Environmental Biochemistry of Chromium: M.E. Losi, C. Amrhein, and W.T. Frankenberger, Jr. T

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 31

Environmental Biochemistry of Chromium

M.E. Losi,* C. Amrhein,* and W.T. Frankenberger, Jr.*t

Contents
I. Introduction 91
II. Occurrence 92
A. Natural Sources 92
B. Industrial Sources 92
III. Environmental Chemistry .. .. .. ... ..... .. .. .... .. .. .. .. .. .. .. .. . 93
A. Oxidation States .. ............ ... .. .. ... .... ... .. .. ... .. .. .. . 93
B. Speciation 95
Cr(VI) 95
Cr(III) 96
C. Reactions and Behavior 97
IV. Analysis in Water and Soil Extracts 101
V. Nutrition and Toxicity............................................................. 103
A. Animals 103
B. Plants 104
C. Microorganisms 104
VI. Microbial Resistance/Transformations 105
VII. Bioremediation Techniques 106
A. Biosorption/Reduction 107
B. Gaseous Bioreduction 112
Summary 114
References 115

I. Introduction
Chromium (Cr) was first discovered in Siberian red lead ore (crocoite) in
1798 by the French chemist, Vauquelin. Named for the bright colors of its
compounds, Cr has since found an extremely wide variety of industrial uses
that exploit these colors, as well as various other characteristics of Cr, such
as strength, hardness and corrosion resistance of the metal, and the oxidiz-
ing capabilities of certain Cr species. Not surprisingly, large volumes of Cr
waste in various chemical forms are generated from industrial processes
and discharged into the environment. While Cr, in trace amounts, is essen-
tial for human life, exposure to some Cr compounds can pose a major
health risk to all forms of life. This review will discuss some of the environ-
mental impacts of chromium, including novel bioremediation techniques

*Department of Soil and Environmental Sciences, University of California, Riverside, CA


92521, U.S.A.
tCorresponding author.
© 1994 by Springer-Verlag New York, Inc.
Reviews ofEnvironmental Contamination and Toxicology, Vol. 136.

91
92 M. Losi, C. Amrhein, and W. Frankenberger, Jr.

currently under development for detoxifying Cr-contaminated water and


soils.

II. Occurrence
A. Natural Sources
Chromium is found in all phases of the environment, including air, water,
soil, and virtually all biota. It ranks 21st among elements in crustal abun-
dance (Krauskopf 1979). The average Cr concentration in the continental
crust has been reported to be 125 mg kg-I, with a commonly observed
range of 80-200 mg kg- I [National Academy of Sciences (NAS) 1974].
As might be expected, the Cr content of soils is largely dependent on
parent materials. Soil Cr concentrations normally range from 10 to 150 mg
kg-I, with an average of 40 mg kg- I (Bertine and Goldberg 1971). How-
ever, in soils derived from sepentinitic parent materials, Cr levels can reach
125,000 mg kg-lor more (Adriano 1986). Cr levels in various geologic
materials differ considerably, averaging from 20 to 35 mg kg-I in granitic
igneous rock, limestones, and sandstones to 220 mg kg- I in basaltic igneous
rock and 1,800 mg kg-I in ultramafic rock (Bowen 1966).
Freshwater Cr concentrations generally range from 0.1 to 6.0 p,g L-I ,
with an average of 1.0 p,g L-I, while values for seawater average 0.3 p,g
L-I and range from 0.2 to 50 p,g L-I (Bowen 1979). Freshwater Cr concen-
trations can actually be higher, depending on the soil Cr levels in the sur-
rounding watershed. Concentrations in the Mississippi River have been
reported to be as high as 84 p,g L-I, and levels have generally been shown
to be higher in the eastern U.S. than in the west (NAS 1974). In addition,
drainage water from irrigated agriculture in areas with elevated amounts of
soil Cr can have high Cr concentrations (as high as 800 p,g L-I), as observed
in various locations within the San Joaquin Valley, California (Gains 1988;
Deverel et al. 1984).
Chromium levels also vary widely within the atmosphere. Pacyna and
Nriagu (1988) cite studies that have measured background concentrations
from 5.0 x 10- 6 to 1.2 X 10- 3 p,g m- 3 in air samples from remote areas
such as Antarctica, Greenland, and the Norwegian Arctic. These levels are
considered to result from natural sources, such as windblown dust and
volcanic activity, and have been estimated to have deposition fluxes of 50
x 103 and 3.9 x 103 tons y-I, respectively (Pacyna 1986). By comparison,
Cr content of air samples collected over urban areas averages from 0.015
to 0.03 p,g m -3, ranging even higher over areas with steel industries (Nriagu
et al. 1988).

B. Industrial Sources
Because anthropogenic discharge of Cr into the environment results from in-
dustrial use, this discussion will focus on production and uses of chromium.
Environmental Chromium 93

Chromium is extracted from chromite ore, which has the formula [(Fe,
Mg)O(Cr, Al, Fe)203]' The largest deposits of chromite are located in South
Africa, the former USSR, the Philippines, southern Zimbabwe, and Turkey
(Mathews and Morning 1980). Although chromite deposits exist in the
U.S., it has become more economical to import the higher-grade foreign
ore, and thus there has been no chromite mining in the U.S. since 1961
(NAS 1974). Selected industrial uses of chromium are shown in Table 1.
The major users include metallurgical, chemical, and the refractory brick
industries (Langard 1980), which are largely responsible for anthropogenic
Cr emissions.
Chromium is a steel-gray metal which is hard, brittle, lustrous, corro-
sion-resistant, and highly polishable (Adriano 1986). One of several types
of ferrochromium or chromium metals are prepared from metallurgic-
grade chromite ore (~50OJo chromic oxide) and then alloyed with Fe, Ni,
or Co (NAS 1974). These alloys are used in the production of a wide variety
of steels, including stainless steel, austenite steel, and high-speed and high-
temperature steels, and in other nonferrous alloys. Chromium emissions
from steel-making processes are almost entirely in the form of particulates,
and although this type of pollution has been reduced significantly (through
the use of various emission control devices), it is still the major contributor
(by mass) to Cr pollution in the U. S. and Europe (Pacyna and Nriagu 1988).
The chemical industry uses lower-grade chromite ore (-45% chromic
oxide) to synthesize sodium chromate and sodium dichromate, from which
most other Cr chemicals are prepared (U.S. Environmental Protection
Agency [USEPA]/ORNL 1978). These chemicals have numerous industrial
applications, including products used in pigment manufacture, plating/
metal finishing, corrosion inhibition, organic synthesis, leather tanning,
and wood preservation. They commonly find their way into environmental
systems in a variety of ways, such as improper disposal of industrial wastes,
spills, or faulty storage containers.
Chromite ore used in production of refractory brick is a lower grade
(- 34% chromic oxide) than that used in the metallurgical and chemical
industries. Although considered to be a major source of Cr emissions, the
use of Cr in refractory processes is reportedly declining (NAS 1974). Other
important anthropogenic sources of Cr in the environment include fuel
combustion, cement production, and sewage sludge incineration/deposi-
tion (USEPA 1984).

III. Environmental Chemistry


A. Oxidation States
Chromium can exist in oxidation states ranging from 2- to 6+. However,
only 3 + and 6+ are normally found within the range of pH and redox
Table 1. Industrial Uses of Chromium Compounds

Name Formula Uses 'R


Cr(VI) compounds
Chromium oxide Cr20 3 Various metallurgical uses including production of ferroalloys used
to make stainless steels and nonferrous alloys or superalloys, such
as those used in jet engines
Barium chromate BaCr04 Pyrotechnics, high-temperature batteries ~
Cadmium chromate CdCr0 4 Catalysts, pigments b
~.
Cadmium dichromate CdCrP7: H p Metal finishing o
Calcium chromate CaCr04 Metal primers, corrosion inhibitors, high-temperature batteries
Copper dichromate CuCr207:2H20 Wood preservatives, catalysts
Magnesium chromate MgCr04:5H20 Corrosion inhibitor in gas turbines, refractories
i
.s'
Mercuric chromate HgCr04 Antifouling formulation ~
c.
Pyridine dichromate (CsH sNH)2Cr207 Photosensitizer in photoengraving, ceramics ~
Strontium chromate SrCr04 Corrosion-inhibiting pigment, plating additive
Cr(lll) compounds ig-
Chromic acetate Cr(OCOCH3)3 : xH 20 Printing and dyeing textiles
O'
Chromic chloride CrCl 3 Chromatizing, Cr metal, organochromium compounds Ja
Chromic chloride, hydrated CrCI 3 : 6H 20 Mordant, tanning, Cr complexes ~
....:-'
Chromic fluoroborate Cr(BF4)3 Chromium plating, catalysts
Chromic napthenate Not definite Textile preservative
Chromic phosphate CrP0 4 Pigments, phosphate coating, wash primers
Copper chromite CuCr204 Catalysts, especially for automobile exhaust
Magnesium chromite MgCr20 4 Refractory

Source: Hartford; see Nriagu (1988).


Environmental Chromium 95

1.2 ...
...
0.8 .......... ...
I .......... ...
I .......... ...

-->
cr3+ I .......... ...
0.4
I
~
I CrOH2+
W I
0 ... I
. . . . . . . l.... ... Cr(OHhO

.......... ...
-0.4 .......... ... Cr{0H)4-
.......... ...
... ..........
-0.8
0 2 4 6 8 10 12 14
pH
Fig. 1. Stability diagram showing aqueous speciation of chromium at various Eh
and pH values (Rai et al. 1989).

potentials common in environmental systems, and of these, Cr3+ is gener-


ally considered to be the more stable form (Mertz 1969).

B. Speciation

Cr(VI). Figure 1 depicts a generalized Eh-pH speciation diagram for


aqueous chromium. Hexavalent Cr is a strong oxidizer and, as a result,
exists only in oxygenated species that are very soluble and pH-dependent
according to the following equilibria (Nieboer and Jusys 1988):
H2Cr04 <=> H+ + HCr04-Kal = 10°·6 (1)

HCr0 4- <=> H+ + CrO/- Ka2 = 10- 5.9 • (2)

From equations (1) and (2), we observe that at very low pH values (near 0),
H 2Cr04 is the dominant species, while between the values of 0 and 5.9,
HCr04- dominates and, at or above pH 6, CrO/- prevails. Since the pH
in environmental matrices would generally not be expected to fall near zero;
only HCr04- and CrO/- should be present in natural systems. Also, at
concentrations greater than O.OIM (520 mg L -I), dimerization of the chro-
mate ion occurs (Beas and Mesmer 1976), yielding the dichromate ion
[equations (3) and (4); Whitten and Gailey 1984J:
96 M. Losi, C. Amrhein, and W. Frankenberger, Jr.

(3)

(4)

This reaction is pH-sensitive as well, with dichromate being favored at a


lower pH. Solving equation (4) using values for [CrO/-] and pH that
would likely be encountered in Cr-contaminated groundwater (5.2 mg L- 1
and 7.0), the ratio of dichromate to chromate would be 0.04. Thus, Cr(VI)
chemistry in environmental systems is largely confined to that of the chro-
mate ion.

Cr(III). Trivalent Cr is the more stable form (Cary 1982; NAS 1974), and
its chemistry is more complex than that of hexavalent Cr. A plethora of
data exist regarding complexation and speciation of Cr(lll). Herein, we
will briefly discuss some of the environmentally relevant facts regarding
Cr(lll) speciation.
Because it has relatively less affinity for O2 , Cr(III) is known to form
numerous complexes, with both organic and inorganic ligands. However,
Rai et al. (1986) report that among the ligands OH-, SO/-, N03-, and
CO/- (evaluated at concentrations commonly encountered in natural envi-
ronments), only OH- was found to significantly complex Cr(lll). Within
environmental systems, major Cr(III) hydroxy species include Cr(OH)2+,
Cr(OH)3° and Cr(OH)4- (Rai and Zachara 1988). The solubility of
Cr(OH)3' the solid species expected to be most prevalent at pH levels en-
countered in J;lature, is known to be very low [see equation (7)].
Certain compounds (notably organics) have been shown to form com-
plexes with Cr(lll), preventing its precipitation at lower pH values. James
and Bartlett (l983a) determined in laboratory studies that citric and fulvic
acids effectively complexed Cr(lll), preventing precipitation up to approxi-
mately pH 7.5. Meanwhile, experimentation in our laboratory indicated
that significant conjugation of Cr(lll) with organic compounds found in
cattle manure did not occur.
In one such study, chromate-spiked water (pH 7-8) was passed through
pots containing soil and cattle manure, and both total [Cr] and [Cr(Vl)] in
the effluent were measured (Losi et al. 1994a). We found that reduction of
Cr(VI) readily occurred, followed by precipitation, and, hence, removal of
Cr from the water [adsorption was shown to be negligible for this soil (Losi
et al. 1994b)]. We can hypothesize that if the reduced species (Cr3+) formed
insoluble complexes with organic compounds, total [Cr] in the efflUent
would be elevated over [Cr(VI)] by some factor reflecting the extent of
Cr3+ solubilization. This was shown to occur to some degree, since effluent
total [Cr] was slightly elevated over [Cr(VI)] over the course of the experi-
ment. These findings are reported in Fig. 2, which shows the relationship
Environmental Chromium 97

2500

2000
...
..j
~ 1500
=l.

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

;;.- 1000
'i:'
:::.
500

500 1000 1500 2000 2500


TOTAL fer], Jlg L·t

Fig. 2. Results from a biofilter experiment in which Cr(VI)-spiked water was passed
through a soil amended with dried cattle manure. Both total [Cr] and [Cr(VI)] were
measured to monitor the extent of solubilized Cr(III) due to organic complexation.
The relationship suggests that some complexation did ocur, but that the majority of
Cr in the effluent was in the hexavalent state (Losi et al. 1994a).

between [Cr(VI)] and total [Cr] in the effluent, giving a measure of organi-
cally complexed/solubilized Cr3+.
Thus, speciation of Cr(VI) and Cr(lll) will generally depend on a variety
of environmental parameters, including pH, concentration, and available
ligands. In most natural systems, hexavalent Cr will be present as CrO/-
and major trivalent Cr species may include hydroxides and various organic
complexes. The behavior of both Cr(VI) and Cr(lll) and the interconver-
sion between these two forms must be understood when considering the
environmental properties of Cr.

C. Reactions and Behavior


Chromium is known to undergo various chemical and biological reactions
in natural systems that govern speciation and, in turn, environmental be-
havior. Important reactions include oxidation/reduction, precipitation/dis-
solution, and adsorption/desorption. Figure 3 illustrates the possible fates
of Cr in soil/water systems.
Both oxidation of Cr(lll) and reduction of Cr(VI) can occur in geologic
and aquatic environments. Hexavalent Cr is a strong oxidizing agent and is
readily reduced in the presence of appropriate electron donors, as shown in
equation (5):
(5)
98 M. Losi, C. Amrhein, and W. Frankenberger, If.

Kinetics of Redox
}=========,.:;I Transformations I~...;:::::::=======l

Predicted Cr
Concentrations

Fig. 3. Possible reactions of Chromium applied to a soil/water system (Rai et aI.


1989).
Environmental Chromium 99

Rai and Zachara (1988) reported that Fe(II) actively reduces Cr(VI), the
reaction rate being dependent on the solubility of the Fe compound. The
presence of organic matter has been shown to enhance Cr(VI) reduction
(Cary et al. 1977b; Bartlett and Kimble 1976b). Experiments in our labora-
tory have confirmed the ability of an organic-amended soil to effectively
reduce Cr(VI) at near-neutral pH, while the same soil reduced much less
Cr(VI) when left unamended (Losi et al. 1993a). From later experiments,
we concluded that chemical and biological processes each accounted for
roughly one-half of the observed reduction (Losi et al. 1993b). Low oxygen
status was also reported to be of importance, giving greater reduction rates
(Bloomfield and Pruden 1980). This observation was later confirmed in
our studies (Losi et al. 1993a,b).
Direct and indirect microbially mediated bioreduction of hexavalent Cr
has been observed. Anaerobic bacterial strains with accelerated Cr(VI)-
reducing capabilities have been isolated from chromate-contaminated water
and sludge (pugH et al. 1990; Komori et al. 1990a; Kvasnikov et al. 1985;
Romanenko and Koren'kov 1977). It is evident that anaerobic bacteria can
effect preferential bioreduction of chromates under reduced conditions.
The existence of bacterial isolates capable of aerobic Cr(VI) bioreduction
(presumably as a detoxification mechanism) has also recently been demon-
strated (Horitsu et al. 1987; Bopp and Ehrlich 1988; Ishibashi et al. 1990,
Losi et al. 1993b). The process is believed to involve a specific, plasmid-
related enzyme, and reaction rates vary in response to available carbon
sources. Microbially mediated reduction of Cr(VI) will be discussed in more
detail later in this review.
Whereas reduction of Cr(VI) is likely to occur in environmental systems
where appropriate electron donors are present, oxidation of Cr(III) appears
to be less likely. Initial studies convincingly showed that, in most cases,
oxidation of Cr(III) does not occur in soils, regardless of the conditions
(Bartlett and Kimble 1976a). However, it has since been determined that
some quantity of Cr(lll) can be oxidized to the hexavalent form in the
presence of Mn4+. Mn4 + serves as the oxidizing agent and is reduced to
Mn2+, as shown in equation (6):
2Cr3+ + 3Mn02 + 2H 20 ~ 2CrO/+ + 3Mn2+ + 4H+. (6)
Bartlett and James (1979) showed that aerobic, fresh field soils can oxi-
dize a quantity of Cr(lll) proportional to the amount of available Mn4+.
The addition of citric acid was later determined to increase the portion
of Cr(VI) formed (James and Bartlett 1983b). Subsequent research has
supported these findings (Amacher and Baker 1982; Eary and Rai 1987).
These conditions, however, are fairly specific, and few cases of Cr(lll)
oxidation are reported in the literature. Furthermore, we analyzed soil and
surface water samples from the New Idria formation in central California,
which is highly serpentinitic, and found that although soil total Cr levels
averaged 770 mg kg-I, there was no detectable Cr(VI) in the water samples
100 M. Losi, C. Amrhein, and W. Frankenberger, Jr.

or in the fresh soils when extracted with phosphate buffer (Losi and Frank-
enberger, unpublished). Although oxidation of Cr(lll) can occur, reduction
of Cr(VI) apparently occurs much more readily (Adriano 1986).
Trivalent Cr precipitates almost completely as Cr(OH)3 [equation (7);
Sposito and Mattigod 1980], often in conjunction with Fe, at pH values
from 5.5 to 12.0, keeping aqueous concentrations generally low (Rai et al.
1989; McGrath and Smith 1990):
Cr3+ + 3H20 <* Cr(OHMs) + 3H+; K eq -- 10- 12 (7)

Unless solubilized by significant organic complexation, Cr(lll) can be ex-


pected to remain in a solid form in most environmental systems and can be
physically immobilized within the soil matrix or be subject to sedimentation
in aqueous environments. On the other hand, the majority of Cr(VI) com-
pounds are much more soluble under environmental conditions, with the
possible exception of BaCr04 (Rai et al. 1989), and thus will most likely
remain in solution under aqueous conditions. A study was conducted in
our laboratory (discussed earlier) in which Cr(VI)-spiked water (pH = 7-
8) was passed through pots containing soil amended with organic material,
favoring reduction of a significant portion of the Cr(VI) to Cr(lll). This
experiment reinforced the view of elevated Cr(VI) solubility/mobility ver-
sus immobilization of Cr(III) (based on precipitation reactions) in neutral
to basic soil/water systems, since virtually all Cr passing through the pots
was in the hexavalent state.
At pH values below 5.5, Cr(lll) will be present in the cationic form and
may be adsorbed on cation exchange sites. Because of cation exchange
reactions, Cr3+ (the cationic form) can be considered relatively immobile in
most soil/water systems, especially those with appreciable amounts of clay
minerals.
Conversely, in systems where levels of organic matter and/or inorganic
electron donors are low or manganese oxide levels are high, hexavalent
chromium may be stable and soluble over a broad pH range. Mobility of
this Cr species will depend on pH, the presence of specific adsorption sites,
and competing anions. Chromates are weakly adsorbed to hydroxyl groups
on Fe/Al oxides and layer silicates (Rai et al. 1989). Chromate adsorption
increases with decreasing pH as OH groups become protonated. Griffin et
al. (1977) demonstrated that chromate adsorption was highly dependent on
pH and type of clay minerals, while Cary et al. (1977b) reported that migra-
tion of Cr(VI) in soil columns was greater at higher pH. Chromate adsorp-
tion by soils can also be influenced by the presence of competing anions.
Bartlett and Kimble (1976b) found that orthophosphate in the equilibrating
solution can totally inhibit Cr(VI) adsorption. Sulfate and carbonate have
also been shown to compete for CrO/- adsorption sites (Zachara et al.
1987,1989).
Although exceptions exist (as noted), there are several general, recurring
themes throughout a large body of literature regarding reactions and behav-
Environmental Chromium 101

Table 2. Methods for Chromium Analysis

Atomic absorption and emission spectrometry


Inductively coupled plasma
X-ray fluorescence
X-ray photoelectron spectrometry
Proton-induced x-ray emission (valence detection)
Neutron activation
Mass spectrometry
Ion chromatography
UV-visible spectrometry
Electrochemical
Chromatography (high-performance liquid chromatogra-
phy, gas chromatography)
Radioactivity, SICr detection
Infrared and raman spectrometry
Colorimetry·
Activation analysis b

Source: "Bartlett (1991); bLove (1983); all others, Fundamen-


tal Reviews; see Shupak (1991).

ior of Cr in the environment. The first is that reduction of Cr(VI) to Cr(IlI)


is much more frequently observed than oxidation of Cr(IlI) to Cr(VI) and,
therefore, Cr(III) is the more stable form in most natural systems. Sec-
ondly, Cr(IlI) is less mobile than Cr(VI) in most soil/water systems due to
the relative insolubility of Cr(IlI) at relevant (> 5) pH values. These factors
are of chief importance in assessing potential environmental hazards and
remediation strategies for ecosystems with high levels of natural or anthro-
pogenic chromium.

IV. Analysis in Water and Soil Extracts


Because of its redox sensitivity, a number of parameters need to be ad-
dressed in measuring the Cr content of soil and water samples if speciation
is a goal of the analysis. These must include sample treatment and storage,
extraction and preparation (in the case of soils), and the actual analytical
technique itself. Gochfeld (1991) alleged that difficulties still exist in distin-
guishing Cr6 + from Cr3+ and that substantial disparities may be observed in
data generated by different laboratories analyzing the same environmental
sample.
A partial list of techniques for detecting and quantifying Cr in a variety
of matrices is given in Table 2. While quantification of Cr in certain biologi-
cal materials may require extremely sensitive detection methodology, analy-
sis of Cr in environmental samples need not always be so rigorous [EPA
National Drinking Water Standard is 50 IJ-g L -1, and soils accepting dis-
102 M. Losi, C. Amrhein, and W. Frankenberger, Jr.

Table 3. Detection Limits for Methods Commonly Used in the


Analysis of Chromium in Water and Soil Extracts

Approximate
detection
limit
Method Species detected (;Lg L -I)

Flame atomic absorption Total Cr 500a


spectrometry
Graphite furnace atomic Total Cr I"
absorption spectrometry
Inductively coupled plasma Total Cr 6_10"
optical emission spectrometry
Inductively coupled plasma Total Cr OS
mass spectrometry
Colorimetric Cr(VI) 50b
(diphenylcarbazide)
High-performance liquid Cr(VI) 92c
chromatography
(single-column
ion-chromatography)

"Detection limits will vary, depending on instrument and settings (consult


manual).
bBartlett, see Gochfeld (1991).
"Mehra and Frankenberger (1989).

posal of waste or sludge can have up to 3,000 mg Cr kg- 1 (USEPA 1976;


1993)]. Presumably because of this, as well as cost restrictions and equip-
ment limitations, the major methods used to analyze soils/sediments and
water are atomic absorption spectrometry (AAS) and inductively coupled
plasma (ICP) spectrometry for total Cr, and colorimetry (diphenylcarbaz-
ide method) and high-performance liquid chromatography (HPLC) for
quantifying Cr6 +. The difference between the amounts of total Cr and Cr6 +
is assumed to be Cr3+. Approximate detection limits for analytical methods
commonly used to measure Cr in water and soil extracts are reported in
Table 3. These analyses will usually be adequate for environmental samples
in most cases. It is important to note that all of these methods are subject
to interferences, which must be evaluated through the use of appropriate
matrix spikes.
In analyzing water, sample preparation is of minor consequence when
determining total Cr and usually amounts to filtration and acidification
(USEPA 1986). Because the redox state may influence signal response in
AAS analyses (Reisenauer 1982), oxidation of Cr in the sample to the
hexavalent state may be recommended prior to aspiration (Perkin-Elmer
Environmental Chromium 103

1982). In ICP analyses, however, this is not necessary since all Cr is reduced
to the trivalent state during the analysis. When determining Cr6 + levels in
water, samples must be stored at 4 °C and analyzed within 24 h (USEPA
1986) due to redox sensitivity. Acidification is not recommended since re-
duction to Cr3+ is likely to occur at low pH [see equation (5)].
Analysis of chromium in soils and sediments represents a more challeng-
ing problem. It is important that samples be field-moist, sieved (4 mm),
well mixed, and stored at 4 °C (Bartlett 1991). Determination of total soil
Cr is relatively straightforward and can be carried out by means of AAS or
ICP following complete digestion, such as the EPA-recommended hot
HNO/HCI and H 20 2 method (USEPA 1986).
Hexavalent Cr may be determined by either colorimetric or HPLC proce-
dures after extraction with 10 mM K2HP0 4/KH 2P04 , pH 7.2 (James and
Bartlett 1983c), to give exchangeable Cr6 + (phosphate-extractable), or with
water, to give water-soluble Cr6 +. Filtration of the extract (0.45 #Lm pore
size) at pH values above 5.5 will remove Cr(lll) precipitates, and analysis
by ICP or AAS should give a reasonable quantification of [Cr6 +]. Several
additional procedures relevant to Cr analysis in soils, including total and
available Cr6 + -reducing capacity by soils, available soil Cr3+, total oxidiz-
able Cr3+, and others, are described by Bartlett (1991).

V. Nutrition and Toxicity


In light of the above observations, it appears probable that Cr(VI) will
remain highly mobile in soil/water systems, especially under conditions
where the clay content of soils is low and pH is neutral to alkaline. Evidence
for this is that Cr(VI) has been detected in drinking water and groundwater
under waste-disposal basins and landfills in many instances (Deutsch 1972;
Griffin et al. 1977; Khasim et al. 1989; Losi et al. 1994a). Thus, contact
with chromates by many terrestrial and aquatic organisms, including hu-
mans, is possible.

A. Animals
Chromium, in trace amounts, is an essential component of human and
animal nutrition (Mertz 1969; Jeejeebhoy et al. 1977). It is most notably
associated with glucose metabolism (Mertz 1969) and was later shown to be
an integral component of the "glucose tolerance factor" (GTF), which has
been synthesized by Mertz (1977). Chromium is also known to be of impor-
tance in fat metabolism in animals (Anderson 1989).
It is universally observed that trivalent Cr is the nutritionally useful
form, while the hexavalent form is toxic and mutagenic. The biotoxicity of
chromate is largely a function of its ability to cross biological membranes
and its powerful oxidizing capabilities (NAS 1974). Humans can absorb
Cr(VI) compounds through inhalation, dermal contact, and ingestion. Ad-
104 M. Losi, C. Amrhein, and W. Frankenberger, Jr.

verse effects due to excessive Cr exposure include ulceration and perfora-


tion of the nasal septum, respiratory cancer, skin ulceration, contact der-
matitis (NAS 1974), and in the event of ingestion, kidney damage (Gad
1989) and damage to various proteins and nucleic acids, leading to mutation
and carcinogenesis (Levis and Bianchi 1982).

B. Plants
Although shown to produce stimulatory effects in plants at appropriate
concentrations (Pratt 1966), Huffman and Alloway (1973) demonstrated
conclusively that Cr is not an essential component of plant nutrition. It has
been proposed that the growth benefits result from limited substitution of
chromium for molybdenum (Warington 1946), but these effects, in general,
are reportedly minimal and difficult to explain (Huffman 1973).
Chromium toxicity in plants is rare in natural systems, probably because
the majority of naturally occurring Cr is in the trivalent state. Elevated
levels, however, have been reported in plants grown in Cr-contaminated
soils (Khasim et al. 1989). The majority of data involving the effects of Cr
on plants have been collected in laboratory studies. The NAS (1974) re-
ported that toxic effects of CrO/- appear in plants at concentrations from
18 to 34 mg kg-I dry weight (d.wt.), while Skeffington et al. (1976) ob-
served that barley grown for 9 d in nutrient solution spiked with 0.5 mg
Cr(VI) L-I had a shoot concentration of 43 mg Cr kg-I (d.wt.), and no
reduction in yield was noted. The latter study showed, however, that when
grown in 5.0 mg Cr(VI) L-I-spiked solution, barley yields dropped by 750/0.
Also of note is that Cr concentrations of root material have generally been
found to exceed those of the shoots by approximately one order of magni-
tude (Skeffington et al. 1976; Cary et al. 1977a,b), which was supported in
our study (Losi et al. 1993a).

C. Microorganisms
While considered essential for the growth of all animals, chromium, along
with nickel, molybdenum and tin, is required only by certain microorgan-
isms for specific metabolic processes. The major functions are related to
glucose metabolism and enzyme stimulation, but beyond that, little is
known (Hughes and Poole 1989).
The toxic and mutagenic effects of chromates on microorganisms are
well documented. Ross et al. (1981) found that 10-12 mg Cr(VI) L-I was
inhibitory to most soil bacteria in liquid media, while, at this level Cr(lll)
had no effect. Ajmal et al. (1984) reported that chrome-electroplating waste
was toxic to saprophytic and nitrifying bacteria, with toxicity increasing
directly with the Cr(VI) content of the waste. In other studies, organisms
exhibiting sensitivity to Cr compounds at environmentally relevant concen-
trations include various tomato pathogenic fungi (Naguib et al. 1984),
mixed bacterial populations (Lester et al. 1979), freshwater algae (Bharti et
Environmental Chromium 105

al. 1979), and estuarine phytoplankton (Frey et al. 1983). These toxic ef-
fects are attributed to altered genetic material and altered metabolic and
physiological reactions (Coleman 1988).

VI. Microbial Resistance/Transformations


Because heavy metals occur naturally at high concentrations in various
environmental systems, resistance/tolerance mechanisms have been evolv-
ing within microbial communities since the advent of life. Consequently,
resistance to virtually all of the toxic heavy metals can be found within a
vast array of terrestrial and aquatic microorganisms, with the various resist-
ance mechanisms, including transformation of the toxic species (e.g., meth-
ylation, demethylation, oxidation, or reduction), and changes in uptake
and/or transport of the metal (Williams and Silver 1984). Chromium-
tolerant microbes have been isolated from water (Simon-Pujol et al. 1979),
sediments (Luli et al. 1983), and soil (Losi and Frankenberger 1994).
Resistance is often reported to be plasmid-related. Summers and Jacoby
(1977) reported that Cr resistance in Pseudomonas aeruginosa was plasmid-
determined, but that the actual biochemical mechanism was unknown. Sim-
ilarly, Bopp et al. (1983) isolated a chromate-resistant strain of Pseudomo-
nas fluorescens and determined that the resistance was plasmid-conferred.
Working with the same strain, Bopp and Ehrlich (1988) later reported that
Cr(VI) reduction was associated with a constitutive, membrane-associated
enzyme that is able to mediate transfer of electrons from nicotinamide
adenine dinucleotide (NADH) to chromate. It has since been shown that
both bacterial reduction and altered uptake appear to function as mecha-
nisms of resistance to chromates.
Bioreduction of chromate can occur directly, as a result of microbial
metabolism, or indirectly, mediated by a bacterial metabolite (such as H2S).
It can occur under both aerobic (Bopp and Ehrlich 1988; Horitsu et al.
1987; Shimada and Matsushima 1983) and anaerobic (Komori et al. 1990c;
Romanenko and Koren'kov 1977) conditions, and, as mentioned, direct
reduction is enzymatic in nature (Bopp and Ehrlich 1988; Horitsu et al.
1987).
Direct Cr(VI)-bioreduction can be carried out as a detoxification mecha-
nism (Wood and Wang 1983) or can possibly result from chromate being
the terminal electron acceptor in the oxidation of organic matter when the
chromate-oxygen ratio is favorable. Various energy sources and environ-
mental conditions have been evaluated as to their effect on chromate biore-
duction. Shimada and Matsushima (1983) studied a chromate-resistant bac-
terium, Pseudomonas K-21, and they found it to aerobically reduce Cr(VI)
while growing on glucose. Komori et al. (1989) reported that Enterobacter
cloacae strain HOI reduced Cr(VI) anaerobically while growing on acetate,
ethanol, malate, succinate, and glycerol, but that its reduction efficiency
was decreased significantly while the organism was growing on glucose,
106 M. Losi, C. Amrhein, and W. Frankenberger, Jr.

citrate, pyruvate, and lactate. Optimal pH and temperature were reported


at 7.0-7.8 and 30-37 °C, respectively. Glucose was determined to have no
effect on the induction of the Cr-reducing enzyme present in Pseudomonas
ambigua G-l, but it protected the enzyme from inactivation upon dialysis
(Horitsu et al. 1987). Apel and Turick (1991) showed significant differences
in the bioreduction abilities of three Cr(Vl}-reducing bacterial strains with
glucose as the carbon source at various Cr concentrations. Moreover, they
found that Cr(VI) reduction was well correlated with the exponential
growth phase, which differed between the organisms over time and was
a function of chromate concentration. Evidently, optimal conditions and
reduction efficiencies can be expected to vary considerably among organ-
isms and matrices. No instances of microbial oxidation of Cr(lll) were
found in the literature.
Decreased uptake was also identified as a plasmid-conferred Cr-
resistance mechanism in bacteria. Horitsu et al. (1983) discovered that
Cr(VI) uptake was about six times less in Cr(VI)-tolerant Pseudomonas
ambigua G-l than in its chromate-sensitive mutant. These findings were
supported by Ohtake et al. (1987), who observed that decreased uptake was
responsible for chromate resistance in Pseudomonas fluorescens. Ohtake et
al. (1987) also noted that chromate was transported into the cell mainly by
the sulfate-active transport system.
Interestingly, bacterial chromate resistance has also been correlated with
increased Cr uptake. Marquez et al. (1982) showed that a chromate-tolerant
Pseudomonas strain accumulated Cr, and they proposed further research
to study biosorption as a possible Cr-recovery technique. Coleman and
Paran (1983) also cultured a Cr-tolerant bacterial isolate that effectively
accumulated Cr. This phenomenon could be explained by what is known as
"intercellular traps for ion removal" (Wood and Wang 1983), or inclusion
bodies. In this process, metals reaching toxic levels are temporarily removed
from solution and incorporated into cellular material for later expulsion.
Mechanisms such as reduction of Cr(VI) to the less toxic Cr(lll) and Cr
accumulation (formation of inclusion bodies) generally serve to detoxify
the microbe's immediate environment. It seems logical that these processes
might also be utilized to reduce or eliminate the toxic effects of chromates
on a larger scale.

VII. Bioremediation Techniques


Traditional techniques for remediating chromate-contaminated water in-
volve reduction of Cr(VI) to Cr(III) by chemical (usually Fe2+) or electro-
chemical means at pH> 5, followed by precipitation and finally filtration
or sedimentation (Eary and Rai 1988). These processes can be extremely
reagent- and/or energy-intensive, as suggested by the fact that present costs
for disposal of one drum containing chromium waste can run as high as
$2,000 (Studt 1993). The discovery of microorganisms that preferentially
Environmental Chromium 107

reduce Cr(VI) has led to applications in the bioremediation field, which are
potentially more cost-effective than traditional methods in the treatment of
wastewater containing Cr.

A. Biosorption/Reduction
Russian researchers first proposed the use of Cr(VI)-reducing bacterial iso-
lates in the removal of chromates from various industrial effluents (Roma-
nenko et al. 1976; Romanenko and Koren'kov 1977). Since then, various
reduction parameters have been evaluated for a diverse group of microor-
ganisms with accelerated Cr(VI)-reducing capabilities with the prospect of
developing commercially viable bioremediation techniques exploiting these
organisms (Table 4).
The major focus of this research is aimed toward the development of
bioreactors, which basically consist of a reduction phase, with Cr(VI)-
reducing bacteria immobilized on inert matrices within the reactor, fol-
lowed by a settling or filtration phase to remove Cr(III) precipitates. Chro-
mate-contaminated water is pumped into the reactor and supplemented
with various carbon sources and nutrient additives; the Cr(VI) is then re-
duced, precipitated, and removed. The main advantages of this system are
considered to be its cost (reagent/energy)-effectiveness and the fact that no
reagents containing other heavy metals (reductants) must be added, thus
yielding a "cleaner" floc. A disadvantage is that batch experiments in the
laboratory have demonstrated that the lowest achievable effluent Cr con-
centrations are probably around 1 mg L -1 (Apel W, see Mattison 1992),
which is considerably higher than the national EPA National Drinking
Water Standard of 0.05 mg L -1. Also, reaction rates may be slow since
CrO/- must diffuse into direct contact with the cells, which maximizes its
toxicity effects as well. Since optimum bioreduction has been correlated
with the exponential growth phase of the bacteria (Apel and Turick 1991),
conditions within the reactor must be adjusted to sustain high growth rates.
Bioreactors such as these are probably best suited to treating wastewater
prior to discharge rather than in situ applications, because reaction time
and flow rates are expected to be relatively slow, and pumping of ground-
water can significantly raise the cost.
We recently investigated a land application method (biofilter) for reme-
diating Cr(VI)-contaminated groundwater (Losi et al. 1994a). The method
consists of passing the contaminated water through an organic matter-
enriched soil where reduction, precipitation, and immobilization would
take place. A glasshouse experiment was conducted to determine the effects
of organic matter loading (cattle manure), cropping (alfalfa), and irrigation
management on Cr(VI) immobilization.
In the experiment, we were able to treat relatively large volumes of water
spiked with 1.0 mg Cr(VI) L -1, yielding outflow Cr levels consistently be-
low 0.02 mg L -1 (Losi et al. 1993a). We proposed that reduction followed
Table 4. Bioremediation of Chromium
0
Mechanism Organism Description and effectiveness Source 00
-
Bioreduction Pseudomonas ambigua G-l Cr(VI) concentration was lowered from 150 to Horitsu et al. (1987)
- 35 mg L-lover 36 hr in liquid media.
~hown to be enzymatic.

Bioreduction Pseudomonas K-21 Reduction rates were proportional to the glu- Shimada and Matsushima ::
cose concentration in the medium. [Cr(VI)] (1983)
t""
dropped from 200 to 100 ppm in the presence 0
::!l.
of 0.2070 glucose and to - 0 ppm at 1% glu- (J
cose after 40 hr.
Bioreduction Pseudomonas fluorescens LB300 Reduced about one-half of CrO/- in media Bopp and Ehrlich (1988) ~
spiked with 40 mg L -1 Cr(VI) over 48 hr. t:r

Biosorption Oscillatoria sp. Algal cultures removed 20% of Cr from water Filip et al. (1979) pi'
"'
III
spiked at levels from 1-20 mg L - 1 ::s
0-
Biosorption Arthrobacter sp. Accumulated Cr with increasing concentra- Coleman and Paran (1983) :e
Agrobacter sp. tion gradient of [Cr(VI)] up 400 mg L -1 (Ar- "T1
throbacter sp.) and 100 mg L -1 (Agrobacter
...
§
sp.) :>I"
::s
"'
Bioreduction/ Pseudomonas aeruginosa S128 Removed from -15-50070 of Cr from liquid Marquez et al. (1982) c'
OQ
biosorption media spiked with 1,000 mg L -1 Cr(VI) over
..."'
72 hr. Glucose enhanced the removal rate. ."'..
....
:-'
Bioreduction Sulfate-reducing bacteria [Cr(VI)] in water was lowered from 11 to DiFilippi and Lupton (1992)
< 0.21 mg L -1 with addition of sulfate and ac-
etate in an anaerobic bioreactor that produced
H 2S, subsequently reducing the Cr(VI).
Bioreduction Enterobacter cloacae HOI Cr(VI) at 182 mg L -I in wastewater was com- Ohtake et al. (1990)
pletely reduced at high cell densities with ap-
propriate nutrient additions.
Bioreduction Enterobacter cloacae HOI 90070 of Cr was removed from water with an Komori et al. (199Ob)
initial [Cr(VI)] of 208 mg L -I. Cultures are
contained within dialysis tubes submerged in
the contaminated water. Chromate diffused
into the tube, was reduced, precipitated and
thus was unable to diffuse out.
Biosorption/ Sulfate-reducing bacteria 100% Cr removal was achieved in water with Roda and Smirnova (1989)
bioreduction [Cr(VI)] up 150 mg L -1.
tIi
Biosorption/ Chlamydomonas sp. An average of 30% Cr removed from water Ayoub and Sayigh (1987) t:l
~.
bioreduction with initial concentrations of 0.2 mg L -I in an o
aerobic bioreactor system.
Bioreduction Pseudomonas putida PRS2000 The isolate reduced the [Cr(VI)] in liquid me- Ishibashi et al. (1990) ie.
dia from 20 JLM to about 7 JLM over 90 hr. Re-
(')
duction was shown to be associated with solu- g-
ble protein in cells. Process was unaffected by o
SO/- or N03 - • e.r::
Chlorella vulgaris; Maximum adsorption was seen at pH 2 and at Zumryie et al. (1990)
S
Biosorption
Zoogloea ramigera 25-50 °C for both organisms. Adsorption in-
creased with [Cr(VI)] up to 200 mg L -1 (C vul-
garis) and 75 mg L -1 (Z ramigera).
Bioreduction Consortium of bacteria Reduced all CrO/- in liquid media at 30 mg Eliseeva et al. (1991)
L -I after 48 hr, and less as [Cr(VI)] was in-
creased.
(continued)

§
Table 4. (Continued)
o
-
Mechanism Organism Description and effectiveness Source
Bioreduction Aeromonas dechromatica Effective reduction observed with appropriate Kvasnikov et aI. (1986, 1987)
carbon sources (propyl alcohols and ethyl ace-
tate), Cu(OH}z, Zno and Zn(OHh promoted
bioreduction, Cd had no effect, and Ni, ~
Fe(III), and Cu inhibited this transformation. t""
o
~.
Bioreduction Unidentified pure bacterial cultures Reduced 40-600/0 of Cr(VI) added to liquid Losi et aI. (1994a)
media at 10 and 25 mg L -I; approximately ()
50% at 60 mg L -I , and 10% at 120 mg L -I
over 18 d with no additional nutrient supple- ~
ment. [
.?
Bioreduction/ Indigenous soil organisms Removed up to 980/0 of Cr(VI) added in irriga- Losi et aI. (1994a) ~
Immobilization tion water with a [Cr(VI)] of 1,000 p.g L -I. Im- Co

in surface soils portant factors shown to be organic matter ~


(Biofilter) loadings, O2 status, and residence time in bio- ~
active zone. ~
~
go
n...
~....
:"
Environmental Chromium 111

Water
... Bacteria

0 Cr(Vl)
@ Cr(lll)

• Cr(OH)3

Fig. 4. Direct, enzymatic reduction of Cr(VI) followed by chemical precipitation. 1.


Biological transformation of Cr(VI)-contact between bacterial cell and Cr(VI)
results in enzymatic reduction to Cr(lll) (this reaction also occurs chemically). 2.
Nonbiological precipitation of Cr(III)-Cr(III) reacts with OH- to form insoluble/
immobile Cr(OH)3'

by precipitation was the dominant removal mechanism and demonstrated


that lower [02] favored this reaction. Subsequent experiments showed that
indigenous microorganisms, in conjunction with a readily available carbon
source (cattle manure), were largely responsible for the high removal rates.
Figure 4 illustrates' the steps involved in this process. In light of this ap-
proach, it appears that the reactive constituents in soils may be useful (and
economical) in optimizing bioremedial systems designed to treat Cr(VI)-
contaminated water. Perhaps more importantly, our study indicates that
biofilters show promise for remediation of Cr-contaminated aquifers. This
method is potentially very cost-effective, requiring only animal waste as an
additive, and has an additional advantage of producing alfalfa with slightly
elevated Cr levels, which may be useful for supplementing the Cr content
of animal feed in areas where soil Cr levels are low (Bartlett and James
1988). The potential to rapidly and efficiently treat enormous quantities of
contaminated water could be expected to offset the cost of pumping the
aquifer.
Another application of direct reduction was demonstrated by Komori et
al. (1990b). In this study, anaerobic Cr-reducing bacterial cultures were
contained within dialysis tubes and submerged in contaminated water.
Chromate diffusing into the tubes was reduced and precipitated, and thus
was unable to diffuse out. Laboratory studies using this system showed
that 900/0 of the Cr was removed from water with an initial [Cr(VI)] of 208
mg L- 1. Advantages of this type of system are again cost- and reagent-
112 M. Losi, C. Amrhein, and W. Frankenberger, Jr.

efficiency. A major disadvantage of this technique is that the process is


gradient-driven, and thus diffusion into the tube will slow as Cr(VI) concen-
trations are lowered. This could lead to relatively long residence times
needed to achieve acceptable removal rates. Another disadvantage is that
the method is not suitable for treating groundwater in situ.
Direct microbial reduction represents a natural mechanism with potential
for use in the detoxification of hexavalent Cr in contaminated water or soil.
Evidence suggests that Cr-reducing microorganisms may be widespread in
soils, and with the addition of an appropriate carbon source, reduction can
be substantially enhanced in a variety of aerobic and anaerobic environ-
ments.
B. Gaseous Bioreduction
Gaseous bioreduction involves microbial reduction of Cr(VI) by a meta-
bolic by-product (in this case, H2S), which is produced in anaerobic envi-
ronments by sulfate-reducing bacteria. This technology has shown some
promise for both anaerobic bioreactor systems and, possibly, for in situ
applications.
DeFillipi and Lupton (1992) designed an anaerobic bioreactor utilizing
marine-derived, sulfate-reducing bacteria, immobilized as a biofilm on
gravel. Their experiments have suggested that, with additional clarification,
effluent Cr levels as low as 0.01 mg L -1 are attainable (Mattison 1992). An
advantage of this system over biosorption/reduction is that the CrO/-
need not come into contact with the cells for reduction to occur. H 2S
diffuses out into the medium, which can increase the reaction rate as well
as protect the cells from the toxic effects of Cr(VI). This could promote
faster reduction rates and may explain the lower effluent concentrations
achieved. A potential disadvantage of this system is that anaerobicity (re-
quiring energy and costs) must be maintained within the reactor.
This technique is also being evaluated as a possible in situ treatment for
immobilization of Cr in chromate-contaminated soils and groundwater.
Such an application would involve production of H 2S in groundwater or
deep within the soil profile by in situ stimulation of sulfate-reducing bacte-
ria with additions of sulfate and other nutrients (Fig. 5). This process was
previously shown to be effective in the reduction and sedimentation of
Cr(VI)-containing tannery wastes entering Otago Harbor in New Zealand
(Smillie et al. 1981).
In addition to obvious cost advantages, this system would have other
benefits as well. First, the H 2S can also reduce Mn4+, which is known to
reoxidize Cr(III) to Cr(VI), thus stabilizing the Cr(lll) precipitate. Sec-
ondly, the H2S could diffuse through soil/water systems, reducing adsorbed
chromates that may otherwise not come into contact with bacterial cells,
resulting in a more complete treatment (Lupton F, see Mattison 1992).
Again, the need for maintaining anaerobicity is a disadvantage of this
system.
Environmental Chromium 113

Water
.. " Bacteria
0 Cr(VI)
~ Cr(lII)

• Cr(OH)3

Fig. 5. Indirect, gaseous reduction of Cr(VI) followed by chemical precipitation. 1.


Hydrogen sulfide is produced by sulfate-respiring bacteria in anaerobic systems. 2.
Hydrogen sulfide gas diffuses into contact with Cr(VI). 3. Cr(VI) is chemically
reduced to Cr(III) with H 2S serving as the electron donor. 4. Cr(lll) reacts with
OH- to form insoluble/immobile Cr(OH)3'

Some additional questions regarding these technologies still remain, pri-


marily with respect to stabilization of precipitated Cr(IJI) following in situ
reduction of Cr(VI). These questions mainly focus on reoxidation of precip-
itated Cr(lll) and the mobilization of organo-Cr complexes. As discussed,
oxidation of Cr(lll) has been shown to occur in the presence of Mn(IV}
(Bartlett and James 1979). While some researchers believe this reaction
could be important and widespread, much of the currently available evi-
dence suggests otherwise. More research is needed to examine reoxidation
of Cr(III) precipitates. Also, it has been suggested that complexation of
Cr(III) with certain organic compounds can prevent precipitation above
pH 5. The extent to which this occurs in nature has not been adequately
addressed.
Of the technologies discussed, gaseous bioreduction technology is the
most promising at present and is furthest along in terms of development
(Mattison 1992). The final test will be to determine whether these tech-
niques can compete with chemical methods in efficiency as well as cost-
effectiveness. If so, they would certainly be environmentally preferred and
could be operational in the near future.
114 M. Losi, C. Amrhein, and W. Frankenberger, Jr.

leaching
plant uptake
/~ adsorption!
HCrQ-4'/ precipitation

OH

¢rR~~.
OH ~

o
¢rR
o

Fig. 6. Cycling of chromium in soil and water (source: Bartlett 1991).

Summary
Chromium is a d-block transitional element with many industrial uses.
It occurs naturally in various crustal materials and is discharged to the
environment as industrial waste. Although it can occur in a number of
oxidation states, only 3+ and 6+ are found in environmental systems.
The environmental behavior of Cr is largely a function of its oxidation
state. Hexavalent Cr compounds (mainly chromates and dichromates) are
considered toxic to a variety of terrestrial and aquatic organisms and are
mobile in soil/water systems, much more so than trivalent Cr compounds.
This is largely because of differing chemical properties: Hexavalent Cr
compounds are strong oxidizers and highly soluble, while trivalent Cr com-
pounds tend to form relatively inert precipitates at near-neutral pH. The
trivalent state is generally considered to be the stable form in equilibrium
with most soil/water systems.
A diagram of the Cr cycle in soils and water is given in Fig. 6 (Bartlett
1991). This illustration provides a summary of environmentally relevant
reactions. Beginning with hexavalent Cr that is released into the environ-
ment as industrial waste, there are a number of possible fates, including
pollution of soil and surface water and leaching into groundwater, where it
may remain stable and, in turn, can be taken up by plants or animals, and
Environmental Chromium 115

adsorption/precipitation, involving soil colloids and/or organic matter.


Herein lies much of the environmental concern associated with the hexava-
lent form.
A portion of the Cr(VI) will be reduced to the trivalent form by inorganic
electron donors, such as Fe2+ and S2-, or by bioprocesses involving organ-
ic matter. Following this conversion, Cr3+ can be expected to precipi-
tate as oxides and hydroxides or to form complexes with numerous ligands.
This fraction includes a vast majority of global Cr reserves. Soluble Cr3+
complexes, such as those formed with citrate, can undergo oxidation when
they come in contact with manganese dioxide, thus reforming hexavalent
Cr.
In trace amounts, Cr is an essential component of animal nutrition,
functioning mainly in glucose metabolism, and possibly in fat metabolism.
While shown to be nonessential for plants, it is required by some microbes,
possibly as a cofactor for specific enzyme systems.
Bacteria with plasmid-conferred resistance to Cr(VI) have been isolated
from water, soil, and sediments, and the resistance mechanisms have been
somewhat characterized. One of the chief mechanisms is bioreduction of
toxic Cr(VI) to the relatively nontoxic Cr(III). This has been shown to
occur directly, by enzymatic processes at the cell membrane, and indirectly,
with microbially produced H 2S acting as the reductant. Since reduction of
Cr(VI) to Cr(III) significantly lowers Cr toxicity and mobility, research is
currently being conducted to assess the potential of these organisms for
biological detoxification and removal/immobilization of Cr from Cr(VI)-
contaminated water and soils. Once reduced, reoxidation to the hexavalent
state is unlikely under most circumstances.
New bioremediation techniques being studied involve bioreactor systems
in which chromates are reduced (either directly or by microbially evolved
H 2S), followed by precipitation, and then removal of the solid by various
filtration/sedimentation methods. In situ bioprocesses are also being as-
sessed, including reduction of Cr(VI) and, hence, immobilization in soils ei-
ther anaerobically, by stimulation of sulfur-reducing bacteria (producing
H 2S), or aerobically in organic matter-amended surface soils. These methods
are potentially more cost-effective than traditional techniques, which typi-
cally involve chemical or electrochemical reduction requiring extensive re-
agent/energy inputs. Also, because only small nutrient inputs are needed,
bioremediation techniques could prove to be more environmentally effective.

References
Adriano DC (1986) Trace elements in the terrestrial environment. Springer-Verlag,
New York, pp 156-180.
Ajmal M, Nomani AA, Ahmad A (1984) Acute toxicity of chrome electroplating
wastes to microorganisms: Adsorption of chromate and chromium (VI) on a
mixture of clay and sand. Water Air Soil Pollut 23:119-127.
116 M. Losi, C. Amrhein, and W. Frankenberger, Jr.

Amacher MC, Baker DE (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. Dept. of Energy. Las
Vegas, NV.
Anderson RA (1989) Essentiality of chromium in humans. Sci Total Environ 86:
15-23.
Apel WA, Turick CE (1991) Bioremediation of hexavalent chromium by bacterial
reduction. In: Smith RW (ed) Mineral bioprocessing. The Minerals, Metals &
Materials Society, Warrendale, PA, pp 377-387.
Ayoub GM, Sayigh BA (1987) The effects and removal of chromium in Chlamydo-
monas sp. Tox Assess: Int Quart 2:253-264.
Bartlett RJ, Kimble JM (1976a) Behavior of chromium in soils: I. Trivalent forms.
J Environ Qual 5(4):379-383.
Bartlett RJ, Kimble JM-(1976b) Behavior of chromium in soils: II. Hexavalent
forms. J Environ Qual 5(4):383-386.
Bartlett R, James B (1979) Behavior of chromium in soils: III. Oxidation. J Environ
Qual 8(1):31-34.
Bartlett R, James B (1988) Mobility and bioavailability of chromium in soils. In:
Nriagu JO, Nieboer E (eds) Chromium in natural and human environments.
John Wiley and Sons, New York, p 276.
Bartlett RJ (1991) Chromium cycling in soils and water: Links, gaps and methods.
Environ Hlth Perspect 92:17-24.
Beas CF Jr, Messmer RE (1986) The hydrolysis of cations. John Wiley and Sons,
New York.
Bertine KK, Goldberg BD (1971) Fossil fuel combustion and the major sedimentary
cycle. Science 171:233-235.
Bharti A, Saxena RP, Pandley GN (1979) Physiological imbalances due to hexava-
lent chromium in fresh water algae. Indian J Environ Hlth 21(3):234-243.
Bloomfield C, Pruden G (1980) The behavior of Cr(VI) in soil under aerobic and
anaerobic conditions. Environ PoIIut Ser A 23:103-114.
Bopp LH, Chakrabarty AM, Ehrlich HL (1983) Chromium resistance plasmid in
Pseudomonasfluorescens strain LB 300. J Bacteriol 155:1105-1109.
Bopp LH, Ehrlich HL (1988) Chromate resistance and reduction in Pseudomonas
f1uorescens strain LB3oo. Arch Microbiol 150:426-431.
Bowen HJM (1966) Trace elements in biochemistry. Academic Press, New York.
Bowen HJM (1979) Environmental chemistry of the elements. Academic Press,
New York.
Cary EE, Alloway WH, Olson OE (1977a) Control of chromium concentrations in
food plants. 1. Absorption and translocation of chromium by plants. J Agric
Food Chern 25:300-304.
Cary EE, Alloway WH, Olson OE (1977b) Control of chromium concentrations in
food plants. 2. Chemistry of chromium in soils and its availability to plants. J
Agric Food Chern 25:305-309.
Cary EE (1982) Chromium in air, soils and natural waters. In: Langard S (ed)
Biological and environmental aspects of chromium. Elsevier Biomedical Press,
Amsterdam, pp 49-64.
Coleman RN, and Paran JH (1983) Accumulation of hexavalent chromium by
selected bacteria. Environ Technol Lett 4: 149-156.
Environmental Chromium 117

Coleman RN (1988) Chromium toxicity: Effects on microorganisms with special


reference to the soil matrix. In: Nriagu JO, Nieboer E (eds) Chromium in natural
and human environments. John Wiley and Sons, New York, pp 335-351.
DeFilippi LJ, Lupton FS (1992) Bioremediation of soluble Cr(VI) using sulfate
reducing bacteria. Allied Signal Research: National R&D Conference on the
Control of Hazardous Materials, pp 138-141.
Deutsch M (1972) Incidents of chromium contamination of groundwater in Michi-
gan. Water quality in a stressed environment. Burgess Publishing Company,
Minneapolis, MN, pp 256-271.
Deverel SJ, Gilliom RJ, Fujii R, Izbicki JA, Fields JC (1984) A real distribution of
selenium and other inorganic constituents in shallow ground water of the San
Luis Drain Service Area, San Joaquin Valley, California: A preliminary study.
U.S. Geological Survey. Water-Resources Investigation Report 84-4319.
Eary LE, Rai D (1987) Kinetics of chromium (III) oxidation to chromium (VI) by
reaction with manganese dioxide. Environ Sci Technol 21:1187-1193.
Eary LE, Rai D (1988) Chromate removal from aqueous waste by reduction with
ferrous iron. Environ Sci Technol 22:972-977.
Eliseeva GS, Klyushnikova TM, Kasatkina TP, Serpokrylov NS (1991) Reduction
of Cr(VI) by microorganisms in media with inedible plant raw material. Khimiya
i Tekhnologiya Vody 13(1):72-75.
Filip D, Peters ST, Adams VD, Middlebrooks EJ (1979) Residual heavy metal
removal by an algae-intermittent sand filtration system. Water Res 13:305-
313.
Frey BE, Riedel GF, Bass AB, Small LF (1983) Sensitivity of estuarine phytoplank-
ton to hexavalent chromium. Est Coast Shelf Sci 17:181-187.
Fujii E, Toda K, Ohtake H (1990) Bacterial reduction of toxic hexavalent chromium
using a fed-batch culture of Enterobacter cloacae strain Hal. J Fermentation
Bioengineering 69(6):365-367.
Gad SC (1989) Acute and chronic systemic chromium toxicity. Sci Total Environ
86: 149-157.
Gaines RW (1988) West San Joaquin Valley Agricultural Setting. Prepared for U.
S. Bureau of Reclamation, Contract No. 7-CS-20-05230.
Gochfeld M (1991) Panel discussion: Analysis of chromium: Methodologies and
detection levels and behavior of chromium in environmental media. Environ
Hlth Perspect (92):41-43.
Griffin RA, Au AK, Frost RR (1977) Effect of pH on adsorption of chromium
from landfill leachate by clay minerals. J Environ Sci Hlth AI2(8):431-449.
Horitsu H, Futo S, Ozawa K, Kawai K (1983) Comparison of hexavalent chromium-
tolerant bacterium, Pseudomonas ambigua G-l, and its hexavalent chromium-
sensitive mutant. Agric Bioi Chem 47(12):2907-2908.
Horitsu H, Futo S, Miyazawa Y, Ogai S, Kawai K (1987) Enzymatic reduction of
hexavalent chromium by hexavalent chromium tolerant Pseudomonas ambigua
G-1. Agric Bioi Chem 51(9):2417-2420.
Huffman E Jr (1973) Chromium: essentiality to plants, forms and distribution in
plants, and availability of plant chromium to rats. Ph.D. dissertation, Cornell
University.
Huffman EWD Jr, Allaway WH (1973) Growth of plants in solution culture con-
taining low levels of chromium. Plant Physiol 52:72-75.
118 M. Losi, C. Amrhein, and W. Frankenberger, Jr.

Hughes MN, Poole RK (1989) Metal toxicity. In: Metals and microorganisms.
Chapman and Hall, London and New York, pp 253-302.
Ishibashi Y, Cervantes C, Silver S (1990) Chromium reduction in Pseudomonas
putida. Appl Environ Microbiol 56(7):2268-2270.
James BR, Bartlett RJ (1983a) Behavior of chromium in soils: V. Fate of organically
complexed Cr(lll) added to soil. J Environ Qual 12:169-172.
James BR, Bartlett RJ (1983b) Behavior of chromium in soils: VI. Interactions be-
tween oxidation-reduction and organic complexation. J Environ Qual 12: 173-176.
James BR, Bartlett RJ (1983c) Behavior of chromium in soils: VII. Adsorption and
reduction of hexavalent forms. J Environ Qual 12: 177-181.
Jeejeebhoy KN, Chu RH, Marliss EB, Greenberg GR, Bruce-Robinson A (1977)
Chromium deficiency, glucose intolerance, and neuropathy reversed by chro-
mium supplementation in a patient receiving long term total parenteral nutrition.
Am J Clin Nutr 30:531-538.
Khasim DI, Nanda Kumar NV, Hussain RC (1989) Environmental contamination
of chromium in agricultural and animal products near a chromate industry. Bull
Environ Contam ToxicoI43:742-746.
Komori K, Wang P, Toda K, Ohtake H (1989) Factors affecting chromate reduction
in Enterobacter cloacae strain HOI. Appl Microbiol BiotechnoI31:567-570.
Komori K, Rivas A, Toda K, Ohtake H (1990a) Biological removal of toxic chro-
mium using an Enterobacter cloacae strain that reduces chromate under anaero-
bic conditions. Biotechnol Bioeng 35:951-954.
Komori K, Rivas A, Toda K, Ohtake H (1990b) A method for removal of toxic
chromium using dialysis-sac cultures of a chromate-reducing strain of Enterobac-
ter cloacae. Appl Microbiol Biotechnol 33: 117-119.
Komori K, Toda K, Ohtake H (199Oc) Effects of oxygen stress on chromate reduc-
tion in Enterobacter cloacae strain HOI. J Fermen Bioeng 69(1):67-69.
Krauskopf KB (1979) Introduction to geochemistry, 2nd ed. McGraw-Hill, New
York.
Kvasnikov EI, Stepanyuk VV, Klyushnikova TM, Serpokrylov NS, Simonova GA,
Kasatkina TP, and Panchenko LP (1985) A new chromium-reducing, gram-
variable bacterium with mixed type of flagellation. Translated from Mikrobio-
logiya 54(1):83-88.
Kvasnikov EI, Serpokrylov NS, Klyushnikova TM, Kasatkina TP, Zukov 1M, To-
kareva LL (1986) Optimization of a nutrient medium for Aeromonas dechromat-
ica reducing Cr(VI). Khimiya i Tekhnologiya Vody 8(3):64-66.
Kvasnikov EI, Serpokrylov NS, Klyushnikova TM, Kasatkina TP, Zukov 1M, To-
kareva LL (1987) Reduction of Cr(VI) by a culture of Aeromonas dechromatica
KS-ll in the presence of certain heavy metals. Khimiya i Tekhnologiya Vody
9(2): 159-162.
Langard S (1980) In: Waldron HA (ed) Metals in the environment. Academic Press,
New York, pp 111-132.
Lester IN, Perry R, Dadd AH (1979) The influence of heavy metals on a mixed
bacterial population of sewage origin in the chemostat. Water Res 13:1055-1063.
Levis AG, Bianchi V (1982) Mutagenic and cytogenic effects of chromium com-
pounds. In: Langard S (ed) Topics in environmental health. Vol. 5. Biological
and environmental aspects of chromium. Elsevier Biomedical Press, New York,
pp 171-206.
Environmental Chromium 119

Losi ME, Amrhein C, Frankenberger WT Jr (1994a) Bioremediation of chromate


contaminated groundwater by reduction and precipitation in surface soils. J
Environ Qual (in review).
Losi ME, Amrhein C, Frankenberger WT Jr (1994b) Factors affecting chemical
and biological reduction of hexavalent chromium in soil. Environ Toxicol Chern
(in review).
Losi ME, Frankenberger WT Jr (1994) Chromium resistant microorganisms isolated
from evaporation ponds of a metal processing plant. Water Air Soil Pollut 74:
1-9.
Love AHG (1983) Chromium-biological and analytical considerations. In: Burrows
D (ed) Chromium: Metabolism and toxicity. CRC Press, Inc., Boca Raton, FL,
pp 1-13.
Luli GW, Talnagi JW, Strohl WR, Pfister RM (1983) Hexavalent chromium-
resistant bacteria isolated from river sediments. Appl Environ Microbiol 46(4):
846-854.
Marquez AM, Espuny MJ, Congregado F, Simon-Pujol MD (1982) Accumulation
of chromium by Pseudomonas aeruginosa. Microbios Lett 21: 143-147.
Matthews NA, Morning JL (1980) In: US Bureau of mines, minerals yearbook
1978-79, Vol 1. Metals and Minerals, 193-205. US Dept of Interior, Washing-
ton, DC.
Mattison PL (1992) Bioremediation of metals-putting it to work. Ch. 3, Chro-
mium. Cognis, Inc., Santa Rosa, CA.
McGrath SP, Smith S (1990) Chromium and nickel. In: Alloway BJ (ed) Heavy
metals in soils. John Wiley and Sons, New York, pp 125-147.
Mehra HC, Frankenberger WT Jr (1989) Single-column ion chromatographic deter-
mination of chromium(VI) in aqueous soil and sludge extracts. Talanta 36(9):
889-892.
Mertz W (1969) Chromium occurrence and functions in biological systems. Physiol
Rev 49:165-239..
Mertz WE, Toepfer W, Polansky MM, Roginski EE, Wolf WR (1977) Preparation
of chromium-containing material of glucose tolerance factor activity from brew-
ers yeast extracts and by synthesis. J Agric Food Chern 25(1):162-166.
Naguib MI, Haikal NZ, Gouda S (1984) Effects of chromium ions on the growth of
Fusarium oxysporum f.sp. lycopersici and Cunninghamella echinulata. Arab
Gulf J Sci Res 2(1):149-157.
National Academy of Sciences (NAS) (1974) In: Chromium. NAS, Washington DC.
Nieboer E, Jusys AA (1988) Biologic chemistry of chromium. In: Nriagu JO, Nie-
boer E (eds) Chromium in natural and human environments. John Wiley and
Sons, New York, pp 21-81.
Nriagu JO (1988) Production and uses of chromium. In: Nriagu JO, Nieboer E
(eds) Chromium in natural and human environments. John Wiley and Sons,
New York, pp 81-105.
Nriagu JO, Pacyna JM, Milford JB, Davidson CI (1988) Distribution and character-
istic features of chromium in the atmosphere. In: Nriagu JO, Nieboer E (eds)
Chromium in natural and human environments. John Wiley and Sons, New
York, pp 125-173.
Ohtake H, Cervantes C, Silver S (1987) Decreased uptake in Pseudomonas fluo-
rescens carrying a chromate resistance plasmid. J BacterioI169(8):3853-3856.
120 M. Losi, C. Amrhein, and W. Frankenberger, Jr.

Ohtake H, Fujii E, Toda K (1990) Reduction of toxic chromate in an industrial


effluent by use of a chromate-reducing strain of Enterobacter cloacae. Environ
TechnoI11:663-668.
Pacyna JM (1986) Atmospheric trace elements from natural and anthropogenic
sources. In: Nriagu JO, Davidson C (eds) Toxic metals in the atmosphere. John
Wiley and Sons, New York, pp 33-52.
Pacyna JM, Nriagu JO (1988) Atmospheric emissions of chromium from natural
and anthropogenic sources. In: Nriagu JO, Nieboer E (eds) Chromium in natural
and human environments. John Wiley and Sons, New York, pp 105-125.
Perkin-Elmer (1982) Analytical methods for atomic absorption spectrophotometry.
Perkin-Elmer Corp., Norwalk, CT.
Pratt PF (1966) Chromium. In: Chapman HD (ed) Diagnostic criteria for plants
and soils. Quality Printing Co. Inc., Abilene, TX, pp 136-141.
Rai D, Zachara JM, Eary LE, Girvin DC, Moore DA, Resch CT, Sass BM, Schmidt
RL (1986) Geochemical behavior of chromium species. Electric Power Research
Institute, Palo Alto, CA, EA-4544.
Rai D, Zachara JM (1988) Chromium reactions in geologic materials. EA-5741.
Research Project 2485-3. Prepared for Electric Power Research Institute. Palo
Alto, CA. Land and Water Quality Studies Program, Environment Division.
Rai D, Eary LE, Zachara JM (1989) Environmental chemistry of chromium. Sci
Total Environ 86:15-23.
Reisenauer HM (1982) Chromium. In: methods of soil analysis, part 2, Chemical
and microbiological properties. Agronomy Monograph no. 9 (2nd ed.) ASA-
SSSA, Madison, WI, p 340.
Roda IG, Smirnova GF (1989) Biochemical treatment of chromium-containing
waste water. Khimiya i Tekhnologiya Vody 11(2): 169-172.
Romanenko VI, Kusnetsov SI, Koren'kov VN (1976) Koren'kov method for biologi-
cal purification of wastewater. USSR patent SU 521,234.
Romanenko VI, Koren'kov VN (1977) A pure culture of bacteria utilizing chromates
and bichromates as hydrogen acceptors in growth under anaerobic conditions.
Institute of Biology of Inland Water, Academy of Sciences of the USSR. Trans-
lated from Mikrobiologiya 46(3):414-417.
Ross DS, Sjogren RE, Bartlett RJ (1981) Behavior of chromium in soils: IV. Toxic-
ity to microorganisms. J Environ Qual 10(2):145-148.
Shimada K, Matsushima K (1983) Isolation of potassium chromate-resistant bacte-
rium and reduction of hexavalent chromium by the bacterium. Bull Fac Agric
Mie Univ 67:101-106.
Shupack SI (1991) The chemistry of chromium and some resulting analytical prob-
lems. Environ Hlth Perspect 92:7-11.
Simon-Pujol MD, Marques AM, Ribera M, Congregado F (1979) Drug resistance
of chromium-tolerant Gram negative bacteria isolated from a river. Microbios
Lett 7: 139-144.
Skeffington RA, Shewry PA, Peterson PJ (1976) Chromium uptake and transport
in barley seedlings (Hordeum vulgare L.). Planta 132:209-214.
Smillie RH, Hunter K, Loutit M (1981) Reduction of chromium(VI) by bacterially
produced hydrogen sulfides. Water Res 15:1351-1354.
Sposito G, Mattigod SV (1980) GEOCHEM: A computer program for the calcula-
tion of chemical equilibria in soil solutions and other natural water systems.
Kearney Foundation of Soil Science, Univ. of California, Riverside, CA.
Environmental Chromium 121

Studt T (1993) Chromium hazard drives replacement R&D. R&D Magazine. Cahn-
ers Publishing, Des Plaines, IL, Sept., p 6.
Summers AO, Jacoby GA (1977) Plasmid-determined resistance to boron and chro-
mium compounds in Pseudomonas aeruginosa. Antimicrob Agents Chemother
13(4):637-6440.
U.S. Environmental Protection Agency (USEPA) (1976) Quality criteria for water.
USEPA, Washington, DC.
USEPA (1984) Health assessment for chromium. Report No. EPA-600/8-83-014F.
Environmental Criteria and Assessment Office, Research Triangle Park, NC.
USEPA (1986) EPA Methods. SW-846. 3rd ed.
USEPA (1993) Standards for use and disposal of sewage sludge: final rules. EPA
40 CFR part 257. Fed Regis, Feb. 19, pp 9248-9415.
USEPA/Oak Ridge National Laboratories (ORNL) (1978) Reviews of environmen-
tal effects of pollutants III. Chromium. Environmental Protection Agency, Oak
Ridge National Laboratories, U.S. National Technical Information, Springfield,
VA.
Warington K (1946) Molybdenum as a factor in the nutrition of lettuce. Ann Appl
BioI 33:249-254.
Whitten KW, Gailey KD (1984) General Chemistry, 2nd ed. Saunders College Pub-
lishing, Philadelphia, p 789.
Williams JW, Silver A (1984) Bacterial resistance and detoxification of heavy met-
als. Enzyme Microb TechnoI6:530-537.
Wood JM, Wang H-K (1983) Microbial resistance to heavy metals. Environ Sci
TechnoI17(12):582A-59OA.
Zachara JM, Girvin DC, Schmidt RC, Resch CT (1987) Chromate adsorption on
amorphous iron hydroxide in the presence of major groundwater ions. Environ
Sci Technol21 :589-594.
Zachara JM, Ainsworth CC, Cowan CC, Resch CT (1989) Adsorption of chromate
by subsurface soil horizons. Soil Sci Soc Am J 53:418-428.
Zumriye A, Sag Y, Kutsal T (1990) A comparitive study of the adsorption of Cr(VI)
ions to C vulgaris and Z ramigera. Environ Technol 11 :33-40.

Manuscript received September 29, 1993; accepted December 18, 1993.

You might also like