You are on page 1of 11

Land Contamination & Reclamation, 18 (1), 2010 DOI 10.2462/09670513.

909

2010 EPP Publications Ltd

Sequestration of heavy metals in soils from two polluted industrial sites: implications for remediation
Krishna R. Reddy, Swapna Danda, Yeliz Yukselen-Aksoy and Ashraf Z. Al-Hamdan

Abstract This study was conducted to determine the distribution of heavy metals in two different contaminated field soils, and to assess its influence on remedial performance. The two field soils were obtained from two different polluted industrial sites in the Metropolitan Chicago area and were characterized for physical and chemical properties. The soils were referred to as silty sand and silty clay, based on their particle-size distribution. A five-step sequential liquidsolid extraction procedure was used to speciate the heavy metals into: (1) easily exchangeable; (2) carbonate bound; (3) Fe/Mn oxide bound; (4) organic bound; and (5) residual fractions. These results showed that all the metals except mercury were predominantly distributed as the Fe/Mn-bound form in silty sand. In contrast, they were distributed as the residual form in silty clay. Results indicated that the metals in both soils were in the forms that are not easily amenable to soil washing using water. Therefore, batch extraction tests with varying concentrations of ethylenediaminetetraacetic acid (EDTA) (0.05, 0.1 and 0.2 M) and acetic acid (0.5, 0.1 and 2 M) were conducted to identify an optimum soil-washing-based remediation system for both silty sand and silty clay. Ethylenediaminetetraacetic acid was found to be effective for the remediation of silty sand, while acetic acid was found to be suitable for the removal of metals that existed in residual form in silty clay. Based on these results, total removal efficiency was found to follow the order: sand + EDTA > clay + acetic acid > sand + acetic acid > clay + EDTA. This study showed that the speciation and remediation of heavy metals in soils depend on the sitespecific soil composition, and this should be carefully considered in the selection of an efficient remedial method. Key words: acetic acid, distribution, EDTA, heavy metals, remediation, sequential extraction, soil, soil washing

INTRODUCTION

The toxicity and potential mutagenity of heavy metals,


Received Oct. 2008; revised and accepted Nov. 2009 Authors Krishna R. Reddy,1* Swapna Danda,2 Yeliz Yukselen-Aksoy3 and Ashraf Z. Al-Hamdan4 1. Professor, Dept. of Civil and Materials Engineering, University of Illinois at Chicago, 842 West Taylor Street, Chicago, Ill. 60607, USA. Email: kreddy@uic.edu. 2. Graduate Research Assistant, Dept. of Civil and Materials Engineering, University of Illinois at Chicago, 842 West Taylor Street, Chicago, Ill. 60607, USA. 3. Assistant Professor, Dept. of Civil Engineering, University of Celal Bayar, Muradiye Campus, 45100 Manisa, Turkey. 4. Lecturer, Dept. of Civil and Environmental Engineering, University of Alabama in Huntsville, 301 Sparkman Drive, Huntsville, Ala. 35899, USA. *Corresponding author

such as arsenic, lead, chromium, cadmium, copper, nickel and zinc in the environment have been well established. According to the United States Environmental Protection Agency (USEPA), there are over 200 000 heavy-metal-contaminated sites in the United States that require urgent cleanup to protect public health and the environment (USEPA 1997). Among the physical and biological processes in the subsurface soils, geochemistry plays a major role in the distribution, speciation, as well as the remediation potential of heavy metals. Fine-grained soil particles are geochemically active, with the soil-mineral surface sites possessing negative or positive charge or being electrically neutral (Evans 1989). Oppositely charged metallic counter-ions from the pore water are attracted to these charged surfaces, depending on the degree of acidity or

13

Land Contamination & Reclamation / Volume 18 / Number 1/ 2010

alkalinity, mineralogical composition, and organic content of the soil. For example, although kaolinite is the least active clay (Suraj et al. 1998), adsorption of metals depends on the pH of the soil and pore water system; metal adsorption is usually accompanied by the release of hydrogen (H+) ions from the edge sites of the soil minerals (Miranda-Trevino and Coles 2003). Among the currently available technologies for soil remediation, soil washing is considered to be a fast and efficient process. Soil washing generally utilizes extracting solutions of acids, bases, chelating agents or other additives (Griffiths 1995). In practice, acid- and chelate-enhanced soil washings are the two most prevalent methods for the removal of metals from soils (Rampley and Ogden 1998). Ethylenediaminetetraacetic acid (EDTA) is the most commonly used chelating agent in soil washing because of its strong chelating ability with different heavy metals such as Pb, Zn, Cu and Cd in contaminated soils (Nowack 2002). A detailed review on the efficiency of EDTA for the remediation of metal-contaminated sites is presented by Peters (1999). The development of soil-washing remediation technology to clean up contaminated soils requires a good understanding of contaminant distribution and retention within the soilwater system. An evaluation of the relative ease of removal of a contaminant from the soil can be made if one obtains a better understanding of how the contaminants are held to or within the soil constituents. The contaminant retention mechanisms for the soil constituents differ, depending on the mineral type, amorphous material, soil organic matter, and the type and nature of pollutants. Contaminant speciation and distribution will help to assess the contaminantretaining mechanisms and will provide a rational approach to engineer the chemistry for enhanced contaminant remediation. Sequential extraction procedures have been widely used to quantify the distribution of heavy metals in contaminated soils (Tessier et al. 1979; Hall et al. 1996). These extraction procedures are operationally defined schemes. The mobility of contaminants in soils depends strongly on their well-defined chemical species (i.e., specific chemical forms). However, welldefined speciation of contaminants in soils is difficult and often virtually impossible to obtain. Therefore, sequential extraction procedures are commonly used because they provide information about the fractiona14

tion of metals binding forms, which can be a good compromise to give data interpretations and, in some instances, sufficient information for contamination risk assessment (Pueyo et al. 2001; Margui et al. 2004). To ensure appropriate geochemical interpretation, it is crucial to combine the application of sequential extraction procedures with detailed soil mineralogical studies. The objective of this study was to evaluate the effect of soil composition on the distribution of heavy metals in contaminated field soils and the influence of such distribution on soil washing remediation. The contaminated soils obtained from two different polluted industrial sites (a silty sand and a silty clay) were characterized for their physical and chemical properties. The silty sand was contaminated by large quantities of heavy metals as well as PAHs, and it was high in organic matter, whereas the silty clay was a clayey soil contaminated with only heavy metals. X-ray diffraction analyses were performed to determine the mineral phases present in both soils. A five-step sequential extraction analysis was performed to determine the distribution of heavy metals in the soils. A series of batch experiments were conducted to assess the effect of EDTA and acetic acid on the enhanced removal of metals from the soils.

MATERIALS AND METHODS Soil characterization methods

Soil samples were collected from two polluted industrial sites in the Metropolitan Chicago area. The grainsize distribution, Atterberg limits, organic content, specific gravity, hydraulic conductivity and pH of the samples were determined according to the American Society for Testing and Materials (ASTM 2006) standard testing procedures. The soils were also analysed for total metals using USEPA Method 6020, except mercury, which was analysed using USEPA Method 7470A. Buffering capacity of the soils was determined by titration, using 0.1 M and 0.5 M nitric acid as the titrant solution for silty sand and silty clay, respectively. High nitric acid concentration was chosen for silty clay because of its high buffering capacity. Soil slurry (20 g dry weight in 200 mL deionized water) was titrated incrementally and mixed using a magnetic stirrer. The deionized water sample was also tested simultaneously as a control sample. The equilibrium pH of

Sequestration of heavy metals in soils from two polluted industrial sites: implications for remediation

the slurry was measured with a pH meter (Thermo Orion model 720 A). The mineral composition of the soils was determined by X-ray diffraction analysis using a Siemens D5000 diffractometer (New York, NY). The soil samples were crushed to a powder with a mortar and pestle, taking care not to grind them. In order to detect any clay minerals that were present in the sample, a suspension was prepared by mixing dry soil with deionized water and stirring it vigorously. This suspension was then placed in an ultrasonic bath for a minute to break up any remaining large particles. The solids were allowed to settle for 20 minutes and liquid was drawn from the top centimetre of the suspension using an eyedropper to isolate particles with an equivalent spherical diameter less than or equal to 2 micrometres. The liquid was applied to a slide and allowed to evaporate, leaving behind a dusting of clay particles for X-ray diffraction analysis.
Sequential extraction procedure

Step 3 (FeMn oxides-bound form) the residual soil sample from Step 2 was treated with 100 mL of 0.04 M hydroxylamine hydrochloride (NH2OH. HCl) in 25% (v/v) acetic acid and heated to 96C, with occasional agitation for 6 hours. The suspension was centrifuged (30 mins, 10 000 rpm) and the supernatant was analysed for metals. The soil residue was washed with 40 mL of deionized water as before. Step 4 (organic-bound form) 15 mL of 0.02 M nitric acid (HNO3) and 25 mL of 30% hydrogen peroxide (H2O2) (pH adjusted to 2.0 with nitric acid) were added to the residual soil sample and heated to 85C for 2 hours with occasional agitation. Then, 15 mL of 30% hydrogen peroxide (H2O2) (pH adjusted to 2.0 with nitric acid) were added and heated again to 85C for 3 hours with occasional agitation. The mixture was cooled to room temperature of 26C and 25 mL of 3.2 M ammonium acetate (NH4OAc) in 20% (v/v) HNO3 were added. Finally, the sample was diluted to 100 mL with deionized water and mixed continuously for 30 mins. The suspension was centrifuged (30 mins, 10 000 rpm) and the supernatant was analysed for metals. Step 5 (residual form) the sum of the four fractions was subtracted from the total metal concentration as determined by USEPA methods 6020 or 7470A (for mercury) to determine the metals present in unextractable residual form.
Soil-washing experiments

The sequential extraction procedure developed by Tessier et al. (1979) was used to determine the speciation of the metal forms in each soil. This procedure speciates the heavy-metal distribution into an easily extractable (exchangeable) form, a form bound to carbonates, a form bound to FeMn oxides, a form bound to organics, and a residual form. The sequential extraction of silty sand and silty clay was performed in duplicate using 5 g of dry soil in the following five steps: Step 1 (exchangeable form) 5 g of soil sample were continuously mixed and reacted with 40 mL of 1 M sodium acetate solution (pH adjusted to 8.2 with acetic acid) for one hour. Then the suspension was centrifuged (30 mins, 10 000 rpm) and the supernatant was analysed for metals. The soil residue was mixed with 40 mL of deionized water and centrifuged (30 mins, 10 000 rpm) and the supernatant was discarded to remove the added chemicals. Step 2 (carbonate-bound form) the residual soil sample from step 1 was again treated with 40 mL of 1 M sodium acetate (pH adjusted to 5.0 with acetic acid) and mixed for 5 hours. The suspension was centrifuged (30 mins, 10 000 rpm) and the supernatant was analysed for metals. The remaining soil residue was then washed with 40 mL of deionized water as before.
15

The soil-washing experiments were conducted in batch mode using varying concentrations of two extractants, di-sodium ethylenediaminetetraacetic acid (Na2EDTA) and acetic acid. Five grams of dry soil (silty sand or silty clay) were placed in a 120 mL glass vial with 20 mL of extractant solution of varying concentration (0.05, 0.1, and 0.2 M Na2-EDTA, or 0.05, 0.1, and 0.2 M acetic acid). The glass vial was first shaken by hand vigorously for five minutes to ensure a homogeneous mixture, and then shaken on a gyratory shaker for 24 hours. The soil mixture was then centrifuged (30 mins, 4000 rpm), and the supernatant was collected by filtering through a 110-m diameter Whatman filter paper. The collected supernatant was stored in 40-mL glass vials and analysed by inductively coupled plasma (ICP) spectrometry for the total metals according to the USEPA methods 6020 and 7470A (for mercury). The

Land Contamination & Reclamation / Volume 18 / Number 1/ 2010

removal efficiency of each heavy metal was then calculated based on its initial mass in the soil prior to extraction and its final mass in the supernatant after extraction. All of the soil-washing experiments were conducted in duplicate.

RESULTS AND DISCUSSION Soil properties

Figure 1 shows the particle-size distributions and Table 1 summarizes the physical and chemical properties of

the two soil samples. Based on the particle-size distribution, these two soils are referred as silty sand and silty clay. The d-spacing values from X-ray diffraction suggested the presence of quartz and traces of kaolinite, illite and dolomite minerals in silty sand; while illite, kaolinite and chlorite were the main minerals present in silty clay. The silty sand mainly consisted of fine sand (99% sand-fraction) and had an average hydraulic conductivity of 1.6 104 cm/s, while silty clay was dominated by fines composed of silt and clay (99.5% fines) and had an average hydraulic conductiv-

Table 1. Properties of the two polluted industrial site soils Property Mineralogy Test method X-ray diffraction Silty sand Quartz Trace kaolinite Trace illite % gravel = 0.7 % sand = 99 % fines = 0.3 Non-plastic Silty clay 84% illite 11% kaolinite 5% chlorite % gravel = 0 % sand = 0.5 % fines = 99.5 LL = 50% PL = 24% PI = 26% Silty clay, CH 2.63% 2.52 cm/s 2.1 108 cm/s 7.68.9

Grain-size distribution

ASTM D422

Atterberg limits

ASTM D4318

USCS classification Organic content Specific gravity Hydraulic conductivity pH

ASTM D2488 ASTM D2974 ASTM D854 ASTM D2434 ASTM D2974

Poorly graded sand/silty sand, SP-SM 9.2% 2.63 1.6 10


4

7.08.1

LL= Liquid limit; PL= Plastic limit; PI = Plasticity index

100 90 80 PERCENT FINER 70 60 50 40 30 20 10 0 100.000 10.000 1.000 0.100 0.010 0.001 Silty Clay Silty Sand

GRAIN SIZE (mm)

Figure 1. Particle-size distributions of two polluted industrial site soils

16

Sequestration of heavy metals in soils from two polluted industrial sites: implications for remediation

9 8 7 6 5
0.1 M HNO3 Used for Titration

0.1M HNO3 Used for Tritration Silty Sand Control

pH
4 3 2 1 0 0 2 4 6 8 10 12 14

16

18

20

Volume of Acid Added (mL)

Figure 2a. Buffering capacity of silty sand, using 0.1 M nitric acid

10 9 8 7 6 0.5M HNO3 Used for Titration

pH

Silty Clay 5 4 3 2 1 0 0 20 40 Control

Volume of Acid Added (mL)

60

80

100

120

140

Figure 2b. Buffering capacity of silty clay, using 0.5 M nitric acid

ity of 2.1 108 cm/s. The organic content of silty sand was found to be higher (9.2%) than silty clay (2.63%). Figures 2a and b show the titration results, which demonstrate that the silty sand possesses a low buffering capacity of 0.012 M and the silty clay possesses high buffering capacity of 0.184 M. Table 2 summarizes the concentrations of different metals found in the soils. Both silty sand and silty clay had elevated concentrations of heavy metals.
17

Sequential extraction results

Figure 3 shows the sequestration of heavy metals in different phases of silty sand. The amount of exchangeable metals in the sequential extraction was quite low. Heavy metals that were detected include Sb, As, Ba, Be, Cd, Cu, Mn, Ni, and Zn. The relatively high concentrations of Na and K were released in Step 1 of the extraction procedure. In contrast, low concentrations of Al, Sb, As, Ba, Be, K, Co, Se, Ag, Na, and Tl were

Land Contamination & Reclamation / Volume 18 / Number 1/ 2010

Table 2. Heavy metals found in two polluted industrial site soils Heavy metals Aluminium Arsenic Barium Beryllium Cadmium Calcium Chromium Cobalt Copper Iron Lead Magnesium Manganese Mercury Nickel Potassium Silver Sodium Vanadium Zinc NA = Not available Concentration in silty sand (mg/kg-dry) 1640 3.81 197 0.3 8.39 47 000 19.5 4.33 263 25 000 1160 19 800 294 9.31 18.6 201 1.48 226 3.53 1100 Concentration in silty clay (mg/kg-dry) 11 900 8.58 66.5 0.697 NA 20 000 20.1 16.2 20.4 21 300 11.7 13 300 718 0.042 35.1 2860 NA 341 20 40.6

found in the carbonate-bound fraction. With the exception of mercury, all the metal levels were relatively high in the Fe/Mn-bound fraction; accounting for more than 50% of the total. Except for Fe and Hg, all the other heavy metals were bound with organic matter. Aluminium, As, Cu, Hg, Tl, and Ag were the only metals not extracted from the soil. All of the Hg was present in the residual form (no mercury was detected in any of the preceding extractions). Copper was found to be 65% in the residual form and Ag was 57% in the residual form. Some of the other metals present in the residual form include Al (10%), As (15%) and Tl (30%). Overall, the order of distribution of heavy metals in silty sand can be represented as Fe/Mn bound > organic-matter bound > exchangeable form > residual form > carbonate-mineral bound. The sequential extraction results for silty clay were different from those of silty sand. Results in Figure 4 show that very few metals were extracted in Step 1, indicating that the metals were strongly sorbed to the soil particles. This was likely to be due to the high clay content of the soil, which is known to strongly bind
18

metals. Even the release of Na and K was reduced significantly under extraction Step 1. Arsenic was not detected in Step 1 supernatant, while Hg was detected, indicating that some exchangeable Hg was present. In the Step 2 extraction, small but measurable amounts of most of the metals were released, indicating the presence of carbonate-bound metals (Figure 4). Comparatively more metals were found to be bound with the Fe/ Mn minerals. Only small amounts of some metals were found in the organic-matter-bound fraction (Figure 4). Without exception, the amount of metals present in the residual form was higher than that observed in any of the extractions. The distribution of heavy metals in the silty clay was: residual form > Fe/Mn bound > organicmatter bound > carbonate-mineral bound > exchangeable form. These sequential extraction results show clearly that the removal of metals from the silty sand may be relatively easy as compared to the silty clay. Despite the presence of organic matter, nearly all metals (except Hg, Cu, Ag and Tl) can be removed from the silty sand using a suitable extractant, because of the low clay con-

Sequestration of heavy metals in soils from two polluted industrial sites: implications for remediation

100% 90% 80% 70% 60% 50% 40% 30% 20% 10% 0%
Al um i An num tim o Ar ny se n Ba ic Be rium ry Ca llium dm i Ca um Ch lciu ro m m iu m Co ba Co lt pp er Iro n M ag Lea ne d M an sium ga ne s M er e cu r N y Po ick e ta ss l Se ium le ni um Si lv So er d Th ium a Va lliu na m di um Zi nc

Percent Fractions

Residual Organic bound Fe-Mn bound Carbonates bound Exchangeable

Figure 3. Sequestration of metals in silty sand

100% 90% 80% 70%

Percent Fractions

Residual
60% 50% 40% 30% 20% 10% 0%

Organic bound Fe-Mn bound Carbonates bound Exchangeable

l ny nic um um um ium ium balt per ron ead ium ese ury ke ium ium lver ium ium ium inc um I Z p n c L es in imo rse ari ylli mi o rc Nic ss len Si od all ad r d Cal rom C ga Me B n Co S Th an A ta Se um Ant Be Ca ag an V Al Ch Po M M

Figure 4. Sequestration of metals in silty clay

tent of the silty sand. On the other hand, most of the metals in the silty clay exist mainly in the residual form except for Be, Co, Mg, Hg, Se, Mn and Zn, which were in exchangeable, carbonate bound, FeMn bound and organic-bound forms. The percentages of the extractable forms were as follows: Be 82%; Co 60%; Mg 45%; Hg 45%; Se 100%; Mn 70%; and Zn 63%. This is attributed to the clayey content, which provides a greater reactive surface area for the metals to bind. The clay minerals in the silty clay react with the metals
19

to form stable complexes and oxides, which are strongly held within the soil.
Soil washing

Soil washing with EDTA Figure 5 compares the extraction of heavy metals from silty sand using EDTA at three different concentrations of 0.05 M, 0.1 M and 0.2 M. Generally, the extraction of heavy metals was observed to increase with increasing EDTA concentration. Removal of Co increased

Land Contamination & Reclamation / Volume 18 / Number 1/ 2010

Removal Efficiencies with EDTA: Silty Sand


100

EDTA-0.05M EDTA-0.1M

80

EDTA-0.2M

% Removal

60

40

20

Figure 5. Removal of metals from silty sand using EDTA at different concentrations

from 42% to 51% as the EDTA concentration increased from 0.05 M to 0.2 M. The removal of Fe increased from 21% to 71% as the EDTA concentration increased from 0.05 M to 0.2 M. Similarly, the removal of Ca, Mg, Ni and V increased significantly with an increase in EDTA concentration. Under the concentration range of EDTA used, Pb, Cd, Zn and Cu were completely removed from the silty sand. The removal efficiencies

of Zn, Pb and Cu were slightly greater than 100%. This error could be attributed to the dilution factor of the highly concentrated samples of Pb, Zn and Cu. The highly concentrated samples are usually diluted before analysis so that their concentrations fit within the standard calibration limits. The inductively coupled plasma (ICP) spectrometry used to determine the metal concentrations in this study is highly sensitive and

Removal Efficiencies with EDTA: Silty Clay 100

80
EDTA-0.05M EDTA-0.1M

% Removal

60

EDTA-0.2M

40

20

Figure 6. Removal of metals from silty clay using EDTA at different concentrations

20

Sequestration of heavy metals in soils from two polluted industrial sites: implications for remediation

Removal Efficiencies with Acetic Acid: Silty Sand 100


HAc-0.5M HAc-1M HAc-2M

80

% Removal

60

40

20

Figure 7. Removal of metals from silty sand using acetic acid at different concentrations

Removal Efficiencies with Acetic Acid: Silty Clay 100


HAc-0.5M HAc-1M HAc-2M

80

% Removal

60

40

20

Figure 8. Removal of metals from silty clay using acetic acid at different concentrations

capable of the determination of a range of metals at concentrations below one part in 1012. The maximum removal of Hg, Co and Ni from silty sand was 5, 65 and 79%, respectively. This indicates that EDTA has a low affinity for the removal of residual metals. Overall, 0.2 M EDTA was an effective extractant for the remediation of nearly all the metals from the highly contaminated silty sand. This result indicated that the metals are present in removable form (i.e. exchangeable, carbonate bound, Fe/Mn bound and organic bound). They
21

are effectively removed by chelation extraction due to the formation of more stable metalEDTA complexes as compared to those metals that existed in residual form. Figure 6 summarizes the results of batch extraction experiments on silty clay using EDTA of varying concentrations (0.05, 0.1, and 0.2 M). These results show that the removal efficiencies were lower for silty clay as compared to silty sand (Figure 5). Lead was not removed under the concentration range of EDTA from

Land Contamination & Reclamation / Volume 18 / Number 1/ 2010

the soil. Removal percentages of Hg, Zn, As and Ag change in the range from 1.4% to 5%. Among the metals, removal of Co (from 41% to 83%), Fe (from 18% to 40%), Ni (from 17% to 36%) and Cu (from 46% to 48%) was achieved. This may be due to the fact that in silty clay most of the metals were present in the residual form, tightly attached to the clay minerals and thus not available for EDTA to form metal complexes. Another factor interfering with the metal complexation may be the high buffering capacity of the soil. At high pH, the stability of metalEDTA complexes in silty clay may be decreased due to the sorption of metal EDTA complexes. In contrast, appreciably high removal of copper and cobalt suggests that the metal EDTA complexes of the same charge but different structure have completely different adsorption behaviours. This may be the reason for the high removal of cobalt and copper in comparison to Pb, Ni, Hg and Zn. Soil washing with acetic acid The initial pH of the silty sand was 7.05 and that of the silty clay was 7.56. It is possible that under such pH conditions, metals may have precipitated, leading to low removal. Therefore, the effect of acidification was investigated by using acetic acid at different concentrations of acetic acid (0.5, 1 and 2 M) in both soils. The pH of the silty sand decreased considerably when acetic acid was added. The pH of the soil was reduced to 5.1, 4.8 and 4.35 with 0.5 M, 1 M and 2 M acetic acid, respectively. Figure 7 shows that the removal efficiencies were higher at an acetic acid concentration of 1 M (for Sb, As, Be, Cd, Mn, Hg, Se, Ag and Tl), whereas the removal was lower for the acetic acid concentrations of 0.5 M and 2 M. The decrease in pH with 0.5 M acetic acid was not adequate to solubilize heavy metals, leading to low metal removal. On the other hand, at a higher acetic acid concentration (2 M), soil pH was lowered (to <4.8), which may have led to a more positive surface charge on the soil surfaces. The positively charged silty sand surfaces probably adsorbed acetic acid ligands or negatively charged heavy-metalacetic acid complexes, resulting in comparatively low removal of metals. The results indicate that the mechanism of acetic acid enhancement of the extraction of heavy metals from the silty sand was ion exchange, dissolution and counterion binding. The pH of 4.8 was an ideal value at which most of the metal acetate complexes seem to be mobile. Therefore, 1 M
22

acetic acid concentration resulted in better removal of metals as compared to the other two concentrations. The removal efficiencies were higher with EDTA than acetic acid. The pH of the silty clay also decreased with the addition of acetic acid. The pH of the soil decreased to 5.3, 4.9 and 4.56 with the addition of 0.5 M, 1 M and 2 M acetic acid, respectively. Figure 8 shows the metal removal efficiencies of acetic acid at different concentrations. The removal efficiencies were higher with 2 M acetic acid, suggesting that lowering the soil pH to 4.5 enhances the release of metals strongly bound with the silty clay. Further, it is found that higher removal efficiencies of metals can be achieved with acetic acid as compared to EDTA. This may be due to the presence of illite and kaolinite in the silty clay, minerals which may undergo a change in surface polarity with the addition of acetic acid. Thus, as pH decreases, metal release is maximized due to formation of stable metal-acetate ions that may not be easily exchanged on the broken edges of the kaolinite.

CONCLUSIONS

This study investigated the distribution of heavy metals in two polluted industrial site soils, designated as silty sand and silty clay, and its implications for remediation using soil-washing methods. Sequential extractions were performed to determine the exchangeable, carbonate-bound, Fe/Mn-bound, organic-matter-bound and residual forms of metals in each soil. The removal of metals using EDTA and acetic acid in soil-washing remediation was investigated. Metals were predominantly bound with Fe/Mn minerals in the silty sand, while they were present in residual form in the silty clay. Ethylenediaminetetraacetic acid was found to be suitable for the removal of all metals present in silty sand, while acetic acid was found to be effective for the removal of heavy metals from silty clay. Overall, this study showed that EDTA is suitable for the removal of metals from sandy soils; however, acetic acid is suitable for releasing metals that are in a tightly bound residual form in clayey soils. A site-specific investigation must be undertaken to select extracting solutions in the soil-washing remediation method on the basis of soil composition and sequestration of heavy metals within the soils.

Sequestration of heavy metals in soils from two polluted industrial sites: implications for remediation

ACKNOWLEDGEMENTS

This study was funded by the Illinois Technology Challenge Grant, and this is gratefully acknowledged.

Nowack, B. (2002) Environmental chemistry of aminopolycarboxylate chelating agents. Environ. Sci. & Technol., 36, 40094016 Peters, R.W. (1999) Chelant extraction of heavy metals from contaminated soils. J. Hazard. Mater., 66, 151210 Pueyo, M., Rauret, G., Lck, D., Yli-Halla, M., Muntau, H., Quevauviller, P. and Lpez-Snchez, J.F. (2001) Certification of the extractable contents of Cd, Cr, Cu, Ni, Pb and Zn in a freshwater sediment following a collaboratively tested and optimised three-step sequential extraction procedure. J. Environ. Monit., 3, 243250 Rampley, C.G. and Ogden, K.L. (1998) Preliminary studies for removal of lead from surrogate and real soils using a water soluble chelator: adsorption and batch extraction. Environ. Sci. & Technol., 32, 987993 Suraj, G., Iyer, C.S.P., Rugmini, S. and Lalithambika, M. (1998) Adsorption of cadmium and copper by modified kaolinites. Appl. Clay Sci., 13, 293306 Tessier, A., Campbell, P.G.C. and Bission, M. (1979) Sequential extraction procedure for the speciation of particulate trace metals. Anal. Chem., 51, 844850 United States Environmental Protection Agency (USEPA) (1997) Recent Developments for In-situ Treatment of Metal Contaminated Soils. EPA-542-R-97-004. Office of Research and Development, Washington, DC

REFERENCES American Society of Testing and Materials (ASTM) (2006) Annual Book of Standards. West Conshohocken, PA, USA Evans, L.J. (1989) Chemistry of metal retention by soils. Environ. Sci. & Technol., 23, 10461056 Griffiths, R.A. (1995) Soil-washing technology and practice. J. Hazard. Mater., 40, 175189 Hall, G.E.M., Vaive, J.E., Beer, R. and Hoashi, M. (1996) Selective leaches revisited, with emphasis on the amorphous Fe oxyhydroxide phase extraction. J. Geochem. Exploration, 56, 5978 Margui, E., Salvado, V., Queralt, I. and Hidalgo, M. (2004) Comparison of three-stage sequential extraction and toxicity characteristic leaching tests to evaluate metal mobility in mining wastes. Anal. Chim. Acta, 524, 151159 Miranda-Trevino, J.C. and Coles, C.A. (2003) Kaolinite properties, structure and influence of metal retention on pH. Appl. Clay Sci., 23, 133139

Apart from fair dealing for the purposes of research or private study, or criticism or review, this publication may not be reproduced, stored in a retrieval system or transmitted in any form or by any means, electronic, mechanical, photographic or otherwise, without the prior permission in writing of the publisher. The views expressed in this and in all articles in the journal Land Contamination & Reclamation are those of the authors alone and do not necessarily reflect those of the editor, editorial board or publisher, or of the authors employers or organizations with which they are associated. The information in this article is intended as general guidance only; it is not comprehensive and does not constitute professional advice. Readers are advised to verify any information obtained from this article, and to seek professional advice as appropriate. The publisher does not endorse claims made for processes and products, and does not, to the extent permitted by law, make any warranty, express or implied, in relation to this article, including but not limited to completeness, accuracy, quality and fitness for a particular purpose, or assume any responsibility for damage or loss caused to persons or property as a result of the use of information in this article.

23

You might also like