Anda di halaman 1dari 52

L1608_C05.

fm Page 161 Friday, July 23, 2004 5:44 PM

5
The Transport and Fate of Cr(VI)
in the Environment

Frederick T. Stanin and Malcolm Pirnie

CONTENTS
5.1 The Presence of Chromium in the Environment .................................163
5.1.1 Antropogenic Sources...................................................................164
5.1.2 Natural Sources .............................................................................165
5.2 Geochemistry of Chromium ....................................................................165
5.2.1 Cr(III)...............................................................................................166
5.2.2 Cr(VI)...............................................................................................167
5.3 Oxidation-Reduction of Chromium........................................................168
5.3.1 Review of Oxidation-Reduction Reactions ...............................168
5.3.2 General Redox Behavior of Chromium
in the Environment .......................................................................169
5.3.3 Oxidation of Chromium...............................................................170
5.3.3.1 Oxidation of Cr(III) to Cr(VI) by Dissolved
Oxygen and Manganese Dioxides ...............................171
5.3.3.2 Oxidation of Cr(III) to Cr(VI) by H2O2 .......................172
5.3.4 Reduction of Cr(VI) to Cr(III) .....................................................172
5.3.4.1 General..............................................................................172
5.3.4.2 Fe(II) (Dissolved Fe(II) and Fe(II)-Containing
Minerals)...........................................................................173
5.3.4.3 Sulfides .............................................................................175
5.3.4.4 Organic Matter ................................................................175
5.3.4.5 Cu(I) ..................................................................................176
5.3.4.6 Hydrogen Peroxide(H2O2).............................................176
5.4 Precipitation/Dissolution Reactions of Chromium .............................176
5.5 Sorption and Desorption Reactions of Chromium ..............................177
5.5.1 General Discussion of Sorption ..................................................177
5.5.2 Sorption of Chromium .................................................................178
5.5.2.1 Sorption of Cr(III) ...........................................................178
5.5.2.2 Sorption of Cr(VI) ...........................................................179

1-5667-0608-4/01/$0.00+$1.50 161
2004 by CRC Press LLC
L1608_C05.fm Page 162 Friday, July 23, 2004 5:44 PM

162 Chromium(VI) Handbook

5.6 General Transport and Fate of Chromium


in Environmental Media...........................................................................180
5.6.1 Chromium in the Atmosphere....................................................181
5.6.2 Chromium in Aquatic Environments ........................................182
5.6.2.1 Surface Waters .................................................................182
5.6.2.2 Groundwater ...................................................................184
5.6.2.3 Plumes of Chromium in GroundwaterCase
Studies ..............................................................................186
5.6.2.3.1 Nassau County, New York..........................186
5.6.2.3.2 Telluride, Colorado.......................................187
5.6.2.3.3 Southwestern MichiganWood
Treatment Plant .............................................187
5.6.2.3.4 Industrial Waste Landfill,
Northern France............................................188
5.6.3 Chromium in Soil..........................................................................188
5.6.3.1 Overview of Metals in Soil ...........................................189
5.6.3.2 Behavior of Chromium in Soil......................................191
5.6.3.2.1 Sorption of Cr(III) and Cr(VI) ....................193
5.6.3.2.2 Reduction of Cr(VI) and Oxidation
of Cr(III) .........................................................193
5.6.4 The Uptake and Transformation of Chromium by Biota.......195
5.7 Utilizing Natural Environmental Processes as a Remedy
for Soil and Groundwater Contaminated with Chromium ...............195
5.7.1 There Are Natural Reductants in the Aquifer..........................197
5.7.2 The Amount of Cr(VI) and Other Reactive Constituents
Do Not Exceed the Reductive Capacity of the Aquifer .........198
5.7.2.1 Mass of Cr(VI) .................................................................198
5.7.2.2 Mass of Cr(III) .................................................................199
5.7.2.3 Reduction Capacity of the Aquifer ..............................199
5.7.2.4 Oxidation Capacity of the Aquifer ..............................200
5.7.3 The Rate of Cr(VI) Reduction to the Target Concentration
Compared to the Rate of Transport of Cr(VI)
from Source to Point of Compliance..........................................200
5.7.3.1 Rates of Oxidation and Reduction...............................201
5.7.3.2 Estimating Reduction from Monitoring
Well Data ..........................................................................201
5.7.3.3 Monitoring Reduction Via Stable
Isotopes of Chromium ...................................................202
Bibliography ......................................................................................... 204

The knowledge of the transport and fate of contaminants in the subsurface


environment is essential to achieving environmental restoration objectives.
Unfortunately, gaining and using this knowledge can be difficult because of
problems and limitations involving complex hydrogeology, hydrochemistry,
and microbiology, or by economical restraints. This chapter discusses the
L1608_C05.fm Page 163 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 163

most important physical, chemical, and biological parameters to understand-


ing the transport and fate of Cr(VI), and chromium in general, and how
these parameters can be measured and utilized in environmental restoration
projects. These discussions are drawn from several sources, as referenced.
The assessment of the transport and fate of any contaminant in the envi-
ronment generally consists of site characterization, risk assessment, and
remediation, which are the three main phases of environmental restoration
programs. Conventional approaches to site characterization may not ade-
quately define the need to obtain enough detailed information about natural
processes affecting the transport behavior and the ultimate fate of contami-
nants. The use of state-of-the-art site characterizations, although more costly
to implement than more conventional means, may ultimately result in sig-
nificant savings because of improved technical effectiveness and efficiency
of site cleanup. Also, proper site characterization methods can aid risk man-
agement decisions (risk assessments) in determining if remediation is even
necessary and/or to what extent, and choosing the proper cleanup technol-
ogies if remediation is needed.
A sound conceptual site model (CSM) is absolutely necessary for transport
and fate assessments. An adequate CSM incorporates information on geo-
logic, hydrologic, chemical, and biological processes to produce an effective
contaminant transport evaluation. The practical use of risk assessments and
remedial technologies is highly dependent on site-specific knowledge of
these processes. Therefore, the processes that govern the subsurface behavior
and treatability of contaminants must be understood. This is particularly
true for Cr(VI).
This chapter discusses the transport and fate of Cr(VI) as well as other
forms of chromium in the environment, most notably Cr(III). These forms
or species of chromium are inexorably linked by many environmental trans-
port and fate processes. First presented is an overview of the presence of
chromium in the environment and its general geochemistry. The most impor-
tant transport and fate processes are discussed, namely oxidation-reduction,
precipitation-dissolution, and sorption-desorption. Then, the general aspects
of the transport and fate of chromium in the atmosphere, aquatic environ-
ments (surface waters and groundwater), and soil, including the uptake and
transformation of Cr(VI) by biota, are summarized. The chapter concludes
with a discussion of how natural attenuation processes can be implemented
as a remedy for sites contaminated by Cr(VI).

5.1 The Presence of Chromium in the Environment


Chromium contamination of soil and groundwater is a significant problem
worldwide. The extent of this problem is due primarily to its use in numerous
industrial processes (i.e., metal plating and alloying, leather tanning, wood
L1608_C05.fm Page 164 Friday, July 23, 2004 5:44 PM

164 Chromium(VI) Handbook

treatment, etc.), but also its natural presence in rocks enriched in chromium
(i.e., ultramafic rocks such as serpentinite). Compared to the results of con-
tamination of soil and groundwater by industrial practices, the naturally-
occurring concentrations of chromium in soil and groundwater are relatively
low. However, relatively high concentrations of naturally-occurring dis-
solved chromium have been observed, usually associated with the very
soluble chromate species (Robertson, 1975). Thus, both anthropogenic (man-
made) and natural sources of chromium can lead to locally elevated levels
in soils and waters.
The presence of chromium in the environment is discussed by several
authors (Davis and Olsen, 1995; Kimbrough et al., 1999; Richard and Bourg,
1991; Kotas and Stasicka, 2000). Their presentations are briefly summarized
below. Chapter 4 gives additional information on naturally-occurring Cr(VI)
in groundwater.

5.1.1 Antropogenic Sources


Chromium is used in several industries, including metallurgy (steel, ferro-
and nonferrous alloys), refractory (chrome and chrome-magnesite), and
chemical manufacturing (pigments, electroplating, tanning and other),
involving numerous commercial processes including electroplating, leather
tanning, pulp production, milling, mining (ore refining), and wood preser-
vation. The industrial use of chromium generally begins with the mining of
chromite (a naturally-occurring ore), usually as ferrous chromite (FeO Cr2O3)
(Hartford, 1983). Then, the ore is either oxidized or reduced during industrial
processing. Sodium chromate (Na2CrO4) has usually been produced by the
oxidation of chromite. Sodium carbonate, calcium oxide, and calcium chro-
mate (CaCrO4) are produced as byproducts. In turn, several substances are
derived from Na2CrO4, including dichromates (Na2Cr2O7 or K2Cr2O7), Cr(VI)
oxide (CrO3), chromic acid (H2Cr2O7), and other oxides of Cr (e.g., K2CrO4),
including chromium pigments (barium, calcium, lead, strontium, and zinc
chromate). Also, chromite ore can be reduced by a variety of methods using
aluminum, silicon, or carbon as reducing agents (Hathway, 1989). Materials
from this reduction are used for producing chromium alloys and chrome
alum (NH4Cr(SO4)2 12H2O).
Most of the chromium consumed by industry in the U.S. is for the pro-
duction of metal alloys, mainly wrought-stainless and heat-resisting steels
(Hartford, 1983). Chromium as part of an iron alloy is insoluble with a zero
oxidation state and therefore is not a form of chromium having an environ-
mental concern. However, Cr can be oxidized and leached from stainless
steel into a water-soluble form (Kimbrough et al., 1999).
Chromium from anthropogenic sources can be released to soils and sedi-
ments indirectly by atmospheric deposition, but releases are more commonly
from dumping of Cr-bearing liquid or solid wastes such as chromate byprod-
ucts (muds), ferrochromium slag, or chromium plating wastes. Such wastes
can contain any combination of Cr(III) or Cr(VI) with various solubilities.
L1608_C05.fm Page 165 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 165

The nature and behavior of various forms of chromium found in wastewaters


can be quite variable. The presence, form and concentration of Cr in dis-
charged effluents depends mainly on the chromium compounds utilized in
the industrial process, on the pH, and on the presence of other organic and
inorganic processing wastes.
Chemicals containing Cr(VI) are principally used for metal plating (which
use H2Cr2O7), as dyes, paint pigments, and leather tanning (Hartford, 1983).
Chromium platers involve the use of H2Cr2O7 to plate chromium onto pieces
of other metals. Thus, Cr(VI) will dominate in wastewater from the metal-
lurgical industry, metal finishing industry (Cr hard plating), refractory indus-
try and production or application of pigments (chromate color pigments and
corrosion inhibition pigments). Cr(III) will be found mainly in wastewaters
of the tannery, textile (printing, dying) and decorative plating industries.
However, there are exceptions to these generalities due to several factors.
For example, in tannery wastewater where Cr(III) is the most expected form,
the redox reactions occurring in sludge can increase the concentration of
Cr(VI). Such transformations are also common in the subsurface, such as
oxidation, reduction, sorption, precipitation, and dissolution, which are all
discussed later in this chapter.

5.1.2 Natural Sources


As previously mentioned, chromium occurs naturally in the environment,
most notably in its most concentrated forms as an ore mineral. Chromium
also occurs naturally as a component of soils (Schacklette and Boerngen,
1984), usually as chromite (FeCr2O4), a relatively insoluble soil mineral
(Schmidt, 1984). The main source of such chromium in natural soils is the
weathering of their parent materials. The average amount of this element in
various kinds of soils ranges from 0.02 to 58 micromoles per gram (Richard
and Bourg, 1991; Coleman, 1988). The chromium concentration will be influ-
enced by the composition of the parent rock. Granite, carbonates and sandy
sediments present the lowest chromium content whereas shales, river sus-
pended matter, and soils typically exhibit highest levels. Highest chromium
contents tend to be associated with finest grain size soils (Robertson, 1975)
and sediments (Salomons and DeGroot, 1978). Thus, the natural concentra-
tion of chromium in the environment varies greatly (Cary, 1982).

5.2 Geochemistry of Chromium


Chromium can exist in several chemical forms with oxidation numbers rang-
ing all the way from 2 to +6. However, in the environment, chromium
commonly exists in only two stable oxidation states, Cr(VI) and Cr(III), which
have greatly contrasting toxicity and transport characteristics. Chromium
speciation in the environment, particularly in groundwater, is affected
L1608_C05.fm Page 166 Friday, July 23, 2004 5:44 PM

166 Chromium(VI) Handbook

primarily by Eh (oxidizing or reducing conditions) and pH (acidic or alkaline


conditions). In general, Cr(VI) predominates under oxidizing conditions, and
Cr(III) predominates under more reducing conditions.
These two different forms of chromium are quite different in their prop-
erties: charge, physiochemical characteristics, mobility in the environment,
chemical and biochemical behavior, bioavailability, and toxicity. Most nota-
bly, Cr(III) is considered to be a trace element essential for the proper func-
tioning of living organisms, whereas Cr(VI) exerts toxic effects on biological
systems. Also, Cr(VI) compounds are generally more soluble, mobile, and
bioavailable in the environment compared with Cr(III) compounds. And, as
previously mentioned, the more toxic and mobile Cr(VI) predominates in
oxidizing environments, while the less toxic and immobile Cr(III) is restricted
to reducing environments. Therefore, it is quite important to distinguish
these forms of chromium rather than discussing this element as total chro-
mium. The geochemistry of these forms are briefly discussed below. A more
comprehensive discussion of the geochemistry of chromium and chromium
compounds is presented in Chapter 2.

5.2.1 Cr(III)
In aqueous systems, Cr(III) can be present as Cr3+, Cr(OH)2+, Cr(OH)2+, and
Cr(OH)4. Additionally, the precipitated phase Cr(OH)3 predominates
between pH 6 and pH 12 (Rai et al., 1987). Under slightly acidic to alkaline
conditions, and if Fe(III) is present, Cr(III) can precipitate as an amorphous
mixed hydroxide CrxFe1x(OH)3 (Eary and Rai, 1988). Amorphous Cr(OH)3
can crystallize as Cr(OH)3 3H2O or Cr2O3 (eskolaite) under different condi-
tions (Palmer and Puls, 1994). With high redox potential, Cr(VI) predomi-
nates with a much higher solubility (Loyaux-Lawniczak et al., 2001). In a
reducing environment, and in the absence of Fe, Cr(III) precipitates readily
to form Cr(OH)3 (Rai et al., 1987).
In relatively low Eh environments, the main aqueous Cr(III) species are
Cr3+, Cr(OH)2+, Cr(OH)30 and Cr(OH)4 (Rai et al., 1986, 1987). The Cr3+ species
is prevalent only at pH lower than about 4. With increasing pH, hydrolysis
of Cr3+ yields Cr(OH)2+ (generally present in groundwater at a pH of 6 to 8
but also in some acidic waters), and Cr(OH)30 and Cr(OH)4 (generally in
alkaline groundwater) (Rai et al., 1987). Polymeric species such as Cr2(OH)24+,
Cr3(OH)45+ and Cr4(OH)66+ are never significant in natural systems (Rai et al.,
1986, 1987). Cr(III) readily forms complexes with a variety of ligands:
hydroxyl, sulfate, ammonium (NH4), cyanide and sulphocyanide, fluoride
and chloride (to a lesser extent), and natural and synthetic organic ligands
(Richard and Bourg, 1991). Only one Cr(III) compound, Cr2O3, is an oxide,
so the role of oxygen is central to the oxidation/reduction (redox) process
for chromium (Kimbrough et al., 1999).
Solubility can significantly limit the concentration of Cr(III) in groundwa-
ter at a pH above 4. The low solubility of the Cr(III) solid phases, Cr2O3 and
L1608_C05.fm Page 167 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 167

Cr(OH)3 (Hem, 1977), is likely the major reason why Cr(III) generally makes
up a small percentage of the total chromium concentration in natural or
contaminated groundwaters. Cr(III) tends to be essentially immobile in most
groundwaters because of its low solubility (Calder, 1988).

5.2.2 Cr(VI)
Chromium(VI) exists in the environment as part of several compounds, Cr(VI)
is present in solution as monomeric forms: H2CrO40, HCrO4 (bichromate),
CrO42 (chromate), and CrO3 (chromium(III) oxide), or as the dimeric ion
Cr2O72 (dichromate) (Palmer and Puls, 1994). Under oxidizing conditions,
aqueous chromium is present in a Cr(VI) anionic form, HCrO4 or CrO42,
depending on the pH (CrO42 at a higher pH) (Richard and Bourg, 1991).
Within the normal pH range in natural waters (i.e., 6 to 8), the CrO42, HCrO4
and Cr2O72 ions are the forms expected. At relatively high Cr(VI) concentra-
tions, the Cr2O72 ion predominates in acidic environments (Richard and
Bourg, 1991).
It should be noted here that the term Cr(VI) is somewhat of a misnomer.
This is because Cr(VI) is not present in the environment as a free cation
(whereas Cr(III) does exist in the environment as previously mentioned). In
fact, as all Cr(VI) species are oxides, they act like a 2 anion (ion2) rather
than a Cr(VI) cation (Kimbrough et al., 1999).
The relative concentration of the various Cr(VI) species depends on the
pH and the total Cr(VI) concentration (Palmer and Puls, 1994). For example,
significant concentrations of H2CrO40 only occur under the extreme condition
of pH around 1. Above a pH of about 6, CrO42 generally dominates (Davis
and Olsen, 1995). Below pH of about 6, HCrO4 dominates when the Cr(VI)
concentrations are relatively low, but Cr2O72 becomes more significant as
Cr(VI) concentrations increase, or it may even dominate when the total
Cr(VI) concentrations are relatively high (Palmer and Puls, 1994). These
species constitute many of Cr(VI) compounds which are quite soluble and
thus mobile in the environment (Kotas and Stasicka, 2000). These species
differ in their solubility and in their tendency to be sorbed by soil or aquifer
materials (Calder, 1988).
There are no significant solubility constraints on the concentrations of
Cr(VI) in groundwater. The chromate (CrO42) and dichromate ions
(Cr2O72) are water soluble at all pH. However, chromate can exist as an
insoluble salt of a variety of divalent cations, such as Ba2+, Sr2+, Pb2+, Zn2+,
and Cu2+, and these salts have a wide range of solubilities. The rates of
precipitation/dissolution reactions between chromate, dichromate
anions, and these cations vary greatly and are pH dependent. An under-
standing of the dissolution reactions is particularly important for assess-
ing the environmental effects of chromium because Cr(VI) often enters
the environment by dissolution of chromate salts (e.g., SrCrO4) (Rai et al.,
1987).
L1608_C05.fm Page 168 Friday, July 23, 2004 5:44 PM

168 Chromium(VI) Handbook

5.3 Oxidation-Reduction of Chromium


Chromium in the environment is altered by oxidation-reduction reactions,
changing its physical and chemical properties. To understand these changes,
it is worthwhile to review the basics of oxidation and reduction.

5.3.1 Review of Oxidation-Reduction Reactions


Oxidation-reduction reactions involve the transfer of electrons. In oxidation-
reduction reactions, some components (charged/uncharged atoms) lose
electrons and some components gain electrons from this transfer. The pro-
cess of removing electrons from a component (loss of electrons) is called
oxidation and results in a more positive oxidation number. After oxidation
has occurred, the component is said to have been oxidized. The process of
adding electrons to an atom (gain of electrons) is called reduction, and results
in a more negative oxidation number. After reduction has occurred, the com-
ponent is said to have been reduced. The component which gains electrons
in an oxidation-reduction reaction is called the oxidizing agent, whereas
the component which loses electrons is the reducing agent. Due to the
conservation of mass, and also the conservation of electrons and oxidation
numbers, electrons lost by the oxidized component are gained by the oxi-
dizing agent. Therefore, oxidation is always accompanied by reduction, and the
oxidation and reduction always takes place to an equal degree. For this reason,
reference is usually made to combined oxidation-reduction reactions (redox
reactions) rather than separate oxidation or reduction reactions. It is some-
times useful to consider the oxidation reaction and reduction reaction sep-
arately as 1/2 reactions.
The potential for an electron transfer is best measured by the redox poten-
tial (Eh), which is sometimes expressed as the redox intensity factor (pe),
the negative log of the electron activity (ae),

pe = log(ae) (5.1)

Also,

Eh = (pe)(2.3RT/F)

= 0.059 pe

where
R = gas constant
T = temperature
F = Faraday constant
L1608_C05.fm Page 169 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 169

essentially the ratio of electron donors (oxidizing agents) to electron acceptors


(reducing agents) (Kimbrough et al., 1999).

5.3.2 General Redox Behavior of Chromium in the Environment


The oxidation and reduction of chromium in the environment is discussed by
several authors, including Kimbrough et al. (1999), Richard and Bourg (1991),
and Calder (1988). Their presentations are summarized below, followed by
separate, more detailed discussions of chromium oxidation and reduction.
The distribution between Cr(III) and Cr(VI) in the environment, including
aquatic environments such as groundwater, will be regulated by redox
reactions and redox conditions (Richard and Bourg, 1991). To understand
the distribution of Cr(III) and Cr(VI), Eh-pH diagrams for chromium in
aqueous environments are usually employed, such as the one discussed
earlier in Chapter 3.
The redox transformation of Cr(III) to Cr(VI), the oxidation of Cr(III), or the
reduction of Cr(VI) to Cr(III), requires another redox couple (of oxidizing/
reducing agent) which accepts or gives the necessary electrons. In natural
aquatic environments, the significant redox couples (reducing agents/oxida-
tion agents) are (Richard and Bourg, 1991):

H2O/O2 (aq)
Mn(II)/Mn(IV)
NO2/NO3
Fe(II)/Fe(III)
S2/SO42
CH4/CO2

In the case where O2 is the oxidizing agent, oxidation of chromium requires


donation of electrons to oxygen, while reduction of oxygen requires the accept-
ing of electrons from chromium. For a given chromium compound, the redox
reactions involving other chemical agents (i.e., oxidizing and reducing agents)
are governed by the agents capacity for donating or accepting electrons. Sev-
eral oxidation and reduction reactions of chromium with common environ-
mental agents are given in several sources, including Kimbrough et al. (1999).
Oxidation Example:

Cr(III) Cr(VI)

2Cr2O3 + 3O2 4CrO3

The concentration of these oxidizing and reducing agents affects the oxi-
dation-reduction of chromium. Many oxidizing agents are known to oxidize
Cr(III) to Cr(VI), but only a few of them are found in the environment (i.e.,
groundwater) in sufficient concentration to do so. On the other hand, the
L1608_C05.fm Page 170 Friday, July 23, 2004 5:44 PM

170 Chromium(VI) Handbook

many reducing agents for Cr(VI) reduction to Cr(III) are typically found at
sufficient levels. For example, ozone (O3) can theoretically oxidize Cr(III) to
chromate (a reaction with Eh = 0.87 volts), but the concentration of ozone in
the environment is usually insufficient (it is relatively unstable) to accom-
plish this oxidation (Grohse et al., 1988). However, the reduction of Cr(VI)
by Fe(II), even though less favored thermodynamically (a reaction with Eh
only 0.56 volts) is feasible because iron concentrations are generally sufficient
in the environment (Rai et al., 1989).
At lower pHs, chromates exist as chromic acid (H2CrO4) and hydrogen
chromate (HCrO4). When the concentration of CrO42 is high, chromates are
transformed to dichromate (H2Cr2O7 or HCr2O7), which are strong oxidizing
agents and are thus rapidly reduced in the presence of reducing agents at
low pH or high Eh. At high pH, chromates exist in the form of CrO42, which
is a poor oxidizing agent and hence, with lower Eh values, are more stable.
Cr(VI) can be transported great distances in groundwater due in part to
its high solubility, whence it may be transformed by reduction to, and pre-
cipitated as, Cr(III) if the transported Cr(VI) enters an area with relatively
low Eh. Cr(VI) can be reduced readily to Cr(III) in the presence of organic
matter, especially where pH is low (Bartlett and Kimble, 1976; Bloomfield
and Pruden, 1980). Cr(VI) can also be reduced by Fe(II) and dissolved sul-
fides (Schroeder and Lee, 1975).
Cr(III) generally is not transported great distances by groundwater because
of its low solubility. However, Cr(III) can be transformed to the more soluble
Cr(VI) if the redox conditions along the transport pathway change from
reducing to oxidizing. Under natural conditions, Cr(III) has been found to
be oxidized to Cr(VI) by manganese (Schroeder and Lee, 1975; Bartlett and
James, 1979). In the laboratory, Cr(III) can exist as highly soluble organic
complexes, particularly under low pH conditions (Bartlett and Kimble, 1976;
James and Bartlett, 1983). Therefore, if Cr(III) is in a complexed form, it could
be present at much higher concentrations in groundwater than if it is uncom-
plexed. However, the existence of Cr(III) complexes has not been docu-
mented under field conditions.

5.3.3 Oxidation of Chromium


There are several sources of oxygen for the oxidation of chromium. In the
environment, water is the most important source; manganese dioxide, ozone,
hydrogen peroxide (H2O2), manganese dioxide, and lead dioxide are other
notable sources of oxygen (Kimbrough et al., 1999). Oxidation of chromium
involving these sources requires the presence of water, thus water chemistry
is important to the understanding of chromium oxidation. Generally, high
values of Eh in water correspond to strongly oxidizing conditions. A sum-
mary of Cr(III) oxidation via dissolved oxygen, manganese dioxides, and
hydrogen peroxides are discussed below, based on papers by Kotas and
Stasicka (2000); Loyaux-Lawniczak et al. (2001); Palmer and Puls (1994);
Davis and Olsen (1995); and Richard and Bourg (1991).
L1608_C05.fm Page 171 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 171

5.3.3.1 Oxidation of Cr(III) to Cr(VI) by Dissolved Oxygen


and Manganese Dioxides
The redox potential of the Cr(VI)/Cr(III) couple is high enough so that only
a few oxidants are present in natural systems capable of oxidizing Cr(III) to
Cr(VI). Only dissolved oxygen and manganese dioxides (MnO2) are known
to oxidize Cr(III) to Cr(VI) (Eary and Rai, 1987) and manganese dioxides are
the more common oxidant; oxidation of Cr(III) by dissolved oxygen without
any mediate species has been reported to be negligible (Schroeder and Lee,
1975; Eary and Rai, 1987), whereas mediation by manganese oxides has been
found to be the effective oxidation pathway in environmental systems
(Schroeder and Lee, 1975; Bartlett and James, 1979; Nakayama et al., 1981;
Saleh et al., 1989; Johnson and Xyla, 1991).
Dissolved oxygen can oxidize Cr(III) into Cr(VI), (Rai et al., 1986; Schroeder
and Lee, 1975; Eary and Rai, 1987; Nakayama et al., 1981) but laboratory
studies indicate that this can be relatively slow requiring several months
(Palmer and Wittbrodt, 1990), especially in slightly acidic to basic environ-
ments (Eary and Rai, 1987). Such slow kinetics enable Cr(III) to be involved
in other reactions (sorption or precipitation) that are much faster. Therefore,
the oxidation of Cr(III) in the environment by dissolved oxygen is unlikely.
Manganese oxides are likely to be responsible for most Cr(III) oxidation
in aquatic environments. Fendorf and Zasoski (1992) suggest that CrOH2+ is
the reactive species in this Cr(III) oxidation. Bartlett and James (1979)
observed a correlation between the amount of Cr(III) oxidized by soils and
the amount of reduced manganese in soils, thereby suggesting the oxidation
of Cr(III) is the result of interaction with manganese dioxides, which has
been verified by laboratory studies. Experimental results indicate that the
oxidation follows the reaction (Palmer and Puls, 1994):

CrOH2+ + 1.5 MnO2 HCrO4 + 1.5Mn2+


(5.2)
2 CrOH2+ + 3 MnO2 2HCrO4 + 3Mn2+

Manganese oxides are present in the subsurface as grain coatings, deposits


in cracks or fractures, or as finely disseminated grains; sometimes this pres-
ence is a result of bacterial activities. The mechanisms for the reaction with
MnO2 occurring at the manganese oxide surfaces (by adsorption of Cr(III)
on active surface sites) are very complex and not yet fully understood (Eary
and Rai, 1987; Fendorf and Zasoski, 1992). The Cr(III) oxidation rate is likely
related to the amount and surface area of manganese oxides (Schroeder and
Lee, 1975; Eary and Rai, 1987), and lab studies indicate this rate to be initially
rapid, but then slowing down significantly. Also, there is an increase in the
rate and amount of Cr(III) oxidation as pH decreases, and the surface area
to solution volume increases.
Richard and Bourg (1991) explain that the oxidation of Cr(III) by manga-
nese dioxides is likely to occur as a result of three sequential steps (Rai et al.,
1986; Schroeder and Lee, 1975; Bartlett and James, 1979; Eary and Rai, 1987;
L1608_C05.fm Page 172 Friday, July 23, 2004 5:44 PM

172 Chromium(VI) Handbook

Amacher and Baker, 1982). First, the Cr(III) would be sorbed onto MnO2
surface sites. Then, the Cr(III) would be oxidized to Cr(VI) by Mn(IV) on the
surface sites, however, all the Mn(IV) reaction sites are probably not acces-
sible to Cr(III) (Rai et al., 1986; Amacher and Baker, 1982). Finally, the reaction
products, Cr(VI) and Mn(II), would be desorbed. Richard and Bourg (1991)
give theoretical stoichiometries that have been suggested for this oxidation:

2Cr3+ + 3 MnO2 (s) + 2H2O = 2HCrO4 + 3Mn2+


+ 2 H+ (Amacher and Baker, 1982) (5.3)

and

Cr(OH)2+ + 3 MnO2 (s) + 3H2O = HCrO4 + 3MnOOH (s)


+ 3H+ (Eary and Rai, 1987) (5.4)

The solid MnOOH(s) would decay into aqueous Mn2+ afterwards.

5.3.3.2 Oxidation of Cr(III) to Cr(VI) by H2O2


In an attempt to mobilize Cr(III) by oxidizing it to Cr(VI), H2O2 was applied
to groundwater in laboratory studies by Davis and Olsen (1995). Also, Pettine
et al. (2002) have studied the oxidation of Cr(III) with H2O2 in basic solutions
(Pettine and Millero, 1990 and 1991). They found that H2O2 controls the rate
of oxidation of Cr(III) in surface waters.

5.3.4 Reduction of Cr(VI) to Cr(III)


5.3.4.1 General
Cr(VI) is a strong oxidant and therefore can be reduced in the presence of
electron donors. The most common forms of chromium dissolved in natural
waters, within the environmentally normal range of pH, are CrO42, HCrO4
and Cr2O72 ions, (Kotas and Stasicka, 2000) which form many of the Cr(VI)
compounds that can be quite readily reduced to Cr(III) forms in the presence
of electron donors like organic matter and inorganic compounds in their
reduced state, many of which are quite common in soil, water, and the
atmosphere (Stollenwerk and Grove, 1985). The major factors in this reduc-
tion to Cr(III) are dissolved Fe(II), minerals with Fe(II), sulfides (reduced
sulfur), and organic matter (Kotas and Stasicka, 2000; Palmer and Puls, 1994;
Wielinga et al., 2001). Studies of reaction kinetics by Fendorf et al. (2001)
indicate that Fe(II) and dissolved sulfides will probably dominate the reduc-
tion of chromate (Wielinga et al., 2001). Loyaux-Lawniczak et al. (2001)
report that photoreduction is another pathway to reduce Cr(VI) in the
environment (Kieber and Heiz, 1992; Hug et al., 1997), but this mechanism
is probably only important in the atmosphere or in upper surface waters.
L1608_C05.fm Page 173 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 173

Much of the information of Cr(VI) reduction is based on laboratory studies,


and many are referenced herein. However, it should always be noted that
the application of these experimental observations to field situations
remains dubious.
Cr(VI) can be reduced by biological and chemical (abiotic) processes. It is
difficult to determine which processes are responsible for the reduction of
metal contaminants. However, it is probable that the reaction rates will deter-
mine the reduction process and its specific pathway. By comparing reduction
rates involving Fe(II) and sulfides with those reported for direct microbial
reduction, the chemical reduction of chromate by Fe(II) is more than 100 times
faster than the observed biological reduction rate, thus chemical reduction of
Cr(VI) will probably be the main process for chromate reduction when either
ferrous iron or sulfide are present, and these are present in the environment
under anaerobic conditions (Wielinga et al., 2001). Yet, both aerobic and
anaerobic reduction by microbes have been observed, the latter being more
common (Palmer and Puls, 1994). The specific mechanisms for Cr(VI) reduc-
tion by these microbes is not well known, but the chromate may actually be
used as an electron acceptor for cell metabolism (Palmer and Puls, 1994).
The major factors in the reduction of Cr(VI) to Cr(III), namely dissolved
Fe(II), minerals with Fe(II), sulfides (reduced sulfur), and organic matter, are
discussed below, based on papers by Pettine et al. (2002); Palmer and Puls
(1994); Wielinga et al. (2001); and Richard and Bourg (1991). The roles of
copper and H2O2 are also discussed.

5.3.4.2 Fe(II) (Dissolved Fe(II) and Fe(II)-Containing Minerals)


Fe(II) is a major factor in the reduction of Cr(VI) to Cr(III)experimental
results of Davis and Olsen, 1995 from column tests were consistent with
other published observations (Schroeder and Lee, 1975) that found Cr(VI)
to be reduced to Cr(III) by Fe(II). Dissolved Fe(II) ions in environmental
waters can be generated by the discharges of some industrial wastes, but
also can result from the weathering of Fe(II)-containing minerals. There are
numerous minerals in geologic materials that contain Fe(II) for Cr(VI)
reduction. These minerals include silicates, oxides, or sulfides (Palmer and
Puls, 1994):

Silicates: Olivine, pyroxenes (augite and hedenbergite), amphiboles


(hornblende, cummingtonite, and grunerite), micas (biotite, phlo-
gopite, and glauconite), chlorites, and clays (the smectite nontron-
ite).
Oxides: Magnetite, ilmenite, and hematite.
Sulfides: Pyrite, in which both the iron(II) and the sulfide are active
in reducing Cr(VI).

Cr(VI) reduction via Fe(II) in silicate minerals (e.g., biotite in solution


rather than at the mineral surface) was reported by Eary and Rai. (1989).
L1608_C05.fm Page 174 Friday, July 23, 2004 5:44 PM

174 Chromium(VI) Handbook

This is described by Palmer and Puls (1994) as a rather complex process. The
presence of Fe3+ increases the reduction rate, and the Fe3+ is reduced at the
mineral surface. The iron in the crystal structure is oxidized, K+ is released
to solution, and Cr(VI) in solution is then reduced by the Fe2+. The Fe3+
resulting from this reduction reaction is then sorbed to the surface of the
biotite where it is again reduced to Fe2+.
Palmer and Puls (1994) state that Cr(VI) reduction in the presence of iron
oxides has been observed in several experiments (Eary and Rai, 1989; White
and Hochella, 1989). In the case of hematite, the reduction is suggested to
occur in solution after the FeO component has dissolved. The reduction of
Cr(VI) via Fe(II) in pyrite is described in work by Lancy (1966), who sug-
gested that pyrite could be used for treating spent cooling waters that contain
Cr(VI) as a corrosion inhibitor, because reduction of Cr(VI) occurs at the
pyrite surface rather than in solution. This reduction was described as occur-
ring even in slightly alkaline solutions; however, the pyrite had to be con-
tinuously abraded to remove surface coatings. Batch testing reported by
Blowes and Ptacek (1992) involving pyrite both in the presence and in the
absence of calcite showed faster removal of Cr(VI) with no calcite. Also,
Loyaux-Lawniczak, et al. (2001) demonstrated that Fe(II)-Fe(III) hydroxysalt
green rusts can reduce Cr(VI); ferrihydrite is the Cr(III)-bearing solid phase
that is formed from this reduction.
Under neutral to alkaline pH conditions, Fe(II) controls the reduction of
Cr(VI) in natural anaerobic systems (Pettine et al., 1998), while at acidic pH
levels, other reductants may be more efficient than Fe(II). The involvement
of Fe(II) in Cr(VI) reduction, where the Fe(II) (as FeO) comes from hematite
or biotite, can be expressed as follows (Richard and Bourg, 1991):

[3FeO] + 6H+ + Cr(VI) (aq) = Cr(III) (aq) + 3Fe(III) (aq) + 3H2O (5.5)

This can be a relatively rapid reaction from the standpoint of environmen-


tal situations; laboratory studies report this reaction being complete in less
than 5 min (Eary and Rai, 1988). In acidic waters, the end products of this
reaction are Fe(III) and Cr(III) (Stollenwerk and Grove, 1985), whereas under
neutral to alkaline conditions, Cr(OH)3 is probably the end product because
of the very low solubility of Fe(OH)3 (Rai et al., 1988). In groundwaters of
pH more than 4, Cr(III) precipitates with the Fe(III) in a solid solution with
the general composition CrxFe1x(OH)3 (Eary and Rai, 1988; Rai et al., 1988;
Sass and Rai, 1987).
Wielinga et al. (2001) reported that the Cr(VI) reduction via Fe(II) (or
sulfide) depends on microbial activity. A diverse and widely distributed
group of bacteria are able to couple the oxidation of organic compounds
or H2 to the reduction of iron (hydr)oxides (Lovely, 1993; Coates et al.,
1996). Thus, in many environments where iron reduction is the predominant
terminal electric accepting process (TEAP) in microbial respiration, the
indirect reduction of Cr(VI) (chromate) via reaction with a respiratory
L1608_C05.fm Page 175 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 175

byproduct is likely a dominant reductive pathway as shown below (Wielinga


et al., 2001):

3C3H5O3 + 12Fe(OH)3 3C2H3O2 + 12Fe2+


+3HCO3 + 8H2O + 21OH (5.6)

3Fe2+ + HCrO4 + 8H2O 3Fe(OH)3 + Cr(OH)3 + 5H+ (5.7)

This reduction of Cr(VI) (chromate) is via a coupled, two-step, biotic-


abiotic reaction pathway in which Fe(II) produced during iron respiration
catalyzes the reduction of Cr(VI). Thus, attenuation of chromate in saturated,
subsurface environments may be in large part attributable to iron reduction.
In addition, the capacity for soils to reduce and immobilize Cr(VI) could be
dramatically underestimated if this biotic-abiotic process is not considered.
Wielinga et al. (2001) emphasize that the implications of these reactions is
importantthe primary terminal electron acceptor is continually regenerated.
Fe(II) produced in the first reaction listed (Equation 5.6) is cycled back to
Fe(III) in the second listed reaction (5.7) thereby acting as an electron shuttle
(a catalytic role) between the bacteria and chromium. Thus, a significant
amount of Cr(VI) could potentially be reduced even with little available Fe.
With the rapid cycling of Fe(II) back to Fe(III), evidence such as high pore
water Fe(II) concentrations in pore water could be hidden.

5.3.4.3 Sulfides
Sulfides (reduced sulfur) can be a major factor in the reduction of Cr(VI) to
Cr(III). Although most sulfides are not soluble, some dissolved sulfides can
be present in the environment due to processes including the discharge of
industrial wastes, the decomposition of organic matter, and sulfate reduction
(a common process in the biodegradation of chlorinated solvent chemicals).
Laboratory studies have reported that the reduction of Cr(VI) involving
sulfides is initially rapid, slows down in a few minutes, but reaches comple-
tion after one day (Schroeder and Lee, 1975). The rates of reduction of Cr(VI)
with H2S have been studied by Pettine et al. (1994) and Pettine et al. (1998).

5.3.4.4 Organic Matter


Organic matter is an important reductant in soils. Much of the organic matter
(i.e., organic carbon) in soil is present as humic and fulvic acids. Organic
matter, important in the reduction of Cr(VI), is also present as simple amino-
acids (Schroeder and Lee, 1975). The reduction of Cr(VI) by soil humic and
fulvic acids has been demonstrated by several researchers as referenced in
Palmer and Puls (1994). This reduction, with an intermediate Cr(V) species,
is favored by acidic conditions (Bloomfield and Pruden, 1980; Stollenwerk
and Grove, 1985; Cary et al., 1977; Grove and Ellis, 1980).
L1608_C05.fm Page 176 Friday, July 23, 2004 5:44 PM

176 Chromium(VI) Handbook

The rate of reduction of Cr(VI) decreases with increasing pH, increases


with the increasing initial Cr(VI) concentration, and increases as the concen-
tration of soil humic substance increases (Palmer and Puls, 1994). At a very
low pH, laboratory studies indicate that the half-life for Cr(VI) reduction
with humic acids is approximately three days, whereas several days are
required within the pH range of 4 to 7 (Eckert et al., 1990).

5.3.4.5 Cu(I)
Reduced copper [Cu(I)] may also play a role in the reduction of Cr(VI) to
Cr(III), discussed by Pettine et al. (2002) especially in atmospheric and sur-
face waters with low pH and low ionic strength (Abu-Saba et al., 2000). The
reduction of Cr(VI) with Cu(I) has been produced in the laboratory by
radiolysis experiments in dilute solutions in the presence of Cr(II).

5.3.4.6 Hydrogen Peroxide (H2O2)


The role of H2O2 in the reduction of Cr(VI) is discussed by Pettine et al.
(2002) who describe H2O2 as an oxidant of Cr(III) at pH > 7.5, a reductant at
lower pH, and its strength as a reductant being greatly increased at low pH.
In acid wastes receiving freshwater, and in atmospheric aqueous media with
pH ranging from about 1 to 5, the reduction of Cr(VI) with H2O2 is thermo-
dynamically possible (Seigneur and Constantinou, 1995), and has been used
in treatment processes for removing Cr(VI) from wastewaters (Eary and Rai,
1988). In the latter, the reduction of Cr(VI) with H2O2 includes a preliminary
conversion of Cr(VI) to Cr(III) and its subsequent precipitation.

5.4 Precipitation/Dissolution Reactions of Chromium


In addition to oxidation-reduction (redox) mechanisms for Cr(VI) reduction
as discussed above, chromium can undergo precipitation dissolution reac-
tions (Bodek et al., 1988), which are governed by the solubility of the chromium
compound and the kinetics of the dissolution. Kimbrough et al. (1999) list
and discuss these reactions.
Most Cr(III) species that are water-soluble do not occur naturally and are
unstable in the environment. The principle Cr(III) reaction in water is the
formation of chromium hydroxides of varying solubilities. The precipitation
of Cr(III) as the mixed iron-chromium hydroxide (Cr,Fe)(OH)3, discussed
previously, enhances the precipitation of Cr(III) in waters with neutral pH
levels. The kinetics of this reaction is rather rapid, making it an important
solubility controlling compound (Sass and Rai, 1987).
The Cr(VI) ions chromate (CrO42) and dichromate (Cr2O72) are water
soluble at all pHs. However, chromate can exist as the insoluble salt of a
variety of divalent cations, such as Ba2+, Sr2+, Pb2+, Zn2+, and Cu2+. The rates
L1608_C05.fm Page 177 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 177

of precipitation/dissolution reactions between chromate, dichromate anions,


and these cations vary greatly and are pH dependent. An understanding of
the dissolution reactions is particularly important for environmental assess-
ments because Cr(VI) often enters the environment by dissolution of such
chromate salts. Dissolution of somewhat soluble chromate salts (e.g., SrCrO4)
is particularly important because they provide a continual source of
chromate anions.

5.5 Sorption and Desorption Reactions of Chromium


5.5.1 General Discussion of Sorption
McLean and Bledsoe (1992) and Calder (1988) give general discussions of
sorption, which are summarized here. It is first important to define and
distinguish some important terms, which are used interchangeably, properly
and improperly, in the literature. The general term sorption actually com-
prises two processes: (1) adsorption, the process by which a solute clings to
a solid surface; and (2) absorption, the process by which the solute diffuses
into a porous solid and clings to interior surfaces.
Sorption is important in the transport and fate of a constituent. An equi-
librium distribution coefficient (Kd) is used in the estimation of the retarda-
tion of a constituents migration in groundwater. As with redox and
precipitation reactions, sorption reactions are highly influenced by the com-
plex environmental conditions inherent in the subsurface. Therefore, general
assumptions about sorption cannot be made. Such variables as pH, surface
area, density of active sites, among others, influence sorption equilibrium
(Kimbrough et al., 1999). Sorption studies also can are used to evaluate the
effect that changing a soil solution parameter, (e.g., adjustment of pH, ionic
strength, addition of competing cations, or addition of inorganic or organic
ligands) has on the retention of a constituent by the aquifer matrix.
Laboratory studies generate sorption isotherms, which describe equilib-
rium conditions of sorption. These isotherms are the relationship between
the amount of constituent sorbed and the equilibrium concentration of the
constituent. If the isotherm is linear, a single coefficient (Kd) can be defined
to describe sorption. For metals, such as chromium, the relationship is sel-
dom linear. Soil processes are never at equilibrium; soil systems are dynamic
and are thus constantly changing. Therefore, other equations with two or
more coefficients must be used. Nonlinear sorption behavior of metals in
soil are usually expressed by the Langmuir and the Freundlich equations,
even though sorption of metals by soils violates many of the assumptions
associated with these equations.
For nonlinear sorption, groundwater transportation equations must be
solved iteratively using a concentration-dependent retardation factor
because the retardation of the contaminant will vary with time due to
L1608_C05.fm Page 178 Friday, July 23, 2004 5:44 PM

178 Chromium(VI) Handbook

changing solution concentrations as the plume of contaminated groundwater


passes a particular portion of the aquifer segment. This makes comparisons
and predictions more difficult than for the linear adsorption model. Equilib-
rium studies predict whether a reaction will occur but give no indication of
the time necessary for the reaction to take place. Therefore, kinetic studies
have been performed to establish the proper time interval for use in equilib-
rium sorption/desorption studies. Many mathematical transport models
now allow a kinetic term for sorption.
Adsorption occurs because dissolved ionic species are attracted to mineral
surfaces that have a net electrical charge due to imperfections or substitutions
in the crystal lattice or chemical dissociation reactions at the particle surface.
This electrical charge varies with pH.
The importance of sorption to the transport and fate of constituents in
groundwater is that it retards the advance of the contaminant with respect
to the groundwater velocity, and can also reduce the contaminant concen-
tration. However, sorption is reversible, meaning that sorbed contaminants
can be released back into the aqueous medium, causing an increase in con-
centrations after periods of decreasing concentrations. Davis and Olsen
(1995) showed that in laboratory experiments with columns containing pre-
dominantly Cr(III) that were not augmented by additives, less than 2.5% of
the total Cr was leached, while in soil bearing primarily Cr(VI), over 80% of
total Cr was leached; Cr(VI) readily dissolved or desorbed from contami-
nated soils, while Cr(III) occurred in a predominantly nonleachable form.

5.5.2 Sorption of C4hromium


Several papers, notably Calder (1988); Richard and Bourg (1991); and
Davis and Olsen (1995) give detailed discussion of the sorption of the
different forms of chromium in the subsurface. Their work is discussed
below. Also, Calder (1988) gives examples of Cr(III) and Cr(VI) partition-
ing ratios (e.g., Kd) that can be used to describe the sorption of chromium
and describes the effect of different values pH and chromium concentra-
tions on sorption.

5.5.2.1 Sorption of Cr(III)


Cr(III) is rapidly, strongly and specifically sorbed in soil by Fe and Mn oxides,
clay minerals, and sand (Bartlett and Kimble, 1976; Schroeder and Lee, 1975;
Korte et al., 1976; Griffin et al., 1977; Rai et al., 1984; Dreiss, 1986). According
to experimental data, this sorption is rapid, with about 90% of chromium
being sorbed by clay minerals and iron oxides in 24 h. Furthermore, the
sorption of Cr(III) increases with increasing pH (Griffin et al., 1977; Rai et al.,
1984) (as the clay surfaces become more negatively charged) and increasing
organic matter content of soils (Paya Perez et al., 1988); whereas the adsorp-
tion of Cr(III) decreases when other inorganic cations or dissolved organic
ligands are present.
L1608_C05.fm Page 179 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 179

Partitioning ratios (Kd) for Cr(III) have been estimated to be very high.
Cr(III) sorption is nonlinear, however, so the Kd sorption model cannot be
used for assessing retardation except for a particular concentration. If these
partitioning ratios were equivalent to Kds, retardation factors of 500 to 6100
could be estimated, indicating the relative immobility of Cr(III) due to sorption
at a pH of around 4. Above this pH, Cr(III) would also be relatively immobile
because of its low solubility. It is probable that Cr(III) mobility would be
enhanced by the formation of complexes due to their decreased sorption or
increased solubility compared with the uncomplexed form of Cr(III).

5.5.2.2 Sorption of Cr(VI)


Chromate ions (the anionic forms HCrO4 and CrO42) can be sorbed by Mn,
Al, and Fe oxides and hydroxides (positively charged surfaces), clay minerals
and natural solids and colloids (Rai et al., 1986; James and Bartlett, 1983;
Stollenwerk and Grove, 1985; Rai et al., 1988; Griffin et al., 1977; MacNaugh-
ton, 1975; Davis and Leckie, 1980; Music et al., 1986; Zachara et al., 1987).
These substances commonly coat aquifer materials (Stumm and Morgan,
1981). Amorphous iron is the adsorbate found at the highest concentrations
in most aquifer materials. The batch experiments of James and Bartlett (1983)
confirm that iron hydroxides strongly sorb to Cr(VI). Batch sorption data
from experiments on Cr(VI) conducted by Davis and Olsen, 1995 conformed
best to a Freundlich isotherm, but Langmuirean behavior at a neutral pH,
and a decreasing Kd of Cr(VI) with increasing concentration, has been
reported (Griffin et al., 1977), probably due to competitive inhibition for
surface sites at higher Cr concentrations. Davis and Olsen, 1995 note that
their observed linear sorption is probably due to the lower range of Cr
concentrations used in their experiments.
This sorption is pH dependent (Richard and Bourg, 1991). At dilute con-
centrations, adsorption of Cr(VI) increases as pH decreases, no matter what
the sorbent (Rai et al., 1986; Bartlett and James, 1979; Rai et al., 1988; Griffin
et al., 1977; Rai et al., 1984; Davis and Leckie, 1980; Zachara et al., 1987). This
suggests that Cr(VI) sorption is favored on sorbents which are positively
charged at low to neutral pH. Interestingly, compared to clay, sandy material
has a greater preponderance of positively charged surfaces over the pH 57.5
range, resulting in a greater affinity for CrO4 and thus a higher Kd of Cr(VI)
on sand than for clay. Lower pH values result in higher Kd values, based on
published Cr(VI) sorption data for sandy soils compared with soils contain-
ing kaolinite and montmorillonite clays.
Sorption of Cr(VI) by clays, soils, and natural aquifer materials is low to
moderate in pH ranges common to groundwater. Sorption of Cr(VI) charac-
teristically decreases with increasing pH due to the decrease in positive
surface charge of the sorbing medium. Furthermore, Cr(VI) sorption has
been found to be nonlinear (Stollenwerk and Grove, 1985; Griffin et al., 1977),
fitting the Langmuir adsorption model. Similar to the discussion of Cr(III)
previously, if Cr(VI) adsorption were linear, the calculated partitioning ratios
L1608_C05.fm Page 180 Friday, July 23, 2004 5:44 PM

180 Chromium(VI) Handbook

would correspond to retardation factors of 2.5 to 329 (much lower than those
calculated for Cr(III)), indicating that Cr(VI) mobility at pH 7 could range
from high to lowlower below a pH of 7 and higher above a pH of 7 (and
above a pH of 8.5, Cr(VI) would be entirely unretarded).
Competing anions have a drastic effect on Cr(VI) sorption, with the effect
being variable, depending on dissolved concentrations of the competing
anion and CrO42, on their relative affinities for the solid surface, and on
surface site concentration (Rai et al., 1986). The competitive sorption of
Cr(VI) with cations and anions has been investigated by much research (Rai
et al., 1986; Stollenwerk and Grove, 1985; Rai et al., 1988; Rai et al., 1984;
Music et al., 1986; Zachara et al., 1987; Benjamin and Bloom, 1981). The
electrostatic sorption of anions is enhanced by cation sorption (due to
enhanced positive surface charge). Also, chromates either increase or have
no effect on the sorption of heavy metals (Cd2+, Co2+, Zn2+); competition for
surface sites is relatively minor (Benjamin and Bloom, 1981). Additionally,
sorption of chromates in the presence of a mixture of ions is lower than in
two-solution systems, particularly when H2SiO42 is present. The effect
appears to be additive (Rai et al., 1986; Zachara et al., 1987).
Richard and Bourg (1991) note that the kinetics of Cr(VI) sorption are not
well documented. The sorption of chromates on soils apparently follows a
two-step reaction rate (Amacher et al., 1988). Also, sorption of Cr(VI) does
not seem totally reversible. Amacher et al. (1986) attributed this lack of
reversibility to reduction of Cr(VI) to Cr(III), possibly by organic matter from
the soil they studied.

5.6 General Transport and Fate of Chromium


in Environmental Media
Kimbrough et al. (1999) discuss a generalized intermedia transport scheme
for environmental chromium. Chromium is directly emitted from industrial
activity either into the air, into water systems (e.g., streams, sewers, lakes,
etc.), or to the ground. Airborne chromium eventually settles out into soil
or water. In a given parcel of soil, there can be a mixture of Cr(VI) and Cr(III),
both naturally occurring and anthropogenic. Cr(VI), but not Cr(III), can be
leached out of the soil and enter groundwater, which in turn can become
part of an aquifer and also migrate to surface waters. As Cr(VI) is leached
from the soil, the remaining Cr(III) can slowly oxidize to Cr(VI) to reestablish
the equilibrium of the soil (Bartlett, 1991). In surface waters, Cr(VI) can
migrate in the dissolved form, while both Cr(III) and Cr(VI) can migrate
while being bound with dissolved organic carbon (DOC) or suspended
particles. And, chromium can migrate from the aqueous phase to sediments
from a dissolved state or with DOC or particles. In the sediment, dissolved
Cr(VI) can be immobilized if it enters a stable anaerobic portion, but Cr(VI)
in aerobic sediments can be redissolved.
L1608_C05.fm Page 181 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 181

The following are discussions of each of these environmental compart-


ments included in the general transport and fate of chromium in environ-
mental. Presented are discussions of chromium in the atmosphere, aquatic
environments (surface waters and groundwater), and soil, and the uptake
of chromium by biota.

5.6.1 Chromium in the Atmosphere


Kotas and Stasicka (2000) and Kimbrough et al. (1999) give detailed discus-
sions of the behavior of chromium in the atmosphere, and their work is
summarized here. The majority of chromium in the atmosphere (approxi-
mately 60 to 70%) is due to anthropogenic sources. Chromium from natural
sources accounts for the remaining amounts (Seigneur and Constantinou,
1995). Human activities that can product chromium in the atmosphere
include metallurgical industries, refractory brick production, electroplating,
combustion of fuels, the production of chromium chemicals (i.e., chromates
and dichromates, pigments, chromium trioxide, chromium salts), the cement
industry, production of phosphoric acid in thermal processes, and combus-
tion of refuse and sludges (Nriagu, 1988). Natural sources of chromium
include volcanic eruptions, erosion of soils and rocks, airborne sea salt par-
ticles, and smoke from forest wildfires (Pacyna and Nriagu, 1988). Average
atmospheric concentrations of chromium range from 1 ng/m3 in rural areas
to 10 ng/m3 in polluted urban areas (Nriagu, 1988).
The atmosphere has become a major pathway for long-range transfer of
chromium to different ecosystems (Nriagu, 1988; Spokes and Jickells, 1995).
Atmospheric chromium-containing particles are transported over varying
distances by the wind, before they fall or are washed out from the air onto
land and water surfaces, and the distance of transport depends on meteo-
rological factors, topography, and vegetation (Nriagu, 1988; Spokes and
Jickells, 1995). Wet precipitation and dry fallout of chromium from the
atmosphere is greatly affected by particle size; the chromium oxidation state
is less important.
The atmospheric transport and fate of chromium largely occurs in the
liquid phase and solids phases (i.e., droplets and particles) or, more generally,
aerosols instead of as a gas (Seigneur and Constantinou, 1995). The size of
the particles is important not only to the transport of chromium in the
atmosphere, but to health effects as well. Only particles with diameters less
than 10 m are respirable; their retention in the lungs can pose a carcinogenic
risk (Friess, 1989).
Chromium in aerosols is generally removed from the atmosphere by both
dry deposition and wet deposition. In dry deposition, the particles settle and
are captured by the soil or surface waters via gravitational sedimentation,
impaction, or interception. Wet deposition is the process where aerosol parti-
cles are actively entrained or scavenged by atmospheric moisture, such as rain,
snow, fog, or dew. Chromium can also be introduced, or reintroduced, into
the atmosphere via wind resuspension of chromium-containing soil particles.
L1608_C05.fm Page 182 Friday, July 23, 2004 5:44 PM

182 Chromium(VI) Handbook

The two stable oxidation states of chromium in the atmosphere are Cr(III)
and Cr(VI). Atmospheric particles do not contribute to the chemical reactions
that control the occurrence and ratio between Cr(III) and Cr(VI). Instead,
precipitation, complex formation, and oxidation reactions influence the
abundance and ratio of Cr(III) and Cr(VI).
Computer simulations by Seigneur and Constantinou (1995) have led to
the conclusion that typical atmospheric conditions favor the Cr(VI) reduction
to Cr(III). This is the likely case because of the presence and concentrations
of reducing agents in the air (i.e., V2+, Fe2+, H2S, HSO3, NO2, and organic
materials) as well as the acidity of the atmosphere. Cr(VI) can be reduced
rapidly in the atmosphere based on theoretical (Seigneur and Constantinou,
1995) and experimental (Grohse et al., 1988) studies. Estimates of atmo-
spheric half-life for Cr(VI) reduction to Cr(III) range from 16 h (Grohse et al.,
1988) to 4.8 days. The few materials capable of oxidizing Cr(III) to Cr(VI) in
the atmosphere, such as ozone, occur in concentrations too low to produce
measurable conversions in the atmosphere.

5.6.2 Chromium in Aquatic Environments


5.6.2.1 Surface Waters
Kotas and Stasicka (2000) and Kimbrough et al. (1999) give detailed discus-
sions of the behavior of chromium in surface aquatic environments, and their
work is summarized here. Chromium in natural waters originates from
natural sources or from manmade pollution. Natural sources include the
weathering of rock constituents, wet precipitation and dry fallout from the
atmosphere, and run-off from the terrestrial systems. Manmade pollution
sources to waters (mostly surface waters such as rivers) include the discharge
of industrial wastewaters (i.e., from the metallurgical, electroplating, tan-
ning, and dying industries), from sanitary landfill leaching, and from water
cooling towers (Nriagu, 1988). The number and type of chromium species
present in effluents depend on the character of the industrial processes using
chromium.
In natural waters, chromium exists in its two stable oxidation states, Cr(III)
and Cr(VI). The presence and ratio between these two forms depend on
various processes, which include chemical and photochemical redox trans-
formation, precipitation/dissolution reactions, and adsorption/desorption
reactions. Simplistically, Cr(III) should be the only form present in anaerobic
or subanaerobic conditions, whereas in aerobic aqueous environments,
Cr(VI) should be the only form present. However, the presence of Cr(III)
and Cr(VI) is also dependent on the pH of the water. Under neutral to
basic conditions, Cr(III) will tend to precipitate out, while under acid
conditions, Cr(III) will tend to solubilize. While Cr(VI) ions (i.e., chromate
and dichromate) are extremely water soluble at all pHs, they can precipitate
with a number of divalent cations. In waters of intermediate pH values, the
Cr(III)/Cr(VI) ratio is largely dependent on the concentration of oxygen.
L1608_C05.fm Page 183 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 183

In contrast to the atmosphere, many aqueous environments do contain


oxidizing agents, such as MnO2 and Mn3+ in sufficiently high concentrations
to produce measurable yields of Cr(VI). In oxygenated surface waters, not
only are pH and oxygen concentration important, but the nature and con-
centrations of reducing agents, oxidation mediators, and complexing agents
play important roles. These factors seem to be responsible for the occurrence
of significant Cr(III) quantities in many oxygenated surface waters (Kieber
and Heiz, 1992; Cranston and Murray, 1978 and 1980; Pettine et al., 1991).
At times, Cr(III) can be the predominant chromium component in oxygen-
ated waters (Chuecas and Riley, 1966). Several mechanisms for this might
include Cr(VI) reduction via Fe(II), Cr(VI) reduction via H2O2, and dissolved
organic matter. Other reducing agents might include hydrogen sulfide (H2S),
sulfur (S), NH4, nitrate (NO3), and vanadium (V2+) (Eary and Rai, 1988; Bodek
et al., 1988). The photochemical generation of Cr(III) has also been suggested
(Kieber and Heiz, 1992; Kaczynski and Kieber, 1993).
Both Cr(III) and Cr(VI) have been shown to bind with naturally occurring
dissolved organic carbon (DOC). Organically-bound Cr(III) can stay in solu-
tion at higher pH than unbound Cr(III) (Palmer and Wittbrodt, 1991) and
organically bound chromium can also sorb to and desorb from the organic
portion of suspended and settled sediments. Therefore, chromium migrates
as either dissolved ions or as attached to particles, or both.
The effect of sunlight can be important in surface water chemistry. The
oxidation and reduction of chromium are affected by sunlight (Kaczynski and
Kieber, 1994). Sunlight appears to degrade organically-bound chromium, and
this process releases inorganic chromium. Also, sunlight acts indirectly by
assisting the reduction of iron, which also results in the formation of H2O2
(Kieber and Heiz, 1992; Beaubien et al., 1994) affecting the oxidation state of
chromium. Additionally, sunlight aids the oxidation of manganese (Bartlett,
1991), also affecting the oxidation state of chromium.
Chromium transport and fate in surface waters can be discussed by using
three subsystems: rivers, lakes, and oceans. The transport pathways are
controlled by specific conditions prevailing in each of these subsystems,
including temperature, depth, degree of mixing, oxidation conditions, and
amount or organic matter. Chromium as a component of suspended particles
is the most important transport mechanism in rivers. Dissolved chromium in
river water decreases during its passage through turbid coastal environments.
Lakes generally have relatively high levels of biologic activity and high
ratios of sediment-to-water surface area, which greatly influence transport
of metals. The high level of organic matter creates the medium for reduction
and the formation of complexes, favoring the reduction of Cr(VI) to Cr(III),
which is afterwards rapidly precipitated or sorbed onto the sediment min-
erals. And, chromium in sediments can be remobilized into the surrounding
pore water via oxidation or solubilization of Cr(III) sediments. The most
complex transport pathways of chromium are in seasonally anaerobic lakes
(Beaubien et al., 1994; Achterberg et al., 1997) where deep basinal water in
the summer months becomes anaerobic due to the coupling of high biological
L1608_C05.fm Page 184 Friday, July 23, 2004 5:44 PM

184 Chromium(VI) Handbook

activity and thermal stratification. Therefore, depth and season heavily influ-
ence the concentration and speciation of chromium. Dissolved chromium
usually decreases in the summer months, and the areas dominated by Cr(VI)
versus Cr(III) become more segregated to surface and deep layers, respec-
tively. This distribution of chromium species is consistent with what would
be expected from seasonal increases in temperature, a decrease in pH, and
the oxygen content in the basinal water. The aerobic regime favors Cr(VI),
and Cr(III) is favored in anaerobic areas.
Chromium generally enters oceans via rivers and from atmospheric fall-
out. Atmospheric inputs result in more homogeneous distribution of chro-
mium in the ocean water compared with river inputs; the latter are the
subject of eustarine removal processes and ocean circulation patterns Spokes
and Jickells, 1995). It has been proposed that chromium sources to the oceans
are mostly as particles (suspended solids from river and aerosols). In ocean
waters, dissolved and precipitated chromium exist together in equilibrium.
Dissolved chromium is generally removed from the aqueous phase and
incorporated into biologic material (i.e., siliceous and carbonaceous skele-
tons) and by adsorption onto sediment particles. This removal occurs both
in the water column and at the sedimentwater interface, resulting in deep
and bottomwater enrichment of dissolved chromium.
Except in estuaries, chromium concentrations in seawater are dominated
by chromates, probably due to the generally oxidizing conditions in the ocean
and low suspended concentration of particles. Reduction of Cr(VI) occurs in
anaerobic basins and the oxygen-free zones, where increased Cr removal
may be due to Cr(III) adsorption onto bottom sediments (Smith et al., 1995).
Chromium cycling in the water column occurs in response to nutrient bio-
geochemistry. When Cr scavenged by particles is deposited on the ocean
floor, diagenetic processes can lead to remobilization of Cr either as chromate
or as organic Cr(III) complexes. The remobilization of Cr(III) from sediment
can also occur by its oxidation, carried out mostly by manganese dioxide.

5.6.2.2 Groundwater
Richard and Bourg (1991) and Calder (1988) give detailed discussions of the
behavior of chromium in groundwater environments, and their work is
summarized here. The mobility of chromium in groundwater depends on
its solubility and its tendency to be sorbed by soil or aquifer materials. These
factors, in turn, depend on the groundwater chemistry and the characteristics
of soil or aquifer material in contact with the chromium-containing ground-
water. Otherwise, much of the basic considerations discussed for surface
waters also apply for the understanding of the transport and fate of chro-
mium in groundwater.
Large plumes of chromium-contaminated groundwater in shallow aqui-
fers have been well documented (Stollenwerk and Grove, 1985; Deutsch,
1972; French et al., 1985; Perlmutter and Lieber, 1970; Wiley, 1983). Sources
of this contamination can be the same as those listed for surface waters. Such
L1608_C05.fm Page 185 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 185

plumes in sand and gravel aquifers have been reported to reach lengths of
up to 1300 m (Perlmutter and Lieber, 1970). Impacts of many of the plumes
have been serious enough to necessitate the abandonment of local ground-
water supplies.
Groundwater contamination by chromium can be extensive in permeable
aquifers (i.e., sand and gravel and fractured rock aquifers) because ground-
water velocities in these materials are relatively high (i.e., about 0.1 and 5 m
per day, respectively). On the other hand, groundwater velocities in aquifers
of much lower permeability (i.e., clayey materials) tend to be low, perhaps
on the order of a few centimeters or less per year. Thus, chromium-contam-
inated groundwater in these settings cannot extend far from the source of
the chromium.
Cr(III) tends to be relatively immobile in most groundwater because of the
precipitation of Cr(III) compounds of low-solubility (e.g., Cr(OH)3(s),
FeCr2O4(s), (Fe1x, Crx)(OH)3(ss)) in neutral to alkaline pH range (i.e., above
pH 4). This results in low Cr(III) dissolved concentrations. Also, in neutral
to slightly acidic conditions (i.e., especially below pH 4), Cr(III) is removed
from solution by sorption. Calder (1988) reported that sorption of Cr(III)
increases with increasing pH. Furthermore, it has been speculated that Cr(III)
may be mobile in groundwater if it is in a complexed form, although this
has not been documented in the field. Precipitation and sorption can be
inhibited by complex formation with dissolved ligands such as natural
organic matter (Gerritsee et al., 1982).
Cr(VI) tends to be moderately to highly mobile in most shallow ground-
water aquifers. This tends to be due to two major factors: (1) the lack of
solubility constraints; and, (2) the low to moderate sorption of Cr(VI)
anionic form in neutral to alkaline waters. In soils or sediments with high
content of Fe and Mn oxides conditions, Cr(VI) should be removed by
sorption processes (Rai et al., 1986; Eary and Rai, 1987; Stollenwerk and
Grove, 1985; Rai et al., 1988; Cary et al., 1977). But sorption is significantly
depressed by competing background anions (Rai et al., 1988) so that Cr(VI)
can expected to be highly mobile. In alkaline environments, sorption is not
strong enough to keep Cr(VI) from migrating through soil or sediments.
Cr(VI) sorption generally increases with decreasing pH, so sorption of
Cr(VI) can be very significant in neutral to acidic groundwater. As previ-
ously discussed, Cr(VI) sorption can be strongly nonlinear, such that sorp-
tion decreases with increasing Cr(VI) concentration. Also, sorption also
appears to be rate-dependent, so the kinetics of the sorption process would
very important, especially in high-velocity groundwater regimes. Also,
Cr(VI) reduced to Cr(III) with subsequent precipitation and sorption is
believed to control the mobility of Cr(VI) (Rai et al., 1988): Cr(VI) reduction
to Cr(III), which is afterwards rapidly precipitated or sorbed. In Fe(II)-rich
and dissolved organic matter-rich environments, the reduction of Cr(VI) is
more likely to occur and the resulting aqueous Cr(III) concentration will be
controlled by the solubility of Cr(III). In such cases, Cr(VI) should not migrate
significantly.
L1608_C05.fm Page 186 Friday, July 23, 2004 5:44 PM

186 Chromium(VI) Handbook

The impact of water chemistry on the presence and movement of chro-


mium in groundwater was demonstrated in field experiments involving
injections of 100 mol/l of Cr(VI) into various zones of a gravel aquifer
(Kent et al., 1989). Some chromium disappeared from the aqueous phase in
the anaerobic part of the aquifer due to reduction to the less soluble Cr(III)
form. Cr(VI) (chromate) generally migrated at about the same rate as
groundwater flow, except in areas with low pH and low concentrations of
anions, where it was retarded due to competition with these forms for
sorption sites.
The relationship of chromium and manganese is particularly interesting
in the transport and fate of chromium in groundwater. Cr and Mn form a
pair of chemical elements with contrasting tendencies (Murray et al., 1983).
Under oxidizing conditions, Cr(VI) is soluble as CrO42 while Mn(IV) is
scavenged as MnO2(s). Under reducing conditions, Cr(III) is removed from
solution as Cr(OH)3(s) while Mn(II) is soluble as Mn2+. These contrasting
tendencies for the solubility of Cr and Mn have been observed in shallow
groundwater of the Western San Joaquin Valley in California (Deverel and
Millard, 1988). In the alluvial-fan geologic zone, dissolved Cr concentrations
are high, whereas dissolved Mn is low. However, in the basin-trough zone,
Cr concentration is low and Mn concentration is high.

5.6.2.3 Plumes of Chromium in Groundwater Case Studies


There are a number of documented cases of Cr plumes in groundwater. The
following are accounts of major occurrences of Cr in groundwater as sum-
marized by Calder (1988) and Loyaux-Lawniczak et al. (2001).

5.6.2.3.1 Nassau County, New York


The best known and the first major published case study in North America
of Cr in water supply wells was in Nassau County, Long Island (Lieber et al.,
1964). A number of investigators have studied the site, and this case dem-
onstrates the types of uncertainties that complicate predictions of Cr migra-
tion in groundwater. The source of Cr was an aircraft plant that used Cr
solutions for anodizing and plating metals. The site is located on a very
permeable sand and gravel aquifer with groundwater velocities estimated
to be approximately 0.15 to 0.5 m per day (Perlmutter and Lieber, 1970). The
estimated length and width of the plume was 1300 m by 300 m, with a
maximum Cr concentration in groundwater of 40 mg/l. The groundwater
pH was 4.6 to 6.2 which is the range where Cr(VI) could be significantly
sorbed by the aquifer materials. Calder (1988) promotes three hypotheses
that can account for the apparent retardation of the Cr plume:

1. Greater sorption as a result of the lowering of the chromium con-


centration due to remediation efforts (i.e., concentration-dependence
of chromium sorption).
L1608_C05.fm Page 187 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 187

2. The slow reduction of Cr(VI) to Cr(III), with subsequent precipita-


tion of Cr(III), particularly in deeper groundwater.
3. Slow kinetics of the sorption process in the more permeable por-
tions of the aquifer, such that chemical equilibrium, and therefore
maximum sorption, would not occur until the aquifer had been
exposed to chromium-containing groundwater over a long period
of time.

5.6.2.3.2 Telluride, Colorado


This is the site of a heavy metal mining and milling operation near Tellu-
ride, Colorado (Grove and Stollenwerk, 1985). The source of chromium
was a tailings pond, which, since 1977, apparently discharged chromium-
containing wastes into the groundwater system. Water in the tailing pond
had chromium concentrations of 8.8 mg/l (Stollenwerk and Grove, 1985).
The USGS initiated a study of the site in October 1978, providing an
excellent opportunity for comparison of field observations with laboratory
experiments and computer simulation (Stollenwerk and Grove, 1985). The
shallow aquifer was gravel and sand alluvium, and the estimated ground-
water velocities were approximately 5 m per day. In 1979, a chromium
plume at least 520 m long was observed, with a maximum chromium
concentration of 2.7 mg/l (Grove and Stollenwerk, 1985). The groundwater
pH was approximately 6.8 (Stollenwerk and Grove, 1985). Laboratory and
field investigations determined that the chromium was retarded by a factor
of 10 relative to the groundwater velocity. Even with such an appreciable
retardation rate, the high groundwater velocity resulted in a relatively
mobile chromium plume.

5.6.2.3.3 Southwestern MichiganWood Treatment Plant


An incidence of groundwater contamination by chromium from a wood
treatment plant was reported in southwest Michigan in 1980 (French et al.,
1985). The wood was treated with a 2% aqueous solution containing 47.5%
chromium trioxide, 34% arsenic pentoxide, and 18.5% copper oxide. Effluent
from a sump pit was discharged to the ground adjacent to the treatment
building until 1980. An effluent sample from the pit contained 1600 mg/l
chromium, of which 1500 mg/l was Cr(VI). The site is located on a permeable
outwash plain consisting of gravelly sands with up to 17% silt and clay-sized
particles. The water table was at a depth of about 8 m. The groundwater
had a moderately alkaline pH of up to 8.4, and groundwater velocities were
estimated at approximately 0.15 m per day. The highest chromium concen-
tration in the plume was 6.58 mg/l, and chromium was detected in the
facilitys supply well as high as 2.5 mg/l. The chromium plume, defined by
total chromium concentrations above 50 parts per billion, extended approx-
imately 600 m from the discharge area, with a width of approximately 200 m
and a vertical thickness of 20 m. The length of the plume was found to be
consistent with estimated groundwater velocities, suggesting that Cr was
L1608_C05.fm Page 188 Friday, July 23, 2004 5:44 PM

188 Chromium(VI) Handbook

essentially unretarded. It was assumed that the chromium was almost


entirely Cr(VI). A purge well-spray irrigation system was established to
restore the aquifer to drinking water standards.

5.6.2.3.4 Industrial Waste Landfill, Northern France


The site, an industrial waste landfill located in northern France, was
operational from 1905 to 1982, producing materials including dichromates,
chromic acid, sulfuric acid, and phosphates. Chromite and pyrite were
the main primary minerals used at the facility, and Cr mineral processing
wastes were collected in a slag heap, which was covered in 1990 by a
geomembrane to limit runoff. The groundwater table is normally located
at 2 m depth, with an annual fluctuation of 1 m, within the infill. The
water table aquifer, in a silt layer, is well separated from deeper aquifers
by a green clay unit. The hydraulic conductivity was estimated to be about
7 107 m/s. The chromium concentrations in the source area were about
210 mg/l, with the migrating plume extending approximately 160 m
downgradient. The plume does not extend to the downgradient boundary
of the site 375 m away. The Fe(II) distribution in groundwater is quite
variable. It is virtually absent in the major area of the Cr plume, then the
Fe(II) concentration abruptly increases near the downgradient end of the
Cr plume (1680 mg/l). Concentrations of total copper, cadmium, zinc, and
sulfate ions show a similar distribution to Fe(II). The pH values in the
groundwater are mostly neutral (6.5 to 7.3) in the major portion of the
plume, becoming more acidic (approximately 4) at the downgradient por-
tion of the Cr plume, possibly due to the oxidation of pyrite that was used
in massive amounts in this area of the site. Loyaux-Lawniczak et al. (2001)
explain the distribution of metals in groundwater by postulating that
Cr(VI) (produced by leaching of the ore residue slag heap) migrates in
groundwater flow and then enters into a reducing zone (with Fe(II)
present), where the Cr(VI) is reduced by Fe(II). It has been widely accepted
that the kinetics of this reaction in solution is fast, and that with excess
Fe(II), all of the Cr(VI) is reduced. Cr(III) is especially immobilized in the
clay fraction of the soil; analyses of this clay fraction revealed that mont-
morillonite flakes are the chromium-bearing mineralogical phase. In sum-
mary, Cr(VI) migration in groundwater is retarded horizontally by a redox
mechanism involving chromate ions and ferrous ions or Fe(II)-bearing
minerals, and vertically by a thick green clay unit.

5.6.3 Chromium in Soil


The discussion of chromium behavior in soil presented here is preceeded
by an overview of the presence and behavior of metals in soil. The discus-
sion specific to the behavior of chromium in soil includes a general over-
view followed by discussions of the sorption, oxidation, and reduction of
chromium.
L1608_C05.fm Page 189 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 189

5.6.3.1 Overview of Metals in Soil


McLean and Bledsoe (1992) present an overview of metals in soil, and their
work is discussed below. Metals are found in soil within one or more soil
pools:

1. Dissolved in the soil solution


2. Occupying exchange sites on inorganic soil constituents
3. Specifically sorbed on inorganic soil constituents
4. Associated with insoluble soil organic matter
5. Precipitated as pure or mixed solids
6. Present in the structure of secondary minerals
7. Present in the structure of primary minerals

Metals exist in soil solution as either free (uncomplexed) metal ions (e.g.,
Cr3+), or in soluble complexes with inorganic or organic ligands, or associated
with mobile inorganic and organic colloidal material. A complex is a molec-
ular unit where a central metal ion is bonded by a number of associated
atoms or molecules ( e.g., Cr(OH)4), and these associated atoms or molecules
are termed ligands (i.e., OH is a ligand). Metals will form soluble complexes
with inorganic ligands such as SO42, Cl, OH, PO43, NO3, and CO32.
Soluble complexes with organic ligands are not as well defined. The free
metal ion is generally the most bioavailable and toxic form of the metal.
With complex formation, the resulting metal species may be positively or
negatively charged or be electrically neutral. The presence of a complex
species of a metal in the soil solution can significantly affect the migration
of metals through the soil relative to the free metal ion. In addition to
dissolved metal complexes, metals also may associate with mobile colloidal
particles (size of 0.01 to 10 m). Colloidal particles include iron and manga-
nese oxides, clay minerals, and organic matter. These surfaces have a high
capacity for metal sorption.
The extent of migration of metals from the ground surface into and through
the subsurface depends on the retention capacity of the soil is exceeded, and
it is directly related to the solution and surface chemistry of the soil and to
the specific properties of the metal and associated waste matrix. The mech-
anisms for retention of metals in soil include sorption and precipitation.
Retention of cationic metals is related to soil properties such as pH, redox
potential, surface area, cation exchange capacity, organic matter content, clay
content, iron and manganese oxide content, and carbonate content. Anionic
metal retention has been correlated with pH, iron and manganese oxide
content, and redox potential. Consideration must also be given to the type
of metal, its concentration, the presence of competing ions and complexing
ligands, and the pH and redox potential of the soil-waste matrix. Also, the
migration of metals can depend on the type of wastes that may be associated
with the metal. Therefore, because of the differing varieties of soils, and the
L1608_C05.fm Page 190 Friday, July 23, 2004 5:44 PM

190 Chromium(VI) Handbook

many different forms of metals themselves and the wastes containing them,
evaluating the extent of metal retention by a soil is site specific, soil specific,
and waste specific.
Precipitation and sorption are the main metal retention mechanisms in
soil. Precipitation is where metals precipitate to form a solid (three-dimen-
sional) phase in soils, which might be a pure solid or a mixed solid (e.g.,
(FexCr1x)(OH)3); the latter forms when various elements co-precipitate. Sorp-
tion of metals is the accumulation of ions at the solid phaseaqueous phase
interface. Sorption of metals in the soil matrix often involves organic matter,
clay minerals, iron and manganese oxides and hydroxides, carbonates, and
amorphous aluminosilicates. Binding of metals to organic matter ranges
from weak forces of attraction to formation of strong chemical bonds. Soil
organic matter can be the main source of soil cation exchange capacity.
Organic matter content generally decreases with depth in soil, so that the
mineral (inorganic) constituents of soil will become a more important surface
for sorption with increasing depth. There have been numerous studies of
the adsorptive properties of clay minerals, in particular montmorillonite and
kaolinite, and iron and manganese oxides. Griffin and Shimp (1978) found
the relative mobility of 9 metals through montmorillonite and kaolinite to
be: Cr(VI) > Se > As(III) > As(V) > Cd > Zn > Pb > Cu > Cr(III). Also, Jenne
(1968) concluded that Fe and Mn oxides are the principal soil surface that
control the mobility of metals in soils and water. In arid soils, carbonate
minerals may immobilize metals by providing sorbing and nucleating sur-
face (Santillan-Medrano and Jurinak, 1975; Cavallaro and McBride, 1978;
McBride, 1980; Jurinak and Bauer, 1956; McBride and Bouldin, 1984; Dudley
et al., 1988, 1991).
Generally, the sorption capacity for anions (some metals form anionic
contaminants) is lower than the cation sorption capacity of soils. The sorption
capacity (both exchange and specific sorption) of a soil is determined by the
number and kind of sites available. Sorption process are affected by various
soil factors (i.e., pH, redox potential, clay, soil organic matter, oxides, and
calcium carbonate content), by the form of the metal added to the soil, and
by the solvent introduced along with the metal. Therefore, interactions of
these influences on sorption may increase or decrease the migration of metals
in soil.
Although the principles affecting sorption and precipitation are similar for
cationic and anionic metals, the following is a list with a brief description
of factors affecting the behavior of cationic metals in soils.
Competing cations: Trace cationic and anionic metals are preferentially
sorbed over the major cations (Na+, Ca2+, Mg2+) and major anions (SO42,
NO3). However, when the specific sorption sites become saturated, exchange
reactions dominate and competition for these sites with soil major ions
becomes important.
Complex formation: Metal cations form complexes with inorganic and
organic ligands, whereby the ligand forms a coordinate bond with the metal
atom. The resulting association has a lower positive charge than the free
L1608_C05.fm Page 191 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 191

metal ion (and might even be uncharged or negatively charged). The effect
of complex formation on sorption is dependent on the type and amount of
metal present, and type and amount of ligands present, soil surface proper-
ties, soil solution composition, pH, and redox. The presence of complexing
ligands may either increase metal retention or greatly increase metal mobil-
ity. Data from the literature that do not consider the presence of complexing
ligands at the site, both organic and inorganic, may lead to significant error
in estimating metal mobility.
pH: The pH of the soil system is a very important parameter, directly
influencing sorption/desorption, precipitation/dissolution, complex forma-
tion, and oxidationreduction reactions. The pH affects several mechanisms
of metal retention of soils both directly and indirectly. Sorption increases
with pH for all cationic metals, but retention does not significantly increase
until the pH gets above 7. As true for all oxyanions (i.e., Cr(VI)), sorption
decreases with pH. The pH dependence of sorption reactions of cationic
metals is due in part to the preferential sorption of the hydrolyzed metal
species in comparison to the free metal ion. Many sorption sites in soils are
pH dependent (i.e., Fe and Mn oxides, organic matter, carbonates, and the
edges of clay minerals). As the pH decreases, the number of negative sites
for cation sorption diminishes while the number of sites for anion sorption
increases. Also, as the pH becomes more acidic, metal cations also face com-
petition for available permanent charged sites by Al3+ and H+. Jenne (1968)
stated that hydrous oxides of Fe and Mn play a principle role in the retention
of metals in soils. Solubility of Fe and Mn oxides is also pH-related. Below
pH 6, the oxides of Fe and Mn dissolve, releasing sorbed metal ions to
solution (Essen and El Bassam, 1981). In soils with significant levels of
dissolved organic matter, increasing soil pH may actually mobilize metal
due to complex formation. A word of caution is warranted here, however.
The complexity of the soil-waste system (several types of surfaces and solu-
tion compositions) may render generalizations just given to be not true. For
example, cationic metal mobility can actually increase with increasing pH
due to the formation of metal complexes with dissolved organic matter.
Oxidation-Reduction: Many metals have more than one oxidation state, and
are directly affected by changes in the oxidation-reduction (redox) potential
of the soil. Redox reactions can greatly affect contaminant transport, in
slightly acidic to alkaline environments. In general, oxidizing conditions
favor retention of metals in soils, while reducing conditions contribute to
accelerated migration.

5.6.3.2 Behavior of Chromium in Soil


Kimbrough et al. (1999); Kotas and Stasicka (2000); and McLean and Bledsoe
(1992) present overviews of the behavior of chromium in soil, and their work
is summarized here. More details concerning the chemistry of chromium in
soils and sediments are provided in review articles by Cary (1982); and
Richard and Bourg (1991).
L1608_C05.fm Page 192 Friday, July 23, 2004 5:44 PM

192 Chromium(VI) Handbook

Chromium commonly exists in two oxidation states in soils, Cr(III) and


Cr(VI). Forms of Cr(VI) in soils are as hydrogen, chromate ion (HCrO4
predominant at pH < 6.5 or CrO42 predominant at pH 6.5), and as dichro-
mate (Cr2O72) predominant at higher concentrations and at pH 2 to 6. The
dichromate ion is more toxic to humans than chromate ion. In neutral-to-
alkaline soils, Cr(VI) is mostly soluble (e.g., Na2CrO4) but also in moderately-
to-sparingly soluble chromates (e.g., CaCrO4, BaCrO4, PbCrO4) (Bartlett and
Kimble, 1976; James, 1996). In more acidic soils (pH < 6), HCrO4 becomes
a dominant form. The chromate ions (CrO42 and HCrO4) are the most
mobile forms of Cr in soils. They can be taken up by plants and easily leached
out into the deeper soil layers causing groundwater and surface water pol-
lution (Calder, 1988; James and Bartlett, 1984; Handa, 1988). Some minor
amounts of Cr(VI) are bound in soils. This binding depends on the miner-
alogical composition and pH of the soil. The CrO42 ion can be sorbed by
goethite, FeO(OH), aluminum oxides and other soil colloids with a posi-
tively charged surface (Richard and Bourg, 1991; James and Bartlett, 1983
and 1988). The HCrO4 ion, which occurs in more acidic soils, may also be
held in soils, or remain soluble (James and Bartlett, 1983). Reviews of the
processes that control the fate of chromium in soil and the effect these
processes have on remediation are given in Bartlett (1991) and Palmer and
Wittbrodt (1991).
In a study of the relative mobilities of 11 different trace metals for a wide
range of soils, Korte et al. (1976) found that clayey soil, containing free iron
and manganese oxides, significantly retarded Cr(VI) migration. Cr(VI) was
the only metal that was highly mobile in alkaline soils. The study also
showed that free iron oxides, total manganese, and soil pH influenced
Cr(VI) immobilization, whereas soil properties such as cation exchange
capacity, surface area, and percent clay had no significant influence on
Cr(VI) mobility.
Chromium in soils is naturally present mostly as insoluble Cr(OH)3 or as
Cr(III) sorbed to soil components, which prevents leaching into the ground-
water or its uptake by plants (Bartlett and Kimble, 1976). The dominant
chromium form depends strongly on pH; in acidic soils (pH < 4) it is
Cr(H2O)63+, whereas at pH < 5.5 it is its hydrolysis products, mainly soluble
CrOH2+ (Ritchie and Sposito, 1995); both these forms are easily sorbed by
clays. The process is intensified by an increase in pH, which can be inter-
preted in part due to an increase of negative charge on the clays. Here humic
acids contain donor groups forming stable Cr(III) complexes. The Cr(III)
sorption to humic acids renders it insoluble, immobile, and unreactive; this
process is the most effective within the pH range of 2.7 to 4.5 (James, 1996).
In contrast, mobile compounds such as citric and fulvic acid, form soluble
Cr(III) complexes which control its oxidation to Cr(VI) in soils (Bartlett and
Kimble, 1976; Bartlett and James, 1979; James and Bartlett, 1983; James, 1996;
Wittbrodt and Palmer, 1995).
Both reduction of Cr(VI) to Cr(III) and the sorption of Cr(VI) can occur in
soil, even simultaneously. Therefore, this causes a problem in assigning one
L1608_C05.fm Page 193 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 193

mechanism of observed attenuation of Cr(VI) in the subsurface (Bartlett,


1991). The following are discussions of each of these mechanisms.

5.6.3.2.1 Sorption of Cr(III) and Cr(VI)


Cr(VI) is an anion, and its association with soil surfaces is thus limited to
positively charged exchange sites, which decrease in number with increasing
soil pH. Therefore, sorption of Cr(VI) decreases with increasing soil pH. Iron
and aluminum oxide surfaces will adsorb chromate ions (CrO42) at acidic
and neutral pH (Davis and Leckie, 1980; Zachara et al., 1987; Ainsworth
et al., 1989). The sorption of Cr(VI) in groundwater by alluvium aquifer
materials has been found to be due to the iron oxides and hydroxides coating
the alluvial particles (Stollenwerk and Grove, 1985). But the sorbed Cr(VI)
was desorbed by adding uncontaminated groundwater, indicating nonspe-
cific sorption of Cr(VI). The presence of chloride and NO3 have little effect
on Cr(VI) sorption, whereas sulfate and phosphate tend to inhibit sorption
(Stollenwerk and Grove, 1985).
Chromates can be sorbed by iron, aluminum oxides, amorphous alumi-
num, hydroxides, organic complexes, and other soil components, which may
protect Cr(VI) from reduction (Bartlett, 1991; Bartlett and James, 1988). Also,
Cr(III) materials can be sorbed onto silicacious components. In the aqueous
phase of soils, Cr(III) that is not sorbed by the solid phase would generally
hydrolyze to the hydroxide and precipitate. Chromate would be far less
likely to precipitate and so would be expected to be more mobile. In this
situation precipitation reactions are closely tied to oxidation and reduction
reactions. In anaerobic sediments, oxidation is unlikely to take place and
chromium hydroxide [Cr(OH)3] could be immobilized as long as the sedi-
ments are physically stable (Eary and Rai, 1989).
Soil pH determines both the speciation of Cr(VI) and the charge charac-
teristic of the surface with which it reacts (James and Bartlett, 1983). Above
pH 6.4, HCrO4 dissociates to CrO42 as the dominant form of Cr(VI) and the
charge characteristic of the surface with which it reacts (James and Bartlett,
1983), and the CrO42 may in turn be sorbed. Sorption of chromates can be
a reversible process suggested by leaching of Cr(VI) from soils (Baron et al.,
1996). However, such reversibility depends on the chemistry of the leachate
and of the soil or sediment.

5.6.3.2.2 Reduction of Cr(VI) and Oxidation of Cr(III)


Oxidation and reduction reactions can convert Cr(III) to Cr(VI) and vice
versa (Bartlett and Kimble, 1976; Bartlett and James, 1979; James and Bartlett,
1983; James and Bartlett, 1988; Wittbrodt and Palmer, 1995; James, 1994;
Powell et al., 1995; Deng and Stone, 1996). These processes depend on pH,
oxygen concentration, presence of appropriate reducers and mediators act-
ing as ligands or catalysts. Mobile forms for Cr(VI) (HCrO4 and CrO42) can
be reduced by different reducers such as Fe(II) or S2.
L1608_C05.fm Page 194 Friday, July 23, 2004 5:44 PM

194 Chromium(VI) Handbook

It has been thought that Cr(VI) can be reduced to Cr(III) under normal
soil pH and redox conditions. However, Bloomfield and Pruden (1980) rein-
vestigated earlier claims that Cr(VI) is readily reduced to Cr(III) under such
normal soil conditions, and they found that the analytical methods used in
previous investigations (Bartlett and Kimble, 1976) were unreliable because
the soil extracts probably contained organic matter capable of reducing
Cr(VI). They also found that the reduction of Cr(VI) in soil of normal pH
was not particularly rapid under aerobic conditions.
The reduction of chromates by Fe, V2+, sulfides, and organic materials is
well demonstrated (Cary, 1982), and the kinetics of Cr(VI) reduction has
been reported to follow a simple first-order reaction kinetics (Amacher and
Baker, 1982; Bartlett and James, 1988). Soil organic matter has been identified
as the important electron donor (i.e., the principal reducing agent) in this
reaction (Bartlett and Kimble, 1976; Bloomfield and Pruden, 1980), and the
reduction of Cr(VI) in the presence of organic matter proceeds at a slow rate
at normal levels of pH and temperatures found in the environment (Bartlett
and Kimble, 1976; James and Bartlett, 1983). This relatively slow rate of
Cr(VI) reduction increases with decreasing soil pH (Bloomfield and Pruden,
1980; Cary et al., 1977). In the absence of soil organic matter, Fe(II)-containing
minerals reduce Cr(VI), however, tests have shown that this reduction occurs
in subsurface soil with a low pH (<5) (Eary and Rai, 1991). Reduction by
Fe(II) is more favorable under anaerobic conditions since oxygen can oxidize
the iron (II) (Fendorf and Li, 1996). However, high concentrations of Cr(VI)
may quickly exhaust the available reducing capacity of the soil, and excess
Cr(VI) may persist for years in soils (Baron et al., 1996). In general, it has
been noted that chromates are relatively stable and mobile in soils that are
sandy or have low organic content (Cary, 1982; Bloomfield and Pruden, 1980;
Frissel et al., 1975).
Under conditions prevalent in some soils, Cr(III) can be oxidized (Bartlett
and James, 1979). Only a few oxidants present in soils and sediments (i.e.,
dissolved oxygen and MnO2) are capable of oxidizing Cr(III) to Cr(IV). The
oxidation of Cr(III) by MnO2 (which serves as an electron acceptor) has been
shown to occur in soils (Eary and Rai, 1987; Johnson and Xyla, 1991; Fendorf
and Zasoski, 1992), and aerobic sediments, but not in anaerobic sediments.
Oxidation of Cr(III) by dissolved oxygen has been found to be insignificant
(Rai et al., 1989) when compared with MnO2, which is the most likely oxidant
of Cr(III) in soils. Thus, if soluble Cr(III) is added to an average soil, a
portion of the soluble Cr(III) will become immediately oxidized by MnO2 to
Cr(VI) (Cary, 1982). The rest of the Cr(III) may remain reduced for long
periods of time, even in the presence of electron-accepting manganese
oxides, perhaps because soluble Cr(III) can form complexes with low-molec-
ular mass organic molecules and then be oxidized where redox conditions
are optimal. Added organic matter also may facilitate oxidation of Cr(III) to
Cr(VI). This has implications to remediation strategies. The addition of
organic residues potentially as a remediation strategy for Cr(VI)-contami-
nated soils containing high levels of oxidized manganese may result in the
L1608_C05.fm Page 195 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 195

formulation of unstable Mn(III) organic complexes that not only temporarily


prevent Cr(III) oxidation but also promote the desired reduction of Cr(VI)
(Bartlett and James, 1988).

5.6.4 The Uptake and Transformation of Chromium by Biota


Kotas and Stasicka (2000) and Kimbrough et al. (1999) present overviews of
the uptake and transformation of chromium by biota, and their work is
summarized here. Chromium can be taken up by biota from the air, water,
and soil. Most studies in this realm do not distinguish between the oxidation
states of chromium. However, it is known that the Cr(VI) form is more
available for living organisms than Cr(III), and the uptake of Cr(VI) by biota
is a main role in the removal of Cr(V) from water and soil systems. The
following discussion of the role of biota in the transport and fate of chromium
is presented for bacteria, plants, aquatic animals, and terrestrial animals.
Microorganisms accumulate chromium (Coleman, 1988) and reduce Cr(VI)
to Cr(III) (Campos et al., 1995; DeLeo and Ehrlich, 1994). Although high
levels of Cr(VI) are toxic to microorganisms (Bartlett, 1991), chromium is
important to yeast metabolism (Coleman, 1988; Anderson et al., 1977). How-
ever, there is not much evidence in the literature of bioaccumulation of
chromium as Cr(VI) in bacteria, since most studies report chromium bioac-
cumulation in terms of total chromium. There are conflicting views concern-
ing the uptake and translocation of Cr(VI) in plants. Also, whether chromium
is an essential element in plants has been debated in the literature. The World
Health Organization says it is unknown whether chromium is an essential
nutrient for all plants, yet all plants contain the element. On the other hand,
Richard and Bourg (1991) suggested that Cr(III) is an essential nutrient in
plant metabolism (amino and nucleic acid synthesis). The literature on chro-
mium bioaccumulation in aquatic animals (e.g., finned fish) suggests that
Cr(VI) is not expected to accumulate and increase in the aquatic food chain.
And, there is no indication of biomagnification of chromium within the
terrestrial animal food chain (soil-plant-animal) (Clay, 1992; ATSDR, 1992).

5.7 Utilizing Natural Environmental Processes as a Remedy


for Soil and Groundwater Contaminated with Chromium
Natural attenuation is a term that describes the naturally-occurring environ-
mental processes that act without human intervention to reduce the mass,
toxicity, mobility, volume, or concentration of contaminants. These processes
can be grouped into two classes, destructive and nondestructive processes.
Destructive processes include biotransformation and abiotic chemical reac-
tions. Nondestructive processes include sorption (sorption and absorption),
dispersion, dilution from recharge, and volatilization. Natural attenuation
L1608_C05.fm Page 196 Friday, July 23, 2004 5:44 PM

196 Chromium(VI) Handbook

is sometimes referred to by several other names, such as intrinsic remedia-


tion, intrinsic bioremediation, natural restoration, or passive bioremediation.
For the purposes of this chapter, the term natural attenuation will be used,
because some of the synonyms used such as intrinsic bioremediation actually
refer to only one of many processes responsible for natural attenuation.
The implementation of natural attenuation processes as a remedy for soil
and groundwater contamination is termed Monitored Natural Attenuation
(MNA). The United States Environmental Protection Agency (USEPA) has
stated its position on the use of MNA for the remediation of contaminated
soil and groundwater in their Final OSWER Directive Use of Monitored
Natural Attenuation at Superfund, RCRA Corrective Action, and Under-
ground Storage Tank Sites (OSWER Directive Number 9200.417P), dated
April 21, 1999. The USEPA defines MNA as the reliance on natural attenu-
ation processes (within the context of a carefully controlled and monitored
clean-up approach) to achieve site-specific remediation objectives within a
time frame that is reasonable compared to that offered by other more active
methods. MNA is generally not seen as a viable stand-alone remedy, but is
more commonly viewed as a possible component of an overall remedial
strategy for a contaminated site. Nonetheless, MNA is increasing being
viewed as a viable alternative for the management of contaminated sites in
the U.S. and other countries.
Palmer and Puls (1994) present an outline of how the natural attenuation
of Cr(VI) in the environment, especially in groundwater, can be evaluated
and how MNA can be implemented as a remedy for contamination. Their
work is summarized here. Also Ellis et al. (2002) and Blowes (2002) give
summaries of the use of stable isotopes of Cr as an evaluation methodology
for the implementation of MNA, and their work is also summarized herein.
Minerals containing reduced forms of iron and sulfur are abundant in
many aquifers, and these minerals reduce Cr(VI) to Cr(III) and promote the
precipitation of insoluble solids such as Cr(OH)3. Also, organic carbon-rich
materials can also reduce Cr(VI). In aquifers where reduced sediments are
abundant and the concentrations of Cr(VI) are low, the attenuation capacity
of the aquifer may be sufficient to prevent chromium migration. This mech-
anism of the natural attenuation of Cr(VI) could be used as a remedial
strategy. The appropriateness of MNA depends on the groundwater flow
system, the rate of chromium migration, and the ability of the aquifer mate-
rials to reduce Cr(VI). It can be difficult, however, to distinguish measured
chromate decreases caused by attenuation reactions from those caused by
mixing or dispersion (a dilution effect) in the aquifer.
However, for such a strategy to be adopted by a regulatory agency, the
likelihood that natural attenuation is likely to occur under the specific con-
ditions at the site being investigated will most likely have to be demon-
strated. There is no single test that can tell us if natural attenuation of Cr(VI)
will occur at a particular site. Several tests are briefly described which have
been utilized to address the factors affecting Cr(VI) transport in the subsur-
face and describe how the results can be utilized in determining the potential
L1608_C05.fm Page 197 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 197

for the natural attenuation of Cr(VI) in the subsurface. Such a demonstration


will likely comprise at least the following three items:

1. There are natural reductants in the aquifer


2. The amount of Cr(VI) and other reactive constituents do not exceed
the reductive capacity of the aquifer
3. The rate of Cr(VI) reduction to the target concentration is greater
than the rate of transport of Cr(VI) from source to point of compliance

Some of these criteria are relatively simple while others require additional
tests and interpretation. Each are discussed below.

5.7.1 There Are Natural Reductants in the Aquifer


In principle, the natural attenuation of Cr(VI) in the subsurface is feasible
as a result of interaction with naturally existing reductants. There are several
natural reductants that can transform Cr(VI) to Cr(III). Potential reductants
of Cr(VI) include

Aqueous species
Sorbed ions
Mineral constituents
Organic matter

During the migration of a Cr(VI) plume in groundwater, there is little


mixing of the waters containing the reducing agents and the Cr(VI). What
mixing does occur will be driven by molecular diffusion at the front or edges
of the plume, and diffusion from lower permeability lenses containing rel-
atively immobile water. Thus, reductants that are primarily dissolved in
groundwater such as Fe2+ are not going to be important in reducing Cr(VI).
The mixing of reductants and Cr(VI) is instead going to occur primarily
through the interactions of the Cr(VI) plume and the immobile soil matrix.
Such interactions include

Desorption of reductants such as Fe2+ from mineral surfaces


Direct and indirect surface redox reactions between Cr(VI) and the
mineral surfaces
Reduction by soil organic matter

The presence of Cr(III) may indicate either active reduction in the soil or
neutralization of acidic waters containing Cr(III) with subsequent precipita-
tion of chromium hydroxides. Therefore, specific reductants in the aquifer
should be identified, and this can be fairly straightforward. Reductants (such
as pyrite, FeS2) can be readily identifiable by visual characteristics or by
L1608_C05.fm Page 198 Friday, July 23, 2004 5:44 PM

198 Chromium(VI) Handbook

standard petrographic techniques (i.e., powder x-ray diffraction, scanning


electron microscopy, polarized light microscopy). Also, testing for organic
carbon can provide a measure of the amount of carbon available for reduc-
tion of Cr(VI).
Knowledge of the specific reductant within the aquifer can be useful in
determining the time scale for the reduction of Cr(VI). This is discussed later.
Based on literature studies, soils containing iron sulfides or organic matter
are more likely to reduce Cr(VI) on the time scales of interest than soils
containing iron(II) silicates.

5.7.2 The Amount of Cr(VI) and Other Reactive Constituents


Do Not Exceed the Reductive Capacity of the Aquifer
Studies demonstrate that groundwater contributes less than 1% of the
oxidation capacities (equivalents of chromium oxidized per gram of soil)
and reduction capacities (equivalents of chromium reduced per gram of
soil) of aquifer systems while the soil matrix contributes the remaining
fraction (Barcelona and Helm, 1991). Thus, any discussion of Cr(VI) reduc-
tion or Cr(III) oxidation in the subsurface must focus on the soil matrix.
Several soil tests can be useful in determining the mass of Cr(VI) and
Cr(III) and also the reduction and oxidation capacities of the aquifer
materials.

5.7.2.1 Mass of Cr(VI)


It must be demonstrated that the amount of Cr(VI) in the aquifer does not
exceed the capacity of the soil for reducing Cr(VI) to Cr(III). Therefore an
important step in evaluating the potential for natural attenuation is to deter-
mine the mass of Cr(VI) in the aquifer. Aqueous samples are most often
obtained from monitoring wells, or water can separated from the soil matrix
either by centrifugation or by squeezing. The pH of the water should be
measured to determine if it is within the proper range (5.5 to 12) to insure
the Cr(III) concentrations are less than 1 mol/l (0.05 mg/l).
Cr(VI) associated with the soil matrix may be sorbed to mineral surfaces
(particularly iron oxides) or precipitated as chromate minerals. There is no
accurate method for determining each of these fractions of Cr(VI); but,
sequential extractions have been used, where an initial water extraction serves
to remove remaining pore water and dissolve highly soluble chromium min-
erals present in the soil or that may have precipitated due to evaporation
during sample handling. Then, a phosphate extraction is used as a measure
of the exchangeable chromate in the soil (Bartlett and James, 1988); the
water is separated from the slurry and Cr(VI) is measured by the DPC Method
(Bartlett and Kimble, 1976; Bartlett and James, 1988). The increase in the
chromate concentrations is the amount of exchangeable chromate. The
phosphate removes chromate by both directly competing for the sorption
sites in the soil and indirectly (in some cases) by increasing the pH. Palmer
L1608_C05.fm Page 199 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 199

and Puls (1994) report that at many sites, the total Cr(VI) associated with the
soil matrix is the sum of BaCrO4 and the phosphate-extractable Cr(VI). This
sum is then added to the aqueous Cr(VI) to get the total concentration of
Cr(VI) in the soil (Cr(VI)tot).

5.7.2.2 Mass of Cr(III)


The presence of Cr(III) in the soil may demonstrate that Cr(VI) reduction is
occurring. Thus, the mass of Cr(III) in the soil could provide a measure of
the amount of reduction that has occurred. The total amount of Cr(III)
present in the soil is the sum of the mass in solution as well as mass associated
with the solid phase. Total chromium in solution can be determined by
atomic absorption spectroscopy (AAS) or inductively coupled plasma spec-
troscopy (ICP). When total chromium is statistically greater that Cr(VI),
Cr(III) can be simply determined by difference.

5.7.2.3 Reduction Capacity of the Aquifer


For the natural attenuation of Cr(VI) to be demonstrated, the soil must
possess sufficient reducing capacity (RC) to reduce all the Cr(VI) in the source
area. Thus, the total mass of Cr(VI) from the source (MO) must be less than
the total mass of Cr(VI) (MTOT) that can be reduced by the aquifer material
between the source and a point of compliance where XC is distance between
them. Where A is the cross-sectional area of the plume normal to the direction
of groundwater flow, and b is the dry bulk density of the aquifer, this
requirement can be expressed as

MO < MTOT = XCAbRC (5.8)

As the distance XC increases, the mass of Cr(VI) that can be reduced also
increases. A key difficulty in applying this criterion is in providing a reason-
able estimate of the total mass of Cr(VI) in the source area (MO).
The total Cr(VI) reducing capacity can be obtained using the classical
Walkley-Black method for determining soil organic carbon (Bartlett and
James, 1988; Walkley and Black, 1934). Although this method has its limita-
tions (Nelson and Sommers, 1982), it is a direct measure of how much Cr(VI)
can be reduced by a soil at extreme acid conditions. Variations on this method
have been described (Barcelona and Helm, 1991; Nelson and Sommers, 1982).
The extreme conditions of pH and temperature used in the total Cr(VI)
reducing capacity test may yield a greater reducing capacity than would be
available under most environmental conditions. The available reducing
capacity test of Bartlett and James (1988) is designed to determine the
reducing capacity at pH values more likely to be encountered in the field.
However in long-term reduction tests at near neutral pH, reduction has been
observed to be occurring after 250 days. Such long-term reduction tests are
not practical at most waste sites.
L1608_C05.fm Page 200 Friday, July 23, 2004 5:44 PM

200 Chromium(VI) Handbook

5.7.2.4 Oxidation Capacity of the Aquifer


As part of a geochemical cycle, both oxidation and reduction of Cr are
occurring simultaneously within the subsurface. As the Cr(III) is oxidized
to Cr(VI) by manganese dioxides in the soil, Cr(VI) can be reduced to Cr(III)
by some reductant such as soil organic carbon or pyrite. Ultimately, a
steady state situation will be reached where the rate of loss of Cr(VI) via
reduction is balanced by the rate of production by the oxidation of Cr(III).
Under certain conditions, oxidation of Cr(III) may be favored over the
reduction of Cr(VI). For example, soil containing Cr(III) formed by the
reduction of Cr(VI) may become a source of Cr(VI). Therefore, a potential
limitation to the use of natural attenuation of Cr(VI) as a remedial option
is the oxidation of the Cr(III) to Cr(VI) by oxidants such as MnO2. If the
oxidizing capacity of the soil is greater than the reduction capacity, then
as the chromium is cycled in the soil, the reductants could be exhausted,
the Cr(III) could oxidized, and ultimately, Cr(VI) could be mobilized in the
soil. It is therefore important to determine the capacity of the aquifer to
oxidize Cr(III).
Bartlett and James (1979, 1988) suggest a simple test for the amount of
Cr(III) that can be oxidized by a soil. Barcelona and Helm (1991) give a
method to measure the oxidation capacity of soils. Each of these methods
has some problems. If the sole Cr(III) oxidation mechanism is by manganese
oxides, then using extraction methods specifically designed for this purpose
may be useful (Chao, 1972; Gambrell and Patrick, 1982).
Fendorf and Zasoski (1992) suggest that CrOH2+ is the reactive species in
the oxidation of Cr(III) by MnO2. The key point to consider when considering
natural attenuation in soils that contain both a reductant and MnO2 is that
as long as the supply of reductant and MnO2 have not been significantly
depleted, (the HCrO4 concentration) does not converge to zero with increas-
ing residence time within the aquifer as one would expect for a first-order
reaction that only considers reduction of Cr(VI). Rather, HCrO4 converges
to some steady-state concentration that is >0 that may or may not be above
the MCL. This steady-state concentration increases with a preference of oxi-
dation over reduction and it varies with pH.

5.7.3 The Rate of Cr(VI) Reduction to the Target Concentration


Compared to the Rate of Transport of Cr(VI) from Source
to Point of Compliance
In principle, if the rate equations are correct and all of the parameters are
known, one could calculate the steady-state Cr(VI) concentration and deter-
mine if natural attenuation could achieve compliance goals. The major lim-
itation to this approach is the lack of information about the rate of oxidation
and reduction of chromium under conditions likely to be encountered by
plumes emanating from chromium sources. Without better information
L1608_C05.fm Page 201 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 201

about these rate processes under a wider range of conditions with respect
to pH, the use of the natural attenuation option for contaminated soils will
continue to be a highly debated issue.

5.7.3.1 Rates of Oxidation and Reduction


If natural attenuation is to be a viable option, the time for the reduction
reaction to decrease the concentration from its initial concentration (CO) to
some target concentration (CS), such as a drinking water standard (i.e., MCL),
should be less than the residence time of the contaminated water in the
portion of the aquifer between the source of the Cr(VI) and the point of
compliance. Difficulties in utilizing this criterion as mentioned above arise
in applying the appropriate rate equation and obtaining the pertinent rate
coefficients. Rates can be obtained from the technical literature, but one must
use reduction rates based on materials that are most likely controlling the
Cr(VI) reduction at the site under evaluation. In addition, because the rates
are concentration-dependent, and are related to the specific reductant and
the pH level, it is important to obtain rate coefficients that were acquired
under conditions similar to the site. As far as reductants are concerned,
literature studies indicate soils containing iron sulfides or organic matter are
more likely to reduce Cr(VI) on the time scales of interest than soils contain-
ing iron(II) silicates.
One method of obtaining the net rate of reduction is through tests on
uncontaminated soils (background soil) at the site that are similar to those
through within the contaminant plume. Cr(VI) can be added to the back-
ground soil in the laboratory and the Cr(VI) concentrations monitored over
time. The reaction vessels must exclude light to prevent photoreduction
reactions and the slurry must have the same pH as the contaminant plume.
A key limitation to such experiments is that they require several months to
a year to complete.

5.7.3.2 Estimating Reduction from Monitoring Well Data


In principle, Cr(VI) reduction can be estimated from the decrease in the mass
of Cr(VI) in the aquifer (Henderson, 1994). The key difficulty in such an
approach is to estimate the mass of Cr(VI) using aqueous concentrations.
The total mass of Cr(VI) in the aquifer is the sum of the mass that is in
solution, the mass that is sorbed to the aquifer matrix, and the mass that is
precipitated within the aquifer. The mass of Cr(VI) in solution (Maq) is
obtained by integrating the Cr(VI) concentrations over the volume of the
contaminated aquifer,

Maq = v CV (5.9)

where V is the volume of aquifer containing a plume with a Cr(VI) concen-


tration of C. (and v represents the porosity). The mass of Cr(VI) sorbed to
L1608_C05.fm Page 202 Friday, July 23, 2004 5:44 PM

202 Chromium(VI) Handbook

the soil matrix (Mads) can be computed from an sorption isotherm. There
is no unique amount of Cr(VI) precipitate for a given Cr(VI) concentration.
Therefore it is impossible to estimate mass of this fraction of Cr(VI) in the
subsurface using only the measured concentrations in monitoring wells.
Thus, natural attenuation of Cr(VI) from mass balances using monitoring
well data can only be used when it can be reasonably demonstrated the
Cr(VI) precipitates cannot form within the aquifer. Even when it is dem-
onstrated that the formation of precipitates within the aquifer is unlikely,
there are inherent problems with any monitoring system, creating uncer-
tainties in the estimated mass of Cr(VI) during a sampling round.

5.7.3.3 Monitoring Reduction Via Stable Isotopes of Chromium


Variations in the isotopic ratios of light elements are sensitive indicators of
chemical processes that occur in natural systems. For example, the 34S/32S
ratio in dissolved sulfate increases when bacteria reduce sulfate to sulfide.
Reduction reactions tend to enrich products in the lighter isotopes because
they preferentially react (Hoefs, 1987), and the residual reactants become
progressively enriched in the heavier isotopes as reduction proceeds (Boettcher
et al., 1990; Thode and Monster, 1965; and Johnson et al., 1999). Ellis et al.
(2002) now show that the 53Cr/52Cr ratio also changes during reduction of
Cr(VI) to Cr(III). They show that abiotic reduction of Cr(VI) resulting from
reaction with the mineral magnetite, estuarine sediments, and freshwater
sediments leads to a consistent 53Cr/52Cr shift. This observation indicates
that 53Cr/52Cr ratios increase systematically with progressive Cr(VI) reduc-
tion in groundwater.
The measurements of chromium isotope ratios provides a new tool for
evaluating the extent and rate of the natural attenuation of chromate. Rela-
tive to conventional methods, Blowes (2002) states that the use of isotopes
has two advantages: (1) Each 53Cr/52Cr determination provides a measure of
the amount of reduction that has already occurred, and there is thus no need
to see whether Cr(VI) mass decreases over time; and (2) the measurement of
reduction integrates spatially over a flow path, whereas analyses of aquifer
solids give information on a much smaller spatial scale.
53Cr/52Cr measurements can also help to evaluate in situ approaches of

remediation, which have been developed for site where natural attenuation
is insufficient to prevent chromate migration (i.e., permeable reactive barri-
ers, injection of chemical reactants to reduce chromate, and injection of a
reductant to react with the aquifer materials to form reduced minerals in the
aquifer). The new finding by Ellis et al. (2002) that 53Cr/52Cr ratios reveal
the extent of abiotic reduction suggests that 53Cr/52Cr measurements can
assist in the evaluation of the effectiveness of all these approaches to chro-
mate reduction in groundwater.
Chromium has four stable isotopes of masses 50 (4.35%), 52 (83.8%),
53 (9.50%), and 54 (2.37%) (Rotaru and Birck, 1992; Handbook of Chemistry
L1608_C05.fm Page 203 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 203

and Physics, 1996). Ellis et al. (2002) measured chromium isotope fraction-
ation during reduction of Cr(VI) by slurries of magnetite and two sediment
samples. Because magnetite is a likely reducing agent in some aquifer sed-
iments (Anderson et al., 1994), the magnetite experiment provided a simple
analog for a natural aquifer. (In these experiments) at pH 6 to 7, sorption of
Cr(VI) was negligible. As reduction proceeded and the Cr(VI) concentrations
decreased, 53 Cr (defined in Chapter 2) values of the remaining unreduced
Cr(VI) increased, indicating preferential reduction of the lighter isotopes.
These experiments showed that reduction of Cr(VI) results in Cr stable-
isotope fractionation.
To estimate initial isotopic compositions of the contaminants, Ellis et al.
(2002) measured 53Cr values in samples of plating baths in active use at
different sites, the chromic acid supply used to make up plating baths,
laboratory chromium reagents, and basaltic rock standards. All of these
samples yielded 53Cr values close to 0%. Therefore, Ellis et al. (2002)
suggested that Cr released as plating waste generally has an initial 53Cr
value slightly greater than zero. If so, detection of Cr(VI) reduction in
groundwater systems would be relatively simple, as the initial 53Cr value
would be known and groundwater values greater than that would directly
indicate the extent of reduction. If, on the other hand, plating wastes have
variable 53Cr values, then it may be possible to distinguish different
contamination sources via their 53Cr values. In studies of groundwater
contamination sites, Ellis et al. (2002) showed that all of the groundwater
Cr(VI) analyses indicated enrichment in the heavy isotope relative to the
plating baths, and Cr(VI) reduction had preferentially removed lighter
isotopes from the groundwater. The variation in 53Cr values at each of
the sites suggested that reduction of Cr(VI) was occurring and had pro-
gressed to different degrees in different parts of the contaminant plumes.
The highest 53Cr values were found in samples with lowest Cr(VI) con-
centration at (the sites studied). This result was to be expected because
the fringe areas of the contaminant plumes likely would have greater
degrees of reduction than the plume cores, where chromium concentra-
tions are high and the reducing power of the aquifer materials had been
depleted.
Stable Cr isotope ratios can thus serve as indicators of the extent of Cr(VI)
reduction in groundwater. Cr(VI) reduction by bacteria or reducing agents
other than those studied by Ellis et al. (2002) could induce greater or lesser
isotopic fractionation than (they) observed. And, processes other than reduc-
tion (i.e., sorption, precipitation, and uptake by plants and algae) could also
be responsible for removing chromium from solution (James and Bartlett,
1983, 1984, and 1988). If such processes and/or Cr(III) oxidation also induce
isotopic fractionation, this could complicate the interpretation of 53Cr mea-
surements. However, as with S and Se isotopes (Johnson et al., 1999) it could
be expected that the dominant cause of chromium isotope fractionation is
reduction.
L1608_C05.fm Page 204 Friday, July 23, 2004 5:44 PM

204 Chromium(VI) Handbook

Bibliography
Abu-Saba, K.E., Sedlak, D.L., and Flegal, A.R., 2000, Mar. Chem., vol. 69, pp. 3341.
Achterberg, E.P., van den Berg, C.M.G., Boussemart, M., and Davison, W., 1997,
Speciation and cycling of trace metals in Esthwaite Water: a productive
English lake with seasonal deep-water anoxia, Geochim. Cosmochim. Acta,
vol. 61, pp. 52335253.
Agency for Toxic Substances and Disease Registry (ATSDR), 1992, Toxicological Profile
for Chromium, Draft, U.S. Dept. of Health and Human Services, Washington,
D.C.
Ainsworth, C.C., Girvin, D.C., Zachara, J.M., and Smith, S.C., 1989, Chromate ad-
sorption on goethite: effects of aluminum substitution, Soil Sci. Soc. Am. J.,
vol. 53, pp. 411418.
Amacher, M.C. and Baker, D.E., 1982, Redox reactions involving chromium, plutonium
and manganese in soils, DOE/DP/04515-1, Institute for Research on Land and
Water Resources, Pennsylvania State University and U.S. Department of Ener-
gy, Las Vegas, NV, 166 p.
Amacher, M.C., Kotuby-Amacher, J., Selim, H.M., and Iskandar, I.K., 1986, Retention
and release of metals by soils, evaluation of several models, Geoderma, vol. 38,
pp. 131154.
Amacher, M.C., Selim., H.M., and Iskandar, I.K., 1988, Kinetics of chromium(VI) and
cadmium retention in soils: a non-linear multireaction model, Soil Sci. Soc. Am.,
vol. 52, pp. 398408.
Anderson, L.D., Kent, D.B., and Davis, J.A., 1994, Environ. Sci. Technol., vol. 28, pp. 178.
Anderson, R.A., Polansky, M.M., Brantner, J.H., and Roginski, E.E., 1977, Chemical
and biological properties of biologically active chromium, in Int. Symp. Trace
Element Metabolism in Man and Animals, Friesing, Federal Republic of Germany,
pp. 269271.
Barcelona, M.J. and Helm, T.J., 1991, Oxidation-reduction capacities of aquifer solids,
Environ. Sci. Technol., vol. 25, pp. 15651572.
Baron, D. Palmer, C.D., and Stanley, J.T., 1996, Identification of two iron-chromate
precipitates in a Cr(VI)-contaminated soil, Environ. Sci. Technol., vol. 30, pp.
964968.
Bartlett, R.J. and James, B., 1979, Behavior of chromium in soils: III: Oxidation, J.
Environ. Qual., vol. 8, pp. 3135.
Bartlett, R.J. and James, B.R., 1988, Mobility and bioavailability of chromium in soils,
in Chromium in the Natural and Human Environments, Nriagu, J.O. and Nieboer,
E., Eds., John Wiley and Sons, New York, NY, pp. 267304.
Bartlett, R.J. and Kimble, J.M., 1976, Behavior of chromium in soils, I. Trivalent forms,
II. Hexavalent forms, J. Environ. Qual., vol. 5, pp. 379386.
Bartlett, R.J., 1991, Chromium cycling in soils and water: links, gaps and methods,
Environ. Health Prespective, vol. 92, pp. 1724.
Beaubien, S., Nriagu, J., Blowes, D., and Lawson, G., 1994, Chromium speciation and
distribution in the Great Lakes, Environ. Sci. Technol., vol. 28, pp. 730738.
Benjamin, M.M. and Bloom, N.S., 1981, Effect of strong binding of anionic adsor-
bates on adsorption of trace metals on amorphous iron oxyhydroxide, in
Adsorption from Aqueous Solution, Tewari, P.H., Ed., Plenum Press, New York,
NY, pp. 4160.
L1608_C05.fm Page 205 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 205

Bloomfield, C. and Pruden, G., 1980, The behavior of Cr(VI) in soil under aerobic
and anaerobic conditions, Environ. Pollut. (Series A), vol. 23, pp. 103114.
Blowes, D.W. and Ptacek, C.J., 1992, Geochemical remediation of groundwater by
permeable reactive walls: Removal of chromate by reaction with iron-bearing
solids, Proc. Subsurface Restoration Conference, June 2124, 1992, Dallas, TX,
pp. 214216.
Blowes, D., 2002, Tracking Hexavalent Cr in Groundwater, Science, vol. 295.
Bodek, I., Lyman, W.J., Reehl, W.F., and Rosenblatt, D.H., 1988, Environmental Inor-
ganic Chemistry Properties, Processes and Estimation Methods, Bodek, I. Ed., Per-
gamon Press, New York, NY.
Boettcher, J., Strebel, O., Voerkelius, S., and Schmidt, H.L., 1990, J. Hydrol., vol. 114,
pp. 413.
Calder, L.M., 1988, Chromium contamination of groundwater, in Chromium in the
Natural and Human Environments, Nriagu, J.O. and Nieboer, E., Eds., John Wiley
and Sons, New York, pp. 215231.
Campos, J., Martinez-Pacheco, M., and Cervantes, C., 1995, Hexavalent-chromium
reduction by chromate-resistant Bacillus sp. Strain, Antione von Leeuwenhoek,
vol. 68, pp. 203208.
Cary, E.E., Allaway, W.H., and Olson, O.E., 1977, Control of chromium concentration
in food plants, II. Chemistry of chromium in soils and its availability to plants,
J. Agric. Food Chem., vol. 25, pp. 305309.
Cary, E.E., Ed., 1982, Chromium in Air, Soil, and Natural Waters, Elsevier Biomedical,
New York, NY, pp. 4963.
Cavallaro, N. and McBride, M.B., 1978, Copper and cadmium adsorption charac-
teristics of selected acid and calcareous soils, Soil Sci. Soc. Am. J., vol. 42,
pp. 550556.
Chao, T.T., 1972, Selective dissolution of manganese oxides from soils and sedi-
ments with acidified hydroxylamine hydrochloride, Soil Sci. Amer. Proc.,
vol. 36, pp. 764768.
Chuecas, D.L. and Riley, J.P., 1966, The spectrophotometric determination of chromi-
um in sea water, Analyt. Chim. Acta, vol. 35, pp. 240246.
Clay, R.E., 1992, Multimedia Environmental Distribution of Gaseous, Dissolved and parti-
cle-Bound Pollutant, Master of Science Thesis, Chemical Engineering Depart-
ment, University of California Los Angeles (UCLA).
Coates, J.D., Phillips, E.J.P., Lonergan, D.J., Jenter, H., and Lovely, D.R., 1996, Appl.
Environ. Microbiol., vol. 62, pp. 1531.
Coleman, R.N., 1988, Chromium toxicity: effects on microorganisms with special
reference to the soil matrix, in Chromium in the Natural and Human Environ-
ments, Nriagu, J.O. and Nieboer, E., Eds., John Wiley and Sons, New York, NY,
pp. 335368.
Cranston, R.E. and Murray, J.W., 1980, Chromium species in the Columbia River
estuary, Limnol. Oceanogr., vol. 25, pp. 11041122.
Cranston, R.E. and Murray, J.W., 1978, Determination of chromium species in natural
waters, Analyt. Chim. Acta, vol. 99, pp. 275282.
Davis, A. and Olsen, R.L., 1995, The Geochemistry of chromium migration and
remediation in the subsurface, Ground Water, vol. 33, pp. 759768.
Davis, J.A. and Leckie, J.O., 1980, Surface ionization and complexation at the
oxide/water interface: III. Adsorption of anions, J. Colloid Interface Sci.,
vol. 74, pp. 3243.
L1608_C05.fm Page 206 Friday, July 23, 2004 5:44 PM

206 Chromium(VI) Handbook

DeLeo, P.C. and Ehrlich, H.L., 1994, Reduction of hexavalent chromium by pseudomonas
fluorescens LB300 in batch and continuous cultures, Appl. Microbiol. Biotechnol.,
vol. 40, pp. 756759.
Deng, B. and Stone, A.T., 1996, Surface-catalyzed chromium(VI) reduction: the
TiO2Cr(VI)-mandelic acid system, Environ. Sci. Technol., vol. 30, pp. 463472.
Deutsch, M., 1972, Incidents of chromium contamination of groundwaters in Mich-
igan, in Water Quality in a Stressed Environment, Pettyjohn, W.A., Eds., Burgess
Publishing Co., Minneapolis, MN, pp. 149159.
Deverel, S.J. and Millard, S.P., 1988, Distribution and mobility of selenium and other
trace elements in shallow groundwater of the Western San Joaquin Valley,
California, Envir. Sci. Technol., vol. 22, pp. 697702.
Dreiss, S.J., 1986, Chromium migration through sludge treated soils, Ground Water,
vol. 24, pp. 312321.
Dudley, L.M., McLean, J.E., Furst, T.H., and Jurinak, J.J., 1991, Sorption of Cd and
Cu from an acid mine waste extract by two calcareous soils: column studies,
Soil Sci., vol. 151, pp. 121135.
Dudley, L.M., McLean, J.E., Sims, R.C., and Jurinak, J.J., 1988, Sorption of copper and
cadmium from the water-soluble fraction of an acid mine waste by two calcar-
eous soils, Soil Sci., vol. 145, pp. 207214.
Eary, L.E. and Rai, D., 1991, Chromate reduction by subsurface soils under acidic
conditions, Soil Sci. Soc. Am. J., vol. 55, pp. 676683.
Eary, L.E. and Rai, D., 1988, Chromate removal from aqueous wastes by reduction
with ferrous iron, Environ. Sci. Technol., vol. 22, pp. 972977.
Eary, L.E. and Rai, D., 1987, Kinetics of chromium (III) oxidation to chromium(VI)
by reaction with manganese dioxide, Environ. Sci. Technol., vol. 21, pp.
11871193.
Au: Pls. Eary, L.E. and Rai, D., 1989, Kinetics of chromate reduction by ferrous ions derived
provide from hematite and biotite at 25C, Am. J. Sci., vol. 289, pp. 180213.
chapter Eckert, J.M., Stewart, J.J., Waite, T.D., Szymezek, R., and Williams, K.L., 1990, Anal.
title for Chim. Acta, vol. 236, pp. 357362.
reference Ellis, A.S., Johnson, T.M., and Bullen, T.D., 2002, Chromium isotopes and the fate of
hexavalent chromium in the environment, Science, vol. 295, pp. 20602062.
Au: Pls. Essen, J. and El Bassam, N., 1981, On the movility of cadmium under aerobic soil
provide conditions, Environ. Pollut. Ser. A., pp. 1531.
chapter Fendorf, S., Wielinga, B.W., and Hansel, C.M., 2001, Int. Geol., vol. 42, pp. 691.
title for Fendorf, S.E. and Li, G., 1996, Kinetics of chromate reduction by ferrous iron, Environ.
reference Sci. Technol., vol. 30, pp. 16141617.
Fendorf, S.E. and Zasoski, R.J., 1992, Chromium(III) oxidation by -MnO2, 1. Char-
acterization, Environ. Sci. Technol., vol. 26, pp. 7985.
French, W.B., Gallagher, M., Passero, R.N., and Straw, W.T., 1985, Hydrogeologic
investigation for remedial action related to a chromium-arsenic-copper dis-
charge to soil and groundwater, in Innovations in Water and Wastewater Fields,
Glysson E.A., Swan D.E., and Way E.J., Eds., Ann Arbor Press, Ann Arbor, MI,
pp. 209229.
Friess, S.L., 1989, Carcinogenic risk assessment criteria associated with inhalation of
airborne particulates containing chromium(VI/III), Sci. Tot. Environ., vol. 86,
pp. 109112.
Frissel, M.J., Poelstra, P., and Reisinger, P., 1975, The Behavior of Chromium in Aquatic
and Terrestrial Food Chains: EUR 5375e, Boite postale 1003, Luxembourg, pp. 2742.
L1608_C05.fm Page 207 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 207

Gambrell, R.P. and Patrick, W.H., 1982, Manganese, in Methods of Soil Analysis, Part
2, Chemical and Microbiological Properties, Page, A. L., Ed., 2nd ed., Agronomy
No. 9, Part 2, American Society of Agronomy, Soil Science Society of America,
Madison, WI, pp. 313322.
Gerritsee, R.G., Vriesema, R., Dalenberg, J.W., and DeRoos, H.P., 1982, Effect of
sewage sludge on trace element mobility in soils, J. Envr. Qual., vol. 11,
pp. 359364.
Griffin, R.A. and Shimp, N.F., 1978, Attenuation of pollutants in municipal landfill leachate
by clay minerals, EPA/600/278/157, U.S. Environmental Protection Agency,
Washington, D.C.
Griffin, R.A., Au, H.K., and Frost, R.R., 1977, Effect of pH on adsorption of
chromium from landfill leachate by clay minerals, J. Envr. Sci. Hlth. Ser. A.,
vol. 12, pp. 431449.
Grohse, P.M., Gutknecht, W.F., Hodson, L., and Wilson, B.M., 1988, The fate of hexava-
lent chromium in the atmosphere: Research Triangle Institute Report on CARB
Contract #A609632, Report Number ARB/R89/379, California Air Resources
Board, Sacramento, CA.
Grove, D.B. and Stollenwerk, K.G., 1985, Modeling the rate-controlled sorption of
hexavalent chromium, Water Resources Res., vol. 21, pp. 17031709.
Grove, J.H. and Ellis, B.G., 1980, Extractable chromium as related to soil pH and
applied chromium, Soil Sci. Soc. Am. J., vol. 44, pp. 238242.
Handa, B.K., 1988, Occurrence and distribution of chromium in natural waters of
India, in Chromium in Natural and Human Environments, Nriagu, J.O. and
Nieboer, E., Eds., Wiley Interscience, New York, NY, pp. 189215.
Handbook of Chemistry and Physics, 1996, 77th ed., CRC Press.
Hartford, W., 1983 Chromium chemicals, in Kirk-Othmer Encyclopaedia of Chemical
Technology, Grayson, M., Ed., 3rd ed., vol. 6, Wiley-Interscience, New York, NY,
pp. 83120.
Hathway, J.A., 1989, Role of Epidemiologic studies in evaluating the carcinogenicity
of chromium compounds, Sci. Total Environ., vol. 86, pp. 169179.
Hem, J.D., 1977, Reactions of metal ions at surfaces of hydrous iron oxide, Geochim.
Cosmochim. Acta, vol. 41, pp. 527538.
Henderson, T., 1994, Geochemical reduction of hexavalent chromium in the Trinity
Sand, Ground Water, vol. 32, pp. 477486.
Hoefs, J., 1987, Stable Isotope Geochemistry, 3rd ed., Springer-Verlag, New York, NY.
Au: Pls. Hug, S.J., Laubscher, H.U., and James, B.R., 1997, Environ. Sci. Technol., vol. 31,
provide pp. 160170.
chapter James, B.R. and Bartlett, R.J., 1983, Behavior of chromium in soils: V. Fate of organ-
title for ically complexed Cr(III) added to soil; VI. Interaction between oxidation-re-
reference duction and organic complexation; VII. Adsorption and reduction of
hexavalent forms, J. Environ. Qual., vol. 12, pp. 169181.
James, B.R. and Bartlett, R.J., 1988, Mobility and bioavailability of chromium in soil, in
Chromium in Natural and Human Environments, Nriagu, J.O. and Nieboer, E.,
Eds., Wiley Interscience, New York, NY, pp. 265305.
James, B.R. and Bartlett, R.J., 1984, Plant-soil interactions of chromium, J. Environ.
Qual., vol. 13, pp. 6770.
James, B.R., 1994, Hexavalent chromium solubility and reduction in alkaline soils
enriched with chromite ore processing residue, J. Environ. Qual., vol. 23,
pp. 227233.
L1608_C05.fm Page 208 Friday, July 23, 2004 5:44 PM

208 Chromium(VI) Handbook

James, B.R., 1996, The challenge of remediation chromium-contaminated soil, Environ.


Sci. Technol., vol. 30, pp. 248251.
Jenne, 1968, Control of Mn, Fe, Co, Ni, Cu, and Zn concentrations in soils and
waterthe dominant role of hydrous manganese and iron oxides, Adv. In Chem.,
vol. 7, pp. 337387.
Johnson, C.A. and Xyla, A.G., 1991, The oxidation of chromium(III) to chromium(VI)
on the surface of manganite (Gamma-MnOOH), Geochim. Cosmochim., vol. 55,
Au: Pls.
pp. 28612866. provide
Johnson, T.M., Herbel, M.J., Bullen, T.D., and Zawislanski, P.T., 1999, Geochim. Cos- chapter
mochim. Acta, vol. 63, pp. 2775. title
Jurinak, J.J. and Bauer, N., 1956, Thermodymanics of zinc adsorption on calcite, colo-
mite and magnesite-type minerals, Soil Sci. Soc. Am. Proc., vol. 20, pp. 466471.
Kaczynski, S.E. and Kieber, R.J., 1993, Aqueous trivalent chromium photoproduction
in natural waters, Environ. Sci. Technol., vol. 27, pp. 15721576.
Kaczynski, S.E. and Kieber, R.J., 1994, Hydrophobic C18 bound organic complexes
of chromium and their potential impact on the geochemistry of chromium in
natural waters, Environ. Sci. Technol., vol. 28, pp. 799804.
Kent, D.B., Davis, J.A., Maest, A.S., and Rea, B.A., 1989, Field and laboratory studies
of transport of reactive solutes in groundwater, in Proc. 6th Int. Conf. Water-Rock
Interaction, Miles, D.L., Ed., Miles, Balkema, Rotterdam, Netherlands, pp.
381383.
Kieber, R.J. and Heiz, G.R., 1992, Indirect photoreduction of aqueous chromium (VI),
Environ. Sci., Technol., vol. 26, pp. 307312.
Kimbrough, D.E., Cohen, Y., Winer, A.M., Creelman, L., and Mabuni, C., 1999, A
critical assessment of chromium in the environment, in Critical Reviews, Envi-
ronmental Science and Technology, vol. 29, pp. 146.
Korte, N.E., Skopp, J., Fuller, W.H., Niebla, E.E., and Aleshii, B.A., 1976, Trace element
movement in soils: influence of soil physical and chemical properties, Soil Sci.,
vol. 122, pp. 350359.
Kotas, J. and Stasicka, Z., 2000, Chromium occurrence in the environment and meth-
ods of its speciation, Environ. Poll. vol. 107, pp. 263283.
Lancy, L.E., 1966, Treatment of spent cooling waters, U.S. Patent No. 3,294,960.
Au: Pls. Lieber, M., Perlmutter, N.M., and Frauenthal, H.L., 1964, Cadmium and hexavalent
provide chromium in Nassau County ground water, J. AWWA, vol. 56, pp. 739747.
chapter Lovely, D.R., 1993, Annu. Rev. Microbiol., vol. 47, pp. 263.
title for Loyaux-Lawniczak, S., Lecomte, P., Ehrhardt, J., 2001, Behavior of hexavalent chro-
reference mium in a polluted groundwater: redox processes and immobilization in soils,
Environ. Sci. Technol., vol. 35, pp. 13501357.
MacNaughton, M.G., 1975, Adsorption of chromium(VI) at the oxide-water interface,
Interim Report: Envir. Chem. Div., Environics Directorate, Air Force Civil Eng.
Center, Tyndal AFB, FL.
McBride, M.B. and Bouldin, D.R., 1984, Long-term reactions of copper(II) in a con-
taminated calcareous soil, Soil Sci. Soc. Am. J., vol. 48, pp. 5659.
McBride, M.B., 1980, Chemisorption of Cd2+ on calcite surfaces, Soil Sci. Soc. Am. J.,
vol. 44, pp. 2628.
McLean, J.E. and Bledsoe, B.E., 1992, Behavior of metals in soils, EPA/540/S-92/018.
Murray, J.W., Spell, B., and Paul, B., 1983, The contrasting geochemistry of manganese
and chromium in the Eastern Tropical Pacific Ocean, in Trace Metals in Seawater,
Wong, C.S., Boyle, E., Bruland, K., Burton, J.D., and Goldberg E.D. Eds., Plenum
Press, New York, NY, pp. 643670.
L1608_C05.fm Page 209 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 209

Music, S., Ristic, M., and Tonkovic, M., 1986, Sorption of chromium(VI) on hydrous
iron oxides, Z. Wass. Abwass. Forsch., vol. 19, pp. 186196.
Nakayama, E., Kuwamoto, T., Tsurubo, S., Tokoro, H., and Fujinaga, T., 1981, Chem-
ical speciation of chromium in sea water: 1. Effect of naturally occurring
organic materials on the complex formation of chromium(III), Analyt. Chim.
Acta, vol. 130, pp. 289294.
Nelson, D.W. and Sommers, L.E., 1982, Total carbon, organic carbon, and organic
matter, in Methods of Soil Analysis, Part Two, Dinauer R.C., Ed., American
Society of Agronomy, Inc. and Soil Science Society of America, Inc., Madison,
WI, pp. 539580.
Nriagu, J.O., 1988, Historical perspectives; and, Production and uses of chromi-
um, in Chromium in the Natural and Human Environments, Nriagu, J.O. and
Nieboer, E., Eds., John Wiley and Sons, New York, NY, vol. 20, pp. 120
and pp. 81104.
Pacyna, J.M. and Nriagu, J.O., 1988, Atmospheric emissions of chromium from nat-
ural and anthropogenic sources, in Chromium in Natural and Human Environ-
ments, Nriagu, J.O. and Nieboer, E., Eds., Wiley Interscience, New York, NY,
pp. 105125.
Palmer, C.D. and Wittbrodt, P.R., 1990, Geochemical characterization of the United
Chrome Products Site, Final Report, in Stage 2 Deep Aquifer Drilling Technical
Report, United Chrome Products Site, Corvallis, Oregon, CH2M Hill, Corvallis,
OR.
Palmer, C.D. and Wittbrodt, P.R., 1991, Processes affecting the remediation of chro-
mium-contaminated sites, Environ. Health Perspectives, vol. 92, pp. 2540.
Palmer, C.D. and Puls, R.W., 1994, Natural attenuation of hexavalent chromium in
groundwater and soils, Ground Water, EPA/540/S94/505.
Paya Perez, A.B., Gotz, L., Kephalopoulos, S.D., and Bignoli, G., 1988, Sorption of
chromium species on soil, in Heavy Metals in the Hydrocycle, Astruc, M. and
Lester, J.N., Eds., Selper, London, pp. 5966.
Perlmutter, N.M. and Lieber, M., 1970, Dispersal of plating wastes and sewage con-
tamination in ground water and surface water, South Farmingdale Massapequa
Area, Nassau County, New York, U.S.G.S. Water Supply Paper 1879-G.
Pettine, M. and Millero, F.J., 1990, Limnol. Oceanogr., vol. 35, pp. 730736.
Pettine, M. and Millero, F.J., 1991, Mar. Chem., vol. 34, pp. 2946.
Pettine, M., Campanella, L., Millero, F.J., 2002, Reduction of hexavalent chromium
by H2O2 in acidic solutions, Environ. Sci. Technol., vol. 36, n. 5.
Pettine, M., DOttone, L., Campanella, L., Millero, F.J., and Passino, R., 1998, The
reduction of chromium(VI) by iron(II) in aqueous solutions, Geochim. Cosmo-
chim. Acta, vol. 62, pp. 15091519.
Pettine, M., Millero, F.J., and La Noce, T., 1991, Chromium(III) interactions in seawater
through its oxidation kinetics, Mar. Chem., vol. 34, pp. 2946.
Pettine, M., Millero, F.J., and Passino, R., 1994, Mar. Chem., vol. 46, pp. 335344.
Powell, R.M., Puls, R.W., Hightower, S.K., and Sabatini, D.A., 1995, Coupled iron
corrosion and chromate reduction: mechanism for subsurface remediation,
Environ. Sci. Technol., vol. 29, pp. 19131922.
Rai, D., Eary, L.E., and Zachara, J.M., 1989, Environmental chemistry of chromium,
Sci. Tot. Environ., vol. 86, pp. 1523.
Rai, D., Sass, B.M., Moore, D.A., 1987, Chromium (III) hydrolysis constants and
solubility of chromium (III) hydroxide, Inorg. Chem., vol. 26, pp. 345349.
L1608_C05.fm Page 210 Friday, July 23, 2004 5:44 PM

210 Chromium(VI) Handbook

Rai, D., Zachara, J.M., Eary, L.E., Ainsworth, C.C., Amonette, J.E., Cowan, C.E.,
Szelmeczka, R.W., Resch, C.T., Schmidt, R.L., Girvin, D.C., and Smith, S.C.,
1988, Chromium reactions in geological materials, Interim Report, Electric Power
Research Institute (EPRI) EA-5741, EPRI, Palo Alto, CA.
Rai, D., Zachara, J.M., Eary, L.E., Girvin, D.C., Moore, D.A., Resch, C.T., Sass, B.M.,
and Schmidt, R.L., 1986, Geochemical behavior of chromium species, Interim Report
Electric Power Research Institute (EPRI) EA EA4544, EPRI, Palo Alto, CA.
Rai, D., Zachara, J.M., Schwab, A.P., Schmidt, R.L., Girvin, D.C., and Rogers, J.E.,
1984, Chemical attenuation rates, coefficients, and constants in leachate migra-
tion, Vol. 1, A critical review, Final Report Electric Power Research Institute
(EPRI) EA EA-3356, EPRI, Palo Alto, CA.
Richard, F.C. and Bourg, A.C.M., 1991, Aqueous geochemistry of chromium: a review,
Wat. Res., vol. 25, pp. 807816.
Ritchie, G.S.P. and Sposito, G., 1995, Speciation in soil, in Chemical Speciation in the
Environment Ure, A.M. and Davidson, C.M., Eds., Blackie Academic and Pro-
fessional, Glasgow, Scotland, pp. 201233.
Robertson, F.N., 1975, Hexavalent chromium in the groundwater in the Paradise
Valley, Arizona, Ground Water, vol. 13, pp. 516527.
Rotaru, M., Birck, J.L., and Allegre, C.J., 1992, Nature, vol. 358, pp. 465.
Au: Pls. Saleh, F.Y., Parkerton, T.F., Lewis, R.V., Huang, J.H., and Dickson, K.L., 1989, Kinetics
provide of chromium transformation in the environment, Sci. Tot. Environ., vol. 86,
chapter pp. 2541.
title for Salomons, W. and DeGroot, A.J., 1978, Pollution history of trace metals in sediments,
reference as affected by the Rhine River, in Environmental Biogeochemistry, Krumbein,
W.E., Ed., Ann Arbor Science, Ann Arbor, MI, pp. 149162.
Santillan-Medrano, J. and Jurinak, J.J., 1975, The chemistry of lead and cadmium in
soils: solid phase formation, Soil Sci. Am. Proc., vol. 29, pp. 851856.
Sass, B.M. and Rai, D., 1987, Volubility of amorphous chromium(III)-Iron(III) hydrox-
ide solid solutions, Inorg. Chem., vol. 26, pp. 22282232.
Schacklette, H.T. and Boerngen, J.G., 1984, Element concentration in soils and other
surficial materials, U.S. Geol. Surv. Prof. Paper 1710, Washington, DC.
Schmidt, R.L., 1984, Thermodynamic properties and environmental chemistry of chromium,
U.S. Dept. Energy, DE-AC06-76RLD-1830, Washington, DC.
Au: Pls. Schroeder, D.C. and Lee, G.F., 1975, Potential transformation of chromium in natural
provide waters, Water Air Soil Pollut., vol. 4, pp. 355365.
chapter Seigneur, C. and Constantinou, E., 1995, Environ. Sci. Technol., vol. 29, pp. 222231.
title for Smith, L.A., Means, J.L., Chen, A., Alleman, B., Chapman, Ch.C., Tixier, J.S., Brauning,
reference S.E., Gavaskar, A.R., and Royer, M.D., 1995, Remedial options for metals-
contaminated sites, Contaminant Fate and Migration, CRS Press, Inc., Boca Raton,
FL, pp. 1734, Chap. 3.
Spokes, L.J. and Jickells, T.D., 1995, Speciation of metals in the atmosphere, in Chemical
Speciation in the Environment, Ure, A.M. and Davidson, C.M., Eds., Blackie
Academic and Professional, Glasgow, Scotland, pp. 137168.
Stollenwerk, K.G. and Grove, D.B., 1985, Adsorption and desorption of hexavalent
chromium in an alluvial aquifer near Telluride, Colorado, J. Environ. Qual.,
vol. 14, pp. 150155.
Stumm, W. and Morgan, J.J., 1981, Aquatic Chemistry, John Wiley and Sons, New York,
NY.
Thode, H.G. and Monster, J., 1965, in Fluids in Subsurface Environmentsa Sympo-
sium, American Association of Petroleum Geologists, Tulsa, OK, pp. 367377.
L1608_C05.fm Page 211 Friday, July 23, 2004 5:44 PM

The Transport and Fate of Cr(VI) in the Environment 211

Walkley, A. and Black, L.A., 1934, An examination of the Degtjareff method for
determining soil organic matter and a proposed modification of the chromic
acid titration method, Soil Sci., vol. 37, pp. 2938.
White, A.F. and Hochella, M.F., 1989, Electron transfer mechanisms associated with
the surface oxidation and dissolution of magnetite and ilmenite, Proc. 6th
Internat. Sympos. On Water-Rock Interaction, pp. 765768.
Wielinga, B., Mizuba, M., Hansel, C., and Fendorf, S., 2001, Iron promoted reduction
of chromate by dissimilatory iron-reducing bacteria, Environ. Sci. Technol., vol.
35, n. 3.
Wiley, K.G., 1983, Hydrogeological investigation examining the identification and
cleanup of chromium contaminated groundwater in Richland Township,
Kalamazoo County, Michigan. M.S. Thesis, Wright State University.
Wittbrodt, P.R. and Palmer, C.D., 1995, Reduction of Cr(VI) in the presence of excess
soil fulvic acid, Environ. Sci. Technol., vol. 29, pp. 255263.
Zachara, J.M., Girvin, D.C., Schmidt, R.L., and Resch, C.T., 1987, Chromate adsorption
on amorphous iron oxyhydroxide in presence of major ground water ions,
Environ. Sci. Technol., vol. 21, pp. 589594.
L1608_C05.fm Page 212 Friday, July 23, 2004 5:44 PM

Anda mungkin juga menyukai