You are on page 1of 12

Environ Geochem Health (2011) 33:91101

DOI 10.1007/s10653-010-9353-3

ORIGIN PAPER

Pilot-scale tests to optimize the treatment of net-alkaline


mine drainage
Min Jang Hyunho Kwon

Received: 20 March 2010 / Accepted: 3 August 2010 / Published online: 3 November 2010
Springer Science+Business Media B.V. 2010

Abstract A pilot-scale plant consisting of an oxidation basin (OB), a neutralization basin (NB), a
reaction basin (RB), and a settling basin (SB) was
designed and built to conduct pilot-scale experiments.
With this system, the effects of aeration and pH on
ferrous oxidation and on precipitation of the oxidized
products were studied systemically. The results of
pilot-scale tests showed that aeration at 300 L/min
was optimum for oxidation of Fe(II) in the OB, and
the efficiency of oxidation of Fe(II) increased linearly
with increasing retention time. However, Fe(II) was
still present in the subsequent basinsNB, RB, and
SB. Results from pilot-scale tests in which neutralization was excluded were used to obtain rate
constants for heterogeneous and homogeneous oxidation. Oxidation of Fe(II) reached almost 100%
when the pH of the mine drainage was increased to
more than 7.5, and there was a linear relationship
between total rate constant, log (Ktotal), and pH.
Absorbance changes for samples from the NB under
different pH conditions were measured to determine
the precipitation properties of suspended solids in the
SB. Because ferrous remained in the inflow to the SB,
oxidation of Fe(II) was dominant initially, resulting

M. Jang (&)  H. Kwon


Institute of Mine Reclamation Technology, Korea Mine
Reclamation Corporation, Coal Center, 30 Chungjin-dong
Street (Susong-dong), Jongno-gu, Seoul 110-727,
South Korea
e-mail: mjang@mireco.or.kr

in increased absorbance, and the rate of precipitation


was slow. However, the absorbance of the suspension
in the SB rapidly dropped when pH was higher than
7.5.
Keywords Pilot-scale  Oxidation  Precipitation 
Net alkalinity  Mine drainage

Introduction
In South Korea, there are approximately 2,500 mines,
including 906 metal mines, 379 coal mines, and 1,173
non-metal mines. Among these mines, approximately
150 abandoned surface and underground coal mines
are responsible for acid drainage discharges into
water streams, and are continuously contaminating
surface and groundwater. An estimated 100,000 tons
per day of mine drainage is being discharged from
these mines into water streams (CIPB 1995; Jung
2003). Acid mine drainage (AMD) from abandoned
metal mines not only pollutes the natural environment, for example surrounding soil, and surface and
ground water, but also has toxic effects on crops and
humans as a result of concentration of the contamination (Jung 1994).
In general, drainage from coal mines is not only of
low pH but also contains high levels of SO42- and
heavy metals, including Fe, Al, and Mn (Sengupta

123

92

1993). These properties of mine drainage disrupts


stream ecosystems and further aggravate the problem
by creating yellow or white sediments (Chon and
Hwang 2000). Thus, mine drainage treatment is
necessary to prevent pollution of watercourses with
iron/aluminum precipitates and acidity.
Basically, mine drainage treatment utilizes the
chemical properties of the main contaminants, for
example Fe, Al, and Mn. Two different conventional
methods are mainly used to treat mine drainage
active or physicochemical treatment (which needs
energy and chemical input) and passive or biological
treatment (which does not need additional input of
energy or chemicals) (Johnson and Hallberg 2005;
Akcil and Koldas 2006). Most active treatment
involves pH adjustment, by adding an alkaline
material, and removal by precipitation as a result of
the formation of oxy/hydroxides. pH adjustment is
helpful for treatment of large quantities of AMD.
Although active treatment enhances treatment efficiency by use of chemicals, there is a large economic
burden, because of the high cost of maintenance and
chemicals, and this process requires continuous
operation (Johnson and Hallberg 2005). Compared
with active treatment, passive systems are economical but require longer retention times and greater
space; they are, therefore, not appropriate for treatment of large-scale mine drainage. Although passive
treatment has been implemented on full-scale sites in
several countries, treatment efficiency can be uncertain, because of seasonal changes in flow rate and
temperature, and the systems are apt to fail during
long-term operation. As the primary mechanism of
metal removal from acidic effluents, passive treatment uses the affinity properties of metal compounds
with sulfur or carbonate ions; metals are also
removed by adsorption or precipitation with organic
compounds (Gazea et al. 1996). Not only biological
treatment using microbial activity but also chemical
treatment using the sensitivity of metal species to
changes in pH is the main approach of the passive
system. Metal species occur as ions under acidic
conditions; however, they are converted to hydroxides when pH increases. In South Korea, sites with
large-scale mine drainage flows occur in uneven and
narrow terrain. Passive treatment of these contaminated sites has the setback of needing large treatment
areas (Hedin 2008). Accordingly, considering the
special status of Korea and two different treatment

123

Environ Geochem Health (2011) 33:91101

concepts with advantages and disadvantages, a complementary new approach should be developed to
minimize maintenance and chemical input and
enhance treatment efficiency.
The objective of this study was to investigate the
applicability of aeration and pH adjustment for
treatment of mine drainage with net alkalinity on a
large scale by conducting pilot-scale tests. A pilotscale (30 m3/day) plant was designed and built to
conduct experiments on the large-scale treatment of
the mine drainage (about 6,000 m3/day). The main
themes were to determine the effects of aeration,
retention time, and pH change on the oxidation of
ferrous, and the effect of ferrous oxidation on
precipitation of the oxidized products. The results
of pilot-scale tests could be expected to contribute to
the design of a full-scale plant on the actual site by
constructing the optimized process on the uneven
ground of the site.

Materials and methods


Pilot-scale plant setup
With a scale of 30 m3/day, a pilot-scale treatment
facility was constructed and set up to treat AMD
which was continuously flowing from Jami (JM) adit
of Hamback coal mine. As shown in Fig. 1, the pilotscale plant mainly consisted of an oxidation basin
(OB, 1 m3), a neutralization basin (NB, 0.4 m3), a
reaction basin (RB, 0.4 m3), and a settling basin (SB,
1 m3). At the bottom of the OB, nine air diffusers
(disk-type) were set up at equal intervals to produce
air bubbles as homogeneously as possible. The rate of
aeration from all the diffusers ranged from 0 to
450 L/min and was controlled by a mechanical
flowmeter. The inlet flow rate was fixed by controlling the pumping pressure of a submersible pump and
a valve in an equalization basin (0.15 m3). An
agitational device (150250 rpm) was set up in both
the NB and RB for homogeneous mixing. The roundshaped SB has a circular sludge collector, including
scraper and sludge hopper, and a surface-loading rate
of 21 m3/m2 per day. The rotation speed of the
scraper was 0.4 rpm. The sludge could be discharged
by means of a sludge pump. To control the pH of
mine drainage in the NB, a pH controller was set up
with a pH probe and a metering pump. The target pH

Environ Geochem Health (2011) 33:91101

93

Fig. 1 Schematic diagram


of pilot-scale mine drainage
treatment process

of mine drainage was automatically adjusted with an


error of 0.1. Lime (Ca(OH)2) was used as neutralizing agent. Lime slurry was prepared with a lime
concentration of 0.7 g/L and injected at different flow
rates (max. 0.25 L/min) into the NB.
In this study, pilot-scale tests were conducted to
estimate not only Fe(II) oxidation efficiency but also
the efficiency of precipitation of oxidation products.
As shown in Table 1, these tests were conducted
under different operating conditions. In the treatment
process, the main operating conditions were rate of
aeration of the OB, retention time, and pH of the NB.
Analysis and physicochemical properties
of mine drainage
Samples obtained from the inlet to the pilot plant and
of the effluent from each basin were analyzed for
physicochemical properties. The pH, ORP, TDS, and
EC of mine drainage and the samples were measured
with a portable pH/ORP and EC/TDS meter (Orion 3
star) and DO meter (Pocket calorimeter, HACH). To
evaluate the effectiveness on the oxidation of Fe(II)
and precipitation of oxidative products, and to
measure total recoverable and dissolved metals of
samples obtained before and after treatment, EPA
method 3005A (acid digestion procedure) was used
(EPA 1992). For analysis of total recoverable metals,
40 mL drainage sample was measured with a measuring cylinder and transferred to a conical tube, and
0.8 mL conc. HNO3 and 2 mL conc. HCl were
injected. The samples were then digested by use of a

heating block at 9095C until the sample volume


reduced to 810 mL. The digested samples were
cooled to room temperature, and the volume was
measured and diluted to 40 mL with de-ionized
water, followed by filtration through a 0.45-lm pore
membrane filter. Concentrations of heavy metals in
the filtrate were measured by ICP-AES (720-ES,
Varian). The alkalinity of samples of pH [ 4.5, was
measured as follows:
addition of an indicator (bromocresol green methyl
red) to the sample;
titration of the sample with H2SO4 solution of
appropriate concentration, by use of a titrator adapted
with a cartridge, until the color of the solution
changed to light pink; then
measurement of the volume of H2SO4 solution
titrated to calculate the alkalinity.
To measure the Fe(II) concentration on-site, the
sample was filtered using a 0.45-lm pore membrane
filter, then ferrozine iron reagent was added into the
filtrate, and after 3 min Fe(II) was measured by UV
visible absorption (DR 2800, Hach) at 510 nm. Table 1
shows the physicochemical properties of the mine
drainage. The mine drainage has a neutral pH (6.6) and
93 mg/L alkalinity. Use of Eq. (1) showed the mine
drainage had net alkalinity (45 mg/L).
Net alkalinity alkalinitymeasured  aciditycalculated
1
The dissolved oxygen (7.8 mg/L) in the mine
drainage was also highly indicative of relative
oxidative condition. Total dissolved solids (TDS) is

123

94

Environ Geochem Health (2011) 33:91101

Table 1 Physicochemical properties of mine drainage


Temp. (C)

pH

ORP (RmV)

TDS (mg/L)

EC (ls/cm)

Fe2? (mg/L) filtered

Alkalinity (mg/L)

DO (mg/L)

11.8

6.6

237.1

245

500

11.15

93

7.78

Table 2 Concentrations of total recoverable metals and anions in mine drainage (unit: mg/L)
Al

As

Ca

Cd

Co

Cr

Cu

Fe

Sb

0.023

0.003

60.9

ND

0.007

0.005

ND

25.28

1.66

0.005

Li

Mg

Mn

Mo

Na

Ni

Pb

Si

0.04

16.28

1.39

ND

0.91

0.01

0.022

ND

69.39

3.55

Sn

Sr

Ti

Zn

F-

Cl-

NO2-

NO3-

0.96

1.03

0.001

0.047

0.001

0.01

0.501

0.919

0.167

ND

the total amount of cations and anions, and has a


large value when the concentration of sulfate in mine
drainage is high. Considering the analytical results
for DO, pH, Fe(II), and TDS, the mine drainage was
not highly contaminated, and had net alkalinity. The
analytical results for total recoverable metals in
mine drainage showed that the main metal species
([1 mg/L) in mine drainage were Fe, Mn, Si, Ca, and
Mg (Table 2). As the main metal species for the
occurrence of turbidity, the concentration of total
recoverable Fe was 25.3 mg/L. On-site measurement
for Fe(II) was 11.15 mg/L so the ratio of Fe(II) to
total recoverable Fe was low44.1%. In other words,
the mine drainage sampled was highly oxidized.

the kinetics the aeration rate was set constant at


200 L/min and the flow rate of mine drainage was
0.23 L/s. In this test the retention time in the OB was
1.2 h. Figure 2 shows the change of pH, DO, and
ferrous oxidation efficiency with time in the OB.
About 1 h was taken to increase pH from 6.75 to 7.14
whereas only half hour was taken to increase DO
from 7 mg/L to 11 mg/L (Fig. 2a, c). Oxidation of
Fe(II) increased linearly to approximately 28% in 1 h
then decreased slightly. Accordingly, pH has the
dominant effect on the ferrous oxidation efficiency.
Therefore, because the ferrous oxidation efficiency
was rather low under these experimental conditions,
different rates of aeration was applied in the OB to
observe the efficiency of oxidation of Fe(II).

Pilot-scale experimental setup


In this study, pilot-scale tests were conducted to
estimate not only Fe(II) oxidation efficiency but also
the efficiency of precipitation of the products of
oxidation. As shown in Table 3, these tests were
conducted under different operating conditions. In the
treatment process, the main operating conditions
were aeration rate, retention time, and pH.
Results and discussion
Kinetics of dissolved oxygen and Fe(II) oxidation
in the OB
Changes in dissolved oxygen (DO) and Fe(II) oxidation were studied as a function of time. In the study of

123

Effect of aeration rate on the efficiency


of oxidation of Fe(II)
To determine the effect of aeration rate on Fe(II)
oxidation efficiency, aeration rate was altered in the
range 0450 L/min and the flow rate of mine drainage
and retention time were fixed at 0.23 L/min and 1.2 h
in the OB. Without aeration in the oxidation basin, the
efficiency of oxidation of the entire process was less
than 10% (Fig. 3). As the aeration rate increased to
300 L/min, the efficiency of oxidation of Fe(II)
increased to approximately 50 and 75% in the OB
and SB, respectively. However, the efficiency of
oxidation of Fe(II) in the OB did not increase further
when the rate of aeration was increased to more than
300 L/min. For all rates of aeration used in the OB,

Environ Geochem Health (2011) 33:91101

95

Table 3 Phases of pilot-scale tests


Phase

Type of test

Fixed conditions

Kinetic test for Fe(II) oxidation efficiency and DO change

Aeration: 200 L/min


Retention time: 1.2 h

II

Dependence of Fe(II) oxidation efficiency in each basin on the aeration


rate (0450 L/min) in the oxidation basin

Retention time: 1.2 h

III

Dependence of Fe(II) oxidation efficiency on retention time (0.8, 1.2, and 2 h)

Aeration rate: 300 L/min

IV

Dependence of Fe(II) oxidation efficiency on the pH (no adjustment,


7.5, 8.0, and 8.5) in the neutralization basin

Aeration rate: 300 L/min

7.2

100

Oxidation efficiency of ferrous (%)

pH

7.0

6.8

Dissolved oxygen (mg/L) Oxidation efficiency of ferrous (%)

Retention time: 1.2 h

40

B
30

SB

80

OB

NB

RB 300 L/min
400 L/min
450 L/min

60

200 L/min

40

20
control

Time (hr)
20

Fig. 3 Fe(II) oxidation efficiency in each basin according to


aeration rate in the oxidation basin
10

the efficiency of oxidation of Fe(II) increased as mine


drainage flowed further into other basins. This might
be because of the increased retention time of mine
drainage (NB: 0.48 h, RB: 0.48 h, and SB: 1.2 h).
Among all cases, the optimum aeration rate for
achieving maximum Fe(II) oxidation in the SB was
300 L/min. Thus, the following pilot-scale tests were
conducted with 300 L/min aeration rate.

C
12

10

Effect of retention time on the efficiency


of oxidation of Fe(II)
0

20

40

60

80

100

Time (min)
Fig. 2 Effect of aeration in the oxidation basin on a dissolved
oxygen, b pH, and c Fe(II) oxidation efficiency (aeration rate:
200 L/min, retention time: 1.2 h)

The efficiency of oxidation of Fe(II) in each basin


was further observed with different retention times
(0.8, 1.2, 2.0 h) (Fig. 4). The aeration rate was fixed
at 300 L/min. The efficiency of oxidation of Fe(II) in

123

96

Environ Geochem Health (2011) 33:91101

sites on the surface (Sung and Morgan 1980; Park and


Dempsey 2005). The kinetics of homogeneous and
heterogeneous oxidation of Fe(II) are highly interdependent. As pH increases, the hydrolyzed iron species
(FeOH? and Fe(OH)2) predominate over Fe2? ions,
because of hydrolysis (Barnes et al. 2008). As
hydrolysis increases, the kinetic rate increases.
In the presence of dissolved oxygen (DO), the
homogeneous oxidation of Fe(II) can be expressed as
shown in Eq. (2). The oxidized species, Fe(III), is
instantly hydrolyzed by the supply of alkalinity to
Fe(III)(oxy)hydroxide compounds which eventually
precipitate (Eq. 3).

Oxidation efficiency of ferrous (%)

100
SB

OB
RB

NB
80

SB
OB
NB

60

RT 0.8hr at OB
RT 1.2hr at OB
RT 2.0 hr at OB

OB
SB

40
NB RB
20

0.0

1.5

3.0

4.5

6.0

Time (hr)
Fig. 4 Fe(II) oxidation efficiency in each basin with different
retention times

the OB and SB was approximately 22 and 38%,


respectively, for 0.8 h RT. As the RT increased to 1.2
and 2.0 h, the efficiency of oxidation of Fe(II)
increased linearly to approximately 50 and 72% in
the OB. This result might be mainly because the pH
increase as a result of aeration enhances the rate of
oxidation of Fe(II) for mine drainage with net alkalinity. Although 100% oxidation efficiency was not
observed in this series of experiments, the maximum
efficiency of oxidation of Fe(II) (91%) in the SB was
obtained when the RT was 2 h. Considering the
treatment for large-scale mine drainage, the volume
of the oxidation reactor will increase in proportion to
increasing retention time. Thus, a long retention time
for aeration could be an economic disadvantage.
Subsequent tests were conducted with a constant
retention time (1.2 h).
Evaluation of homogeneous and heterogeneous
oxidation
Abiotic oxidation of Fe(II) takes place by homogeneous and heterogeneous processes simultaneously
(Singer and Stumm 1968; Millero et al. 1987). In a
suspended solution containing Fe(III) minerals, the
surface of Fe(III)(oxy)hydroxide minerals can act as a
catalyst for Fe(II) oxidation (Azher et al. 2008). In the
heterogeneous oxidation of Fe(II), the soluble Fe(II) is
adsorbed on the surface of Fe(III)(oxy)hydroxide
minerals and the adsorbed Fe(II) is oxidized by an
electron-transfer reaction enhanced by electron-poor

123

1
1
Fe(II) O2 H2 O Fe(III) OH
4
2

Fe(III) 3H2 O Fe(OH)3 s) 3H

In the oxidation of Fe(II), oxidation of dissolved


Fe(II), including Fe2?, FeOH?, and Fe(OH)02, is
called as homogeneous oxidation and the oxidation of
adsorbed Fe(II) as heterogeneous oxidation.
The rate of homogeneous oxidation of Fe(II) was
expressed by Stumm and Lee (1961) as:
Homogeneous rate 

dFeII k1 Fe2 O2 

dt
fH g2
4

where, k1 is the rate constant for homogeneous


oxidation of Fe(II).
Sung and Morgan (1980) and Tamura and Nagayama (1976) expressed the rate of heterogeneous
oxidation in terms of [H?] (Eq. 4):
Heterogeneous rate 

dFeII k2 Fe3 O2 

dt
fH g

The homogeneous and heterogeneous oxidation


processes can be affected by pH in the range between
5 and 8, and the overall oxidation rates can be
expressed by Eq. 5.
Overall abiotic rate for pH [ 5 

k1 k2 Fe3 H Fe2 O2 


fH g2

dFeII
dt
6

If all conditions are assumed to be constant except


the amount of Fe(II), this system follows pseudo-first
order reaction kinetics. The total rate constant, ktotal,
can be obtained by use of Eq. 7:

Environ Geochem Health (2011) 33:91101


0.0004

where, Vreactor, Qtotal, Fe(II)influent, and Fe(II)effluent are


the reactor volume, flow rate, Fe(II) concentration in
the influent, and Fe(II) concentration in the effluent,
respectively. By coupling with Eq. 6, Eq. 8 is
obtained.
2
dFe(II)]
dt H 

Fe2 O2 
Since ktotal

k2 Fe3 H  k1

dFedtII
Fe2 

-1

Total reaction rate (Ktotal, s )

FeIIinfluent  FeIIeffluent  Qtotal


Fe(II)effluent Vreactor

0.0003

0.0001

0.0000

10

15

20

25

30

Fe (III) conc. (mg/L)


3e-19

ktotal H 2
k2 Fe3 H  k1
O2 
By plotting [Fe3 H  against

y = 1E-05x - 2E-05
R2 = 0.3066

0.0002

; the Eq: 9 could be obtained

9
ktotal H 2
O2 

the slope

(k2) and intercept (k1) were obtained by linear


regression (Fig. 5).
The result of linear regression for these data showed
the rate constant for heterogeneous oxidation was
7 9 10-14 (mg/L)-1 s-1, which was lower than literature values (0.167 9 10-8 to 3.1 9 10-8 (mg/L)-1 s-1).
For homogeneous oxidation, the rate constant was
2 9 10-20 mol L-1 s-1, which was also lower than
literature values (1 9 10-14 to 1 9 10-12 mol L-1 s-1;
Tamura and Nagayama 1976; Sung and Morgan 1980;
Millero et al. 1987; Liang et al. 1993; Stumm and Morgan
1996; Ames 1998; Dempsey and Dietz 2001). These
discrepancies might be because the determination coefficient (0.134) of linear regression was much lower than
0.7. In general, a determination coefficient (R2) of more
than 0.7 in linear regression analysis can be regarded as
indicative of a significant relationship (Bernick et al.
1995; Kalnicky and Singhvi 2001), and R2 = 1 is an ideal
case without any scattered data. Accordingly, this
analysis might not be a reliable means of obtaining rate
constants for homogeneous and heterogeneous oxidation.
Dempsey and Dietz (2001) measured the rate
constant of Fe(II) oxidation using a sludge return
method in which a broad range of Fe(III) (up to
2,000 mg L-1) was used to obtain the rate constants.
Therefore, the low reliability obtained in this study
might be because of the low concentration of Fe(III)
applied in the system. Thus, more tests will be
conducted using sludge to obtain more reliable and
exact rate constants.

Ktotal[H+]2/[O2]

ktotal

97

y = 7E-14x + 2E-20
R2 = 0.134

2e-19

1e-19

1e-6

1e-6

2e-6
+

[Fe (III)][H ]
Fig. 5 Relationships between Fe(III) concentration and Ktotal
(a) and between Ktotal[H?]2/[O2] and [Fe(III)][H?] (b)

Figure 6 shows relationships between pH, alkalinity, and dissolved oxygen measured throughout pilot
tests in which neutralization was excluded. As shown
in Fig. 6, pH and alkalinity were in the range 6.76.8
and 92105 mg/L, respectively (average. 93 mg/L as
CaCO3) for non-aeration. However, when the OB was
aerated, pH increased to 6.857.2 and alkalinity
decreased to 7085 mg/L (avg. 75 mg/L as CaCO3).
Based on stoichiometric consideration of Eqs. 1 and
2, 1.8 mg/L alkalinity as CaCO3 could be removed by
1 mg/L Fe(II) oxidation and precipitation of the
oxidation products. Considering the oxidation of
approximately 90% of total Fe(II) which corresponds
to approximately 10 mg L-1 Fe(II), alkalinity should
be reduced by 18 mg/L. Thus, this result is a good
match with the stoichiometric concept.

123

Environ Geochem Health (2011) 33:91101


110

7.4
DO vs pH
DO vs alkalinity

pH

90
7.0
80
6.8

Alkalinity (mg/L)

100
7.2

70
aeration

6.6

10

12

60

Oxidation efficiency of ferrous (%)

98
100

NB

60

RB
SB

40

20

pH not adjusted
pH 7.5
pH 8.0
pH 8.5

OB

Dissolved oxygen (mg/L)


Fig. 6 Relationship between pH, alkalinity, and dissolved
oxygen measured throughout the pilot tests (neutralization
excluded)

80

Time (hr)
Fig. 7 Fe(II) oxidation efficiency in each basin for different
pH in the neutralization basin

Effect of pH adjustment in the NB


on the efficiency of oxidation of Fe(II)

0
-

K = 0.0551x1014[OH ]2

123

-1

(Azher et al., 2008)


-

14
2
K = 0.061x10 [OH ]

log(Ktotal, s-1)

In this test, the efficiency of oxidation of Fe(II) in each


basin was observed when the pH in the NB was
adjusted to the predetermined pH by use of lime.
When the pH in the NB was increased to more than
7.5, the efficiency of oxidation of Fe(II) in the NB and
RB reached more than 95% and almost 100% in the
SB (Fig. 7). Accordingly, the oxidation speed of
Fe(II) became rapid because of the increase of the pH
of the mine drainage. On the basis of this result,
oxidation of Fe(II) in each basin was significantly
affected by pH. After obtaining log (Ktotal) at the
corresponding pH, fitting of the data by linear
regression analysis was conducted to determine the
correlation between pH and rate constant of Fe(II)
oxidation. The determination coefficient (0.8395)
obtained by this analysis was higher than 0.7,
indicating there is a significant relationship between
the overall rate constant and pH. The slope (2.1701) of
the plot fitting the data seems to match well with
literature reports that the rate constant increases by
two orders as pH increases by 1.0 unit (Sung and
Morgan 1980; Azher et al. 2008). In general, the
kinetic constant is very dependent on ionic strength
and dissolved counteranions, and decreases when
ionic strength and anion concentration increase (Emmenegger et al. 2008). Accordingly, the alkalinity and
sulfate present at high levels in the mine drainage
might reduce the rate constant, as shown in Fig. 8.

(Sung and Morgan., 1980)

-2

-3
log(Ktotal) = 2.1701 pH - 19.208

-4

-5
6.5

7.0

7.5

8.0

8.5

9.0

pH
Fig. 8 Correlation between pH in the NB and log (Ktotal)
(R2 = 0.8395)

Change in the absorbance properties


of suspensions in the NB with pH variation
To determine the precipitation properties of suspended
solids, approximately 10 mL suspension obtained in
the RB under different pH conditions was placed in a
cell and the absorbance of the sample was measured
kinetically by UVvisible absorbance (DR 2800,
Hach) at 560 nm (Lenter et al. 2002). Before measurement of the suspension, the blank was the filtrate
obtained by use of a 0.45 lm-pore filter. Generally, the
absorbance increased with increasing suspended solids

Environ Geochem Health (2011) 33:91101

99

(SS) and decreased with precipitation of the SS. By


measuring SS and absorbance for specific suspensions
it was found they had a linear relationship (SS =
58.6 Absat 560nm - 1.08, R2 = 0.996). The objective
of this test was to observe not only precipitation of the
SS but also the effect of Fe(II) oxidation on precipitation (Fig. 9).
After measuring the absorbance of the suspension
kinetically, the agglomeration equation was used to
obtain the rate of removal of SS (Stumm and Morgan
1996). In the agglomeration equation, N represents
the number of particles in solution. In this study,
however, the rate of removal (ka) was obtained by
applying the absorbance instead of the particle
numbers, on the basis of the assumption that the
absorbance increases linearly as the number of
particles increases. The initial absorbance (N0) and
absorbance at different times (N) were used.
dN
ka N 2
dt
1
1

ka t
N N0

10
11

The absorbance of suspension when the pH was not


adjusted in the NB increased from approximately 0.26
to 0.48 in 1.5 h. Although the absorbance decreased
with further increase of time, it was even higher (0.35)
at 4 h than initially. The absorbance decreased to 0.19
with 8 h of settling time. Using the data obtained from

Absorbance at 560 nm

0.5

without pH adjustment
pH 7.5
pH 8.0
pH 8.5

0.4

0.3

0.2

0.1

0.0
0

12

Time (hr)
Fig. 9 Kinetics of absorbance at 560 nm of effluents under
different pH conditions in the neutralization basin

1.5 to 8 h of settling time, the ka obtained was


0.5441 h abs-1. This might have happened because
oxidation of the remaining Fe(II) (only about 55% of
total Fe(II) was oxidized) was dominant at the initial
period and then precipitation of oxidized products
took place in the suspension. For suspensions with pH
adjusted to 7.5 and higher in the NB, the absorbance
dropped rapidly without the increase in absorbance.
The ka was obtained after 3 h of settling time. Among
suspensions with the pH adjusted to 7.5 and higher,
the removal rate was highest (7.897 h abs-1) for the
suspension adjusted to pH 7.5, and was approximately
14.5 times faster than for unneutralized suspension.
The absorbance of the suspension adjusted to pH 7.5
decreased to zero after 4 h of settling time.

Conclusions
The efficiency of Fe(II) oxidation, and precipitation
of the oxidized products, as a result of aeration of
the OB and pH variation in the NB was investigated
in this study. When aeration rates in the OB were
varied (0450 L/min) and the RT was fixed (1.2 h),
pilot-scale tests on the oxidation of Fe(II) showed
that aeration at 300 L/min resulted in the optimum efficiencyapproximately 50% oxidation. With
higher aeration rates, oxidation efficiency did not
increase. With increasing RT in the OB, the efficiency
of oxidation of Fe(II) increased linearly and efficiency
was approximately 72% when the RT was 2 h.
However, Fe(II) still remained in subsequent basins
(neutralization, reaction, and settling basins). Because
of the limited volume of the oxidation basin and the
limited geological configuration of the site, it was not
possible to increase the volume of aeration unit or its
retention time in order to achieve 100% Fe(II)
oxidation. Results from pilot-scale tests in which
neutralization was excluded were used to obtain the
rate constants of heterogeneous and homogeneous
oxidation. The results showed the rate constants were
lower than literature values. Because a very low
determination coefficient resulted from the modeling,
however, the analysis was not suitable for obtaining
reliable values of the rate constants. Enhancement of
the oxidation of Fe(II) by changing the pH in the NB
was also investigated. Oxidation of Fe(II) reached
almost 100% when the pH of the mine drainage was
increased to more than 7.5, and there was a linear

123

100

relationship between log (Ktotal) and pH, the slope of


which was approximately 2.1. Thus, the rate of
oxidation of Fe(II) increased by two orders of
magnitude when the pH of the suspension was
increased by one unit.
Changes in the absorbance of samples obtained
from the NB under different pH conditions were
measured to determine the precipitation properties of
suspended solids. Consideration of all the results
revealed the trend of changes in absorbance with pH
was closely related to oxidation of Fe(II) and precipitation of the oxidized products. Because ferrous
remained in the inflow to the SB, oxidation of Fe(II)
was predominant initially, leading to increased absorbance, and the rate of precipitation was slow. However, the absorbance of the suspension in the SB
dropped rapidly, without a subsequent increase, when
the pH of the mine drainage was adjusted to more than
7.5. This might be because only precipitation
occurred, because all the Fe(II) was oxidized before
the settling basin. Compared with the system in which
Ca(OH)2 was used, the precipitation speed for the
control (suspension not treated with Ca(OH)2) was
much lower. Therefore, the density of floc was lower,
because there was no heterogeneous material, for
example Ca(OH)2, to enhance the floc density, and the
chance of polymerization and enlargement of the
oxidized products was reduced because of the low
collision frequency under the laminar conditions
characteristic of the settling basin, For suspensions
of pH higher than 7.5, however, the absorbance
decreased directly without an initial increase of
absorbance, because there was no Fe(II) left to
oxidize. As shown above, more enhanced treatment
could be designed to treat large-scale mine drainage
with net alkalinity if appropriate aeration and
subsequent pH adjustment could be found. These
results from pilot-scale tests could be expected to
contribute to the design of full-scale plant on the
actual site by constructing an optimized process on the
uneven ground of the site.

References
Akcil, A., & Koldas, S. (2006). Acid mine drainage (AMD):
Causes, treatment and case studies. Journal of Cleaner
Production, 14, 11391145.

123

Environ Geochem Health (2011) 33:91101


Ames, R. P. (1998). Iron oxidation, gas transfer, and solids
formation in passive treatment systems for mine drainage.
Pennsylvania: The Pennsylvania State University.
Azher, N. E., Gourich, B., Vialb, C., Soulami, M., & Ziyad, M.
(2008). Study of ferrous iron oxidation in Morocco
drinking water in an airlift reactor. Chemical Engineering
and Processing, 47, 18771886.
Barnes, A., Sapsford, D. J., Dey, M., & Williams, K. P. (2008).
Heterogeneous Fe(II) oxidation and zeta potential. Journal of Geochemical Exploration, 100, 192198.
Bernick, M. B., Getty, D., Prince, G., & Sprenger, M. (1995).
Statistical evaluation of field-portable X-ray fluorescence
soil preparation methods. Journal of Hazardous Materials, 43, 111116.
Chon, H. T., & Hwang, J. H. (2000). Geochemical Characteristics of the Acid Mine Drainage in the Water System
in the Vicinity of the Dogye Coal Mine in Korea. Environmental Geochemistry and Health, 22, 155172.
CIPB (1995) Study on the environmental improvement in
mining areas (wastewater and waste rocks) (in Korean):
Coal Industry Promotion Board.
Dempsey, B. A., & Dietz, J. M. (2001). A demonstration of
modified passive abiotic treatment of mine drainage.
University Park: Pennsylvania State University.
Emmenegger, L., King, D. W., Sigg, L., & Sulzberger, B. (2008).
Oxidation kinetics of Fe(II) in a eutrophic Swiss lake.
Environmental Science and Technology, 32, 29902996.
EPA. (1992). Acid digestion of waters for total recoverable or
dissolved metals for analysis by FLAA or ICP spectroscopy: SW846, method 3005A. Washington, DC: US EPA,
Office of Solid Waste and Emergency Response.
Gazea, B., Adam, K., & Kontopoulos, A. (1996). A review of
passive systems for the treatment of acid mine drainage.
Minerals Engineering, 9, 2342.
Hedin, R. S. (2008). Iron removal by a passive system treating
alkaline coal mine drainage. Mine Water and the Environment, 7, 200209.
Johnson, D. B., & Hallberg, K. B. (2005). Acid mine drainage
remediation options: A review. Science of the Total
Environment, 338, 314.
Jung, M. C. (1994). Sequential extraction of heavy metals in
soils and a case study. Korean Social Economic and
Environmental Geology, 27, 469477.
Jung, M. C. (2003). Environmental assessment for acid mine
drainage by past coal mining activities in the Youngwol,
Jungseon and Pyungchang areas, Korea (in Korean).
Economic and Environmental Geology, 36, 111121.
Kalnicky, D. J., & Singhvi, R. (2001). Field portable XRF
analysis of environmental samples. Journal of Hazardous
Materials, 83, 93122.
Lenter, C. M., McDonald, J. L. M., Skousen, J. G., &
Ziemkiewicz, P. F. (2002). The effects of sulfate on the
physical and chemical properties of actively treated acid
mine drainage floc. Mine Water and the Environment, 21,
114120.
Liang, L., Mcnabb, J. A., Paulk, J. M., Gu, B., & Mccarthy, J. F.
(1993). Kinetics of Fe(II) oxygenation at low partial
pressure of oxygen in the presence of natural organic
material. Environmental Science and Technology, 27,
18641870.

Environ Geochem Health (2011) 33:91101


Millero, F. J., Sotolongo, S., & Izaguirre, M. (1987). The
oxidation kinetics of Fe(II) in seawater. Geochimica
et Cosmochimica Acta, 51, 793801.
Park, B., & Dempsey, B. A. (2005). Heterogeneous oxidation
of Fe(II) on ferric oxide at neutral pH and a low partial
pressure of O2. Environmental Science and Technology,
39, 64946650.
Sengupta, M. (1993). Environmental impacts of mining:
Monitoring, restoration, and control. Boca Raton: Lewis
publishers.
Singer, P.C. and Stumm, W. (1968) Kinetics of the oxidation of
ferrous iron. In the Second Symposium of Coal Mine
Drainage. pp. 12-34.

101
Stumm, W., & Lee, G. F. (1961). Oxygenation of ferrous iron.
Industrial & Engineering Chemistry Research, 53,
143146.
Stumm, W., & Morgan, J. J. (1996). Aquatic chemistry. New
York: Wiley.
Sung, W., & Morgan, J. J. (1980). Kinetics and products of
ferrous iron oxygenation in aqueous systems. Environmental Science and Technology, 14, 561568.
Tamura, H., & Nagayama, M. (1976). Effect of anions on the
oxygenation of ferrous ion in neutral solutions. Journal of
Inorganic and Nuclear Chemistry, 38, 113117.

123

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

You might also like