Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Ecotoxicology and Environmental Safety 172 (2019) 281–289

Contents lists available at ScienceDirect

Ecotoxicology and Environmental Safety


journal homepage: www.elsevier.com/locate/ecoenv

In situ bioremediation of hexavalent chromium in presence of iron by dried T


sludge bacteria exposed to high chromium concentration
Neetu Bansal , Johan J. Coetzee, Evans M.N. Chirwa

Department of Chemical Engineering, University of Pretoria, Pretoria 0002, South Africa

ARTICLE INFO ABSTRACT

Keywords: Stability of chromium in the ferrochrome slag dumps and leachate are affected by pH, redox potential and the
In situ bioremediation presence of other metallic species in the slag. It is desirable to keep chromium in slag dumps in the trivalent [Cr
Chromium stabilisation (III)] state because trivalent chromium is 1000 times less toxic to living organisms than the hexavalent form [Cr
Biocatalysis (VI)]. Due to the low toxicity and low mobility of Cr(III), it is recommended to convert Cr(VI) to Cr(III) wherever
Ferrochrome slag
possible to protect the health of living organisms. In this study, the role of Cr(VI) reducing organisms for sta-
Chromium reducing bacteria
Inhibition kinetics
bilising chromium in slag dumps was evaluated in the presence of iron [oxidation states Fe(II) and Fe(III)]. The
study showed that stabilisation of chromium species in the trivalent state was most favourable under aerated
conditions. Up to 100 mg/L Cr(VI) was reduced in less than 24 h by cultures grown under aerobic conditions in
the presence of Fe(III). A much shorter time (6 h) was required to reduce the same amount of Cr(VI) in the
presence of Fe(II). When oxygen was completely excluded, it was only possible to reduce 20 mg/L in about 48 h
which was much slower than the removal of 100 mg/L in less than 24 h under aerated conditions. Fe(II) con-
tributed directly to catalytic reduction of Cr(VI) reduction whereas Fe(III) was beneficial to Cr(VI) reduction up
to an initial Cr(VI) concentration of 75 mg/L. Evaluation of Cr(VI) reduction kinetics showed that Cr(VI) re-
duction under aerobic conditions followed the non-competitively inhibited mixed-order reaction. Cr(VI) re-
duction in sealed reactor vessels, under anaerobic conditions, followed a modified non-competitive inhibition
reaction model. The results indicate that chromium stabilisation in ferrochrome slag dumps would require
maintenance of a fully aerated dump supplemented by a culture of Cr(VI) reducing organisms.

1. Introduction (Satarupa and Paul, 2013). The release of Cr(VI) into the surrounding
environment can lead to bioaccumulation in humans and mammals,
Chromium in the environment exists in a range of oxidation states while also causing major health issues (Coetzee et al., 2018; Guertin
ranging from Cr(-II) to Cr(+VI). However, the most prevalent forms in et al., 2016). This especially problematic in situations where the lea-
the environment are chromium oxide [Cr(III)2O3] found in rocks and chate water enters agricultural supply water. Under these conditions,
chromite [FeCr(III)2O4] which is a significant component of geological there is heightened risk of contamination of food products and bioac-
chrome ore seams. Chromite ore contains up to 80% chromium mainly cumulation in fish and livestock products (Chirasha and Shoko, 2010).
in the trivalent state (Dhal et al., 2013; Erdem et al., 2005). On the Chromium, in its trivalent state, readily precipitates as chromium
other hand, chromium emanating from chromium refining is dis- hydroxide, Cr(OH)3(s), at near neutral to alkaline pH conditions
charged mainly as slag, sludge and/or dust slurry containing high levels (pH > 6). The tendency of Cr(III) to precipitate makes Cr(III) less mo-
of chromium in the hexavalent form [Cr(VI)] (van Staden et al., 2014). bile in the water/leachate which makes it much easier to manage its
High temperatures in kilns and chemicals used during refining convert ecological impacts as against Cr(VI) (Dogan et al., 2011; Madhavi et al.,
Cr(III) from the rocks to Cr(VI), which consequently produces waste 2013). Furthermore, the Cr(VI) is highly mobile in water and is known
product streams with high levels of Cr(VI). The slag portion of the waste to be carcinogenic in mammalian cells (Federal Register, 2004;
makes up the largest volume and is typically discharged into slag dumps Zhitkovich, 2011).
(slag heaps) sometimes together with rock residue tailings (Panda et al., Chromium in leachate and dust slurry is usually treated using con-
2013). Leachate from the long–term storage of waste and slag pose a ventional physical-chemical processes by applying chemical reducing
perpetual risk to the higher order organisms living in receiving waters agents to reduce Cr(VI) to Cr(III) followed by precipitation of Cr(III) as


Corresponding author.
E-mail address: neetu.bansal@uqconnect.edu.au (N. Bansal).

https://doi.org/10.1016/j.ecoenv.2019.01.094
Received 8 June 2018; Received in revised form 25 January 2019; Accepted 28 January 2019
Available online 01 February 2019
0147-6513/ © 2019 Elsevier Inc. All rights reserved.
N. Bansal et al. Ecotoxicology and Environmental Safety 172 (2019) 281–289

a chromium hydroxide at a higher pH (< 8.0). This process leaves flasks containing 200 mL LB broth and covered with cotton wool plugs.
behind a toxic chemical sludge that is even more costly to treat (Fiúza Anaerobic cultures were grown in 100 mL serum bottles that were
et al., 2010; Hawley et al., 2004; Stasinakis et al., 2003). Alternatively, purged with pure nitrogen gas (99% pure grade) and sealed with sili-
a more passive biological process could be used to facilitate conversion cone rubber stoppers and aluminium caps.
of the Cr(VI) to ensure long-term protection of the environment. The Pure cultures were grown by transferring 1 mL of a serial diluted
viability of in situ biotransformation of Cr(VI) to Cr(III) has been de- sample onto a petri dish containing LB Agar. The cell was then in-
monstrated in both laboratory scale and pilot scale systems for effective cubated at 30 °C to develop separate identifiable colonies. The in-
removal of Cr(VI) from water but has not been applied at full scale dividual colonies were transferred using a sterile wired loop and the
(Bharagava and Mishra, 2018; Francisco et al., 2002; Jobby et al., 2018; process was repeated three times until there were no more separate
Liu et al., 2018). identifiable colonies.
In this study, the reduction of toxic hexavalent chromium by natural
occurring bacteria in the presence of Fe(II) and Fe(III), a common co- 2.2.3. Culture characterization
pollutant that exists in most chrome refinery waste and residue from ore Individual colonies developed were characterized by using phylo-
extraction, is evaluated. The hypothesis is tested, if the ferrous/ferric genetic characterization of the cells from the 7th to the 10th tube in the
redox couple may act as a cyclic catalyst for Cr(VI) reduction to the serial dilution. Genomic DNA was extracted from the pure cultures
trivalent state, and as a stabilizer of chromium in the less toxic trivalent using a DNeasy tissue kit (QIAGEN Ltd, West Sussex, UK) as per man-
state. A Cr(VI) reducing culture isolated from Cr(VI) contaminated ufacturer's instructions. The 16S rRNA genes of isolates amplified by
waste water plant sludge was used as the biocatalyst to achieve the reverse transcriptase-polymerase chain reaction (RT-PCR) using pri-
stabilisation of chromium as Cr(III) in aqueous solution. mers pA and pH1 (Primer pA corresponds to position 8–27; Primer pH
to position 1541–1522 of the 16S gene) (Coenye et al., 1999). An in-
2. Materials and methods ternal primer pD was used for sequencing (corresponding to position
519–536 of the 16S gene). Species identification was conducted by
2.1. Chemical reagents BLAST search of the NCBI database. Phylogenetic tree diagrams were
then constructed using the neighbour-joining method (Tamura et al.,
Analytical grade reagents with high purity level (99% and above) 2013). Confidence in the tree topology was determined by the bootstrap
were purchased from Sigma-Aldrich Pty Ltd (Industrial Park, Jet Park, analysis based on 100 re-samplings (Felsenstein, 1985).
Johannesburg, South Africa) used in the experiments without further
purification. Standard solutions of chromium were prepared daily by 2.3. Ferrochrome slag samples
diluting 1 g/L of chromium stock solution. 1,5-diphenylcarbazide (DPC)
solution was prepared by dissolving specific amount of DPC in HPLC Ferrochrome slag samples for evaluation of chromium and iron
grade acetone and stored in a brown bottle covered with aluminium species content in tailings was collected from waste dumps located at
foil. 0.85% NaCl was prepared in distilled water and sterilized by au- separate open and closed furnace ferrochrome smelters located in the
toclaving at 125 °C and 30 psi for 15 min. Iron species were added as North West Province (South Africa). Only samples from mixed waste
FeCl3·6H2O for Fe(III) and Mohr's salt for Fe(II). Mineral Salt Media tailing dumps were used during this study.
(MSM) was prepared by dissolving: 10 mM NH4Cl, 30 mM Na2HPO4,
0.8 mM Na2SO4, 0.2 mM MgSO4, 50 μM CaCl2, 25 μM FeSO4, 0.1 μM 2.4. Batch studies
ZnCl2, 0.2 μM CuCl2, 0.1 μM NaBr, 0.05 μM Na2MoO2, 0.1 μM MnCl2,
0.1 μM KI, 0.2 μM H3BO3, 0.1 μM CoCl2, and 0.1 μM NiCl2 into 1 L of 2.4.1. Aerobic reduction experiments
distilled water. The MSM solution was sterilized at 125 °C for 15 min Aerobic cultures were grown in 1 L Erlenmeyer flasks containing
and 1 g/L of D-glucose was added to act as carbon source for the bac- 400 mL LB Broth spiked with 50 mg/L Cr(VI) for 24 h. Cells were har-
teria. vested by centrifuging at centrifugal force of 3220×g for 10 min at 5 °C.
Luria–Bertani (LB) broth and Luria-Bertani (LB) agar media were The supernatant was decanted and the remaining pellet was washed
prepared by dissolving (in 1 L distilled water), 25 g and 35 g of the twice in a sterile saline solution (0.85% NaCl) while centrifuging.
respective commercial media powder purchased from Sigma–Aldrich Aerobic Cr(VI) reduction experiments were conducted in 250 mL
(Johannesburg, South Africa). Solid agar media was allowed to cool Erlenmeyer flasks containing 20, 50, 100, 200, and 400 mg/L Cr(VI).
down to 30 °C after sterilisation at 121 °C for 15 min before introducing An initial 1 mL sample was taken to measure initial concentration of Cr
samples for streaking and/or colony development. (VI) before the cells were inoculated. The flasks were covered with
cotton wool plugs to prevent contamination, while still allowing oxygen
2.2. Bacterial culture transfer. The batches was incubated at 30 ± 2 °C with continuous
shaking on a lateral shaker (Labotec, Gauteng, South Africa). All ex-
2.2.1. Sources of chromium reducing bacteria periments were conducted in duplicate and performed using cells har-
The mixed culture of bacterial culture was collected from the sand vested during the log-growth phase (16–24 h incubation). 1 mL with-
drying beds and dried sludge from the belt filter press at the Brits drawn at regular intervals were stalled for Cr(VI) and total chromium
Wastewater Treatment Works in Brits (North West Province, South determination over time. The effect of Fe(II) and Fe(III) on the reduc-
Africa). The samples were stored in sterile containers at 4 °C. The source tion of Cr(VI) was evaluated at the optimum initial Cr(VI) concentration
of organisms was chosen because the Wastewater Treatment Works had of 50 mg/L using added Fe(II) and Fe(III) concentrations at 5, 20, 50,
received periodic flows of accidental discharges from an abandoned 75, and 100 mg/L.
Ferrochrome processing foundry, which discharged varying loadings of
sodium dichromate into the municipal system. 2.4.2. Anaerobic reduction experiments
Anaerobic cultures were grown in 100 mL serum bottle containing
2.2.2. Culture isolation 100 mL LB Broth (purged with 99% pure N2 gas for 15 min) spiked with
Initial cultures were grown in Luria-Bertani (LB) broth by in- 50 mg/L Cr(VI) for 24 h. 5 mL of the cells were then transferred anae-
oculating the solution with a grain (0.2 g) of collected sludge sample. robically to 500 mL serum bottles containing 400 mL LB broth and also
The inoculated broth was also spiked with 50 mg/L Cr(VI) and in- grown for 24 h. Cells were harvested by centrifugation at 3220×g for
cubated for 24 h at 30 ± 2 °C in a lateral shaker (Labotec, Gauteng, 10 min at 5 °C. The supernatant was decanted and the remaining pellet
South-Africa). Aerobic cultures were grown in 500 mL Erlenmeyer was washed twice in a sterile saline solution (0.85% NaCl) after each

282
N. Bansal et al. Ecotoxicology and Environmental Safety 172 (2019) 281–289

centrifugation cycle. Anaerobic Cr(VI) reduction experiments were by fusion with lithium metaborate. 0.3 g of sample was carefully mixed
conducted in 100 mL serum bottles containing 20, 50, 100 mg/L Cr(VI) with 4 g lithium metaborate in platinum crucible. The crucible con-
in absence of iron and 5, 20, 50, 75, 100 mg/L Cr(VI) in presence of Fe taining samples was heated to 1000 °C for 4 h. After cooling the sample,
(II) or Fe(III). An initial 1 mL sample was taken to measure initial fused mass was heated with C HCl at boiling point until dissolved
concentration of Cr(VI) before the cells were inoculated. The serum (Burman et al., 1978; FDA USGS, 2004). The solution was filtered
bottles containing the suspended cells were then purged with 99% pure through Whatman paper No 42 having diameter 2.5 cm. The final vo-
N2 gas and sealed with a silicon stopper and aluminium cap and then lume of the solution was made up to 100 mL by adding distilled water
incubated at 30 ± 2 °C with continuous shaking on a lateral shaker for analysis by ICP–AES.
(Labotec, Gauteng, South Africa). Cells were harvested during the log-
growth phase to ensure maximum viability of the cells used in the ex- 2.6. Biomass analysis
periment The effect of Fe(II) and Fe(III) on the reduction of Cr(VI) was
performed like aerobic batches except in anaerobic conditions. Quantitative analysis of biomass was conducted gravimetrically.
Biomass samples were collected by withdrawing 5 mL from the ex-
2.4.3. Abiotic experiments perimental batches at regular time intervals (0, 6, 14, 24 h). The sample
Abiotic experiments were conducted by heat killing grown cells was centrifuged at a centrifugal force of 3220×g for 10 min. The re-
before inoculation to determine the abiotic reduction of Cr(VI). maining pellet was then re-suspended in 1 mL distilled water. The re-
Cultures were harvested by centrifuging for 10 min at centrifugal force suspended solution was then filtered through a pre-weighed Whatman
of 3220×g, and then washed twice in a sterile saline solution (0.85% filter paper (No. 1). The filter paper was then dried at 75–80 °C and
NaCl). The remaining pellet was used for the reduction experiments. cooled in desiccator and weighed. The samples were dried until a
The supernatant from the biomass analysis was used for pH measure- constant weight was obtained. Total suspended cell weight was calcu-
ments of the experimental batches. lated from the difference in weight of the filter paper before and after
filtration (Molokwane, 2010).
2.5. Analytical methods
3. Reduction kinetics and model development
2.5.1. Cr(VI) analysis
Aliquote samples were centrifuged in 2 mL Eppendorf tubes for The reaction rate used in this study was developed earlier based on
10 min at 3220×g and 5 °C in a Minispin® Microcentrifuge (Eppendorf, the assumption that each Cr(VI) reducing living cell in the reactor
Hamburg, Germany) to remove suspended cells. The supernatant was produces a certain amount of enzyme (E′) and that deactivating the cell
used for Cr(VI) concentration analysis. Cr(VI) concentration in the su- deprives the system of that particular amount of enzyme (Shen and
pernatant was measured using a UV/Vis Spectrophotometer (WPA, Wang, 1994). Assuming that the remaining active enzyme Ea is pro-
Light Wave II, and Labotech, South Africa) after reacting acidified portional to the amount of live cells (X) in the system (Ea∝X), then the
samples with 0.2 mL 1,5 DPC to yield a bright yellow colour. The so- reaction rate catalysed by the enzymes could be considered as being
lution was scanned at the wavelength γ = 540 nm across a 10 mm light directly proportional to the reaction rate catalysed by cells. Based on
path (Eaton et al., 2005). Fe(II) and Fe(III) is known to cause inter- these assumptions, Shen and Wang (1994) derived the Cr(VI) reduction
ference to the spectrophotometric measurement of Cr(VI) using the rate law of a closed batch system, represented by the specific Cr(VI)
reactive dyes (Mahesvari and Balasabramanian, 1996). The interference reducing activity of living cells in the system:
was eliminated by acidifying the samples by adding 1.0 mL of 1 N sul-
furic acid to the 10 mL flasks before dilution of the sample to the 10 mL dC k C Co C
= mc Xo
mark. A similar method was used (Mahesvari and Balasabramanian, dt Kc + C Rc (1)
1996) to eliminate Fe(II) and Fe(III) interference in a spectro- −3
where C = Cr(VI) concentration (ML ) at any time t (T), kmc =
photometric method based on ion-pair formation with the cationic
maximum specific Cr(VI) reduction rate coefficient (T−1), Xo = initial
dye–rhodamine 6G. The elimination of the interference was verified by
viable cell concentration (ML−3), Kc = half velocity constant (ML−3),
preparation of standard curves of both species in medium without Cr
and the term (Co-C)/Rc represents cells killed due to exposure to Cr(VI)
(VI) which showed no linear correlation between Fe(II), Fe(III) and
such that the term brackets is the concentration of active cells available
absorbance after the above treatment.
at time t:

2.5.2. Total chromium analysis X = Xo


Co C
Total chromium was measured at a wavelength of 359.9 nm using a Rc (2)
Varian AA–1275 Series Flame Atomic Adsorption Spectrophotometer
where X = viable cell concentration (ML−3) at any time t, Co = initial
(AAS) (Varian, Palo Alto, CA) equipped with a 3 mA chromium hollow
Cr(VI) concentration (ML−3), and Rc = finite Cr(VI) reduction capacity
cathode lamp. Before analysis, samples were diluted 5–50 time with
(grams Cr(VI) reduced/gram of cells deactivated) (MM−1). However, it
distilled water to bring the concentration of samples in calibration
was observed that the Cr(VI) reduction rate at different initial Cr(VI)
range of the instrument. The lower Instrument Detection Limit (IDL) of
concentrations (Co) produced different values of the maximum specific
the AAS is 0.1 mg/L. Cr(III) was determined as the difference between
Cr(VI) reduction rate coefficient (kmc).
total chromium measured by AAS and Cr(VI) concentration measured
One possible explanation for this phenomenon is that, due to the
by UV/Vis spectrophotometer using the 1,5 DPC method. The AAS was
fast cell growth rate under aerobic conditions, electrons are released in
calibrated prior to total chromium analysis using 1–5 mg/L Cr(VI)
excess such that Cr(VI) reduction in such a system is non–competitive.
concentration prepared from Cr(VI) stock.
A non–competitive Cr(VI) reduction expression was therefore proposed
in which increased Co values decreased the impact of the kmc re-
2.5.3. Characterization of metals in the slag dumps
presented by Eq. (3) (Molokwane, 2010):
The composition of metallic species and chromium levels in slag,
dust and leachate sample was measured to evaluate the need for de- dC kmc C Co C
= Xo
velopment of remediation steps to prevent mobilization of Cr(VI) in dt C
1 + Ko Kc + C Rc
(3)
waste rock dumps. Metallic species in leachate were measured using the
I

−3
Inductively Coupled Plasma–Atomic Emission Spectroscopy (ICP–AES) where KI = coefficient of inhibition (ML ). When cells grown without
(Perkin–Elmer, Massachusetts, USA). The solid samples were digested oxygen, the cell growth rate was much slower. Eq. (3) applied under

283
N. Bansal et al. Ecotoxicology and Environmental Safety 172 (2019) 281–289

non–inhibitory conditions, but after a certain value Cr was exceeded, Table 1


the reactor was affected by the exponential function of inhibition Ferrochrome slag analysis results by ICP-AES.
coefficient K as shown in the modified Eq. (3) (Eq. (4)): Element Slag Dust Leachate
(mg/g) (mg/g) (mg/g)
dC kmc C Co C
= (1 Xo
dt K Cr / Co) (K
c C) Rc (4) Ti 2.9 2.7 n/a
Al 44.1 48.4 n/a
where, K = competitive inhibition factor (dimensionless), Cr = the Cr Fe 75.9 79.9 n/a
(VI) inhibition threshold concentration (ML−3). In this case, the in- Mn 2.9 3.8 0.3
Mg 74.9 122 97
hibition kinetic is still non-competitive until the concentration exceeds
Ca 20 19.5 162
a certain threshold concentration (Cr). Above the threshold concentra- Cr 97.2 93.4 n/a
tion Cr, the reaction rate responded to the initial Cr(VI) concentration Ni 2.6 1.3 n/a
(Co) as shown in Eq. (4). Zr 18.9 0.9 n/a
Ba 16.0 29.3 n/a
Cu 0.2 0.4 n/a
4. Results and discussion Zn 9.9 22.9 0.2
Pb 45.9 48.3 n/a
4.1. Bacterial culture identification
*n/a: below detection limit.
Culture purification and species identification was conducted using
the 16S rRNA genotype fingerprinting at independent labs in Pretoria
researchers also reported Pseudomonas aeruginosa is a promising chro-
(South Africa). The sequenced genomic DNA was compared to se-
mium reducing bacteria at high chromium concentration (Ozturk et al.,
quences in the NCBI (National Centre for Biotechnology Information)
2012; Xu et al., 2009). Klebsiella oxytoca also showed the high tolerance
gene bank using the BLAST search. The results from the genotype
and reduction, potential to reduce Cr(VI) under aerobic (Garavaglia
characterization of natural consortium formed under aerobic and
et al., 2010) as well as anaerobic conditions (Baldi et al., 2001). Alca-
anaerobic conditions showed the predominance of three aerobic
ligenes faecalis isolated from chromium containing waste showed the Cr
Serratia marescens (X1,X2), Pseudomonas aeruginosa (X3), Alcaligenes
(VI) reduction capability under wide range of pH and temperature
faecalis (X5) and one anaerobic Klebsiella oxytoca (X4), gram-negative
(Biswas et al., 2015; Shakoori et al., 2010). Serratia marescens has also
bacteria (Fig. 1).
been reported to reduce Cr(VI) anaerobic conditions (De Bruijn and
It was assumed here that the gram-negative, bacteria played a cri-
Mondaca, 2000).
tical role in the Cr(VI) reducing activity of the sludge culture. Earlier

Fig. 1. Phylogenetic tree of bacteria identified in the BLAST search using the 16S rRNA sequences.

284
N. Bansal et al. Ecotoxicology and Environmental Safety 172 (2019) 281–289

Fig. 2. Evaluation of aerobic Cr(VI) reduction at initial Cr(VI) ranging between


20 and 400 mg/L in the absence of Fe species (a); Aerobic reduction of 50 mg/L
Cr(VI) reduction in the presence of Fe(II) ranging between 5 and 100 mg/L (b);
Aerobic reduction of 50 mg/L Cr(VI) in the presence of Fe(III) ranging between
Fig. 3. Anaerobic reduction of Cr(VI) in initial concentrations of Cr(VI) ranging
5 and 100 mg/L (c).
between 20 and 100 mg/L in the absence of Fe species (a); Anaerobic reduction
of 50 mg/L Cr(VI) in the presence of Fe(II) ranging from 5 to 100 mg/L (b);
4.2. Mass balances Anaerobic reduction of 50 mg/L Cr(VI) in the presence of Fe(III) ranging from 5
to 100 mg/L (c).
Total chromium was measured in sacrificial runs of 48 h to check
the mass balance between added Cr(VI) and total chromium after 48 h. 4.3. Characterization of metals in tailing dump slag, dust, leachate
The results showed total recovery of all chromium species as total
chromium after 48 h with an error range of ± 5%. For cultures where The elemental analysis of metal in slag, dust and lechate samples
Cr(VI) was completely removed but almost 100% total chromium was was conducted to determine the typical range of chromium and iron
indicated, this confirmed that chromium was still present in the system concentration in an actual slag dump to serve as guide for concentra-
as Cr(III). Losses within the first hour of incubation were constant re- tions to be used in the experiments. The values determined from an
gardless of the initial concentration used and were attributed to ad- actual slag dump samples for chromium and iron are shown in Table 1.
sorption to cell surfaces. The metals in Table 1 were detected as oxides of the parent metals.

285
N. Bansal et al. Ecotoxicology and Environmental Safety 172 (2019) 281–289

Table 2
Optimum kinetic parameters using the non-competitive inhibition model (Eq. (3)) for aerobic conditions and using the competitive inhibition model (Eq. (4)) under
anaerobic conditions.
Conditions Concentrations (mg/L) Aerobic Anaerobic

kmc 1/h Kc mg/L Rc mg/mg Ki mg/L Xo mg/L kmc 1/h Kc mg/L Rc mg/mg K Cr mg/L Xo mg/L

a 20 0.8 700 0.022 7.5 5780 0.12 180 0.01 10 2.5 2240
50 0.8 700 0.022 7.5 5940 0.12 180 0.01 10 2.5 2240
100 0.8 700 0.022 7.5 6220 0.12 180 0.01 10 2.5 2160
200 0.8 700 0.022 7.5 5720 0.12 180 0.01 10 2.5 2195
b 5 0.85 1200 0.02 6.8 5000 0.12 180 0.01 10 2.5 2240
20 0.85 1200 0.02 6.8 5000 0.3 280 0.025 13 20 2360
50 0.85 1200 0.02 6.8 6000 0.1 280 0.028 13 20 2310
75 0.85 1200 0.02 6.8 6000 0.1 280 0.015 13 20 2340
100 0.85 1200 0.02 6.8 6500 0.1 280 0.012 13 20 2280
c 5 0.26 1280 0.008 9.5 4000 0.1 280 0.012 13 20 2260
20 0.26 1280 0.008 9.5 3800 0 400 0 13 20 2360
50 0.26 1280 0.008 9.5 4000 0 400 0 13 20 2360
75 0.26 1280 0.008 9.5 4000 0 400 0 13 20 2360
100 0.26 1280 0.008 9.5 3900 0 400 0 13 20 2360

a: Cr (VI) concentrations, b: Fe(II) concentrations with 50 mg/L chromium, Fe(III) concentrations with 50 mg/L chromium.

Looking at iron and chromium in the Table, it is shown that the prac- showing an inhibitory effect or competition for electrons and Fe(III)
tical experimental range should span the 75 mg/L Fe and 100 mg/L Cr. thermodynamically requires electrons to convert to Fe(II) under redu-
cing conditions (Coetzee et al., 2018). The Cr(VI) reduction capacity
4.4. Mixed culture batch performance beyond the 100 mg Cr(VI)/L threshold was attributed to an inhibiting
effect of Fe(III) in the system or the depletion of carbon source.
4.4.1. Aerobic reduction experiments
Chromium reduction experiments without the presence of iron 4.5. Abiotic reduction
showed a 100% reduction of up to 100 mg/L initial concentration of Cr
(VI) within a period of 24 h. Higher concentrations of Cr(VI) were re- Abiotic reduction of Cr(VI) with heat-killed cultures showed that
duced for a period up to 48 h before reduction slowed down sig- there was no abiotic reduction of Cr(VI) taking place aerobically or
nificantly (Fig. 2a). This can be attributed to the toxicity level of the anaerobically. The results were confirmed using azide killed and cell
chromium bacteria being reached. The relatively low concentration of free controls which also showed negative results.
carbon source used can also indicate that the bacterial cells were de-
pleted of an energy source after about 24 h. 4.6. Batch pH measurements
For aerobic reduction experiments in the presence of Fe(II) a large
initial decline of Cr(VI) was seen after the first couple of seconds of the The pH of the resultant mixture was measured over a period of 24 h.
experiment after Fe(II) was added. This can be attributed to the oxi- It was found that in the reduction experiment with and without iron the
dation-reduction that occurs between Cr(VI) and Fe(II) as both ions are pH between the different batches were similar and remained constant
highly reactive, where Fe(II) act as electron donor. Higher initial Fe(II) (6–7) over the course of the experiment. Earlier, researchers showed, Cr
concentrations proved to have a greater effect on the total reduction. (OH)3 is the predominant species in the pH range 5.5–10.5, this cor-
An initial Cr(VI) concentration of 50 mg/L was fully reduced within six relates with the area where the majority of the biological reaction occur
hours in 50–100 mg/L Fe(II) and within 24 h in lower concentrations (Ball and Nordstrom, 1998; Richard and Bourg, 1991).
(Fig. 2b). Aerobic reduction of Cr(VI) in the presence Fe(III) indicated
that different concentrations of ferric ion did not influence overall Cr 4.7. Kinetic results
(VI) reduction. It is, however observed that the overall reduction was
much slower, which can indicate that the presence of Fe(III) in the 4.7.1. Parameter estimation
system has an inhibiting effect on Cr(VI) reduction (Fig. 2c). Parameter estimation on the kinetic equations was conducted by
using the Computer Program for the Identification and Simulation of
4.4.2. Anaerobic reduction experiments Aquatic Systems AQUASIM 2.01 (Reichert, 1998). The model compar-
It was observed during the anaerobic reduction that 100% of Cr(VI) ison with experimental data was conducted by minimizing the sum of
at an initial concentration of 20 mg/L was achieved within 48 h. Higher squares error between the measured data points and the calculated data
initial concentrations showed a fast initial drop in Cr(VI) reduction, points using the Secant function in AQUASIM.
before the reduction slowed dramatically (Fig. 3a). This can be attrib-
uted to the toxicity levels of the Cr(VI) reducing bacteria being reached 4.7.2. Aerobic batch kinetics
(Chirwa and Wang, 2000; Kourtev et al., 2006) or depletion of carbon After fitting the model to the anaerobic and aerobic batch data,
source. optimum kinetic parameters were obtained for different Fe(II) and Fe
As with the anaerobic reduction of Cr(VI) a fast initial drop in Cr(VI) (III) concentrations as shown in Table 2. The results indicate that the
was observed in the presence of Fe(II). It was observed that higher maximum specific reaction rate coefficient was consistent over the
initial concentrations of Fe(II) allowed faster overall reduction of Cr(VI) range of Cr(VI) concentrations. Parameter estimation for aerobic study
to take place with full reduction of 50 mg/L Cr(VI) being achieved in described the suggested model fits the data for the current consortium.
24 h in the presence of 100 mg/L Fe(II) (Fig. 3b). Under anaerobic This model can then be used to predict behaviour and reduction at
conditions, Cr(VI) reduction rate increased from the lowest Fe(III) ex- other concentrations also. Fig. 4(a,b,c) show modelled plot fits well
posure of 10 mg Fe(III)/L up to the value of 75 mg Fe(III)/L (Fig. 3c). with the experimental value for the aerobic batch experiments. The
After 75 mg/L Fe(III), Cr(VI) reduction rate stabilised. Earlier studies results showed that all constants are constant for various concentration
indicated a decrease in Cr(VI) reduction rate after 100 mg/L Fe(III) of Cr(VI), Fe(II) and Fe(III) for respective aerobic study (Table 2).

286
N. Bansal et al. Ecotoxicology and Environmental Safety 172 (2019) 281–289

Fig. 5. Anaerobic reduction at different concentrations in the absence of Fe


species (line plots indicate model simulations) (a); Anaerobic reduction at fixed
Cr(VI) concentration (50 mg/L) and varying Fe(II) concentration (modelled
plots) (b); Anaerobic reduction of Cr(VI) at different concentrations of Fe(III)
(modelled plots) (c).
Fig. 4. Aerobic Cr(VI) reduction at different concentrations in the absence of Fe
species (line plots indicate model simulations) (a); Aerobic Cr(VI) reduction at a parameters between the different Fe(II) concentrations were similar,
fixed Cr(VI) concentration (50 mg/L) and varying Fe(II) concentrations (mod- with a good fit of the experimental data. This indicates that the re-
elled plots) (b); Aerobic Cr(VI) reduction at a fixed Cr(VI) concentration duction rate is governed by biological reaction only, and that the che-
(50 mg/L) and varying Fe(III) (modelled plots) (c). mical effects are negligible. The modelled maximum specific rate
coefficient, kmc, was significantly lower for the aerobic reduction of Cr
Schlautman and Han (2001) showed that the oxidation–reduction (VI) in the presence of Fe(III). This may indicate that the presence of Fe
reaction between Cr(VI) and Fe(II) can be a fast reaction occurring (III) in the system has an inhibition effect on the reduction of Cr(VI).
within a matter of seconds. A large initial drop in Cr(VI) was observed This lower rate of reduction can also be attributed to a much lower
after Fe(II) was added to the system. Analytical methods to determine initial biomass concentration found in the system. The difference in
Fe(II) concentration showed that there was no Fe(II) available in the initial Cr(VI) concentrations are due to the rapid reaction with Fe(II)
system as fast chemical reaction between Cr(VI) and Fe(II) oxidises present in the system.
approximately all Fe(II) into Fe(III). This is expected as the Cr(VI) is in
excess and the Fe(II), the limiting reagent in the chemical reduction. It
was thus predicted that the initial drop in Cr(VI) was due to the che- 4.7.3. Anaerobic batch kinetics
mical reaction and the remaining Cr(VI) was reduced biologically. The cell inactivation model showed that the pseudo mixed-order
The kinetic results of Fe(II) batches indicated that the maximum non-competitive reaction kinetic model fitted the experimental anae-
specific rate coefficient, kmc was similar to the pure batch. The robic data well. Like aerobic parameter estimation, anaerobic kinetic
results also showed the well fitted graphs with the suggest model

287
N. Bansal et al. Ecotoxicology and Environmental Safety 172 (2019) 281–289

Fig. 5(a,b,c). The modelled maximum specific rate coefficient, kmc, for Congress Sustainable Future, Helsinki, Finland.
the anaerobic reduction experiment was much lower than the aerobic Chirwa, E.N., Wang, Y.T., 2000. Simultaneous chromium(VI) reduction and phenol de-
gradation in an anaerobic consortium of bacteria. Water Res. 34 (8), 2376–2384.
reaction (Table 2). This is expected due to the lower initial biomass https://doi.org/10.1016/S0043-1354(99)00363-2.
concentration that was achieved after growth. Coenye, T., Falsen, E., Vancanneyt, M., Hoste, B., Govan, J.R., Kersters, K., Vandamme, P.,
Similar reaction rate parameter data was observed with anaerobic 1999. Classification of Alcaligenes faecalis-like isolates from the environment and
human clinical samples as Ralstonia gilardii sp. nov. Int. J. Syst. Evolut. Microbiol. 49
Cr(VI) reduction in the presence of Fe(II) which indicates that the (2), 405–413.
chemical effects of Fe(II) only plays a role in the initial reduction of Cr Coetzee, J.J., Bansal, N., Chirwa, E.M.N., 2018. Chromium in environment, its toxic effect
(VI). Anaerobic reduction in the presence of Fe(III) show a decrease in from chromite-mining and ferrochrome industries, and its possible bioremediation.
Expo. Health. https://doi.org/10.1007/s12403-018-0284-z.
the maximum specific rate coefficient, kmc. This indicates that the De Bruijn, J.P.F., Mondaca, M.A., 2000. Chromate reduction by Serratia marcescens
presence of Fe(III) in the system may have an inhibition effect on the immobilized on activated carbon. Toxicol. Environ. Chem. 76 (3–4), 125–135.
reduction performance of the anaerobic Cr(VI) reducing bacteria. The https://doi.org/10.1080/02772240009358923.
Dhal, B., Thatoi, H.N., Das, N.N., Pandey, B.D., 2013. Chemical and microbial remedia-
modelled data between the different Fe(III) concentrations are similar
tion of hexavalent chromium from contaminated soil and mining/metallurgical solid
and fit the experimental data well, which indicates that the reaction is waste: a review. J. Hazard. Mater. 250–251, 272–291. https://doi.org/10.1016/j.
governed by bacterial activity only. jhazmat.2013.01.048.
Dogan, N.M., Kantar, C., Gulcan, S., Dodge, C.J., Yilmaz, B.C., Mazmanci, M.A., 2011.
Chromium(VI) bioremoval by Pseudomonas bacteria: role of microbial exudates for
5. Conclusion natural attenuation and biotreatment of Cr(VI) contamination. Environ. Sci. Technol.
45 (6), 2278–2285. https://doi.org/10.1021/es102095t.
In this study, aerobic and anaerobic microbial cultures dominated Eaton, A., Clesceri, L.S., Rice, E.W., Greenberg, A.E..and Franson, M., 2005. APHA:
Standard Methods for the Examination of Water and Wastewater. USA.
by Pseudomonas aeruginosa, Serratia marescens, Alcaligenes faecal, Erdem, M., Altundoğan, H.S., Turan, M.D., Tümen, F., 2005. Hexavalent chromium re-
Klebsiella oxytoca were found to withstand high level of Fe(II) and Fe moval by ferrochromium slag. J. Hazard. Mater. 126 (1–3), 176–182. https://doi.
(III), and were capable of reducing Cr(VI) at concentrations, up to org/10.1016/j.jhazmat.2005.06.017.
FDA USGS, 2004. NIST: Multi-agency radiological laboratory analytical protocols manual
50 mg/L Cr(VI) in 6 h for aerobic and 24 h in anaerobic conditions. (MARLAP), FDA, USGS, NUREG-1576. Chapter 13, EPA 402-B-04-001C, NTIS
50 mg/L Cr(VI) was fully reduced in the presence of 50–100 mg/L Fe(II) PB2004-105421, Alexandria, VA, USA, 2.
for aerobic condition and in presence of 100 mg/L Fe(II) for anaerobic Federal Register, 2004. Occupational Exposure to Hexavalent Chromium: Occupational
Safety and Health Administration (OSHA). Department of Labor, U.S.A.
condition. Presence of Fe(III) did not influence the reduction of Cr(VI) Felsenstein, J., 1985. Confidence limits on phylogenies: an approach using the bootstrap.
in both aerobic and anaerobic conditions. The results demonstrate that Evolution 39 (4), 783–791.
introducing such cultures in a Cr(VI) contaminated environment such Fiúza, A., Silva, A., Carvalho, G., de la Fuente, A.V., Delerue-Matos, C., 2010.
Heterogeneous kinetics of the reduction of chromium (VI) by elemental iron. J.
as the ferrochrome mining tailing dams could help stabilize the chro-
Hazard. Mater. 175 (1), 1042–1047. https://doi.org/10.1016/j.jhazmat.2009.10.
mium species in the tailing dam. This could be achieved by the re- 116.
ductive potential of ferrous ion which serves as the electron donor in Francisco, R., Alpoim, M., Morais, P., 2002. Diversity of chromium‐resistant and‐reducing
the Fe/Cr redox couple. In South Africa, total chromium in the range of bacteria in a chromium‐contaminated activated sludge. J. Appl. Microbiol. 92 (5),
837–843.
18–1104 mg/L has been observed in mining dump site (Olobatoke and Garavaglia, L., Cerdeira, S.B., Vullo, D.L., 2010. Chromium (VI) biotransformation by β-
Mathuthu, 2016) of which 30–100 mg/L could be Cr(VI). In the aerobic and γ-Proteobacteria from natural polluted environments: a combined biological and
system 100% immobilization of Cr(VI) was observed within 24 h of chemical treatment for industrial wastes. J. Hazard. Mater. 175 (1–3), 104–110.
https://doi.org/10.1016/j.jhazmat.2009.09.134.
100 mg/L Cr(VI), while complete reduction of 20 mg/L Cr(VI) was Guertin, J., Jacobs, J.A., Avakian, C.P., 2016. Chromium(VI) handbook.
achieved within 48 h in the anaerobic system. In the current study, a Hawley, E.L., Deeb, R.A., Kavanaugh, M.C., Jacobs, J.A., 2004. Treatment Technologies
minimum carbon source concentration of 1 g/L was used. In real site for Chromium (VI). CRC Presss, Florida, U. S. A.
Jobby, R., Jha, P., Yadav, A.K., Desai, N., 2018. Biosorption and biotransformation of
applications, only a minimum of carbon sources would be needed to hexavalent chromium [Cr(VI)]: a comprehensive review. Chemosphere 207,
avoid further polluting the environment with the carbon sources which 255–266. https://doi.org/10.1016/j.chemosphere.2018.05.050.
could create a high oxygen demand in receiving water bodies. Kourtev, P.S., Nakatsu, C.H., Konopka, A., 2006. Responses of the anaerobic bacterial
community to addition of organic C in chromium(VI)- and iron(III)-amended mi-
The study conducted under anaerobic conditions is most relevant to
crocosms. Appl. Environ. Microbiol. 72 (1), 628–637. https://doi.org/10.1128/AEM.
the target application in mine tailing dumps site as the environment 72.1.628-637.2006.
would be mostly cut off from contact with atmospheric water, even Liu, L., Li, W., Song, W., Guo, M., 2018. Remediation techniques for heavy metal-con-
taminated soils: principles and applicability. Sci. Total Environ. 633, 206–219.
though operation under anaerobic conditions showed a lower Cr(VI)
https://doi.org/10.1016/j.scitotenv.2018.03.161.
reduction rate coefficient and a higher susceptibility to Cr(VI) toxicity. Madhavi, V., Reddy, A.V.B., Reddy, K.G., Madhavi, G., Prasad, T.N.K.V., 2013. An
In spite of the slower removal rate under anaerobic conditions, the overview on research trends in remediation of chromium. Res. J. Recent Sci. 2 (1),
principle removal mechanisms appear to be the same as in the aerobic 71–83.
Mahesvari, V., Balasabramanian, N., 1996. Spectrophotometric determination of chro-
systems. The trends of removal and the limited removal capacity as Cr mium based on ion-pair formation. Chem. Anal. 41, 569–576.
(VI) reduction is evident in both systems. Molokwane, P.E., 2010. Simulation of In Situ Bioremediation of Cr (VI) in Groundwater
Aquifer Environments Using A Microbial Culture Barrier. University of Pretoria,
South Africa(Retrieved from: 〈http://hdl.handle.net/2263/28191〉).
References Olobatoke, R.Y., Mathuthu, M., 2016. Heavy metal concentration in soil in the tailing
dam vicinity of an old gold mine in Johannesburg, South Africa. Can. J. Soil Sci. 96
Baldi, F., Minacci, A., Pepi, M., Scozzafava, A., 2001. Gel sequestration of heavy metals by (3), 299–304. https://doi.org/10.1139/cjss-2015-0081.
Klebsiella oxytoca isolated from iron mat. FEMS Microbiol. Ecol. 36 (2–3), 169–174. Ozturk, S., Kaya, T., Aslim, B., Tan, S., 2012. Removal and reduction of chromium by
https://doi.org/10.1111/j.1574-6941.2001.tb00837.x. Pseudomonas spp. and their correlation to rhamnolipid production. J. Hazard. Mater.
Ball, J.W., Nordstrom, D.K., 1998. Critical evaluation and selection of standard state 231, 64–69. https://doi.org/10.1016/j.jhazmat.2012.06.038.
thermodynamic properties for chromium metal and its aqueous ions, hydrolysis Panda, C.R., Mishra, K.K., Panda, K.C., Nayak, B.D., Nayak, B.B., 2013. environmental
species, oxides, and hydroxides. J. Chem. Eng. Data 43 (6), 895–918. https://doi.org/ and technical assessment of ferrochrome slag as concrete aggregate material. Constr.
10.1021/je980080a. Build. Mater. 49 (Supplement C), 262–271. https://doi.org/10.1016/j.conbuildmat.
Bharagava, R.N., Mishra, S., 2018. Hexavalent chromium reduction potential of cellulo- 2013.08.002.
simicrobium sp. isolated from common effluent treatment plant of tannery industries. Reichert, P., 1998. AQUASIM. Swiss Federal Institute for Environmental Science and
Ecotoxicol. Environ. Saf. 147 (Supplement C), 102–109. https://doi.org/10.1016/j. Technology (EAWAG) CH - 8600 Dubendorf Switzerland, ISBN 3-906484-17-3.
ecoenv.2017.08.040. Richard, F.C., Bourg, A.C.M., 1991. Aqueous geochemistry of chromium: a review. Water
Biswas, G., Das, R., Kazy, S., 2015. Chromium bio-remediation by Alcaligenes faecalis Res. 25 (7), 807–816. https://doi.org/10.1016/0043-1354(91)90160-R.
strain P-2 newly isolated from tannery effluent. J. Environ. Res. Dev. 9, 840–848. Satarupa, D., Paul, A.K., 2013. Hexavalent chromium reduction by aerobic heterotrophic
Burman, J.O., Ponter, C., Bostrom, K., 1978. Metaborate digestion procedure for in- bacteria indigenous to chromite mine overburden. Braz. J. Microbiol. 44, 307–315.
ductively coupled plasma-optical emission spectrometry. Anal. Chem. 50 (4), Schlautman, M.A., Han, I., 2001. Effects of pH and dissolved oxygen on the reduction of
679–680. https://doi.org/10.1021/ac50026a043. hexavalent chromium by dissolved ferrous iron in poorly buffered aqueous systems.
Chirasha, J., Shoko, N.R., 2010. Zimbabwe alloys ferro chromium production: from cradle Water Res. 35 (6), 1534–1546. https://doi.org/10.1016/S0043-1354(00)00408-5.
to grave sustainably paper presented at the Twelfth International Ferroalloys Shakoori, F.R., Tabassum, S., Rehman, A., Shakoori, A., 2010. Isolation and

288
N. Bansal et al. Ecotoxicology and Environmental Safety 172 (2019) 281–289

characterization of Cr6+ reducing bacteria and their potential use in bioremediation https://doi.org/10.1093/molbev/mst197.
of chromium containing wastewater. Pak. J. Zool. 42 (6), 651–658. van Staden, Y., Beukes, J.P., van Zyl, P.G., du Toit, J.S., Dawson, N.F., 2014.
Shen, H., Wang, Y.T., 1994. Modeling hexavalent chromium reduction in Escherichia coli Characterisation and liberation of chromium from fine ferrochrome waste materials.
33456. Biotechnol. Bioeng. 43 (4), 293–300. Miner. Eng. 56, 112–120. https://doi.org/10.1016/j.mineng.2013.11.004.
Stasinakis, A.S., Thomaidis, N.S., Mamais, D., Karivali, M., Lekkas, T.D., 2003. Chromium Xu, W.-H., Liu, Y.-G., Zeng, G.-M., Li, X., Song, H.X., Peng, Q.Q., 2009. Characterization of
species behaviour in the activated sludge process. Chemosphere 52 (6), 1059–1067. Cr(VI) resistance and reduction by Pseudomonas aeruginosa. Trans. Nonferrous Met.
https://doi.org/10.1016/S0045-6535(03)00309-6. Soc. China 19 (5), 1336–1341. https://doi.org/10.1016/S1003-6326(08)60446-X.
Tamura, K., Stecher, G., Peterson, D., Filipski, A., Kumar, S., 2013. MEGA6: molecular Zhitkovich, A., 2011. Chromium in drinking water: sources, metabolism, and cancer risks.
evolutionary genetics analysis version 6.0. Mol. Biol. Evol. 30 (12), 2725–2729. Chem. Res. Toxicol. 24 (10), 1617–1629. https://doi.org/10.1021/tx200251t.

289

You might also like