0.9

2B 66.3 ± 0.01 26.2 ± 0.1 6.5 ± 0.7 11.0 ± 1.2 7.6 ± 0.1 3.5 ± 1.3 62.7 ± 1.3

3T 67.3 ± 1.0 25.4 ± 0.8 10.1 ± 0.1 17.2 ± 0.2 7.3 ± 0.1 9.9 ± 0.2 57.4 ± 1.1

3B 66.4 ± 0.06 26.6 ± 0.1 5.8 ± 0.3 9.9 ± 0.5 7.1 ± 0.03 2.7 ± 0.5 63.7 ± 0.6

4T 67.3 ± 0.7 22.8 ± 0.3 14.0 ± 1.4 23.7 ± 2.4 9.9 ± 0.4 13.8 ± 2.7 53.5 ± 0.3

4B 62.8 ± 0.5 29.8 ± 0.2 6.9 ± 0.3 11.8 ± 0.5 7.4 ± 0.2 4.4 ± 0.8 58.4 ± 1.2

5T 64.9 ± 0.6 27.9 ± 0.3 8.4 ± 1.8 14.4 ± 3.0 7.2 ± 0.4 7.1 ± 3.4 57.7 ± 4.0

5B 56.2 ± 0.03 38.2 ± 0.05 5.2 ± 0.06 8.9 ± 0.1 5.6 ± 0.02 3.3 ± 0.1 52.9 ± 0.1

1
Dried at 105° C for 14 hrs; 2 Dried at 500° C for 3 hrs ; 3 Measure with a Carlo-Erva 1108 elemental

analyzer; 4 Conversion factor 1.7 for organic carbon to organic matter 5 Assumed as the

difference between mass remaining at 105° and 500° C; 6 Total Carbon minus Non-Volatile OM;
7
Water and Volatiles minus Volatile OM

50
Table A3. Total Carbon Content
Sample Identification Measurements Results

Total Sample Area Detected Detected % Avg. % (wet),


Sample Sediment
Mass (mg) (0.1*uV*s) Carbon (mg) Carbon (wet)† Std Dev.

01A SD019901-T 4.01 300042 0.2532763 6.3161166 8.1415917


01B SD019901-T 1.53 157144 0.1524961 9.9670668 2.5816116
02A SD019901-D 6.44 414500 0.3339989 5.1863176 5.4464119
02B SD019901-D 5.67 399697 0.3235589 5.7065062 0.3678289
03A SD019822-T 8.32 1085714 0.8073788 9.7040724 8.4478268
03B SD019822-T 6.32 585372 0.4545079 7.1915812 1.7765995
04A SD019822-B 4.57 281128 0.239937 5.2502626 5.2063577
04B SD019822-B 4.62 279098 0.2385053 5.1624529 0.0620908
05A SD9821-T 5.3 700415 0.5356431 10.106473 10.108977
05B SD9821-T 3.93 504371 0.3973812 10.111481 0.0035412
06A SD9821-B 7.55 582869 0.4527427 5.9965916 5.8033317
06B SD9821-B 4.63 309216 0.2597463 5.6100718 0.2733108
07A 4N-T 5.08 876095 0.6595431 12.983131 13.952218
07B 4N-T 6.79 1377490 1.0131566 14.921305 1.3704957
08A 4N-B 5.56 506216 0.3986824 7.1705472 6.9426809
08B 4N-B 9.84 877789 0.6607378 6.7148147 0.3222515
09A 6N-T 5.88 825902 0.624144 10.614693 10.262391
09B 6N-T 4.9 629450 0.4855944 9.9100889 0.4982306
10A 6N-B 7.72 703734 0.5379838 6.9687026 6.4898421
10B 6N-B 3.28 220474 0.1971602 6.0109816 0.677211
std01 (check) acetanilide 0.591 548445 0.4284648 72.498278 71.26462
std07 (check) acetanilide 0.923 835078 0.6306154 68.322365 2.559109
std08 (check) acetanilide 0.968 942507 0.7063807 72.973217
% carbon of pure acetanilide = 71.09%
Calibration curve for carbon mass (mg) vs. Area (0.1*uV*s): mass = (7x10-07) x (Area) + 0.0417, R2 = 0.9998
†- 100 x [measured carbon mass]/ [total wet sample mass]

51
Table A4. Total Sediment PAH Concentrations and Batch Equilibrium Data
1 Top 1 Bottom 2 Top 2 Bottom 3 Top 3 Bottom 4 Top 4 Bottom 5 Top 5 Bottom
SD-99-01 Top SD-99-01 Bot SD-00-6N Top SD-00-6N Bot SD-98-21 Top SD-98-21 Bot SD-00-4N Top SD-00-4N Bot SD-98-22 Top SD-98-22 Bot
Csed Caq Csed Caq Csed Caq Csed Caq Csed Caq Csed Caq Csed Caq Csed Caq Csed Caq Csed Caq
(µg/ (µg/ (µg/ (µg/ (µg/ (µg/ (µg/ (µg/ (µg/ (µg/ (µg/ (µg/ (µg/ (µg/ (µg/ (µg/ (µg/ (µg/ (µg/ (µg/
Name gm) mL) gm) mL) gm) mL) gm) mL) gm) mL) gm) mL) gm) mL) gm) mL) gm) mL) gm) mL)
Ethylbenzene 275 219 750 0.711 397 0.737 744 1.292 148 0.849 1031 1.27 236 0.949 524.5 1.522 88.91
m,p-Xylenes 74.5 98.9 271 0.26 187 0.33 322 0.523 85.3 0.442 468 0.554 126 0.465 112.6 0.3 84.32
o-Xylene 105 103 274 0.275 173 0.262 325 0.589 59 0.351 333 0.444 129 0.355 154.8 0.457 100.41
1,3,5-Trimethylbenzene 77.9 29.1 91.7 0.049 83 0.036 114 0.064 29.2 0.048 291 0.057 68 0.049 122.2 0.072 38.62
1,2,4-Trimethylbenzene 181 110 420 0.149 252 0.17 585 0.367 111 0.264 592 0.282 160 0.229 296.1 0.204 56.23
Naphthalene 3573 4016 12428 10.95 7245 10.63 6162 8.242 2537 11.51 16739 13.32 5324 16.14 7754.6 16.19 1890
2-Methylnaphthalene 2196 2066 6010 1.491 3433 1.619 4283 1.896 1228 1.684 8873 2.241 2613 2.041 3708.8 2.457 822.17
Acenaphthalene 268 161 632 0.096 409 0.067 560 0.116 174 0.11 906 0.074 317 0.119 410.8 0.114 89.54
Acenaphthene 1301 1417 4112 0.827 2517 0.724 1991 0.603 804 0.682 6853 0.891 1988 1.015 2709.0 1.136 654.04
Fluorene 599 684 1999 0.256 1291 0.231 835 0.168 424 0.218 3144 0.26 1032 0.311 1216.8 0.332 327.23
Phenanthrene 1681 1932 5972 0.364 3827 0.252 2248 0.194 1167 0.236 8832 0.291 3102 0.317 3482.8 0.423 941.77

52
Anthracene 469 477 1409 0.101 987 0.066 680 0.054 569 0.075 7528 0.083 1174 0.086 1737.3 0.111 239.39
Fluoranthene 570 659 2121 0.113 1396 0.035 642 0.028 410 0.038 4800 0.044 1161 0.045 1283.6 0.081 385.32
Pyrene 765 879 2891 0.112 1897 0.037 1002 0.017 573 0.04 6407 0.041 1515 0.057 1733.6 0.078 475.98
Benzo(a)anthracene 324 381 1186 0.016 848 0 418 0.004 251 0 2563 0 618 0 673.3 0.025 222.14
Chrysene 299 331 928 0.011 656 0 449 0 269 0 2149 0 522 0 592.9 0 198.44

Solids Content (fsolid) 0.31 0.34 0.32 0.34 0.33 0.34 0.33 0.37 0.35 0.44
Loss at 500 o C (%) 20.9 17.7 24.3 22.4 22.5 21.4 30.3 19.9 20.5 12.7
Total AH (mg / g) † 37.1 32.4 99.9 58.8 78.3 24.6 220 43.5 62.7 14.9
Total Tar (mg / g) ‡ 61.8 54 167 98 130 41 367 72.5 104 24.9
Total Tar (mL / g) * 0.06 0.05 0.17 0.1 0.13 0.04 0.37 0.07 0.1 0.03
†- These data represent the sum of all measured aromatic hydrocarbons, plus the estimated mass of all unidentified peaks (assuming average
FID response and MW equal to phenanthrene). [mg aromatic hydrocarbons]/ [g wet sediment].
‡- These data represent "Total AH" data divided by 0.6 (to account for 40% non-extractable pitch), [mg coal tar]/ [g wet sediment].
*- These data represent ["Total tar" (mg tar/ g wet sediment) data] x [0.001 g/ mg] x [1.0 mL tar/ 1.0 g tar]. Units are [mL tar]/ [g wet sediment].
Table A5 - Grand Calumet River Annual
Mean Stream Flow Data, mMeasured in
Hammond, Indiana (USGS)
Year cfs cfd L/day
1992 23.8 2.06E+06 5.82E+07
1993 60.7 5.24E+06 1.49E+08
1994 46.8 4.04E+06 1.14E+08
1995 30.8 2.66E+06 7.54E+07
1996 43.2 3.73E+06 1.06E+08
1997 80.4 6.95E+06 1.97E+08
1998 52.9 4.57E+06 1.29E+08
1999 16.2 1.40E+06 3.96E+07
2000 14.1 1.22E+06 3.45E+07
average: 41.0 3.54E+06 1.00E+08

53

You might also like