You are on page 1of 6

Environ. Sci. Technol.

2001, 35, 4054-4059

estimated 33 000 kg of trichloroethylene (1). In contrast, H2O2


Hydrogen Peroxide Effects on can be used at lower levels (3-300 mM) to deliver oxidizing
Chromium Oxidation State and equivalents to support in situ bioremediation (2).
Such uses of H2O2 raise concern about possible side
Solubility in Four Diverse, reactions with co-contaminants. A common co-contaminant
in hazardous wastes is Cr, which is reported at 43% of the
Chromium-Enriched Soils hazardous waste sites on the National Priorities List in the
United States (3). The ability of H2O2 to act both as an
MELANIE L. ROCK,† oxidizing agent and as a reducing agent is well-known. Under
BRUCE R. JAMES,‡ AND appropriate conditions H2O2 can oxidize Cr(III) to Cr(VI) or
G E O R G E R . H E L Z * ,† reduce Cr(VI) to Cr(III) (4, 5). What its effect, if any, might
Department of Chemistry and Biochemistry and Water be on Cr at hazardous waste sites is an open question.
Resources Research Center and Soil Chemistry Program, Chromium(VI) is a human carcinogen that predominates in
Department of Natural Resources Science and Landscape soils and natural waters as mobile anions, HCrO4- or CrO42-.
Architecture, University of Maryland, In contrast, Cr(III) is an essential trace element for humans
College Park, Maryland 20742 and is relatively immobile in the same environments because
of low solubility and propensity to sorb to natural solids (6).
Cr contamination in soils has resulted from the disposal of
industrial wastes produced by several different processes
High concentrations of H2O2 are being tested for in situ (e.g., electroplating, leather tanning, chromite ore mining
oxidation and remediation of buried organic contaminants and processing). Cr-contaminated sites are therefore chemi-
in soils and groundwater. Peroxide is being considered cally diverse with respect to forms, oxidation state, and
as a direct chemical oxidant in Fenton-type reactions or as solubility of this element.
a source of oxidizing equivalents in bioremediation We explore in this paper the potential for solubilization
schemes. How H2O2 affects the oxidation state and solubility of Cr by aqueous H2O2 in selected soils that incorporate some
of Cr(III) and Cr(VI), common co-contaminants with of the chemical and pedologic diversity found at contami-
organic chemicals, is explored here in four chemically nated sites. Differences include chromium deposition pro-
cesses, total Cr levels, soil pH, redox conditions, chromate/
diverse soils containing elevated levels of Cr. Soil
bichromatespeciation,andorganic/inorganicCr(III)speciation.
contaminated with soluble Cr(VI) from chromite ore Low millimolar H2O2 levels were chosen for the experimental
processing residue and soil containing high levels of treatments based on the concentration range being con-
recently reduced Cr (III) from electroplating waste both sidered for enhanced bioremediation treatments (2).
released dissolved Cr(VI) after single applications of up to
24 mM H2O2. In no case was there evidence that H2O2 Site Descriptions
reduced preexisting Cr(VI) to Cr(III), even though this would
Sampling. A similar sampling protocol was followed at each
be allowed thermodynamically. Chromate in the leachates site: an undisturbed area about 1 m2 was cleared of leaf and
exceeded the U.S. EPA drinking water standard for plant covering, a pit was dug, soil horizons were marked and
total dissolved Cr (2 µM) by a factor of 10-1000. Anaerobic identified, and samples were taken from each horizon.
conditions in an organic-rich, tannery waste-contaminated Samples were kept intact as large blocks (∼2 L volume), sealed
soil protected Cr(III) from oxidation and mobilization. in 0.004 in thick plastic bags, transported to the laboratory
Mineral forms of Cr in serpentinitic soil near a former in coolers, and stored in a refrigerator at 4° C in the field-
chromite mine also resisted oxidation on the time scale of moist condition (0 to ∼0.1 kPa water potential). Intact blocks
days. Mobilization of Cr(VI) could be a hazardous of soil from individual horizons were prepared for the
consequence of using H2O2 for in situ remediation of experiments by passing them through a polyethylene sieve
using gentle hand pressure to obtain a relatively homoge-
chemically complex wastes, but H2O2 could prove attractive
neous subsample from each horizon. Soil from the chromite
for ex situ treatment (i.e., soil washing). This paper ore processing site was prepared using a 0.40 cm sieve; a 0.25
demonstrates marked differences among Cr-contaminated cm sieve was used on all other soils.
soils in their capacity to release Cr(VI) upon chemical Electroplating Waste Site. National Chromium, Inc. (a
treatment with H2O2. small electroplating facility near Putnam, CT) discharged
chromic acid-containing wastewater directly into an adjacent
mixed hardwood wetland and flood plain from 1939 to 1975
Introduction (7). The peat-like uppermost horizon contained the highest
total Cr levels of any soil in this study (Table 1). Green
Hydrogen peroxide is a potentially useful reagent for in situ
chromium(III) hydroxide coatings were evident on fallen
waste remediation. It can be used with high Fe(II) concen-
branches and plant debris surrounding the site, and some
trations to effect Fenton-type oxidation of refractory organic
soil samples contained as much as 60 g of Cr/kg of soil. The
contaminants in soils and groundwater. In one field test, 436
soil is very poorly drained, and its pH is low: 4-5. Even
m3 of 50% H2O2 was injected into an aquifer to oxidize an
though its high levels of organic matter (200 g of C/kg of soil)
and low pH might be expected to reduce Cr(VI) to Cr(III), it
* Corresponding author e-mail: gh17@umail.umd.edu; phone: nonetheless contains leachable Cr(VI) (60-90 µM) in the
(301)405-1797; fax: (301)314-9121.
† Department of Chemistry and Biochemistry and Water Resources uppermost horizon. The X-ray diffraction spectrum indicated
Research Center. the presence of quartz and Cr(III)-rich chlorite. Soluble Cr-
‡ Soil Chemistry Program, Department of Natural Resources (VI) was not detected in the middle horizon at this site but
Science and Landscape Architecture. did appear in the glacial till below 100 cm. Mattuck (8)
4054 9 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 35, NO. 20, 2001 10.1021/es010597y CCC: $20.00  2001 American Chemical Society
Published on Web 08/30/2001
TABLE 1. Soil Properties
horizon soil organic carbon total Cra total Cr(VI)b soluble Cr(VI)c soluble Cr(III)d
soil designation (cm) pH (g/kg of soil) (g/kg of soil) (g/kg of soil) (µM) (µM)
electroplating 0-14 5.4 280d 61 ( 5 0.071 ( 0.007 95 ( 6 11 ( 1
waste 14-40 5.0 118d 21 ( 1 0.079 ( 0.005 <0.1 <1.0
>100 4.9 5.2d 0.40 ( 0.03 0.079 ( 0.003 32 ( 1 <1.0
ore processing residue surface 8.8 66 8.6 ( 0.5 0.914 ( 0.022 940 ( 40 <1.0
tannery 0-20 6.7 154d 1.3 ( 0.1 <0.00005 <0.1
20-40 5.7 202d 4.7 ( 0.2 <0.00005 <0.1 2.0 ( 0.2
chromite mine 53-75 6.5 4.2d 2.5 ( 0.2 <0.00005 <0.1 2.0 ( 0.2
a Atomic absorption analysis, H SO -H O -HF digestate. b DPC colorimetry, heated, strongly alkaline, OH--CO 2- leach. c DPC colorimetry,
2 4 2 2 3
aqueous leach, 10:1 solution:soil ratio. d Measurements by Typrin (36).

reported Cr(VI) levels of up to 950 µM in groundwater from only 0.5% of the samplestoo little to observe by X-ray
a well about 10 m downslope from the sampling locality. diffraction.
Chromite Ore Processing Site. Soil incorporating chromite
ore processing residue in Kearney, NJ, had a comparatively Methods
high pH (8.8), a relatively large Cr(VI):Cr(III) ratio (0.10), and Procedures. Batch experiments on soils were conducted in
high ambient levels of soluble Cr(VI) (940 µM) (9-11). During triplicate in 50 mL, polycarbonate centrifuge tubes. Suspen-
ore processing, chromite ore was crushed to less than 100 sions containing solution:soil ratios of 10:1 by mass were
mesh size, mixed with soda ash, and roasted to produce prepared by placing 3.00 g of field-moist soil in each tube
soluble Na2CrO4. The Na2CrO4 was leached from the roasted and equilibrating for 24 ( 1 h at 25 °C on an orbital shaker
material several times with water, after which the alkaline with 30.0 mL of 0.01 M NaNO3. After equilibration, H2O2 was
residue was discarded in low-lying areas and wetlands. The added to the soil suspensions in small volumes to obtain the
residue, which still releases Cr(VI) decades later (9), retains desired initial concentration in the soil suspension while not
residual levels of sparingly soluble salts including CaCrO4, diluting significantly the original Cr concentrations. To
Ca4Al2CrO10, and FeOHCrO4 as well as insoluble Cr(III) monitor soluble Cr(VI), samples were extracted, centrifuged
resistant to the original processing (10). Soil samples were (15000g, 15 min, 25 °C), and analyzed as described below.
from a flood plain about 600 m from the Hackensack River Experiments were also conducted on synthetic chromium-
where the depth to groundwater was 2-3 m. Powder X-ray (III) (hydr)oxides. Solutions containing 280 µM Cr(NO3)3 and
diffraction patterns revealed only quartz and calcitesno 0.01 M NaNO3 were prepared and neutralized with 0.005 M
crystalline Cr minerals (12). NaOH to give total OH-/Cr ratios of 0, 1.0, 2.0, 3.0, and 4.0.
Tannery Site. The Aberjona watershed near Woburn, MA, To minimize high, localized OH- concentrations during
was the site of over 100 chrome tanning operations between preparation, NaOH was added just below the liquid surface
1838 and 1988. Inorganic arsenical and lead-based insecti- at a rate of 2 mL min-1 while stirring. Formation of particles
cides were manufactured in the same locale from the 1860s (blue-green flocs) was observed only in the 3:1 and 4:1 OH:
until the 1920s (13, 14). In the tanning process, Cr(VI) was Cr solutions. The solutions were aged with gentle swirling
reduced with organic acids over the surface of animal hides, on an orbital shaker (50 cycles min-1) for at least 1 week,
forming stable Cr(III) complexes that preserved the leather. after which time the pH values remained stable for several
Waste Cr(III,VI) was mainly discharged into lagoons that were months. To avoid possible effects that have been noted from
used to dispose of the sludge from leather production (15). the interaction of buffers with Cr/H2O2 intermediate species
Because of the high levels of animal-derived organic waste (5, 20) and to observe pH changes in the systems we
at these sites, natural biodegradation processes produce investigated, solutions were prepared without buffers. Reac-
highly reducing environments that evolve H2S and CH3SH tions of these Cr(III) solutions were studied with a procedure
(13). like that used with the soil suspensions: peroxide was added
Two horizons of a riverbank soil were sampled at the to the OH-/Cr(III) solutions, and the production of Cr(VI)
Superfund site downstream from the tannery disposal was monitored over time.
lagoons. Sulfidic conditions in the shallow groundwater near Analysis. The pH was measured with an Orion flat surface
the site could be expected to reduce Cr(VI) (16). Although Cr combination pH electrode inserted directly into the super-
appears exclusively as Cr(III) in these sites (Table 1), it has natant solution in each centrifuge tube. Soluble Cr(VI) was
a surprising degree of mobility in the groundwater (13), measured by the diphenylcarbazide (DPC) colorimetric
probably due to the enhanced solubility of Cr(III) organic method (21) using a 4.5 mL aliquot of the centrifugate.
complexes formed with dissolved organic C provided by the Diphenylcarbazide reacts with Cr(VI) in acidic solution to
decomposing hides (17). form a Cr(III)-diphenylcarbazone complex that absorbs at
Mine Site. Unlike the three other sites, the Maryland 540 nm (detection limit 0.1 µM). At concentrations higher
serpentine barrens, about 20 km northwest of Baltimore, than 10-5 M, H2O2 causes a negative interference with the
contain only naturally elevated levels of Cr. This site lies DPC determination of Cr(VI) by competitively reducing Cr-
within the belt of serpentinites scattered in the Piedmont (VI) under the pH 1.8 conditions of the test. Pettine et al. (22)
Plateau from New Jersey to Alabama (18). The site first became found that the negative interference of H2O2 can be avoided
a source of industrial chromite around 1840. Soils here have at H2O2 < 10-4 M by increasing reagent concentrations. In
been studied extensively to determine the cause of their low our experiments, however, much higher concentrations of
soil fertility (19). Soil was sampled at Soldier’s Delight Natural H2O2 were used (up to 0.1M). Removal of peroxide by adding
Environmental Area, 100 m from an old mine. Samples used catalytic (10-8 M) amounts of the enzyme, catalase, to a
in this study contained 2.5 g of total Cr/kg of soil (Table 1). sample and waiting 30 min prior to analysis was found to be
X-ray diffraction revealed only antigorite, a serpentine effective. Where catalase was used, Cr(VI) standards were
mineral. No chromite was identified, even though this mineral also prepared with 10-8 M catalase to correct for the slight
is reportedly the Cr host phase at the site. If all the Cr in our absorbance of catalase at 540 nm. Results for standards using
sample were present in chromite, this phase would comprise DPC and catalase corresponded well with results obtained

VOL. 35, NO. 20, 2001 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 9 4055
FIGURE 1. Dissolved Cr(VI) released from the 0-14 cm (peaty) horizon FIGURE 2. Dissolved Cr(VI) release from the 14-40 cm and >100
of the electroplating waste site upon treatment with various cm horizons of the electroplating waste site upon treatment with
concentrations of H2O2, which was added at time ) 0. Solution to H2O2. Analyses made in triplicate.
soil ratio, 10:1. Peroxide decayed rapidly, as indicated in Table 2.
Analyses were made in triplicate; the scatter at each sampling TABLE 2. Decay H2O2 upon Contact with Several Soilsa
point in the figure is contributed both by soil heterogeneity and by
analytical error. soil initial 1.5 h 24 h 48 h
electroplating waste site, 0-14 cm 24 11.7 <0.005 <0.005
using a direct optical method (23) at 350 nm for HCrO4-. electroplating waste site, 14-40 cm 24 18.5 2.6 <0.005
Total Cr(VI) in soil samples was determined using a heated ore processing residue 24 10.1 <0.005 <0.005
carbonate-hydroxide aqueous extraction method with quan-
tification by DPC colorimetry (24). Slight optical interference (VI) was observed after the water-soluble fraction of Cr(VI)
due to solubilized humic materials from the soils was had initially dissolved. Increases in Cr(VI) levels peaked 4 h
corrected by subtracting the absorbance measured in soil after H2O2 application. Thereafter, Cr(VI) drifted back toward
extracts containing no DPC. Soluble total Cr in the soil was pre-peroxide levels over a period of 3 weeks. Within 90 min
determined by atomic absorption spectroscopy following a after application, ∼50% of the H2O2 in samples spiked with
dilute salt extraction (0.01 M CaCl2, 1:5 soil:solution ratio, 1 24 mM had disappeared, and peroxide was not detectable
h equilibration) (36). Soluble Cr(VI) was determined by DPC in any sample after 1 day (Table 2). Thus, the decline in
colorimetry, and soluble Cr(III) was estimated by difference Cr(VI) after 4 h probably reflects reduction of Cr(VI) by organic
between soluble total Cr and soluble Cr(VI). Total Cr in soil matter in the peaty soil after the H2O2 has disappeared. On
was determined by atomic absorption following a digestion a molar basis, the maximum amount of Cr(VI) released from
procedure using H2SO4-H2O2-HF dissolution of oven-dried the soil was approximately 1% of the H2O2 lost, indicating
(105 °C), crushed (35-mesh) soil samples (25). that these processes are not linked stoichiometrically.
A modification of the 4-aminoantipyrene horseradish At the two highest peroxide doses, the maximum Cr(VI)
peroxidase method, reviewed in (26), was used to determine concentration exceeded all the Cr(VI) in the soil sample, based
H2O2 concentrations. H2O2 oxidatively couples with 4-ami- on total Cr(VI) (Table 1). As discussed later, this suggests that
noantipyrene (AAP) and phenol in the presence of horserad- H2O2 was oxidizing Cr(III) in the sample. Alternately, it would
ish peroxidase to produce a quinoneimine dye with a be possible that H2O2 is simply a more efficient extractor of
maximum absorption at 505 nm. The linear range was 5-300 preexisting Cr(VI) in the soil than is the hot alkaline leach
µM H2O2; for many of our assays, it was necessary to dilute used to quantify total Cr(VI) (Table 1), although this seems
the sample into this range before analysis. High concentra- less likely since H2O2 contains no ions capable of desorbing
tions of phenol provided stearic hindrance to complexation or dissolving Cr(VI).
of AAP with organic hydroperoxides in soil solution, ensuring As shown in Figure 2, Cr in the 14-40 cm horizon at the
specificity for H2O2 (26). The modified aqueous reagent was electroplating site was dissolved by a 12 mM H2O2 dose, but
mixed in a 500 mL volumetric flask as follows: 0.0010 g of to a very limited extent. This sample initially contained no
horseradish peroxidase (type VI) from Sigma (about 2 × 10-8 soluble Cr(VI), but total Cr(VI) was similar to that in the
M), 0.50 g of AAP, 1.17 g of phenol, 5.0 mL of 0.1 M phosphate surface layer at this site (Table 1). Chromium(VI) levels
buffer (pH 6.9), and 100 µL of 0.01 M H2O2 (or 2 µM) to give reached 5-6 µM. H2O2 took longer to disappear in the 14-40
more stable readings. Reagent (2.0 mL) and aqueous sample cm horizon as compared to the 0-14 cm horizon (Table 2).
(3.0 mL) were vortexed; absorbance readings at 505 nm Application of only 3 mM peroxide to soil from the >100
reached a maximum in 30-120 s, after which time they were cm horizon at the electroplating site produced an interesting
no longer stable. Sample color faded at a rate that varied effect (Figure 2). Initially, chromate levels fell in this low
with H2O2 concentrations, so it was necessary to make organic carbon soil. Chromate extracted from this soil was
measurements promptly. Further details on methods and determined by direct optical absorption (350 nm) rather than
tabular presentations of the data are available in Rock (12). by DPC. The initial loss of chromate absorption does not
necessarily indicate reduction of Cr(VI) to Cr(III). Conceiv-
Results ably, the loss of HCrO4- absorbance was due to the formation
Soil Experiments. As shown in Figure 1, H2O2 additions to of another, nonabsorbing Cr(VI) species, for example, a
the surface soil at the electroplating site produced rapid, peroxo-Cr(VI) complex. We will return to this possibility later.
dose-dependent increases in soluble Cr(VI) in the uppermost, Peroxide in the >100 cm soil took about 2 d to reach the
peaty horizon. Without added H2O2, little rise in soluble Cr- detection limit of 5 µM.

4056 9 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 35, NO. 20, 2001
FIGURE 3. Dissolved Cr(VI) concentrations released from ore FIGURE 4. Production of dissolved Cr(VI) from synthetic Cr(III)
processing residue (10:1 solution/soil by mass) upon single hydrolysis products aged 1 week before addition of 100 µM H2O2.
applications of H2O2 at time ) 0. Peroxide decayed rapidly and was Time measured from H2O2 addition. Bold numbers indicate the initial
undetectable after 1 day. Analyses made in triplicate. stoichiometric ratios of NaOH to Cr(NO3)3. Indicated pH measure-
ments made at day 17; pH started higher in each case but stabilized
by day 17. Oxidation of Cr(III) by H2O2 generates acidity.
The soil sample enriched with chromite ore processing
residue released 900 µM Cr(VI) when leached for 24 h simply
with 0.01 M NaNO3 containing no H2O2 (Figure 3). As much
Discussion
as 200 µM additional Cr(VI) was released over 24 h by To summarize our observations, responses to H2O2 were
leachates containing up to 24 mM H2O2. Peroxide dropped varied. The highly reducing tannery soil and the soil
by more than 50% within 90 min after addition and was developed on a natural mineral deposit at the mine site
undetectable after 24 h (Table 2). In contrast to the released no Cr(VI), while soils at the electroplating and ore
electroplating site (Figure 1), Cr(VI) in the ore processing processing sites released large amounts of Cr(VI). However
residue did not decline appreciably even over a period of 1 in the acidic, peaty electroplating site surface soil, Cr(VI) was
week after H2O2 was undetectable. The slight Cr(VI) decline lost from solution within days after H2O2 had decayed. In the
observed in the most concentrated sample in Figure 3 was alkaline soil with moderate organic C at the ore processing
only 6%, nearly within analytical uncertainty. It seems likely site, most of the Cr(VI) remained dissolved after H2O2 had
that the very high soluble Cr(VI) concentrations and the decayed. In no case was there evidence that H2O2 was
moderate organic C content of this soil overwhelmed a re- reducing preexisting Cr(VI) to Cr(III), even though this would
adsorption mechanism like that observed at the electroplating be allowed thermodynamically. In the samples where H2O2
site. treatments had any effect, that effect was always a higher,
Release of Cr(VI) was not observed in the highly reducing not a lower, Cr(VI) yield as compared to treatment with a
environment of the Massachusetts tannery waste site soil at redox-inert electrolyte.
peroxide application levels of up to 100 mM. Samples from Total Cr and total Cr(VI) proved poor predictors of Cr(VI)
the chromite ore site also did not produce detectable soluble dissolution after H2O2 treatment. For example, in the 14-40
Cr(VI) after addition of up to 100 mM H2O2 (data not shown cm horizon at the electroplating site, with 2.1% total Cr, very
for either soil). little Cr(VI) was released (Figure 2); whereas, the >100 cm
horizon at the same site, with 0.04% total Cr, released more
Treatment of Synthetic Chromium(III) Hydroxides. To
Cr(VI) despite treatment with 25% as much H2O2. These two
prove that H2O2 can oxidize Cr(III) to Cr(VI) under conditions
horizons were essentially identical with respect to initial Cr-
similar to those used in our soil leaching experiments, we
(VI) content and soil pH (Table 1).
prepared chromium(III) (hydr)oxides by neutralizing 280 µM
In our limited group of samples, the best predictor of
Cr(NO3)3 with 1-4 equiv of NaOH. At OH/Cr ratios of 1and
Cr(VI) release was the presence of soluble Cr(VI) in the soil
2, the pH was 3-4 and all of the Cr(III) remained dissolved
prior to H2O2 treatment. The tannery and mine soils as well
as cationic hydroxide complexes (e.g., CrOH2+). At OH/Cr
as the 14-40 cm horizon at the electroplating waste site
ratios of 3 and 4, blue green flocs were present. The Cr(III)
contained no Cr(VI) that was extractable in the aqueous leach
solubility reached a minimum at OH/Cr ) 3, corresponding
(Table 1). Each of these soils released little or no Cr(VI) upon
to the stoichiometry of Cr(OH)3(s); at higher OH/Cr ratios,
subsequent treatment with H2O2. Each of the three soils that
some Cr(III) redissolved, presumably as anionic hyrdoxo-
contained substantial leachable Cr(VI) prior to H2O2 treat-
chromates (e.g., Cr(OH)4-).
ment also released much more Cr(VI) after treatment.
Figure 4 shows Cr(VI) production by oxidation of the Cr- Several explanations for observed increases in Cr(VI) after
(III) in these mixtures after 100 µM H2O2 was added. H2O2 treatment can be proposed: H2O2 may have oxidized
Production continued slowly over 4 months, although most Cr(III) in the soil or it may have simply solubilized existing
of the Cr(VI) appeared within 3 weeks. The greatest extent Cr(VI). The first can be represented as follows:
of oxidation occurred under mildly alkaline conditions at
OH-/Cr ) 4, where at the end of 115 days, 71% of the 3H2O2(aq) + 2Cr(OH)3(s) f 2HCrO4- + 2H+ + 4H2O
stoichiometric yield of Cr(VI) expected from 100 µM H2O2 (1)
was obtained. The smallest yield occurred in the 3:1 mixture
where Cr(III) solubility was minimal. Note that the reversal The second might occur through formation of aqueous
of Cr(VI) production seen in Figure 1 does not occur in these chromium(VI) peroxide species (5, 27-30), as represented
synthetic mixtures, presumably due to the absence of a by the following reaction in which a solid Cr(VI) source is
reducing agent or sorption surface. represented by chromatite:

VOL. 35, NO. 20, 2001 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 9 4057
2H2O2(aq) + CaCrO4(s) + H+ a contact with our soils was on the order of hours. As noted
- 2+ above, the quantity of peroxide lost greatly exceeded the
Cr(O)(O2)2(OH) + Ca + 2H2O (2) amount of Cr(VI) released in all samples. A recent investiga-
Additionally, Cr(VI) mobilization might be accomplished tion of peroxide decomposition in a number of natural soils
simply physically, as oxidation of organic matter promotes (34) has shown that the principal product of the H2O2 reaction
disintegration of soil structure, thereby increasing access of with ordinary soils is O2, possibly indicating pervasive catalase
the solution to Cr(VI) mineral phases. activity. This implies that most H2O2 introduced into soils
Although we were unable to observe direct evidence for disproportionates without oxidizing soil components. Nev-
formation of chromium(VI) peroxide complexes, several lines ertheless, measurable amounts of hydroxyl radical can be
of evidence favor this complexation phenomenon as an produced (34). At high concentrations, dissolved Cr may
important process at the ore processing site. First, as already catalyze H2O2 decomposition (5, 12). In this process, Cr(VI)
noted, this soil is known to contain sparingly soluble Cr(VI) oxidizes H2O2 to O2 while lower-valent Cr products reduce
salts such as chromatite (CaCrO4). Second, Cr(III) in this soil H2O2 to H2O.
has been found previously to be resistant to other oxidants. It should be kept in mind that our experiments involved
James (11) found that neither the Cr(III) in ore residue soils only one application of peroxide. Most remedial peroxide
nor the soluble Cr(III) added to these soils oxidizes when treatments call for continuous delivery of peroxide over many
exposed to manganese(III,IV) (hydr)oxides, despite favorable days and at much higher concentrations than those used in
conditions (e.g., high pH and low organic matter content). these experiments. The extent and persistence of Cr(III)
Third, if sparingly soluble Cr(VI) salts dissolve through oxidation under such conditions would be much greater than
complexation with H2O2, subsequent decomposition of the we have observed, as would be the possibility of forming
complex would release CrO42- into the soil solutions as H2O2 reactive peroxochromium intermediate species.
levels dropped. In Figure 3, Cr(VI) concentrations appear to Our results raise a cautionary flag regarding the use of
continue rising beyond the first day, by which time all the H2O2 for in situ remediation of contaminated soils associated
free H2O2 is gone, suggesting slow release of CrO42- as with hazardous waste sites. On the other hand, H2O2 leaching
peroxochromates decompose. (At the electroplating site could be attractive under the conditions of ex situ remediation
(Figure 1) this phenomenon, if it existed, was obscured by (i.e., soil washing; 35) provided that recovery of leachates
the rapid removal of Cr(VI) by organic matter after H2O2 was could be assured. Our work shows that peroxide mobilizes
gone.) more anthropogenic Cr than can be obtained with a neutral
On the other hand, complexation of Cr(VI) by peroxide aqueous leach (NaNO3 solution) and that it does not mobilize
cannot be the sole process explaining our results. Oxidation tightly bound Cr in natural forms such as chromite (FeCr2O4).
of Cr(III) by H2O2 is thermodynamically feasible and has been Excess H2O2 decomposes to harmless products, H2O and O2.
demonstrated in Figure 4 over the pH range relevant to our Given the great variability we have observed from one site
soils. A mass balance analysis indicated that some Cr to another, the suitability of H2O2 in any remediation strategy
oxidation must have taken place in the surface horizon at would have to be evaluated site specifically.
the electroplating site, at least at the highest H2O2 treatment
levels. Soluble Cr(VI) extracted from this soil surpassed the Acknowledgments
initial Cr(VI) concentrations reported in Table 1. Further This work was partially supported by the Maryland Water
support for Cr oxidation at the electroplating site is provided Resources Research Center with grants from the U.S.
by the pH, which declined from 5.5 to as low as 4.6 after Geological Survey and from the National Science Foundation
applying H2O2, despite the buffering capacity of the soil. As (DGE9354935).
shown by reaction 1, oxidative dissolution of Cr(III) lowers
pH. However, oxidation of other soil components (e.g.,
Literature Cited
organic matter) could also contribute to acid generation.
(1) In Situ Remediation Technology: In Situ Chemical Oxidation;
A key point demonstrated by the experiments reported
EPA/542/S-98/008; Office of Solid Waste and Emergency
in Figure 4 is that the yield of Cr(VI) obtained by H2O2 Response, U.S. Environmental Protection Agency: Washington,
treatment depends on the history of the Cr(III) starting DC, 1998.
material. The aged Cr(III) solutions used in this study likely (2) Pardieck, D. L.; Bouwer, E. J.; Stone, A. J. Contam. Hydrol. 1992,
contained mixtures of soluble and insoluble oligomers, e.g., 9, 221-242.
Crn(OH)3n-22+ and Crn(OH)3n0 (31). At OH-/Cr(III) ) 3:1, flocs (3) U.S. Environmental Protection Agency. http://www.epa.gov/
superfund/sites/ (Accessed December 2000).
of Crn(OH)3n formed. Figure 4 shows that the 3:1 mixture was (4) Pettine, M.; Millero, F. J. Limnol. Oceanogr. 1990 35, 730-736.
relatively unreactive to H2O2. The soluble Cr(III) products, (5) Perez-Benito, J. F.; Arias, C. J. Phys. Chem. A 1997, 101, 4726-
stable at both lower and higher OH-/Cr(III) ratios, were more 4733.
easily oxidized by H2O2. It is likely that yields of Cr(VI) from (6) James, B. R. Environ. Sci. Technol. 1996, 30, 248A-251A.
soils are also dependent upon such factors as the pH of the (7) Nikolaidis, N. P.; Robbins, G. A.; Scherer, M.; McAninch, B.;
soil, pH of the Cr-bearing contaminant, redox status, and Binkhorst, G.; Asikainen, J.; Suib, S. L. Ground Water Monit.
Remed. 1994, 14, 150-159.
aging time. (8) Mattuck, R. Master’s Thesis, University of ConnecticutsStorrs
Soil organic matter clearly plays important roles in Cr (Technical Report ERI-94.05), 1994.
mobility. At the electroplating waste site, a reducing fraction (9) Burke, T.; Fagliano, J.; Goldoft, M.; Hazen, R. E.; Iglewicz, R.;
of the high levels of organic matter in the surface horizon McKee, T. Environ. Health Perspect. 1991, 92, 131-137.
probably accounted for the decline of soluble Cr(VI) after (10) Weng, C. H.; Huang, C. P.; Allen, H. E.; Cheng, A. H. D.; Sanders,
H2O2 decay. Wittbrodt and Palmer (32) observed the reduction P. F. Sci. Total Environ. 1994, 154 (1), 71-86.
(11) James, B. R. J. Environ. Qual. 1994, 23, 227-233.
of Cr(VI) by soil humic and fulvic acids across a pH range (12) Rock, M. L. Ph.D. Dissertation, University of Maryland, College
of 2-7. Nakayasu et al. (33) found that gallic and tannic acids Park, 1999.
(polyphenols that originate in decaying leaves, likely precur- (13) Davis, A.; Kempton, J. H.; Nicholson, A.; Yare, B. Appl. Geochem.
sors to humic substances) reduced Cr(VI) at pH 5 at even 1994, 9, 569-582.
faster rates than humic and fulvic acids. The refractory (14) Spliethoff, H. M.; Hemond, H. F. Environ. Sci. Technol. 1996,
30, 121-128.
character of Cr at the tannery site is also likely due to organic
(15) Interim Report on the Groundwater Quality of East and North
matter. Woburn, Massachusetts; Contract 68-01-6056, TDD F1-8010-
An issue not answered by our work is the fate of H2O2. The 04B, U.S. Environmental Protection Agency: Washington, DC,
limited data in Table 2 indicate that the half-life of H2O2 in 1981.

4058 9 ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 35, NO. 20, 2001
(16) Thornton, E. C.; Amonette, J. E. Environ. Sci. Technol. 1999, 33, (28) Witt, S. N.; Hayes, D. H. Inorg. Chem. 1982, 21, 4014-4016.
4096-4101. (29) Dickman, M. H.; Pope, M. T. Chem. Rev. 1994, 94, 569-584.
(17) James, B. R.; Bartlett, R. J. J. Environ. Qual. 1983, 12, 169-172.
(18) Pearre, N. C.; Heyl, A. V. Chromite and Other Mineral Deposits (30) Zhang, L. B.; Lay, P. A. Inorg. Chem. 1998, 37, 1729-1733.
in Serpentine Rocks of the Piedmont Upland, Maryland, (31) Stunzi, H.; Marty, W. Inorg. Chem. 1983, 22, 2145-2150.
Pennsylvania and Delaware; Bulleting 1082-K; U.S. Geological (32) Wittbrodt, P. R.; Palmer, C. D. Eur. J. Soil Sci. 1996, 47, 151-162.
Survey: Reston VA, 1960.
(19) Rabenhorst, M. C.; Foss, J. E.; Fanning, D. S. Soil Sci. Soc. Am. (33) Nakayasu, K.; Fukushima, M.; Sasaki, K.; Tanaka, S.; Nakamura
J. 1982, 46, 607-616. H. Environ. Toxicol. Chem. 1999, 18, 1085-1090.
(20) Beck, M. T.; Nagy, I. P.; Székely, G. Int. J. Chem. Kinet. 1991, 23, (34) Petigara B. R. Hydrogen Peroxide Decomposition Mechanisms
881-896. in Natural Soil Systems. Ph.D. Dissertation, University of
(21) Bartlett, R.; James, B. J. Environ. Qual. 1979, 8, 31-35. Maryland, College Park 2000, 182 pp.
(22) Pettine, M.; La Noce, T.; Liberatori, A.; Loreti, L. Anal. Chim. (35) Rock, M. L.; Kearney, P. C.; Helz, G. R. In Pesticide Remediation
Acta 1988, 209, 315-319. in Soils and Groundwater; Kearney, P. C., Roberts, T., Eds.; John
(23) Hug, S. J.; James, B. R.; Laubscher, H.-U. Environ. Sci. Technol. Wiley Sons: Chichester, England, 1998.
1996, 31, 160-170.
(24) James. B. R.; Petura, J. C.; Vitale, R. J.; Mussoline, G. R. Environ. (36) Typrin, L. R. Master’s Thesis, University of Maryland, College
Sci. Technol. 1995, 29, 2377-2381. Park, 1998.
(25) Bowman, R. A. Soil Sci. Soc. Am. J. 1988, 52, 1301-1304.
(26) Frew, J. E.; Jones, P.; Scholes, G. Anal. Chim. Acta 1983, 155, Received for review January 31, 2001. Revised manuscript
139-150. received July 16, 2001. Accepted July 16, 2001.
(27) Funahashi, S.; Uchida, F.; Tanaka, M. Inorg. Chem. 1978, 17,
2784-2789. ES010597Y

VOL. 35, NO. 20, 2001 / ENVIRONMENTAL SCIENCE & TECHNOLOGY 9 4059

You might also like