You are on page 1of 14

This article was downloaded by: [The Library, University of Witwatersrand]

On: 05 March 2012, At: 00:23


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Toxicological & Environmental


Chemistry
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/gtec20

Characterization of cyanide in a
natural stream impacted by gold mining
activities in the Witwatersrand Basin,
South Africa
a a a a
E.N. Bakatula , E.M. Cukrowska , L. Chimuka & H. Tutu
a
School of Chemistry, University of the Witwatersrand,
Johannesburg, South Africa

Available online: 08 Nov 2011

To cite this article: E.N. Bakatula, E.M. Cukrowska, L. Chimuka & H. Tutu (2012): Characterization
of cyanide in a natural stream impacted by gold mining activities in the Witwatersrand Basin, South
Africa, Toxicological & Environmental Chemistry, 94:1, 7-19

To link to this article: http://dx.doi.org/10.1080/02772248.2011.638637

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-


conditions

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation
that the contents will be complete or accurate or up to date. The accuracy of any
instructions, formulae, and drug doses should be independently verified with primary
sources. The publisher shall not be liable for any loss, actions, claims, proceedings,
demand, or costs or damages whatsoever or howsoever caused arising directly or
indirectly in connection with or arising out of the use of this material.
Toxicological & Environmental Chemistry
Vol. 94, No. 1, January 2012, 7–19

Characterization of cyanide in a natural stream impacted by gold mining


activities in the Witwatersrand Basin, South Africa
E.N. Bakatula, E.M. Cukrowska, L. Chimuka and H. Tutu*

School of Chemistry, University of the Witwatersrand, Johannesburg, South Africa


Downloaded by [The Library, University of Witwatersrand] at 00:23 05 March 2012

(Received 4 April 2011; final version received 27 October 2011)

Water systems in the Witwatersrand Basin are heavily contaminated with


pollutants from gold-mine tailing leachates and spillages. Cyanide speciation was
studied in relation to the Natalspruit, a natural stream within a gold mining area
of Johannesburg. The part of the studied stream was adjacent to an abandoned
gold-mine tailings dump. The results of this study pointed to a variation in
surface and groundwater chemistry. The groundwater had low pH and contained
elevated concentrations of pollutants including secondary cyanide species, e.g.
thiocyanate, cyanate, and ammonium. Metal-cyanide complexes that are
weak-acid dissociable (CNWAD) and strong-acid dissociable (CNSAD) cyanides
were found to dominate the speciation of cyanide, with iron cyanide complexes
being most prevalent in groundwater. Decreases in the concentration of 63% for
CuðCNÞ2 2 4
3 , 83% for NiðCNÞ4 , 69% for FeðCNÞ6 , and 47% for FeðCNÞ6 were
3

observed from groundwater to surface water largely due to dilution and


dissociation (in the case of the first two complexes). Speciation modeling was
used to estimate the concentrations of CNWAD and CNSAD based on total cyanide
(CNT), total metal, and total anion concentrations. Other cyanide complexes as
well as dominant sulfate complexes were also predicted.
Keywords: cyanide; metal-cyanide complexes; weak-acid dissociable cyanide;
strong-acid dissociable cyanide; Witwatersrand Basin

Introduction
Gold mining in the Witwatersrand Basin of South Africa began in 1886. Initially, gold was
extracted using the mercury amalgam method, but as mining operations became deeper,
more non-oxidized ore containing pyrite FeS2 was encountered which interfered with the
extraction. In the 1890s, the MacArthur–Forrest process of gold extraction using cyanide
was developed and successfully applied to Witwatersrand ores. This meant that the gold
that was not completely recovered by the previous metallurgical processes (and even small
quantities of around 0.5 g ton1) that remained in tailings could now be profitably
recovered. Although some deep shaft mines are still operational, this has been the most
dominant type of gold mining activity for the last two decades (Tutu, McCarthy, and
Cukrowska 2008).
The matrix of the Witwatersrand ores typically contains about 70–90% quartz,
10–30% phyllosilicates consisting mainly of sericite (KAl2(AlSi3O10)(OH)2), 3–5% pyrite
(with about 1.6% S in the conglomerate), and lesser amounts of a wide variety of other

*Corresponding author. Email: Hlanganani.tutu@wits.ac.za

ISSN 0277–2248 print/ISSN 1029–0486 online


ß 2012 Taylor & Francis
http://dx.doi.org/10.1080/02772248.2011.638637
http://www.tandfonline.com
8 E.N. Bakatula et al.

sulfide and oxide minerals, in addition to gold. Some 70 different ore minerals have been
identified in the conglomerates (Feather and Koen 1975), the most abundant of which,
after pyrite, are uraninite (UO2), brannerite (UTi2O6), arsenopyrite (FeAsS), cobaltite
(CoAsS), galena (PbS), pyrrhotite (FeS), gersdofite (NiAsS), and chromite (FeCr2O4).
At the time of deposition, gold tailings slurries employing the cyanidation technique
contain alkali metal cyanides which are very soluble in water and readily dissociate into
cations (Naþ or Kþ) and anions (CN) when released in water (Smith, Dehrmann, and
Pullen 1984). The CN released can leach and form some new metal-cyanide complexes
and contaminate the subsurface environment of the tailings.
The slurries are generally alkaline due to the addition of lime during gold extraction,
Downloaded by [The Library, University of Witwatersrand] at 00:23 05 March 2012

but the buffering capacity of the material is usually insufficient to neutralize the acid
generated from the weathering of sulfides. Residual sulfide minerals, especially pyrite, in
the tailings dumps are unstable when exposed to atmospheric oxygen and undergo
oxidation resulting in the generation of acid mine drainage (AMD) and the subsequent
release of metalloids and trace metals. The oxidation of pyrite can be described by the
following reaction (Blowes et al. 1998):
FeS2ðsÞ þ 7=2O2 þ H2 O ! 2SO2
4 þ Fe

þ 2Hþ ð1Þ
Similar oxidation reactions involving other minor minerals release dissolved U, As, Cu,
Ni, Pb, Co, and Zn. The flux of the acid drainage from tailings impoundments can
continue for many decades (Blowes et al. 1998).
Natural degradation of cyanides in soil and tailings can be influenced by variables such
as the cyanide species in solution and the relative concentration, pH, aeration, sunlight, the
presence of bacteria, pond size, depth, turbulence (Ritcey et al. 2005), and chemical
mechanisms which include volatilization, adsorption, precipitation, oxidation, and
hydrolysis (Chatwin, Zhang, and Gridley 1988; Church and Boyle 1990; Botz and
Mudder 2000).
Cyanide is acutely toxic to humans. Liquid or gaseous hydrogen cyanide and alkali
salts of cyanide can enter the body through inhalation, ingestion or absorption through the
eyes and skin. The toxicity of cyanide to aquatic life is caused by hydrogen cyanide that
has ionized, dissociated, or photochemically decomposed from compounds containing
cyanide. The sensitivity of aquatic organisms to cyanide is highly species specific, and is
also affected by water pH, temperature, and oxygen content, as well as the life stage and
condition of the organism. Concentrations of free cyanide in the aquatic environment
ranging from 5.0 to 7.2 mg L1 reduce swimming performance and inhibit reproduction in
many species of fish (ICMI 2011)
The South African Department of Water Affairs and Forestry stipulates an allowable
limit for free cyanide in drinking water of 50.02 mg L1 (DWAF 1996). The South
African Cyanide Code set a limit of 0.5 mg L1 for CNWAD in the non-returning mine
effluents (Chamber of Mines South Africa 2001).
Some environmental pollution problems associated with cyanide have been reported
with the failure of gold tailings dams such as the Merriespruit tailings dam in South Africa
in 1994 and a tailings dam in Baia Mare, Romania, in 2000 (McPhail, Stuart, and Venter
1998; Korte, Spiteller, and Coulston 2000; van Niekerk and Viljoen 2005).These accidents
caused significant environment damage and have raised public concerns over the use and
management of cyanide in mining operations.
The solid waste remaining after processing generally contains significant concentra-
tions of various pollutants including cyanide. In one study (Young and Jordan 1995),
Toxicological & Environmental Chemistry 9

waste containing up to 120 mg L1 free CN (CNfree) and 400 mg L1 total CN (CNT),
including various cyanide complexes with heavy metals was reported. These can leach
through mine drainage and therefore cause contamination of the land and of surface and
ground water. Another study (Zagury, Oudjehani, and Deschenes 2004) was conducted on
the characterization of cyanide in solid mine tailings located in northern Quebec, Canada.
The findings revealed an average of 19.5 mg kg1of CNT for the fresh tailings (3 months
old) and 3.2 mg kg1 for the old discharge (6–9 years old). The study indicated that the
more reactive cyanide species initially associated with the solid tailings degraded naturally
within the mine tailings impoundment area, resulting primarily from volatilization
(decrease in pH), leaching, and bacterial degradation.
The aim of this article was to characterize the cyanide released from the gold tailings by
Downloaded by [The Library, University of Witwatersrand] at 00:23 05 March 2012

assessing and determining primary species (CNfree, CNT, weak-acid dissociable (CNWAD),
strong acid dissociable (CNSAD), and metal-cyanide complexes), secondary species
(thiocyanate (SCN), cyanate (CNO), and ammonium (NHþ 4 )) of cyanide, as well as
metals in the surface and ground water in a natural stream in the vicinity of gold mine
tailings.

Study area
The Natalspruit (Figure 1) was selected for this study because its headwaters lie in an area
in which gold tailings dumps abound. The stream is perennial over all the reaches sampled
and also susceptible to severe flash flooding during summer thunderstorms. During the dry
season, discharge in the stream is sustained by ground water emerging through the bed of
the stream and by seepage through the banks (Naicker, Cukrowska, and McCarthy 2003).

Figure 1. Map of study area and sampling points.


10 E.N. Bakatula et al.

Materials and methods


Sampling
Sampling was carried out along the Natalspruit, adjacent and downstream to the mining
area. Both surface water and groundwater samples were collected in August 2007 (end of
the dry season) according to commonly accepted sampling protocols (Hermond and
Fechner-Levy 2000). The surface water samples (labeled as S) were taken from the middle
of the stream; seepage water was sampled at the site where it had formed shallow puddles
on the bank which were sufficiently deep to collect samples. Ground water samples
(labeled as GW) were collected in holes drilled using an auger.
The water samples were placed in 1 L polypropylene bottles, wrapped in black plastic
Downloaded by [The Library, University of Witwatersrand] at 00:23 05 March 2012

bags to avoid contact with light, filtered with a 0.45 mm cellulose nitrate filter paper, and
stored in the refrigerator at 4 C prior to preparation and analysis.
At the time of sampling, efflorescent crusts were observed along the capillary fringe on
the stream bank. These were scraped off carefully using a plastic spatula and collected into
a polypropylene container.

Field measurements
The field parameters, namely pH, redox potential (Eh), electrical conductivity (EC), and
temperature, were measured in the water samples. These measurements were carried out
with a portable kit (Multi Line F/Set 3, WTW, Germany), equipped with a pH electrode,
an integrated temperature probe (SenTix 41), a standard conductivity cell (Tetra Con 375),
and an oxidation–reduction potential probe (SenTix ORP). The pH electrode was
calibrated according to IUPAC recommendations against two buffer solutions of pH 4.0
and pH 7.0 and had an uncertainty of 0.1 pH units. Redox potentials were obtained from
Pt electrodes versus Ag/AgCl. The electrodes were checked using a standard buffer
solution and all the reported potentials were corrected relatively to the standard hydrogen
electrode (SHE). The uncertainty was 0.03 V. For EC, the uncertainty was
0.002 mS cm1.

Preliminary treatment and preparation of water samples


Each filtered water sample was divided into three portions: one portion was acidified with
concentrated nitric acid to pH 2 for the analysis of metals; the other had its pH raised to 12
(by the addition of NaOH) for the analysis of CNfree and metal-cyanides; the third portion
(for the analysis of CNO, SCN, S2 O2 þ 2 
3 , NH4 , SO4 , and Cl ) was not treated. The
samples for the analysis of CNfree were further treated as follows. Sodium thiosulfate
(Na2S2O3) was added to each sample to prevent species like Cl from decomposing the
cyanide. Thereafter, lead carbonate, Pb(CO3)2, was added to eliminate the effect of S2, as
this would lead CN conversion to SCN.

Preparation of efflorescent crusts


The efflorescent crusts were dried in an oven at 40 C. Two grams of the crusts were
weighed and dissolved in 100 mL of deionized water and the solution was filtered using a
Whatman Grade No. 41 filter paper to remove sediment residue. Sufficient solutions were
Toxicological & Environmental Chemistry 11

made and subjected to the treatment similar to the water samples mentioned above for the
analysis of primary and secondary cyanide species; S2 O2 2 
3 , SO4 , Cl as well as metals.

Analytical methods
Water samples were analyzed for CNfree, CNO, and SCN according to standard
methods (Clesceri, Greenberg, and Rhodes 1998). S2 O2 3 was determined by the
spectrophotometric method (Koh 1990; Koh and Okabe 1992). Metal-cyanide complexes
were determined by reversed-phase ion interaction high-performance liquid chromatog-
raphy with UV detection (Grigorova, Wright, and Josephson 1987; Haddad and
Downloaded by [The Library, University of Witwatersrand] at 00:23 05 March 2012

Kalambaheti 1991; Giroux and Dwight 1994). SO2  þ


4 , Cl , and NH4 were determined by
ion chromatography (Metrohm 761 Compact Ion Chromatograph, Switzerland). Metals
were determined by inductively coupled plasma-optical emission spectroscopy (ICP-OES)
(Spectro Genesis, Kleve, Germany). The detection limits are shown in Tables 1 and 2.
For quality assurance, all analyses were performed in triplicate. Certified reference
standards (Industrial Analytical, South Africa) and a procedure blank were used.
Percentage recoveries above 95% were obtained with those for free cyanide around 97%.
All the reported results had relative standard deviations (RSD) 510%.

Speciation modeling
The modeling was based on total cyanide (CNT), total metal concentrations, total anion
concentrations, as well as the geochemical parameters (pH, temperature, Eh, and EC). The
species were allowed to react until equilibrium was established.
The CNT was calculated as
CNT ¼ CNfree þ CNWAD þ CNSAD ð2Þ
CNfree was from the analytical results while CNWAD and CNSAD were from mass balances
based on the determined metal-cyanide complexes.
Modeling was done using the U.S. EPA Visual MINTEQ ver. 2.32 geochemical
modeling code.

Results and discussion


Physical–chemical characterization
The results for field parameters and cation and anion concentrations are presented in
Table 1. The results pointed to a variation in the chemistry of surface and groundwater.
Samples collected upstream and at the seepage areas had near-neutral pH, relatively low
conductivities, and low redox potential.
From the trend it is evident that the water samples cluster into three groups, namely
Group 1 (S1–S4): surface water samples collected upstream of the mining area and the
seepage area, Group 2 (S5–S8): surface water samples collected downstream with low pH
(4.26  4.48), and Group 3 (GW9 & GW10): groundwater with pH 3.55 and 3.56.
Samples collected downstream have low pH, relatively high conductivities, and high
redox potential. These samples generally have elevated sulfate, iron, and trace metal
concentrations. The ground water had much higher dissolved solid concentration and
conductivities, and lower pH relative to surface water.
Downloaded by [The Library, University of Witwatersrand] at 00:23 05 March 2012

12
E.N. Bakatula et al.

Table 1. Physical-chemical characteristics, anion, and cation concentrations in water samples from the Natalspruit.

Temp. Eh EC S2 O2
3 SO2
4 Cl Na K Mg Ca Fe Mn Co Zn Ni Cu
Sample Description ( C) pH (V) (mS cm1) (mg L1) (mg L1) (mg L1) (mg L1) (mg L1) (mg L1) (mg L1) (mg L1) (mg L1) (mg L1) (mg L1) (mg L1) (mg L1)

S1 Surface water 12.1 6.8 0.32 0.24 36.0 331 n.d 3.8 2.0 19.4 30.6 0.1 2.6 0.01 0.01 0.01 n.d
S2 Surface water 15.5 6.4 0.29 0.25 33.0 180 n.d 4.2 2.2 21.3 31.7 0.1 7.9 0.1 0.04 0.04 n.d
S3 Seepage water 11.8 6.7 0.27 0.52 85.0 2475 n.d 4.2 5.9 14.8 30.4 7.0 25.1 0.3 0.5 0.3 0.01
S4 Surface water 17.9 6.3 0.27 0.39 320 2066 n.d 2.3 1.5 11.1 52.5 4.5 8.6 0.3 0.2 0.2 0.02
S5 Surface water 15.6 4.5 0.51 0.89 370 5601 n.d 4.9 4.6 34.6 62.2 14.5 37.0 3.2 2.0 2.5 0.2
S6 Surface water 16.2 4.4 0.51 0.93 320 4695 n.d 5.0 3.6 36.1 63.5 13.7 37.4 3.3 2.2 2.6 0.2
S7 Surface water 14.9 4.5 0.51 0.80 321 4530 n.d 4.8 4.2 30.6 79.7 7.1 30.7 2.5 1.7 2.0 0.2
S8 Surface water 16.4 4.3 0.54 0.81 315 3945 n.d 4.7 3.0 30.5 74.2 6.9 30.7 2.6 1.7 2.0 0.2
GW9 Ground water 11.1 3.6 0.48 4.86 530 7712 20 13.9 5.1 112 173 126 199 25.7 18.6 22.0 4.2
GW10 Ground water 11.6 3.6 0.47 4.61 540 7751 22 11.3 4.5 110 171 157 172 22.9 16.8 19.3 3.0
LOD 3.2 0.3 0.5 0.01 0.02 0.003 0.05 0.004 0.003 0.005 0.006 0.02 0.003
MQL 10.5 0.9 1.7 0.02 0.08 0.001 0.18 0.01 0.01 0.02 0.02 0.06 0.01

Note: LOD – limit of detection; MQL – minimum quantifiable level; n.d – not detected.
Toxicological & Environmental Chemistry 13

Table 2. Cyanide species in the Natalspruit water samples.

Sample CNfree SCN CNO NHþ


4 CuðCNÞ2
3 NiðCNÞ2
4 FeðCNÞ4
6 FeðCNÞ3
6 CoðCNÞ4
6

S1 0.02 n.d 17.6 1.2 0.5 0.6 0.9 n.d n.d


S2 0.02 n.d 18.2 3.9 0.4 0.7 1.6 n.d n.d
S3 0.8 n.d 17.4 9.6 7.1 1.7 14.9 n.d 0.5
S4 0.3 n.d 15.8 2.3 6.5 1.8 9.9 n.d n.d
S5 0.02 n.d 16.9 1.6 4.8 2.4 10.1 10.7 n.d
S6 0.04 n.d 17.4 5.6 4.5 2.0 12.0 9.5 n.d
S7 0.01 n.d 18.4 8.7 4.1 1.6 7.3 5.9 n.d
S8 0.02 n.d 18.4 5.7 4.7 2.4 7.5 6.7 n.d
GW9 0.3 1.8 23.7 22.9 10.9 9.3 22.1 10.3 1.1
Downloaded by [The Library, University of Witwatersrand] at 00:23 05 March 2012

GW10 0.3 2.3 19.8 31.9 10.6 8.5 23.9 11.1 1.2
LOD 0.009 0.039 0.003 0.006 0.033 0.04 0.030 0.027 0.039
MQL 0.03 0.13 0.01 0.02 0.11 0.12 0.10 0.09 0.13

Note: LOD – limit of detection; MQL – minimum quantifiable level; n.d – not detected.
Concentrations are in mg L1.

From previous studies (Naicker, Cukrowska, and McCarty 2003; Tutu et al. 2003), the
groundwater recharging the Natalspruit in this section was found to be originating from
the adjacent tailings dumps that are undergoing reprocessing. The water leaching from
such mine workings tends to be characterized by low pH and high dissolved salt loads. The
stream acts as a mixing zone in this case, with all the above-mentioned water types
accounting for the observed chemical variability.
Upstream of the mining area, the surface water samples S1 and S2 had near-neutral pH
with low metal concentrations and relatively low EC. This was expected as there are no
tailings facilities and reprocessing activities in the vicinity upstream that are likely to
contribute to pollution in that area.
Seepage water (S3) in the study area was characterized by near-neutral pH, high sulfate
concentration, as well as moderately high metal concentrations and low thiosulfate
concentration. This could possibly be indicative of ingress of a mixed-type groundwater. It
is likely that at that point the polluted groundwater plume could be diluted by unpolluted
groundwater which would be unlikely for samples GW9 and GW10.
Adjacent to the mining area, the pH of water falls to slightly above 4 (e.g. in samples
S5–S8). This is attributed to the ingress of acidic groundwater, and consequently heavy
metal and sulfate concentrations are observed to increase. Groundwater is known to
preserve its composition, meaning that pollution plumes from tailings can persist over long
distances and their effects manifest for prolonged time scales (Tutu, McCarthy, and
Cukrowska 2008). As the groundwater seeps to the surface, it is diluted by surface water.
As such, average decreases in concentration of 94% for Fe, 92% for Co, Zn, and Ni,
and 97% for Cu were observed from groundwater (GW9 and GW10) to surface
water (S5–S8).
Results for thiosulfate showed low concentrations in the upstream samples (S1 and S2)
and at the seepage point. Elevated concentrations in the downstream surface water and
groundwater (ranging from 315 to 370 mg L1) were observed. Thiosulfates, together
with sulfates, are from the oxidation of pyrite. Sulfate concentrations displayed a
trend similar to that for thiosulfates that is increasing downstream and with
almost twice the concentration on average in groundwater than in surface water due to
dilution.
14 E.N. Bakatula et al.

Cyanide species
The concentrations of cyanide species are presented in Table 2. The trend generally
followed that for metals and sulfate species, that is higher concentrations in groundwater
than in the surface water. There was a marked decrease in concentrations of CNfree in
surface water compared to groundwater and seepage water. This could be attributed to the
effect of two processes, namely dilution and volatilization. The latter process could, in
part, be enhanced by increase in the temperature of surface water as well as exposure to a
larger surface area. At low pH values, most of the CNfree exists as HCN, which is volatile
and is likely to be lost as groundwater mixes with surface water.
The results in Table 2 show a mixed trend with respect to CNfree concentrations. There
Downloaded by [The Library, University of Witwatersrand] at 00:23 05 March 2012

is an increase in CNfree at point S3 which could be attributed to the ingress of


groundwater. The CNfree is likely to be sustained in solution by the dissolution of CNWAD
and the slightly elevated pH. Then there is a drop in concentration at point S4 owing to
dilution and a subsequent drop in concentrations in the successive surface water samples.
This could be attributed to a drop in pH which could lead to volatilization of CNfree. This
could likely be the case with CNfree in groundwater. The CNfree could be from excess
CNfree present in the original plume and from the dissociation of cyanide complexes,
particularly CNWAD.
An elevated CNfree concentration in S3 than in GW9 and GW10 and elevated metal
concentrations in the latter than in the former were observed. This could suggest that while
the main pathway of metal ingress into the stream could be contaminated groundwater,
that for cyanide could likely be the stream draining from the mine dumps through the
wetland.
Thiocyanates were observed only in the groundwater. These tend to form strong
complexes with Fe and can persist in groundwater as a result of slow decomposition
kinetics (Dzombak, Rajat, and Wong-Chong 2006). Elevated cyanate concentrations were
observed in all samples. Cyanate results from the hydrolysis of cyanide and as such could
be derived from CNfree (Dzombak, Rajat, and Wong-Chong 2006). However, the reason
for the seemingly even distribution of its concentration is not apparent. Ammonium
(NHþ 4 ) concentrations were higher in seepage water and ground water than in surface
water, and this could be due to dilution. There was an average of about 79% difference in
concentrations in groundwater than in the immediate surface water. NHþ 4 results from the
acidification of CNO (Dzombak, Rajat, and Wong-Chong 2006; Zagury, Oudjehani, and
Deschenes 2004).
2CNO þ H2 SO4 þ 4H2 O ! ðNH4 Þ2 SO4 þ 2HCO
3 ð3Þ
This would likely occur in the source plume with elevated sulfate concentrations. The

concentrations of NHþ 4 would most likely vary depending on the available CNO as well.
The metal-cyanide complexes showed a trend of elevated concentrations in ground-
water and lower concentrations in surface water. For the CNWAD, namely CuðCNÞ2 3 and
NiðCNÞ24 , the trend is typical of a combination of dilution and dissociation. CN WAD
dissociate under the conditions 4 5 pH 5 6. Concentrations of NiðCNÞ2 4 dropped by
about 73% between groundwater and immediate surface water. CoðCNÞ4 6 was only
detected in seepage water and groundwater. The trend is likely due to dilution as this
complex is a CNSAD and would dissociate at pH 5 2 (Dzombak, Rajat, and Wong-Chong
2006). The iron-based cyanides also showed a similar trend but it is likely that this was due
to a different combination of processes to that for CNWAD. While there could be
dilution of ingressing groundwater, there could also be oxidation of Fe2þ to Fe3þ by
Toxicological & Environmental Chemistry 15

Table 3. Cyanide species predicted by mass balances and speciation modeling.

CNWAD CNWAD CNSAD CNSAD CNT CNT


Sample (mass balance) (modeled) (mass balance) (modeled) (analyzed) (modeled)

S1 0.05 0.04 0.07 0.05 0.16 0.11


S2 0.04 0.06 0.12 0.09 0.22 0.16
S3 4.9 14.4 11.3 1.9 30.7 16.7
S4 4.8 8.9 7.3 5.7 20.9 14.7
S5 4.2 9.0 15.3 13.2 28.5 22.2
S6 3.8 7.2 15.9 15.1 26.8 22.4
S7 3.3 7.8 9.7 7.7 20.8 15.4
S8 4.2 7.5 10.5 9.0 22.1 16.6
Downloaded by [The Library, University of Witwatersrand] at 00:23 05 March 2012

GW9 11.9 15.7 24.7 20.3 52.3 36.2


GW10 11.3 15.1 26.6 22.1 52.9 37.5
LOD 0.005 0.006 0.005
MQL 0.015 0.020 0.018

Note: LOD – limit of detection; MQL – minimum quantifiable level.

ultraviolet light. Since Fe2þ is more mobile at reducing conditions which are prevailing in
the original plume coming through the groundwater, most of the Fe-cyanides could be
coming through as FeðCNÞ4 3
6 , and the FeðCNÞ6 could be derived from the oxidation of
this ferrocyanide. This reaction would occur as follows (Meussen, Keizer, and de Haan
1992; Dzombak, Rajat, and Wong-Chong 2006):

FeðCNÞ4 3
6 þ uv ! FeðCNÞ6 þe

ð4Þ
The overall CNWAD and CNSAD were predicted using speciation modeling based on
mass balances of the analyzed species. The results are presented in Table 3.
From the results, some discrepancies were observed in some cases, thus pointing to
certain species which were not analyzed but that could be existing in the samples.
For instance, CNWAD was higher in modeled than in analytical results (a difference of over
30%). This could be attributed to the prediction by the model of the formation of other
CNWAD species such as ZnðCNÞ2 4 which was not analyzed. Mercury which is also known
to form CNWAD was not found in the analyzed samples.
The results for the predicted species are presented in Table 4.
Sulfates were shown to dominate in the samples, the most prevalent being CaSO4. The
model also predicted prevalence of different nickel-cyanide complexes. Thiocyanates were
also predicted in groundwater, accounting for the slight concentrations that were detected
in those samples. The model further predicted the existence of double cyanide salts, namely
Ca2 FeðCNÞ06 and ðNH4 Þ2 FeðCNÞ2 6 . Together with the metal-cyanides determined by
analysis, the existence of these complexes substantiate the fact that metal-cyanide
complexes typically dominate aqueous speciation of cyanide in groundwater systems, with
iron-cyanide complexes being most abundant, as observed by Ghosh et al. (1999).
Potassium also tends to form such double salts with Fe, but its low concentrations
compared to Ca and NHþ 4 may have caused it to be outcompeted. It would only be
speculated here that any available potassium would contribute to the formation of jarosite
(KFe(SO4)2(OH)6).
Efflorescent crusts were observed in this study as well as in previous studies in areas
impacted by mining activities in the Witwatersrand (Naicker, Cukrowska, and McCarthy
16 E.N. Bakatula et al.

Table 4. Other species predicted from speciation modeling.

Sample S3 S5 S8 GW9 GW10

Predicted species ZnðCNÞ02 NiH2 ðCNÞ04 NiH2 ðCNÞ04 CoSCNþ ZnðCNÞ 3


ZnðCNÞ3 NiH3 ðCNÞþ4 NiH3 ðCNÞþ4 CuSCNþ Zn(CN)2
CaSO4 ZnðCNÞ02 NiHðCNÞ4 FeSCNþ CoSCNþ
CaSO4 CaSO4 Ca2 FeðCNÞ06 Ca2 FeðCNÞ06
ðNH4 Þ2 FeðCNÞ2
6 ðNH4 Þ2 FeðCNÞ2
6
NiSCNþ CuSCNþ
0 þ
NiH2 ðCNÞ4 FeSCN
NiH3 ðCNÞþ 4 NiSCNþ
Downloaded by [The Library, University of Witwatersrand] at 00:23 05 March 2012

NiHðCNÞ 4 NiH2 ðCNÞ04


FeSO4 NiH3 ðCNÞþ 4
MgSO4 NiHðCNÞ 4
CaSO4 FeSO4
ZnSO4 MgSO4
NiSO4 CaSO4
MnSO4 ZnSO4
CoSO4 MnSO4
CuSO4 CoSO4
CuSO4

2003; Mphephu et al. 2004; Tutu, McCarthy, and Cukrowska 2008). The crusts result from
capillary evaporation of shallow groundwater. While their main constituent has been
found to be gypsum (CaSO4  2H2O), they have also been found to contain a variety of
other sulfate-based minerals, e.g. jarosite (KFe(SO4)2(OH)6), melanterite (FeSO4  7H2O),
copiapite ((Fe, Mg) Fe4(SO4)6(OH)2), goslarite (ZnSO4  7H2O), epsomite
(MgSO4  7H2O), and carnotite (K2(UO2)2V2O8  3H2O). Concentrations of most heavy
metals (e.g. Ni, Co, Zn, and Cu) in the crusts were found to exceed 2000 mg kg1 in some
cases. Because of their solubility, the crusts have been implicated as secondary sources of
pollutants (Winde 2004; Tutu, McCarthy, and Cukrowska 2008). Their overall contribu-
tion to pollution of the stream in this study was not ascertained. However, their potential
to release pollutants was assessed and the results are presented in Figure 2.
The results showed a decrease in the pH of receiving water (deionized water at pH 7 in
this case) on the addition of salt crusts. This was accompanied by an increase in EC. This is
evidence that the crusts largely contain acid-releasing minerals and high concentrations of
metals and anions. This composition typifies that of the groundwater. The salt crusts result
from capillary evaporation of groundwater and deposit along a capillary fringe on the
banks of the stream. The results for the cyanide species in a composite of crusts collected
along the stream bank are presented in Table 5.
The results pointed to elevated concentrations of cyanide species except for CNfree.
This could be attributed the to loss of most CNfree as a result of volatilization during
dissolution of the crusts. During this process, the pH is lowered by about 3 orders of
magnitude and falls in the range in which most CNWAD would dissociate as well.

Conclusion
The Natalspruit is contaminated by pollutants released from the gold mining activities in
the vicinity. Despite the presence of a wetland between the stream and the abandoned mine
Toxicological & Environmental Chemistry 17

8 16
pH
7 EC 14

6 12

EC (mS cm–1)
5 10

pH
4 8

3 6

2 4

1 2
Downloaded by [The Library, University of Witwatersrand] at 00:23 05 March 2012

0 0
0 0.01 0.02 0.03 0.04 0.05
Solid: liquid ratio (g mL–1)

Figure 2. Potential effects of salt crusts on receiving water bodies.

Table 5. Cyanide species in a composite of salt crusts.

Species Concentration (mg kg1)

CNfree 3.13  0.15


SCN 80.2  1.35
CNO 300.6  5.9
NHþ4 78.09  1.02
CuðCNÞ23 18.65  0.83
NiðCNÞ2
4 24.37  1.17
FeðCNÞ3
6 73.59  2.03
CNT 198.4  3.9

works, elevated concentrations of metals, sulfates and cyanide species were observed.
The results show that these pollutants could be transported largely through groundwater.
The contamination of water was enhanced downstream with very low pH, high
concentration of metals, as well as metal-cyanide complexes. The groundwater was
found to contain significant amounts of cyanide species, sulfates, and metals. The metal-
cyanides, particularly the CNWAD, dissociate under variable conditions to yield CNfree,
resulting in sustained pollution from ground water to surface water. Speciation modeling
results yielded some important information about the overall chemical composition of the
water. The models for groundwater corroborated the results obtained for the salt crusts,
thus showing that these features are simply evaporated groundwater. Concentrations of
cyanide species in the studied samples far exceeded those stipulated by the governing
bodies and thus pose a serious threat to poor informal settlements which may be using this
water for their livelihood.

Acknowledgments
The authors would like to acknowledge the National Research Foundation and the Carnegie
Corporation for the financial support.
18 E.N. Bakatula et al.

References

Blowes, D.W., C.J. Ptacek, S.G. Benner, K.R. Waybrank, and J.G. Bain. 1998. Permeable reactive
barriers for the treatment of mine tailings drainage water. Proceedings of the international
conference and workshop on uranium mining and hydrology, Vol. 2, 113–19, in Freiberg,
Germany.
Botz, M.M., and T.I. Mudder. 2000. Modeling of natural cyanide attenuation in tailings
impoundments. Mineral Metal Process 17: 228–33.
Chatwin, T.D., J. Zhang, and G.M. Gridley. 1988. Natural mechanisms in soil to mitigate cyanide
release. Proceedings Superfund’88, the 9th national conference, Hazardous Materials Control
Research Institute, Washington, DC, 467–79.
Downloaded by [The Library, University of Witwatersrand] at 00:23 05 March 2012

Chamber of Mines South Africa. 2001. South African guideline on cyanide management for gold
mining, 69–70. Johannesburg, South Africa.
Church, R.H., and J.J. Boyle. 1990. A study of long term cyanide residuals at an historically
significant site. Proceedings of the western regional symposium on mining and mineral processing
wastes, Society for Mining Metallurgy, and Exploration Inc, Berkeley, CA, 105–14.
Clesceri, S., E. Greenberg, and R. Rhodes. 1998. Standards methods for the examination of water and
wastewater. 20th ed. Washington, DC: American Public Health Association.
DWAF. 1996. South African water quality guidelines. Vol. 1: Domestic use. Pretoria, South Africa:
Department of Water Affairs and Forestry.
Dzombak, A., S. Rajat, and M. Wong-Chong. 2006. Cyanide in water and soil: Chemistry, risk, and
management, Vol. 602. New York: CRC, Taylor & Francis.
Feather, C.E., and G.M. Koen. 1975. The mineralogy of the Witwatersrand Reefs. Minerals Science
Engineering 7: 189–224.
Ghosh, R.S., D. Dzombak, R. Luthy, and J. Smith. 1999. In situ treatment of cyanide-contaminated
groundwater by iron cyanide precipitation. Water Environment Research 71: 1217–28.
Giroux, L., and D.J. Barkley. 1994. Separation of metal-cyanide complexes by reversed-phase
ion-interaction high-performance liquid chromatography. Canadian Journal of Chemistry 72: 269–73.
Grigorova, B., S.A. Wright, and M. Josephson. 1987. Separation and determination of stable
metallo-cyanide complexes in metallurgical plant solution and effluents by reversed-phase ion-pair
chromatograph. Journal of Chromatography 410: 419–26.
Haddad, P.R., and C. Kalambaheti. 1991. Advances in ion chromatography: speciation of mL1
levels of metallo-cyanides using ion-interaction chromatography. Analytica Chimica Acta 250:
21–36.
Hermond, H.F., and E.J. Fechner-Levy. 2000. Chemical fate and transport in the
environment. 2nd ed. 433. San Diego: Academic Press.
ICMI. 2011. Environmental and health effects of cyanide. International Cyanide Management Code
for the Gold Mining Industry. http://www.cyanide.org/cyanide_enivronmental.php (accessed June
30, 2011).
Koh, T. 1990. Analytical chemistry of polythionates and thiosulphate: A review. Analytical Sciences
6, no. 1: 3–14.
Koh, T., and K. Okabe. 1992. Separation and spectrophotometric determination of thiosulfate,
sulfite and sulfide in their mixtures. Analytical Sciences 8: 285–291.
Korte, F., M. Spiteller, and F. Coulston. 2000. Commentary. The cyanide leaching gold recovery
process is a non-sustainable technology with unacceptable impacts on ecosystems and humans:
The disaster in Romania. Ecotoxicology and Environmental Safety 46: 241–5.
McPhail, G.L., R.J. Stuart, and D. Venter. 1998. The disaster of Merriespruit and its consequences:
Remediation-making safe. Proceedings of the fifth international conference on tailings and mine
waste, Fort Collins, Colorado, 957–9.
Meussen, J.L., M.G. Keizer, and F. de Haan. 1992. Chemical stability and decomposition rate of
iron cyanide complexes in soil solutions. Environmental Science & Technology 26: 511–16.
Mphephu, N.F., M.J. Viljoen, H.Tutu, E.M. Cukrowska, and K.Govender. 2004.
Mineralogy and geochemistry of mine tailings in relation to water pollution on the Central
Toxicological & Environmental Chemistry 19

Rand, South Africa. Proceedings to the conference on environmental issues and waste
management in energy and mineral production, Antalya, Turkey, 445–9.
Naicker, K., E.M. Cukrowska, and T.S. McCarthy. 2003. Acid mine drainage arising from gold
mining activities in Johannesburg, South Africa and environs. Journal of Environmental Pollution
122: 29–40.
Ritcey, G.M. 2005. Tailings management in gold plants. Hydrometallurgy 78: 3–20.
Smith, A., A. Dehrmann, and R. Pullen. 1984. The effects of cyanide-bearing gold tailings disposal
on water quality in the Witwatersrand, South Africa. Proceedings of the conference on cyanide
and the environment, Tuscon, Arizona, 221–30.
Tutu, H., E.M. Cukrowska, T. McCarthy, N.F. Mphephu, and R. Hart. 2003. Determination and
modelling of geochemical speciation of uranium in gold mine polluted land in South Africa.
Proceedings of the 8th International Mine Water Association Congress, Johannesburg, 137–55.
Downloaded by [The Library, University of Witwatersrand] at 00:23 05 March 2012

Tutu, H., T.S. McCarthy, and E.M. Cukrowska. 2008. The chemical characteristics of acid mine
drainage with particular reference to sources, distribution and remediation: The Witwatersrand
Basin, South Africa, as a case study. Applied Geochemistry 23: 3666–84.
van Niekerk, H.J., and M.J. Viljoen. 2005. Causes and consequences of the Merriespruit and other
tailings-dam failures. Land Degradation and Development 16: 201–12.
Winde, F. 2004. Uranium contamination in fluvial systems – Mechanisms and processes. Part I:
Geochemical mobility of uranium along the water path – The Koekemoerspruit (South Africa) as
a case study. Cuardernos de Investigacion Geografica 28: 49–57.
Young, C.A., and T.S. Jordan. 1995. Cyanide remediation: Current and past technologies.
Proceedings of the 10th annual conference on hazardous waste research, Great Plain/Rocky
Mountain Hazardous Substance Research Center, Kansas State University, Kansas, 104–28.
Zagury, G.J., K. Oudjehani, and L. Deschenes. 2004. Characterization and availability of cyanide in
solid mine tailings from gold extraction plants. Science of the Total Environment 320: 211–24.

You might also like