Sie sind auf Seite 1von 7

ASSESSMENT AND MODELLING OF THE ENVIRONMENTAL

CHEMISTRY AND POTENTIAL FOR REMEDIATIVE TREATMENT


OF CHROMIUM-CONTAMINATED LAND

J.G. FARMER1,∗ , M.C. GRAHAM1 , R.P. THOMAS1 , C. LICONA-MANZUR1,2,


E. PATERSON2 , C.D. CAMPBELL2 , J.S. GEELHOED2, D.G. LUMSDON2 , J.C.L.
MEEUSSEN2 , M.J. ROE2 , A. CONNER3 , A.E. FALLICK4 and R.J.F. BEWLEY5
1 Department of Chemistry, University of Edinburgh, West Mains Road, Edinburgh EH9 3JJ,
Scotland; 2 The Macaulay Land Use Research Institute, Craigiebuckler, Aberdeen AB15 8QH,
Scotland; 3 East of Scotland Water, 4 Marine Esplanade, Edinburgh EH6 7LU, Scotland; 4 Scottish
Universities Research and Reactor Centre, East Kilbride, Glasgow G75 0QF, Scotland; 5 Dames &
Moore, Blackfriars House, St. Mary’s Parsonage, Manchester M3 2JA, England
(∗ author for correspondence, e-mail: j.g.farmer@ed.ac.uk)

Received 11 June 1999; accepted 23 September 1999

Key words: chromium, chromite ore processing residue, speciation

1. Introduction

Significant areas of Rutherglen and Cambuslang to the south–east of Glasgow,


Scotland, have been infilled by chromite ore processing residues arising from a
former chromium works in operation from c. 1830–1968. These areas are heavily
contaminated with hexavalent chromium, a known carcinogen of significant mo-
bility, with concentrations often in the thousands of mg kg−1 . During the 1990s a
series of trials have been undertaken by various organisations to assess the per-
formance of several process-based remedial techniques for addressing hexavalent
chromium, most notably in situ bioremediation, ex situ physicochemical stabil-
isation, and ‘baseline’ treatment involving non-proprietary chemical or biological
methods (Bewley et al., 1998). Such treatments often aim to reduce mobile anionic
hexavalent chromium (e.g. as CrVI O2− 4 ) to cationic trivalent chromium, which is
much less toxic and is strongly retained in the solid phase (e.g. as CrIII (OH)3 )
(James, 1996). The overall aim of this specific project is to provide an integrated
assessment, including linked physical and chemical modelling, of (i) how indigen-
ous physicochemical conditions at such sites, together with a perturbation in any
particular characteristic, especially a change in pH, redox conditions or influx of
organic matter, influence chromium behaviour; (ii) the implications of this for the
selection of an appropriate remedial technique; and (iii) how, in a post-remedial
scenario, residual concentrations of the chromium, or remediation products, and
biological activity will also be influenced by physicochemical conditions and their
perturbations.
Conditions at the landfill sites are still heavily influenced by the nature of the
chromite ore processing that was undertaken at J. & J. White’s Shawfield Chemical

Environmental Geochemistry and Health 21: 331–337, 1999.


© 2000 Kluwer Academic Publishers. Printed in the Netherlands.
332 J.G. FARMER ET AL.

Works in Rutherglen. The ore, in which the chromite, FeCr2 O4 , is in a spinel [(Cr,
Fe, Al)2 O3 (Fe, Mg)O] structure of approximate composition (post-World War I):
Cr2 O3 – 35%, Fe2 O3 – 20%, Al2 O3 – 15%, FeO – 12%, MgO – 12%, was ground
and then mixed with alkali carbonate before roasting in a furnace at 1150◦ C. Up
until 1880 K2 CO3 was used but, thereafter, Na2 CO3 was employed, along with
K2 CO3 until the early 20th century, and then exclusively after World War I.
4 FeCr2 O4 + 8 Na2 CO3 + 7 O2 → 8 Na2 CrO4 + 2 Fe2 O3 + 8 CO2 .
As liquid chromate, molten above 850◦ C, can prevent access of air, a diluent was
added to ensure that there was sufficient air to provide the oxygen necessary for
the oxidation of chromite to chromate. In the high lime process adopted at the
Shawfield Works until the 1950s, the diluent used was quicklime, CaO, in the 19th
century, but dolomite, CaMg(CO3 )2 , thereafter.
The Na2 CrO4 was extracted with water from the melt and H2 SO4 added to yield
the dichromate
2 Na2 CrO4 + H2 SO4 → Na2 Cr2 O7 + Na2 SO4 + H2 O,
and potassium dichromate could be produced by double decomposition:
Na2 Cr2 O7 + 2 KCl → K2 Cr2 O7 + 2 NaCl,
and potassium chromate by:
K2 Cr2 O7 + 2 KOH → 2 K2 CrO4 + H2 O.
At White’s, production of dichromate was ∼6000 tons per year by 1880 and, at
the time of the plant’s closure in 1968, ∼30,000 tons of dichromate equivalent per
year. For each ton of dichromate produced, the high lime process yielded 2.1 tons
(dry weight) of waste residue. The corresponding figure for the low lime process
was 1.15 tons. As no solids left White’s site except for sale or as waste residues
to landfill, it is reckoned that the total of residue landfilled from the sites between
1830 and 1968 must have been about 2,500,000 tons (dry weight).
Of the initial starting materials, about 30% by weight of the chromite ore was
converted to dichromate, the remainder (Fe2 O3 etc) constituting residue for landfill.
This also contained CaO and MgO formed from the dolomite diluent. The alkaline
raw liquor from the leaching stage was treated with acid to a pH of 7.5 to precipitate
out Al(OH)3 , while subsequent acidification to pH 4.5 yielded anhydrous Na2 SO4 ,
which crystallised out. Both of these chemicals constituted part of the chromium
ore processing residue disposed to landfill. After 1921, White’s manufacture of
chromic acid from Na2 Cr2 O7 and H2 SO4 for the chromium plating industry res-
ulted in chromium-contaminated NaHSO4, acidic waste capable of reacting with
the highly alkaline furnace residue in landfill and of solubilising some of the chro-
mates (e.g. CaCrO4 , Na2 CrO4 ) present in the residue. Some of the CaCrO4 in later
years would have been present as a result of slaked lime (Ca(OH)2 ) treatment of
ASSESSMENT AND MODELLING OF THE ENVIRONMENTAL CHEMISTRY 333

the neutral liquor in the melt leaching stage. Also in later years mother liquor
containing NaCl and some K2 Cr2 O7 from its double decomposition production
would have gone to landfill. So there seems little doubt that all process wastes
landfilled certainly contained some chromate in one form or another. In addition,
it appears that some trivalent chromium (not that in undecomposed ore) may have
been present in acid-soluble form in fresh residue (Farquhar, 1998).

2. Materials and methods

Three sites are under investigation in this study: Dukes Road Playing Fields (Site
2), Cambuslang; Rutherglen Glencairn Football Ground (Site 4), Rutherglen; and
Rosebery Park (Site 7), Rutherglen. During the period 15 April to 15 October
1998, soil and groundwater samples were collected at the three sites, along with
groundwater at an additional contaminated site (Myrtle Park, Site 14) and surface
water samples from the Malls Mire Burn and River Clyde at Richmond Park, as
listed in Table I.
A certain amount of each solid sample collected was retained in field moist
condition for speciation analysis. For each sample, the remainder of the material
was then placed on trays in an oven at 30◦ C and, after drying, passed through
a 5.6 mm sieve and homogenised by coning and quartering. Soil solutions were
collected by centrifugation and filtered through 0.45 µm membrane filters. Water
samples were filtered through 0.45 µm membrane filters before transport to the
laboratory.
Solid phase samples (dry) were digested in hot aqua regia for total chromium
and (moist) extracted with 0.28 M Na2 CO3 /0.5 M NaOH for 1 h at 95◦ C for hexa-
valent chromium (James et al., 1995). Total chromium was determined at 267.7 nm
by ICP-OES and hexavalent chromium by UV-visible spectrophotometry at 540 nm
following acidification with H2 SO4 to pH 2 and complexation with 1,5-diphenyl-
carbazide (DPC) (US EPA, 1992). Water samples were acidified (2 M HNO3 ) prior
to analysis for total chromium.
Solid-state characterisation was carried out using X-ray diffraction (XRD), infra-
red spectroscopy (IRS), and scanning electron microscopy(SEM)/energy dispers-
ive X-ray microanalysis (EDX).
The modelling approach is based on compilation of a set of chemical equilib-
rium and kinetic reactions, relevant for the behaviour of chromium under natural
conditions, from literature data and then using the set of reactions to define a
computer model within the modelling framework ORCHESTRA (Object Repres-
entation of Chemical Speciation and Transport) developed at MLURI.

3. Results

The pH at the sites ranged from 7.1 to 12.5. A summary of the chromium con-
centration data is shown in Table I. Total chromium concentrations in the solid
334 J.G. FARMER ET AL.

TABLE I
Summary of pH and chromium concentrations at the contaminated sites

Site Sample n pH Cr(VI) Cr


(mg kg−1 ) (mg kg−1 )

2 ‘Soil’ 20–40 cm (9) 11.8 1250 ± 12 16200 ± 700


yellow/green nodule (2) 36400 ± 700
grey/green nodule (2) 14700 ± 400
brown nodule (2) 16400 ± 600
Soil solution (2) 11.6 17.8 ± 0.1 24.6 ± 0.1
Groundwater 3 7.1–8.6 2.4–30.0 2.4–30.0

4 ‘Soil’ terracing (0–50 cm) 5 8.9–12.1 290–4670 13100–25000


adjacent ground (0–30 cm) 3 10.9–11.0 520–1950 5800–14500
Soil solution (terracing 15–50 cm) (2) 11.9 79.5 ± 0.2 125 ± 1
Groundwater 2 7.7–12.5 0–1.6 0–15.2

7 ‘Soil’ terracing (5–8 cm) (12) 8830 ± 180


purple nodule 1 22500
pink layer (2) 8810 ± 330
orange layer (2) 16700 ± 700
terracing (8–15 cm) (2) 11.9 8500 ± 20 21600 ± 1100
terracing (15–40 cm) (2) 12.3 3520 ± 40 17600 ± 1100
Soil solution (terracing 8–15 cm) (2) 11.7 26.3 ± 0.1 37.7 ± 0.1
(terracing 15–40 cm) (2) 12.3 12.1 ± 0.1 16.7 ± 0.2
Groundwater 1 7.5 <0.01 <0.01

14 Groundwater 3 11.7–12.2 7.4–33.3 7.4–33.3


Surface waters 3 7.8–8.0 0.2–6.5 0.2–6.5

Brackets ( ) denote replicate analyses of the same sample. Concentrations are in mg kg−1 dry
weight for solid phase samples and in mg l−1 for aqueous samples.

phase were as high as 2.50% (w/w, dry weight), with individual coloured nodules
reaching 3.64%. The base extraction removed 1.4–32.3% (mean 12.0 ± 9.4%) of
the total chromium, with 48.8–97.5% (mean 80 ± 15%) of that extracted occurring
in the hexavalent form. Soil solution chromium concentrations ranged from 16.7
to 125 mg l−1 , with an average 69.6 ± 4.2% apparently in the hexavalent form.
Groundwater and surface water chromium concentrations were as high as
33.3 mg l−1 , all in the form of Cr(VI) with the exception of one sample at Site 4.
In the bulk of the waste material, phases detected by XRD and IRS included
chromite (FeCr2 O4 ), magnesiochromite (MgCr2 O4 ), Mg-containing minerals
(MgO, a high-temperature alteration product, and Mg(OH)2 , an in situ product),
Ca-containing minerals (Ca(OH)2 , CaCO3 , CaMg(CO3 )2 ) and SiO2 . The SEM/
ASSESSMENT AND MODELLING OF THE ENVIRONMENTAL CHEMISTRY 335

EDX work yielded a back-scattered electron image of a chromite particle, in an


aggregate from Site 2, etched outside the central core and with a rim of differ-
ent composition. The chemistry of the rim was characterised by a very high cal-
cium content, with significantly less chromium relative to iron and little or no
magnesium when compared with the chemistry of the core.
Initial modelling has shown that the processes that are important at the studied
sites are (i) the speciation of Cr(III) and Cr(VI) in solution, (ii) the precipitation and
dissolution of solid phases of Cr(III) and Cr(VI), (iii) reaction kinetics, and (iv)
reduction, especially from the remediative treatment viewpoint (Geelhoed et al.,
1999).

4. Discussion

The problem of chromium contamination and continuing release of Cr(VI) to re-


ceiving waters (Whalley et al., 1999) is exemplified by the data of Table I. Although
the base digest apparently reveals the presence of hexavalent chromium in the
solid material, the specific source of the continuing supply of Cr(VI), however,
is not immediately clear. One might expect that chromate salts (Cr(VI)) present
in the residue at the time of deposition would have been readily leached out be-
fore now, especially as the pH dependence of Cr(VI) adsorption indicates that
adsorption of chromate on amorphous iron and aluminium (hydr)oxides in the
processing residue is not an important process under the prevailing conditions
(Mesuere and Fish, 1992). In chromite and magnesiochromite, chromium is present
as Cr(III). Another Cr(III) solid phase which may be present is a solid solution
of amorphous Cr(III) and Fe(III) hydroxide. The aqueous Cr(III) concentration in
equilibrium with the solid solution is even lower than for Cr(III) hydroxide (Sass
and Rai, 1987). Furthermore, while oxidation of solution-phase Cr(III) to Cr(VI)
can take place at a considerable rate at low pH and small Cr(III) concentration
using Mn-oxide as oxidising agent, Cr(III) will precipitate and become unavailable
for oxidation at the higher pH (>5) and potentially larger Cr(III) concentrations at
these sites (Fendorf and Zasoski, 1992).
As modelling suggests that Cr(VI) will not be produced under ambient condi-
tions (i.e. the Cr(VI) pool may be finite), it is important to establish the form of the
Cr(VI)-containing phases now existing in order that they may be represented in the
model and with a view to determining the most effective means of restricting the
mobility of Cr(VI). In the waste material itself, solid phases have been identified
that are not in equilibrium with the environmental conditions at the sites. These
phases may have been formed at the very high temperature used in the ore pro-
cessing and may still be present because of slow dissolution kinetics, although the
possibility of alterations occurring at the sites cannot yet be ruled out. In order to
establish whether the outer rim of chromite particles, perhaps containing Cr(VI)
in a slightly less soluble form, is the continuing supply of Cr(VI), solid-phase
336 J.G. FARMER ET AL.

speciation techniques such as X-ray photoelectron spectroscopy (Olazabal et al.,


1997; Szulczewski et al., 1997) will be tried.

5. Ongoing and future work

1. Proprietary and non-proprietary remediation of soil/waste material (Bewley


et al., 1998).
2. Application of solid-phase characterisation techniques to the investigation of
unremediated and remediated material.
3. Speciated isotope dilution mass spectrometry experiments (Huo et al., 1998;
Kingston et al., 1998) to investigate the chromium speciation integrity of
various extraction methods and subsequent DPC complexation for both un-
remediated and remediated material.
4. Batch equilibration experiments to yield thermodynamic and kinetic inform-
ation prior to column experiments.
5. Column experiments (Weng et al., 1994) to simulate changing conditions
for modelling purposes and to provide reaction media for non-proprietary
remediation.
6. Development of the computer model, incorporating solid-phase information,
results from column experiments and ways of representing remediation treat-
ments.
7. Development and application of bioassay methods and ecotoxicological
testing.

Acknowledgement

This work is supported by the U.K. Natural Environment Research Council UR-
GENT (Urban Regeneration and the Environment) Thematic Programme under
Grant No. GST/02/1991.

References

Bewley, R.J.F., Jeffries, R., Watson, S. and Sansom, M.: 1998, Field trials for evaluation of remedial
technologies for chromium-contaminated soils in Glasgow. In: H.R. Fox, H.M. Moore and A.D.
McIntosh (eds), Land Reclamation: Achieving Sustainable Benefits, Balkema, Rotterdam, pp.
457–463.
Fendorf, S. and Zasoski, R.J.: 1992, Chromium(III) oxidation by δ-MnO2 . 1. Characterization,
Environmental Science and Technology 26, 79–85.
Geelhoed, J.S., Meeussen, J.C.L., Lumsdon, D.G., Roe, M.J., Thomas, R.P., Farmer, J.G. and
Paterson, E.: 1999, Processes determining the behaviour of chromium in chromite ore processing
residue used as landfill, Land Contamination and Reclamation 7, 271–279.
ASSESSMENT AND MODELLING OF THE ENVIRONMENTAL CHEMISTRY 337

Huo, D., Lu, Y. and Kingston, H.M.: 1998, Determination and correction of analytical biases and
study on chemical mechanisms in the analysis of Cr(VI) in soil samples using EPA protocols,
Environmental Science and Technology 32, 3418–3423.
James, B.R.: 1996, The challenge of remediating chromium-contaminated soil, Environmental
Science and Technology 30, 248A–251A.
James, B.R., Petura, J.C., Vitale, R.J. and Mussoline, G.R.: 1995, Hexavalent chromium extraction
from soils: a comparison of five methods, Environmental Science and Technology 29, 2377–2381.
Kingston, H.M., Huo, D., Lu, Y. and Chalk, S.: 1998, Accuracy in species analysis: speciated iso-
tope dilution mass spectrometry (SIDMS) exemplified by the evaluation of chromium species,
Spectrochimica Acta Part B 53, 299–309.
Mesuere, K. and Fish, W.: 1992, Chromate and oxalate adsorption on goethite. 1. Calibration of
surface complexation models, Environmental Science and Technology 26, 2357–2364.
Olazabal, M.A., Nikolaidis, N.P., Suib, S.A. and Madariaga, J.M.: 1997, Precipitation equilibria
of the chromium(VI)/iron(III) system and spectroscopic characterization of the precipitates,
Environmental Science and Technology, 31, 2898–2902.
Sass, B.M. and Rai, D.: 1987, Solubility of amorphous chromium(III)–iron(III) hydroxide solid
solutions, Inorganic Chemistry 26, 2228–2232.
Szulczewski, M.D., Helmke, P.A. and Bleam, W.F.: 1997, Comparison of XANES analyses and
extractions to determine chromium speciation in contaminated soils, Environmental Science and
Technology 31, 2954–2959.
US EPA: 1992, Colorimetric method for the determination of chromium(VI) in water, ‘soil’ extracts
and digests. USEPA Method 7196A (SW-846, 1992). U.S. Environmental Protection Agency,
Washington.
Weng, C.H., Huang, C.P., Allen, H.E., Cheng, A.H.-D. and Sanders, P.F.: 1994, Chromium leach-
ing behavior in soil derived from chromite ore processing waste, The Science of the Total
Environment 154, 71–86.
Whalley, C., Hursthouse, A., Rowlatt, S., Iqbal-Zahid, P., Vaughan, H. and Durant, R.: 1999, Chro-
mium speciation in natural waters draining contaminated land, Glasgow, U.K., Water, Air, and
Soil Pollution 112, 389–405.

Das könnte Ihnen auch gefallen