43
REDUCTION OF Cr(VI) 35 CHAPTER 3: REDUCTION OF Cr(VI) 3.1 Introduction: The reduction of highly toxic and mobile Cr(VI) to less toxic, less mobile Cr(III) is likely to be useful for remediation of Cr(VI) contaminated soils and waters. Owing to the high toxicity of Cr(VI) and disadvantages associated with conventional chemical methods of Cr(VI) removal, focus of environmentalists has been shifted towards development of biological methods for detoxification of Cr(VI) (Timmes and Pieper, 1999). Microbial mediated reduction of Cr(VI) to Cr(III) was first reported by Romanenko and Korenkoc (1977) and there after number of Cr(VI) reducing bacterial strains have been isolated and studied (Ackerley et al., 2004; Camargo et al., 2004; Elongovan et al., 2008). The earliest reports associated with isolation and characterization of Cr(VI) reducing bacteria were of anaerobic bacterium, Enterobacter cloacae isolated from industrial wastewater. E. cloacae was found to use chromate/dichromate as terminal electron acceptor during its anaerobic growth (Wang et al., 1989; 1990). Anaerobic reduction of Cr(VI) has been subsequently described in several organisms such as Shewanella oneidensis, Escherichia coli and several sulphate reducing bacteria like Desulfovibrio vulgaris and Desulfovibrio fructosovorans (Lowe et al., 2003; Mclean and Beveridge, 2001; Shen and Wang, 1993; Michel et al., 2001). In addition to anaerobic Cr(VI) reducing microorganisms, a number of bacteria belonging to genera Bacillus, Ochrobactrum, Pseudomonas, Escherichia and Parracoccus have been reported for their Cr(VI) reduction ability under aerobic condition (Sultan and Hasnain, 2007; Hussein et al., 2005; Garbisu et al., 1998). The Cr(VI) reducing micro-organisms can be categorized into aerobic or anaerobic depending on their ability to reduce Cr(VI) under aerobic or anaerobic conditions. Anaerobic bacteria mainly use Cr(VI) as terminal electron acceptor or reduce Cr(VI) in the periplasmic space by hydrogenase or reduced cytochrome (Figure 3.1). For example, in D. vulgaris and D. fructosovorans involvement of Cyt C 3 and Ni-Fe dehydrogenase in Cr(VI) reduction has been reported by Chardin et al. (2003) and Michel et al. (2001). Several oxidoreductases with unrelated metabolic functions have been reported to catalyse Cr(VI) reduction in bacteria. Examples of such enzymes include, quinone reductases, iron reductases, iron reductases, nitro-reductases, flavin reductases and several NADH/NAD(P)H dependent reductases of unknown function (Clark, 1994; Gonzalez et al., 2005; Gonzalez et al., 2003; Kwak et al., 2003; Mazoch et al., 2004; Suziki et al., 1992). Chromate reductases from P. putida and E. coli have been characterized in detail (Ackerley et al., 2004).

CHAPTER 3: REDUCTION OF Cr(VI) - Shodhgangashodhganga.inflibnet.ac.in/bitstream/10603/7231/8/08_chapter 3.pdf · CHAPTER 3: REDUCTION OF Cr ... Table 3.1 illustrates the oxidation-reduction

Embed Size (px)

Citation preview

REDUCTION OF Cr(VI)

35

CHAPTER 3: REDUCTION OF Cr(VI)

3.1 Introduction:

The reduction of highly toxic and mobile Cr(VI) to less toxic, less mobile Cr(III) is likely to be

useful for remediation of Cr(VI) contaminated soils and waters. Owing to the high toxicity of

Cr(VI) and disadvantages associated with conventional chemical methods of Cr(VI) removal, focus

of environmentalists has been shifted towards development of biological methods for detoxification

of Cr(VI) (Timmes and Pieper, 1999). Microbial mediated reduction of Cr(VI) to Cr(III) was first

reported by Romanenko and Korenkoc (1977) and there after number of Cr(VI) reducing bacterial

strains have been isolated and studied (Ackerley et al., 2004; Camargo et al., 2004; Elongovan et

al., 2008).

The earliest reports associated with isolation and characterization of Cr(VI) reducing bacteria were

of anaerobic bacterium, Enterobacter cloacae isolated from industrial wastewater. E. cloacae was

found to use chromate/dichromate as terminal electron acceptor during its anaerobic growth (Wang

et al., 1989; 1990). Anaerobic reduction of Cr(VI) has been subsequently described in several

organisms such as Shewanella oneidensis, Escherichia coli and several sulphate reducing bacteria

like Desulfovibrio vulgaris and Desulfovibrio fructosovorans (Lowe et al., 2003; Mclean and

Beveridge, 2001; Shen and Wang, 1993; Michel et al., 2001).

In addition to anaerobic Cr(VI) reducing microorganisms, a number of bacteria belonging to genera

Bacillus, Ochrobactrum, Pseudomonas, Escherichia and Parracoccus have been reported for their

Cr(VI) reduction ability under aerobic condition (Sultan and Hasnain, 2007; Hussein et al., 2005;

Garbisu et al., 1998). The Cr(VI) reducing micro-organisms can be categorized into aerobic or

anaerobic depending on their ability to reduce Cr(VI) under aerobic or anaerobic conditions.

Anaerobic bacteria mainly use Cr(VI) as terminal electron acceptor or reduce Cr(VI) in the

periplasmic space by hydrogenase or reduced cytochrome (Figure 3.1). For example, in D.

vulgaris and D. fructosovorans involvement of Cyt C3 and Ni-Fe dehydrogenase in Cr(VI)

reduction has been reported by Chardin et al. (2003) and Michel et al. (2001).

Several oxidoreductases with unrelated metabolic functions have been reported to catalyse Cr(VI)

reduction in bacteria. Examples of such enzymes include, quinone reductases, iron reductases, iron

reductases, nitro-reductases, flavin reductases and several NADH/NAD(P)H dependent reductases

of unknown function (Clark, 1994; Gonzalez et al., 2005; Gonzalez et al., 2003; Kwak et al., 2003;

Mazoch et al., 2004; Suziki et al., 1992). Chromate reductases from P. putida and E. coli have been

characterized in detail (Ackerley et al., 2004).

REDUCTION OF Cr(VI)

36

On the basis of sequence homology; the aerobic chromate reductases characterized till date has been

categorized into two main classes i.e

Class I chromate reductases

Class II chromate reductases

Figure 3.1 Mechanism of enzymatic Cr(VI) reduction under aerobic and anaerobic conditions

(Cheng and Gu, 2007). (MR: membrane chromate reductase; SR: Soluble chromate reductase).

Class I chromate reductase: Two of the most commonly studied class I enzyme include ChrR

(chromate reductase) of P. putida and YieF of E.coli (Accession number and 6058522, respectively)

(Ackerley et al., 2004b). Both these enzymes are dimeric flavoproteins of mass 50 KDa containing

signature sequences of the NADH_dh2 protein family (LFVTPEYNXXXXXXLKNAIDXXS)

representating putative flavin binding quinone reductases. The finding that ChrR and YieF are

members of flavin binding quinone reductase is consistent with the proposal made by Ishibashi et

al. (1990) that chromate reduction might be a secondary property of oxido-reductases with different

primary roles. With all these similarities, the two enzymes differ from each other in the way the

Cr(VI) is reduced to Cr(III).

Generally NADH_dh2 protein family consists of proteins displaying two electron reduction of

various electrophiles but ChrR is an one electron reducer enzyme which leads to the formation of

Cr(V) as intermediate. In contrast, enzyme YieF catalyses the direct reduction of Cr(VI) to Cr(III)

REDUCTION OF Cr(VI)

37

through transfer of four electrons. Out of four electrons, three electrons are consumed in reducing

Cr(VI) to Cr(III) whereas one electron is transferred to molecular oxygen, subsequently generating

H2O2. Hence YieF provide E. coli an effective mechanism of Cr(VI) reduction by forming lower

quantity of reactive oxygen species (Ackerley et al., 2004b).

Class II chromate reductase: The class II chromate reductases posses nitro-reductase activity and

bear no sequence homology to class I enzymes. The only member of the class II family include,

NfsA protein of E. coli (Accession number: 730007). NfsA is an obligate two electron reducer of

nitro compounds and quinones. However, the Cr(VI) reduction by NfsA involves a one electron

transfer and so more than 25% of the available electrons are consumed in ROS production and

Cr(V) is formed (Ackerley et al. 2004a).

Impact and application of redox mediators in bacterial biotransformation of contaminants

Redox mediators also referred as electron shuttle are organic molecules that can reversibly be

oxidized and reduced therefore conferring the capacity to serve as an electron carrier in multiple

redox reactions (Zee and Cervantes, 2009). Redox mediators are capable of transferring electrons in

redox reactions between a wide variety of both organic and inorganic compounds (Burgos et al.,

2003; Liu et al., 2009; Lovley et al., 1998; Maithreepala and Doong, 2009; Wang et al., 2009). The

most commonly studied redox mediators which support reductive biotransformation of various

contaminants include humic acid and their quinoid analogues (Lovley et al., 1998; Santos et al.,

2004a,b

).

Figure 3.2 shows some of the redox mediators reported for their role in reductive biotransformation

of various organics and inorganic contaminants (e.g azo dye, hydrocarbons and several metal ions)

(Bond and Lovley 2002; Ling et al., 2009; Liu et al., 2009).

Among all these redox mediators, flavin based compounds such as flavin adenide mononucleotide

(FMN) and riboflavin as well as quinone based compounds such as anthra quinone sulphonate

(AQS), anthra quinone disulphonate (AQDS) and lawsone have been extensively investigated for

their application in degradation of environmental pollutants (Zee and Cervantes, 2009).

The term quinone collectively refers to organic structures that occur in three oxidation states and are

linked by one electron redox reactions. The three oxidation states are:

i. Hydroquinone with three protonation levels is the fully reduced form;

ii. The p-semiquinone radical with two protonation levels is an intermediate oxidation state;

iii. p-benzoquinone with one protonation level is the fully oxidized form.

REDUCTION OF Cr(VI)

38

Figure 3.2 Structure of main redox mediators (Zee and Cervantes, 2009).

The exact mechanism behind the electron shuttling ability of various quinoid redox mediators is

still a subject of investigation however, the studies carried out to demonstrate the role of redox

mediators in reductive biotransformation of azo dyes and hydrocarbon have proposed a two-step

mechanism of redox mediator mediated biotransformation processes (Burgos et al., 2003; Santos et

al., 2004a). The first step involves the reduction of redox active groups by an appropriate electron

donor while second step involves subsequent oxidation of hydro-quinone/semi-quinone coupled to

reduction of organic/inorganic contaminants. The first step i.e, reduction of redox mediator in most

studies has been shown to be enzyme catalyzed while second step has been proposed to be a purely

non-enzymatic chemical reaction. Hence, the overall reaction rate depends primarily on the

enzyme-catalysed transfer of electrons to redox mediator rather than on the ability of redox

mediators to flux the electrons chemically to contaminants.

An electron shuttling compound will only be effective as a redox mediator for reduction of any

recalcitrant compound if it lowers the reactions activation energy (Zee and Cervantes, 2009). It

means that, the oxidation-reduction potential (ORP) of any effective redox mediator should ideally

be intermediate to those of the two half reactions; (i) the reduction of the recalcitrant

organic/inorganic contaminants (for e.g, azo dye) and (ii) the oxidation of a primary electron donor

REDUCTION OF Cr(VI)

39

(NADH). Other wise the redox mediator will be ineffective or less effective in mediating the

reductive transformation of target pollutant.

Table 3.1 illustrates the oxidation-reduction potential (ORP) of several redox mediators. Although

the studies revealing the impact of redox mediators on reductive biotransformation of Cr(VI) are

rare, studies with other contaminants reveal that quinone compound are preferred redox mediators

owing to their high abundance in humus, the least expensive and easily available organic source in

the biosphere (Wang et al., 2009).

Table 3.1 Oxidation Reduction Potential (ORP) of redox mediators (Zee and Cerventes, 2009).

Redox Mediators ORP (mv)

Anthra Quinone Sulphonate (AQS) -218

Anthra Quinone Disulphonate (AQDS) -184

Lawsone -134

Benzyl viologen -360

Methyl viologen -440

Ethyl viologen -480

Flavin Adenine Dinucleotide (FAD) -219

1,4 Benzaquinone +263

Menadione -19

Cyanocobalamin -530

The present chapter describes the studies on Cr(VI) reduction by two bacterial strains

(Ochrobactrum intermedium and Bacillus sp. BT1) isolated from chromium contaminated soil

samples.

The role of AQS on Cr(VI) reduction by these two bacterial isolates and association of AQS-

mediated reduction with activities of intracellular quinone and chromate reductases has also been

investigated. Finally, the suitability of Bacillus sp. BT1 for reducing Cr(VI) in soil microcosms has

been demonstrated.

REDUCTION OF Cr(VI)

40

3.2 Materials and Methods

3.2.1 Chemicals

Luria Bertani (LB) broth and redox mediators (namely, Anthraquinone sulphonate, Ethyl viologen,

Benzyl viologen and Methyl viologen) were purchased from Hi-Media Chemicals, India. Diphenyl

carbazide and potassium dichromate (K2Cr2O7) were procured from Qualigens, India.

3.2.2 Cr(VI) reduction by bacterial cultures

The 250 mL Erlenmeyer flasks containing 100 mL Luria Bertani broth amended with Cr(VI) (100

to 500 mg/L) were inoculated with overnight grown cells of bacterial culture (1.0 OD600 nm).

Uninoculated controls were used to determine abiotic Cr(VI) reduction. The inoculated cultures

along with un-inoculated controls were incubated and 1.0 mL sample was withdrawn at regular time

intervals to monitor growth and Cr(VI) reduction.

3.2.3 Effect of physico-chemical conditions on Cr(VI) reduction efficiency of bacterial isolate

Effect of different parameters including Cr(VI) concentration (100 to 500 mg/L), temperature (25 to

40 ºC), medium pH (5.5 to 9.0) and metal ions (Cu2+

, Co2+

, As2+

, Ni2+

) at concentration of 50 mg/L

on Cr(VI) reduction by bacterial cultures was investigated. Experiments were performed in 250 mL

Erlenmeyer flasks containing 100 mL LB broth amended with 100 mg Cr(VI)/L or as mentioned in

text. The desired initial pH of medium was adjusted using 1 N NaOH/ HCl. The overnight grown

culture of bacterial isolate was used as inoculum and added to experimental media so as to obtain an

initial cell density of 0.05 (600 nm). In each experiment, the residual Cr(VI) and bacterial growth

was monitored as described above in previous section.

3.2.4 Mediated Cr(VI) reduction by bacterial cultures in batch mode

Bacterial culture was inoculated in 150 mL Cr(VI) (100 to 500 mg/L) amended Luria-Bertani broth

augmented with AQS as redox mediator (1 mM). Cr(VI) reduction was monitored at regular time

interval in 1 mL sample withdrawn aseptically. Control experiments were performed in the same

manner except that no AQS was added to Luria Bertani broth. Un-inoculated media served as

abiotic controls for corresponding experiment.

The Cr(VI) reduction by bacterial culture in presence of different concentration of AQS was studied

wherein, Cr(VI) (100 mg/L) amended Luria Bertani broth was supplemented with AQS in the

concentration range of 0 to 5.0 mM. Cr(VI) reduction was monitored from samples withdrawn at

different time intervals.

REDUCTION OF Cr(VI)

41

3.2.5 Effect of temperature on AQS mediated Cr(VI) reduction

The Cr(VI) amended LB broth supplemented either with or with out 1 mM AQS was inoculated

with overnight grown bacterial culture and incubated at various temperatures (25 to 35 °C). The

samples were withdrawn at regular time interval to monitor residual Cr(VI).

The activation energy of the AQS mediated and non-mediated Cr(VI) reduction by bacterial culture

was calculated by employing Arrhenius equation as follows

ln k = −Ea/R T + lnAo ……………………………………………………………………(3.1)

Where k is the first order rate constant (h-1

), Ea is activation energy and Ao is constant. The value of

Ea can be determined from the Slope (-Ea/R) of ln k versus 1/T plot (Santos et al., 2004).

3.2.6 Repeated cycles of AQS mediated Cr(VI) reduction in fed batch mode

200 mL of Luria Bertani broth supplemented with or without AQS (1mM) was spiked with 100 mg

Cr(VI)/L. The Cr(VI) reduction was monitored immediately after inoculation of bacterial culture.

The samples were withdrawn and Cr(VI) was estimated. On complete reduction of Cr(VI), the

medium was spiked with another aliquot of Cr(VI) and its reduction was monitored. The cycles of

Cr(VI) spiking and reduction were repeated till Cr(VI) reduction ceased.

3.2.7 Preparation of cell free lysate

Bacterial culture was grown in 200 mL Luria Bertani broth for 24 h at 30 C. The cell pellet

obtained upon centrifugation (8,603 g for 15 min) was resuspended in 3 mL of phosphate buffer

(100 mM, pH: 7.0). The resuspended cells were disrupted by sonication for 15 min (Sonics &

Materials, Inc., USA) in cold conditions. The resultant homogenate was centrifuged at 8,603 g for

30 min at 4 ºC to remove cell debris and the clear supernatant was used as cell free lysate for

enzyme assays.

3.2.8 Characterization of quinone and chromate reductase

The influence of pH, temperature and substrate concentration on chromate and quinone reductase

activity of bacterial culture was assessed.

For determination of optimal pH, the quinone and chromate reductase activity was assayed at 30 °C

using 50 mM sodium acetate buffer (pH: 5.0 to 5.5), potassium phosphate buffer (pH: 6.0 to 7.5)

and tris-HCl buffer (7.5 to 8.5).

Optimal temperature was determined wherein the chromate and quinone reductase activity was

assayed at optimum pH with varying temperatures (20 °C to 50 °C) using temperature controlled

thermostatic cuvette holder (ELICO DL 198 biospectrophotometer, Hyderabad, India).

REDUCTION OF Cr(VI)

42

3.2.9 Cr(VI) bioremediation experiment: soil microcosm studies

The soil used for microcosm studies, was collected from agricultural field. Immediately after

sample collection, the soil was sieved to remove gravel and plant residues. 20 g of dry soil was

taken in each 100 mL Erlenmeyer flasks for microcosm experiments. The flasks were plugged with

non-adsorbent cotton and sterilized for three consecutive days by autoclaving (121 ºC, 30 minutes).

The cells of isolate BT1 grown overnight in LB were washed with sterile distilled water in order to

remove any media particles and subsequently resuspended in distilled water to be used as inoculum

for microcosm experiments. The inoculum was added to soil amended with 100 mg Cr(VI)/ kg soil

to achieve a cell density of 104

cfu/g soil. The moisture content of soil in microcosm experiment

was maintained at 50 %.

Abiotic control was kept in order to verify all the non-biological Cr(VI) reduction. Both un-

inoculated and inoculated soil microcosms were vigorously mixed and incubated at 30 ºC for 15

days under aseptic conditions. The water content of the soil was maintained at 50 % level by

weighing the microcosms and adding sterile distilled water to compensate for any weight loss.

3.2.10 Extraction of Cr(VI) and measurement of bacterial growth in soil microcosms.

Each of the soil microcosms was sampled by removing soil sample (0.5 g) with a sterile spatula on

day 3rd

, 6th

, 9th

, 12th

and 15th

. For extraction of Cr(VI), 0.5 g of soil sample was suspended in 5 mL

of distilled water, vortexed thoroughly followed by centrifugation (8,603, 10 minutes, room

temperature). The supernatant was used for Cr(VI) estimation. Likewise, the 0.5 g of soil

resuspended in 5 mL of distilled water was mixed thoroughly and allowed to settle down. Finally,

standard 10-fold serial dilution plate count method was carried out to monitor the total colony

forming units (cfu/g soil) of isolate BT1.

3.2.11 Influence of various factors on Cr(VI) reduction efficiency of isolate BT1 in soil

microcosms.

The efficiency of isolate BT1 was investigated using soils spiked with two different concentrations

of Cr(VI) (viz. 100 and 300 mg/kg soil) and residual Cr(VI) at different time intervals was

monitored as described above.

Effect of malt extract amendment (0.05 to 1.0 %, w/w) on Cr (VI) reduction by isolate BT1 was

evaluated.

3.2.11 Seed germination assay

The seed germination tests were carried out in petriplates by employing two types of test seeds

[mustard (Brassica nigra), variety type: western seed research and mungbean seeds (Vigna

REDUCTION OF Cr(VI)

43

radiata), variety type: K-851-certified]. In order to determine toxicity of Cr(VI), 15 seeds were

placed in each petriplate containing whattman filter paper-1 soaked with pure Cr(VI) solution (50,

100, 200 or 300 mg/L). The plates were incubated at 30 ºC for 48 h and monitored for seed

germination.

In order to determine effectiveness of Cr(VI) bioremediation by isolate BT1 in soil microcosm,

seeds germination experiments were performed using the soil extract obtained from microcosm

experiments. The soil extract was prepared by suspending 10 g of soil from each microcosm system

in 10 mL of distilled water, followed by vortexing and subsequently filtering through whattman

number 1.0 filter paper. The seed germination experiments using microcosm soil extracts were

performed as described above. In all seed germination experiments, seeds sterilized with 0.1 %

(w/v) HgCl2 solution were employed.

3.2.12 Analytical Methods

3.2.12.1 Enzyme assays

The quinone and chromate reductase were assayed as described previously by Puzon et al. (2002).

Briefly, quinone and chromate reductase activity was assayed spectrophotometrically at constant

temperature of 30 ºC following the oxidation of NADH at 340 nm (molar absorption coefficient

6.22 mM-1

cm-1

). The reaction was initiated by addition of cell free lysate to reaction mixture

containing 50 mM phosphate buffer (pH 6.0) and 0.1 mM substrate (Lawsone as quinone reductase

substrate and K2Cr2O7 as chromate reductase substrate).

One unit of enzyme activity was defined as the amount of enzyme required for oxidation of 1 µmole

of NADH per min under standard assay conditions (Santos et al., 2004a).

3.2.12.2 Quantification of growth and Cr(VI)

Bacterial growth was monitored turbidometrically. The cell pellet obtained upon centrifugation of

1.0 mL of culture was resuspended in 1.0 mL distilled water and its absorbance was measured at

600 nm.

The Cr(VI) concentration in the cell-free supernatant was measured using the Diphenylcarbazide

(DPC) reagent (Ishibashi et al., 1990). Briefly, the hexavalent chromium containing samples (in the

range of 1.0 to 10 µg) were acidified by adding 330 µL of 6N sulfuric acid. To this acidified

solution of the hexavalent chromium, DPC was added at concentration of 0.25 % and final volume

was made upto 10 mL. The mixture was incubated for 10 min at room temperature and the colour of

DPC:Cr(VI) complex was measured by reading the absorbance at 540 nm.

REDUCTION OF Cr(VI)

44

3.2.12.3 Protein estimation

The protein concentration of the cell-free extract was estimated using Folin-phenol reagent by

reading absorbance at 720 nm, following the principle of Lowry et al. (1951).

3.3 Results and Discussion

Part A: Reduction of hexavalent chromium by Ochrobactrum intermedium BCR400 isolated

from a chromium-contaminated soil

A hexavalent chromium reducing bacterial strain designated as BCR400 was isolated from the

Cr(VI) contaminated soil collected from the landfill site of chemical industry, Vadodara, Gujarat,

India.

On the basis of the carbon substrate utilization pattern (employing GN2 as well as GENIII plates of

BioLog, USA) and 16s rDNA sequence, isolate BCR400 was identified as O. intermedium with

similarity index of 0.74 and 99% probability (Accession number: JN033212).

O. intermedium belongs to ά–proteobacteria subclass and representatives of this taxa have been

isolated previously from chromium contaminated soils, by several researchers (He et al., 2009;

Ozdemir et al., 2003; Sultan and Hasnain, 2007; Thacker and Madamwar, 2005). O. intermedium

BCR400 could grow in medium containing 500 mg Cr(VI)/L.

Figure 3.3 Profile of Cr(VI) reduction and growth of O. intermedium BCR400 in LB broth

amended with 100 mg Cr(VI)/L.

REDUCTION OF Cr(VI)

45

3.3.1 Time-course of Cr(VI) reduction by O. intermedium BCR400

As shown in Figure 3.3, after a lag of ~30 min growth, Cr(VI) reduction initiated simultaneously in

an agitated batch culture of O. intermedium BCR400. The complete reduction of 112 mg Cr(VI)/L

occurred within 72 h of incubation with initial reduction rate of 1.98 mg Cr(VI)/L/h. Thus, O.

intermedium BCR400 not only showed resistance to Cr(VI) but also possessed ability to reduce

Cr(VI), similar observations have been reported by Branco et al. (2004) in their studies on Cr(VI)

reduction by O.tritici.

The resting cells of O. intermedium BCR400 did not show any significant reduction of Cr(VI). The

growth associated Cr(VI) reduction suggests the role of actively metabolizing cells in Cr(VI)

reduction.

3.3.2 Effect of temperature and pH

The Cr(VI) reduction by O. intermedium BCR400 was found to increase with increase in incubation

temperature from 25 to 37 ºC. Further increase in temperature above 37 ºC caused decrease in both

extent of Cr(VI) reduction as well as its rate. Hence, O. intermedium BCR400 was found to reduce

Cr(VI) in temperature range of 25 to 40 ºC with optimum activity at 37 ºC (Figure 3.4a).

Figure 3.4 Effect of temperature (a) and pH (b) on Cr (VI) reduction rate (mg/L/h) of O.

intermedium BCR400 grown in LB broth amended with 100 mg Cr (VI) /L.

O. intermedium BCR400 exhibited Cr(VI) reduction over a pH range of 6.0 to 8.0, with optimum

activity at pH 7.0 (Figure 3.4b). Our results are in agreement with the optimum pH and temperature

of Bacillus sp. ev3 (pH: 7.0, Temperature: 37 ºC), Brucella sp. (pH: 7.0, Temperature: 37 ºC) and

REDUCTION OF Cr(VI)

46

O. intermedium (pH: 7.0, Temperature: 37 ºC) reported by Rehman et al. (2008), Thacker et al.

(2005), Sultan and Hasnain (2007), respectively.

3.3.3 Effect of initial Cr(VI) concentration on growth and Cr(VI) reduction

O. intermedium BCR400 was able to grow up to 500 mg Cr(VI)/L, while growth rate was found to

decrease with increasing Cr(VI) concentration. The initial Cr(VI) reduction rate was found to

increase up to 100 mg Cr(VI)/L without influencing the Cr(VI) reduction efficiency. However,

further increase in Cr(VI) concentration up to 500 mg/L, negatively affected initial Cr(VI) reduction

rate as well as extent of Cr(VI) reduction by O. intermedium BCR400 (Table 3.2). The decrease in

amount of biomass produced with increasing Cr(VI) concentration may be attributed to

consumption of increasing amount of reducing power generated by cell towards reductive

detoxification of Cr(VI) (Branco et al., 2004 ; Megharaj et al., 2003). This means, electrons

generated by oxidation of organic carbon sources which otherwise would be utilized for

biosynthesis of cell components, seems to be diverted towards Cr(VI) reduction, thus slowing down

growth. Similar observations on decrease in Cr(VI) reduction rate with increasing concentration of

Cr(VI) have been reported by several researchers using different bacterial cultures (Garbisu et al.,

1998; Megharaj et al., 2003; Sultan and Hasnain, 2006).

Table 3.2 Effect on initial Cr(VI) (50 to 500 mg/L) on Cr(VI) reduction and growth of O.

intermedium BCR400.

Initial Cr(VI)

Concentration

(mg/L)

Cr(VI)

Reduction

(mg/L)

Initial Cr(VI)

reduction rate

(mg/L/h)

Maximum

Biomass

(g/L)

Growth rate

(g/L/h)

50 48.15 ± 2.61 1.29 ± 0.148 1.71 ± 0.001 0.0239 ± 0.002

100 99 ± 4.24 1.98 ± 0.034 1.38 ± 0.002 0.0206 ± 0.001

200 140 ± 8.48 1.20 ± 0.118 1.29 ± 0.001 0.0196 ± 0.001

300 131.5 ± 6.36 0.97 ± 0.002 0.93 ± 0.002 0.0132 ± 0.003

400 125 ± 5.65 0.844 ± 0.006 0.012± 0.0003 0.009 ± 0.0007

500 127 ± 0.0 0.730 ± 0.034 0.012± 0.0002 0.007 ± 0.0005

REDUCTION OF Cr(VI)

47

3.3.4 Effect of metal ions on Cr(VI) reduction by O. intermedium BCR400

The industrial wastewaters would contain various other heavy metals apart from Cr(VI), which

though at lower concentration might influence the activity of chromium reducing bacteria (Agrawal

et al., 2008). Thus, Cr(VI) reduction by O. intermedium BCR400 was studied in presence of other

metallic ions at concentration of 50 mg/L. Table 3.3 shows that Cr(VI) reduction increases slightly

in presence of copper (50 mg/L) whereas it decreases in presence of same concentration of cobalt,

arsenic, nickel, zinc and cadmium. The, level of inhibition was different with different metal ions;

cadmium was found to exert maximum inhibitory effect followed by zinc, nickel, cobalt, arsenate at

50 mg/L concentration. Inhibition of Cr(VI) reduction by O. intermedium in presence of metal ions

like Zn2+

, Cd2+

and Pb2+

has been reported by Sultan and Hasnain (2007) however it is interesting to

note they also observed increase in Cr(VI) reduction in presence of metal ions like Cu2+

, Co2+

, Mn2+

and Ni2+

. The main mechanism behind stimulatory effect of certain metals on Cr(VI) reduction is

not clear how ever it may be attributed to their role as cofactors for reductases involved in Cr(VI)

transformation.

Table 3.3 Maximum Cr (VI) reduction and Cr (VI) reduction rate of O. intemedium BCR400 in LB

broth amended with 100 mg/L Cr (VI) supplemented with co-metal ions at concentration of 50

mg/L.

Co-Metal ions

50 (mg/L)

Maximum Cr(VI)

Reduction (mg/L)

Cr(VI) Reduction

Rate (mg/L/h)

None 95.5±0.707 1.98±0.014

Copper (Cu2+

) 100±0.707 2.09±0.014

Cobalt (Co2+

) 65±4.24 1.35±0.088

Arsenic (As5+

) 65±5.65 1.35±0.117

Nickel (Ni2+

) 59.5±3.5 1.23±0.073

Zinc (Zn2+

) 54.5±2.12 1.13±0.044

Cadmium (Cd2+

) 44±2.82 0.91±0.058

3.3.5 AQS-mediated reduction of Cr(VI) by O. intermedium BCR400

Inclusion of low molecular weight redox mediator along with the metal is being currently

considered in order to enhance efficiency of metal reduction by micro-organisms (Bond and Lovley,

2002; Lovley et al., 1998; Lovley et al., 1999). In the present study, O. intermedium BCR400

exhibited higher initial Cr(VI) reduction rate in presence of 1.0 mM AQS, ethyl and methyl

REDUCTION OF Cr(VI)

48

viologen. In presence of 1.0 mM AQS, O. intermedium BCR400 reduced Cr(VI) at 1.4 fold higher

rate (2.88 mg Cr(VI)/L/h) than in absence of any redox mediators (1.98 mg Cr(VI)/L/h).

AQS has been reported as powerful mediator for reductive biotransformation of several organic

recalcitrants (Liu et al. 2010; Zee and Cervantes 2009). Although ethyl and methyl viologen also

enhanced Cr(VI) reduction rate, they were found to inhibit growth of O. intermedium BCR400

(Table 3.4). Benzyl viologen severely inhibited the growth of O. intermedium BCR400.

Quinones are known for their potential role as redox centres in humic acid and therefore the Cr(VI)

reduction rates in presence of AQS may attributed to their electron shuttling ability between O.

intermedium BCR400 and Cr(VI). Zee and Cervantes (2009) suggested that the oxidation reduction

potential of any mediator should not be much lower than -0.320 V, which is the lowest oxidation

reduction potential of cofactor (NADPH) in the cell otherwise it would not be reduced significantly.

Hence, the transfer of electrons from AQS to Cr(VI) seems to be thermodynamically favourable

phenomenon as the standard redox potential (E°) of AQS is -0.218 V which is significantly lower

than the redox potential of CrO42-

(1.28 V) and higher than -0.320 V (Fredrickson et al., 2006).

According to the above hypothesis the poor mediating ability of ethyl viologen and methyl viologen

may be explained by their much lower oxidation reduction potential (- 0.480 and – 0.440 V

respectively) than – 0.320 V.

Table 3.4 Effect of redox mediators on Cr(VI) reduction and growth of O. intermedium BCR400.

Redox mediators

(1.0 mM)

Cr(VI)

Reduction

(mg/L)

Initial Cr(VI)

reduction

rate (mg/L/h)

Maximum

Biomass

(g/L)

Growth rate

(g/L/h)

None 101 1.98 1.01 0.038

AQS 101 2.88 1.35 0.051

Ethyl Viologen 102 2.12 0.825 0.022

Methyl Viologen 103 2.14 0.768 0.021

Benzyl Viologen 65.52 0.91 0.045 0.0012

Furthermore, O. intermedium BCR400 exhibited higher initial Cr(VI) reduction rate in presence of

1.0 mM AQS at all Cr(VI) concentrations tested (Table 3.5). The initial Cr(VI) reduction rates in

batch cultures of O. intermedium BCR400 were found to be 1.29, 1.98, 1.20, 0.97, 0.844 and 0.73

mg Cr(VI)/L/h corresponding to the initial Cr(VI) concentration of 50, 100, 200, 300, 400 and 500

REDUCTION OF Cr(VI)

49

mg/L, respectively. However, in batch cultures augmented with 1.0 mM AQS, initial Cr(VI)

reduction rates were found to be approximately 1.81, 2.88, 1.44, 1.167, 1.01 and 0.876 Cr(VI)/L/h

corresponding to initial Cr(VI) concentration of 50, 100, 200, 300, 400 and 500 mg Cr(VI)/L

(Figure 3.5). Hence, initial Cr(VI) reduction rate of O. intermedium BCR400 was found to decrease

with increasing concentration of Cr(VI) irrespective of AQS supplementation in the growth

medium.

Figure 3.5 Cr(VI) reduction by O. intermedium BCR400 in the presence or absence of AQS. The

cells were cultured in Luria Bertani broth amended with 50 to 500 mg Cr(VI)/L supplemented with

or without AQS (1.0 mM).

3.3.6 Effect of AQS concentration on Cr(VI) reduction

The optimum AQS:Cr(VI) ratio for Cr(VI) reduction by O. intermedium BCR400 was determined

by varying the concentration of AQS while maintaining Cr(VI) concentration constant. It is clearly

evident from Table 3.5, that initial Cr(VI) reduction rate of O. intermedium BCR400 increased with

increase in AQS concentration from 0.1 mM to 1.0 mM. Further increase in AQS concentration to

2.0 mM caused decrease in initial Cr(VI) reduction rate as well as growth of O. intermedium

BCR400. The lower Cr(VI) reduction rate at sub-optimal AQS concentrations may be explained by

distribution of electrons between AQS mediated reduction and direct Cr(VI) reduction. On the other

hand higher AQS concentrations (2.0 mM) itself may be inhibitory to cell growth thus negatively

influencing Cr(VI) reduction (Ling et al., 2009). This is because at higher AQS concentrations, the

accumulation of reduced AQS (AQH2S), in absence or limiting concentration of suitable electron

REDUCTION OF Cr(VI)

50

acceptor, would transfer electrons to O2, resulting in generation of ROS and thereby causing cell

death (Sedlacek and Kucera, 2010). The maximum reduction rate of Cr(VI) as well as growth of O.

intermedium BCR400 at optimal AQS (1.0 mM) in medium containing 100 mg Cr(VI)/L may be

due to efficient relay of electrons from primary electron donor to AQS to Cr(VI) mediated through

action of reductases. Rau et al. (2002) also observed highest reduction of azo dye amaranth in

presence of 1.0 mM AQS.

Table 3.5 Effect on AQS concentration (0 to 2.0 mM) on Cr(VI) reduction and growth of O.

intermedium BCR400.

Redox

mediator

(mM)

Cr(VI)

Reduction

(mg/L)

Initial

Cr(VI)

reduction

rate

(mg/L/h)

Maximum

Biomass

(g/L)

Growth rate

(g/L/h)

0 99.5 1.98 1.01 0.038

0.1 86.1 1.53 0.984 0.030

0.2 95.7 1.70 1.038 0.032

0.5 98 2.0 1.08 0.033

1.0 98.5 2.88 1.334 0.051

2.0 84 1.52 0.90 0.028

3.3.7 Intracellular quinone and chromate reductase activities of O. intermedium BCR400

induced in the presence of AQS

The intracellular quinone and chromate reductase activities in O. intermedium BCR400 grown in

LB were found to be 0.0063 ± 0.0010 and 0.0042 ± 0.0001 U/mg, respectively (Table 3.6). The

quinone and chromate reductase activities were found to increase in O. intermedium BCR400 cells

grown in presence of Cr(VI) (100 mg/L) as well as 1.0 mM AQS. Furthermore, the extent of

quinone reductase induction was found to be higher in cells grown in presence of AQS than in

presence of Cr(VI); whereas the extent of chromate reductase induction was not profoundly

influenced by Cr(VI) over AQS.

REDUCTION OF Cr(VI)

51

Table 3.6 Quinone and chromate reductase activities (U/mg) of O. intermedium BCR400 grown in

Luria Bertani (LB) broth amended with either Cr(VI) or AQS or both Cr(VI) and AQS.

Growth medium Specific Activity (U/mg protein)

Quinone Reductase Chromate Reductase

LB 0.0063±0.0010 0.0042±0.0001

LB + Cr(VI) 0.0139±0.0028 0.013±0.00018

LB + AQS 0.026±0.004 0.016±0.0005

LB + AQS + Cr(VI) 0.025±0.0036 0.020±0.00054

It is noteworthy to mention that the presence of AQS + Cr(VI) synergistically influenced the level

of intracellular chromate reductase but not quinone reductase. This suggests that Cr(VI) reduction in

O. intermedium BCR400 may occur directly by chromate reductase as well as by reduced AQS

(AQH2S) formed upon action of quinone reductase. Furthermore, both types of reductases seem to

have relaxed substrate specificity; chromate reductase being more specific for Cr(VI), while

quinone reductase being more specific for quinoid compounds (Puzon et al., 2002; Rau et al.,

2002).

Figure 3.6 Effect of AQS concentration (0.1 to 1.0 mM) on intracellular levels of quinone and

chromate reductases of O. intermedium BCR400 grown in LB containing 100 mg Cr(VI)/L

When O. intermedium BCR400 was grown in the presence of varying AQS concentration (0.1 to

1.0 mM) the intracellular quinone and chromate reductase activities were found to increase from

REDUCTION OF Cr(VI)

52

0.0025 ± 0.00019 to 0.030 ± 0.0024 U/ mg and 0.0027 ± 0.00002 to 0.027 ± 0.0005 U/mg total

cellular proteins, respectively (Figure 3.6).

Alternately, when O. intermedium BCR400 was grown in presence of varying Cr(VI) concentration

(0 to 500 mg Cr(VI)/L) in LB augmented with 1.0 mM AQS, the intracellular levels of quinone

reductase increased from 0.02 ± 0.0041 to 0.044 ± 0.0026 U/mg protein, while chromate reductase

increased from 0.016 ± 0.0005 to 0.0368 ± 0.002 U/mg protein (Figure 3.7).

Figure 3.7 Effect of Cr(VI) concentration (50 to 500 mg/L) on intracellular levels of quinone and

chromate reductases of O. intermedium BCR400 grown in LB augmented with 1 mM AQS.

Figure 3.8 Effect of pH and temperature on quinone and chromate reductase activity (U/mg) of O.

intermedium BCR400.

REDUCTION OF Cr(VI)

53

A strong and non linear positive co-relation between the concentration of AQS/Cr(VI) and activities

of both the enzymes (γ > + 90) was observed. This suggests the role of AQS in induction of quinone

and chromate reductases in O. intermedium BCR400.

Both quinone and chromate reductases were found to follow similar activity profile over a pH range

of 5.0 - 8.5 with maximum activities at pH 7.0. The quinone and chromate reductase also exhibited

apparently similar activity profiles over a temperature range of 25 to 50 °C (Figure 3.8). On the

basis of relative induction of chromate and quinone reductase by AQS and Cr(VI) both enzyme

activities seems to be independent of each other.

3.3.8 Effect of temperature on AQS mediated Cr(VI) reduction

The rate of AQDS (Anthra Quinone Di- Sulfonate) mediated azo dye reduction was significantly

higher in comparison to non-mediated azo dye reduction suggesting the role of AQDS in lowering

the activation energy (Ea) of azo dye reduction (Santos et al., 2004b). Therefore, we monitored

initial Cr(VI) reduction rate of O. intermedium BCR400 cells in presence and absence of 1.0 mM

AQS at different temperatures (25 to 35 ºC).

The non mediated Cr(VI) reduction rate of O. intermedium BCR400 increased from 0.005 to 0.021

h-1

with increase in temperature from 25 to 35ºC. Similarly, the AQS-mediated Cr(VI) reduction

rate increased from 0.007 to 0.034 h-1

with increase in temperature from 25 to 30 ºC (Table 3.7).

Table 3.7 Cr(VI) reduction rate of non-mediated and AQS-mediated Cr(VI) reduction.

Temperature (ºC) Cr(VI) reduction rate (h-1

)

Non Mediated AQS- Mediated

25 0.005 0.007

30 0.020 0.022

35 0.021 0.034

Furthermore the activation energy of AQS mediated Cr(VI) reduction (120.69 KJ/mol) was found

to be 1.1-fold lower than non mediated Cr(VI) reduction (133.86 KJ/mol). Santos et al. (2004)

reported that activation energy of AQS mediated azo dye (reactive red 2) reduction (22.9 KJ/mol)

by anaerobic sludge was 1.2-fold lower than non mediated dye reduction (27.9 KJ/mol) in

anaerobic sludge.

There have been many studies on the role of redox mediators such as AQS, AQDS and lawsone in

bacterial azo dye reduction under anaerobic condition. Rau et al. (2002) proposed that mechanism

for such redox mediator dependent reduction of azo dyes consist of two independent reaction steps:

REDUCTION OF Cr(VI)

54

First, the quinones are enzymatically reduced to the corresponding hydroquinone (Ling et al., 2009;

Rau and Stolz, 2003) and second, the hydroquinones cleave the azo dyes in purely chemical

reaction. Following the proposed mechanism by Rau et al. (2002), Ling et al. (2009) suggested that

entire reaction rate depends on enzymatic reduction of redox mediators. Accordingly, the

effectiveness of any mediator depends on the presence of membrane bound or intracellular

reductase with mediator reducing ability.

Part B: Reduction of Cr(VI) in liquid media and soil microcosms by Bacillus sp BT1 isolated

from chromium contaminated soil of tanning industry

A hexavalent chromium resistant bacterial strain designated as BT1 was isolated from the Cr(VI)

contaminated soil collected from the vicinity of tannery industry, Ranipet, Tamilnadu, India. The

carbon substrate utilization pattern of BT1 (using Biolog identification system) and 16s rDNA

sequencing revealed that BT1 was closely related to Bacillus sp. (Accession number: JN093017).

The Bacillus sp. have been reported for their occurrence in metal stressed environment as well as

ability to reduce Cr(VI) (Camargo et al., 2004; Campos et al., 1995; Elangovan et al., 2006; Liu et

al., 2006). Bacillus sp. BT1 showed similar Cr(VI) reduction under both agitated as well as static

growth conditions (Data not shown). Hence, all the experiments were carried out under static

growth conditions.

3.3.9 Time-course of Cr(VI) reduction by Bacillus sp. BT1

Growth as well as Cr(VI) reduction was monitored at initial Cr(VI) concentration of 100 mg/L at

pH 7.0 and 30 °C under static condition. The Cr(VI) reduction by Bacillus sp. BT1 was found to be

growth associated as reduction efficiency increased with increase in the growth of culture. The

complete reduction of 100 mg Cr(VI)/L was observed after 132 h of incubation at 30 °C (Figure

3.9). Liu et al. (2006) also demonstrated that there exists relationship between growth and Cr(VI)

reduction as indicated by difference in the growth curve pattern corresponding to various

concentration of Cr(VI).

REDUCTION OF Cr(VI)

55

0

20

40

60

80

100

0 50 100 150Time (h)

Resid

ual

Cr(

VI)

(m

g/L

)

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Gro

wth

(A

600n

m)

Residual Cr(VI) (mg/L) Growth (A600nm)

Figure 3.9 Profile of Cr(VI) reduction and growth of Bacillus sp. BT1 in LB broth amended with

100 mg Cr(VI)/L.

3.3.10 Effect of temperature and pH

An increase in temperature from 25 to 30 ºC led to 1.4-fold increase in Cr(VI) reduction rate, which

remained approximately same at 35 ºC. The considerable decrease in Cr(VI) reduction rate was

observed at incubation temperature above 35 ºC (Figure 3.10a). At optimum temperature (30 ºC)

Bacillus sp. BT1 displayed Cr(VI) reduction over a broad pH range with maximum reduction rate at

pH 7.0 (0.7 ± 0.04) (Figure 3.10b).

Our observations are in agreement to reports of Elongovan et al. (2006, 2008) where maximum

Cr(VI) reduction in Arthrobacter rhombi-RE and Bacillus sp. was found to be at 30 °C and pH 6.0.

Sultan and Hasnain (2006) reported maximum Cr(VI) reduction by Ochrobactrum intermedium at

37 °C and pH 7.0.

3.3.12 Effect of metal ions on Cr(VI) reduction by Bacillus sp. BT1

The Cr(VI) contaminated industrial wastewaters may contain several other heavy metals as well,

which though at lower concentrations might influence the efficiency of chromium reducing bacteria

(Ganguli and Tripathi, 2002). Therefore, the effect of other metal ions on Cr(VI) reduction by

Bacillus sp. BT1 was evaluated. The presence of Co2+

, As5+

or Ni2+

ions interfered with the

chromium reducing ability of Bacillus sp. BT1 as evidenced by the reduced Cr(VI) reduction rate,

whereas Cu2+

(50 mg/L) did not have any significant effect (Table. 3.8). The similar observations

were made for O. intermedium BCR400.

REDUCTION OF Cr(VI)

56

Camargo et al. (2003) reported higher Cr(VI) reduction rate in presence of Cu2+

using cell free

extract of Bacillus sp. ES29 and correlated it with role of Cu2+

as prosthetic group for reductases.

Figure 3.10 Effect of temperature (a) and pH (b) on Cr(VI) reduction rate (mg/L/h) of Bacillus sp.

BT1 grown in LB broth amended with 100 mg Cr(VI) /L.

Figure 3.11 Cr(VI) reduction rate (mg/L/h) of Bacillus sp. BT1 in LB broth amended with varying

concentration of Cr(VI) (100 to 300 mg/L).

REDUCTION OF Cr(VI)

57

Table 3.8 Maximum Cr(VI) reduction and Cr(VI) reduction rate of Bacillus sp. BT1 in nutrient

medium amended with 100 mg/L Cr(VI) supplemented with co-metal ions at concentration of 50

mg/L.

Co-Metal ions (50 mg/L) Cr (VI) Reduction Rate (mg/L/h)

None 0.715±0.048

Copper (Cu2+

) 0.771±0.063

Cobalt (Co2+

) 0.575±0.031

Arsenic (As5+

) 0.574±0.089

Nickel (Ni2+

) 0.528±0.058

3.3.13 AQS mediated reduction of Cr(VI) by Bacillus sp. BT1

The initial Cr(VI) reduction rates in batch cultures of Bacillus sp. BT1 were found to be 0.74, 0.61

and 0.041 mg Cr(VI)/L/h corresponding to the initial Cr(VI) concentration of 100, 200 and 300

mg/L respectively. However, in batch cultures augmented with 1.0 mM AQS, initial Cr(VI)

reduction rates were found to be approximately 4.03, 1.53 and 1.64 Cr(VI)/L/h corresponding to

initial Cr(VI) concentration of 100, 200 and 300 mg Cr(VI)/L (Figure 3.12).

The extent of Cr(VI) reduction was also enhanced in cultures augmented with 1.0 mM AQS. For

instance, 72 % and 55.3 % of initial 200 and 300 mg Cr(VI)/L was reduced by Bacillus sp. BT1 in

LB augmented with 1.0 mM AQS, respectively. However in LB without AQS, Bacillus sp. BT1

reduced only 36 % of initial 200 mg Cr(VI)/L, which was 2.25-fold lower in comparison to AQS

augmented cultures. In absence of AQS, Bacillus sp. BT1 failed to grow in LB containing 300 mg

Cr(VI)/L.

Our results are in agreement with observations of Liu et al. (2010) who reported enhanced Cr(VI)

reduction by resting cells of E. coli in presence of quinone redox mediators like lawsone,

menadione, AQS and AQDS.

REDUCTION OF Cr(VI)

58

Figure 3.12 Cr(VI) reduction by Bacillus sp. BT1 in the presence or absence of AQS. The cells

were cultured in Luria Bertani broth amended with 100, 200, 300 mg Cr(VI)/L supplemented with

or without AQS (1.0 mM).

The enhanced bioremediation potential of many bacterial species in pure cultures amended with

quinone as an electron shuttle has been well documented (Ling et al., 2009; Liu et al., 2009; Santos

et al., 2004b). Ling et al. (2009) demonstrated very high reduction rate of sulfonated azo dye (0.2

mM) by Sphingomonas xenophaga in presence of 0.05 mM AQS (70.96 µmol/g.h) in comparison to

the AQS deficient assays ( 36.11 µmol/g.h). Quinones are known for their potential role as redox

centres in humic acid and therefore, the higher Cr(VI) reduction rate by Bacillus sp. BT1 in

presence of AQS is presumably a result of facilitated flux of electrons toward Cr(VI). Furthermore,

the transfer of electrons from AQS to Cr(VI) is a thermodynamically feasible phenomenon as the

standard redox potential (Eº) of AQS is -0.218 V which is significantly lower than the redox

potential of CrO42-

(1.28 V). Similar possibilities were investigated by Fredrickson et al. (2000),

who demonstrated that Deinococcus. radiodurans could couple the reduction of anthra quinone di-

sulfonate (AQDS) to the reduction of Fe oxides like hydrous ferric oxide, ferric pyrophosphate,

ferric citrate and goethite. D. radiodurans exhibited higher rates of Cr(VI) reduction in presence of

AQDS using lactate as electron donor under anaerobic conditions. It also exhibited reduction of

U(VI) and Tc(VII) to insoluble U(IV) and Tc(IV) only in presence of AQDS. One additional

property of AQS which supports its ability to act as redox mediator is its permeability across

bacterial cell membrane (Zee and Cerventes, 2009).

REDUCTION OF Cr(VI)

59

3.3.14 Effect of AQS concentration on Cr(VI) reduction by Bacillus sp. BT1

Zee and Cerventes (2009) stated that redox mediator concentration used in batch experiment may be

a limiting factor for AQS-mediated biotransformation of any organic inorganic pollutant. The

biotransformation rate increases with increasing mediator concentration. Therefore, effect of

increasing AQS concentration (0 to 5.0 mM) on efficiency and rate of bacterial Cr(VI) reduction

was tested (Figure 3.13).

The increasing AQS concentrations in comparison to its absence was found to have a positive effect

on both efficiency and rate of Cr(VI) reduction (100 mg/L). Bacillus sp. BT1, in the presence of 1.0

mM AQS was found to reduce 100% of Cr(VI) in 24 h with a specific Cr(VI) reduction rate of 3.89

mg Cr(VI)/L/h. This was found to be 5.67-fold higher in comparison to Cr(VI) reduction rate in

absence of AQS. The impact was disconcerting beyond 1.0 mM AQS as at higher AQS

concentrations (2.0 and 5.0 mM), only 15.32 % and 7.97 % of Cr(VI) was reduced with initial

reduction rate of 0.299 and 0.15 mg Cr(VI)/L/h, respectively. This may be attributed either to the

AQS toxicity at higher concentration or to the limiting concentrations of electron donor

(Sendelbach, 1989).

Figure 3.13 Effect of AQS concentration on Cr(VI) reduction by Bacillus sp BT1. The cells were

cultured in Cr(VI) amended (100 mg Cr(VI) L-1

) Luria Bertani broth supplemented with 0, 0.05,

0.2, 0.5, 1.0, 2.0 and 5.0 mM AQS.

REDUCTION OF Cr(VI)

60

3.3.15 Reduction of repeated spikes of Cr(VI) in batch cultures

Bacillus sp. BT1 reduced 100 mg Cr(VI)/L in 120 h with initial reduction rate of 0.83 mg

Cr(VI)/L/h in absence of AQS. However, in presence of AQS, Bacillus sp. BT1 reduced total of

400 mg Cr(VI)/L (during 100 h of incubation) in four spiking of 100 mg Cr(VI)/L each with initial

reduction rate of 4 mg Cr(VI)/L/h (Figure 3.14). The failure of Cr(VI) reduction upon further

spiking of Cr(VI) may be attributed to exhaustion of primary electron donor. Liu et al. (2010)

reported four rounds of 100 mg Cr(VI)/L reduction by resting cells of E. coli in presence of lawsone

indicating the stability and persistence of mediated Cr(VI) reduction.

Figure 3.14 Repeated cycles of Cr(VI) reduction by Bacillus sp. BT1 in presence and absence of

1.0 mM AQS. The Cr(VI) (100 mg/L) was spiked every time after previously added Cr(VI) was

reduced completely.

3.3.16 Effect of AQS on quinone and chromate reductases in the cell-free extract (CFE) of

Bacillus sp. BT1.

Chromate reductases from P. putida, E. coli and Paracoccus denitrificans have been demonstrated

to exhibit quinone reductase activity as well (Ackerley et al., 2004; Gonzalez et al., 2005; Sedlacek

and Kucera 2010). Liu et al. (2009) demonstrated that over expression of quinone reductase in E.

coli JM109 accelerated quinone dependent azo dye AR27 reduction. E. coli JM109 does not reduce

azo dye AR27 in absence of quinone. Since Cr(VI) reduction activity of Bacillus sp. BT1 was found

to be enhanced in presence of quinone, both chromate and quinone reductase activities were

assayed in cells cultured in the presence/ absence of AQS and/ or Cr(VI).

REDUCTION OF Cr(VI)

61

Table 3.9 Quinone and chromate reductase activities (U/mg) of Bacillus sp. BT1 grown in Luria

Bertani (LB) broth amended with either Cr(VI) or AQS or both Cr(VI) and AQS.

Growth medium Specific Activity (U/mg)

Quinone Reductase Chromate Reductase

LB 0.035±0.0002 0.031±0.0012

LB + Cr(VI) 0.0445±0.0005 0.0335±0.0025

LB + AQS 0.3925±0.057 0.0535±0.0235

LB + AQS + Cr(VI) 2.755±0.255 0.6855±0.17

The intracellular quinone and chromate reductase activities of Bacillus sp. BT1 grown in LB were

found to be 0.035±0.0002 and 0.031±0.0012, respectively. The quinone and chromate reductase

activities were found to increase in Bacillus sp. BT1 cells grown in presence of Cr(VI) (100 mg/L)

as well as 1.0 mM AQS (Table 3.9). It is noteworthy to mention here that both the activities were

highest in cells grown in presence of both Cr(VI) (100 mg/L) and AQS (1.0 mM)

When Bacillus sp. BT1 was grown in LB containing Cr(VI) (100 mg/L) and varying AQS

concentration (0 to 1.0 mM), the chromate and quinone reductases (U/mg protein) in its cell-free

extract (CFE) displayed high, non linear positive co-relation with the concentration of AQS (γ > +

90).

The specific chromate and quinone reductase activities in Bacillus sp. BT1 grown in LB amended

with 100 mg Cr(VI)/L were 0.033 and 0.044 U/mg protein, respectively. Both these enzyme

activities were found to increase in cells grown in LB augmented with increasing concentration of

AQS while keeping initial Cr(VI) (100 mg/L) concentration constant.

The intracellular levels of quinone reductase were found to increase by 4.8-, 22.2- and 62.6-fold in

Bacillus sp. BT1 grown in presence of 0.1 mM, 0.5 mM and 1.0 mM AQS in LB amended with 100

mg Cr(VI)/L, respectively in comparison to control. It was interesting to observe that along with

quinone reductase, even chromate reductase levels in cells grown in presence of 0.1 mM, 0.5 mM

and 1.0 mM AQS in LB amended with 100 mg Cr(VI)/L were found to increase by 2.16-, 7.0- and

22.8-fold, respectively in comparison to control (Figure 3.15).

Alternately, when Bacillus sp. BT1 was grown in presence of 50 mg and 100 mg Cr(VI)/L in LB

augmented with 1 mM AQS, the intracellular levels of quinone reductase increased by 2.37- and

7.06-fold, while chromate reductase increased by 1.83- and 12.93-fold, respectively (Figure 3.16).

It can be inferred that while both AQS as well as Cr(VI) caused increase in intracellular levels of

quinone and chromate reductase, AQS induced levels of both enzymes to the similar extent whereas

REDUCTION OF Cr(VI)

62

Cr(VI) seems to be a better inducer of chromate reductase than for quinone reductase in Bacillus sp.

BT1 (Figure 3.15 & 3.16).

Figure 3.15 Effect of AQS concentration on intracellular levels of quinone and chromate reductases

of Bacillus sp. BT1 grown in LB containing 100 mg Cr(VI)/L.

Figure 3.16 Effect of Cr(VI) concentration on intracellular levels of quinone and chromate

reductases of Bacillus sp. BT1 grown in LB augmented with 1.0 mM AQS.

REDUCTION OF Cr(VI)

63

Both quinone and chromate reductases in CFE of Bacillus sp. BT1 were found to show maximum

activity at pH 6.0 and temperature 30 ºC (Figure 3.17). However, the CFE of Bacillus sp. BT1

retained approximately 80 % of its maximum quinone and chromate reductase activity at pH 7.5

and 40 ºC. The results are consistent with the pH and temperature optima of Cr(VI) reduction

activity from CFE of Exiguobacterium sp and Bacillus sp. (Sarangi and Krishnan., 2007).

Figure 3.17 Effect of pH (a) and temperature (b) on chromate and quinone reductase activities of

Bacillus sp BT1

3.3.17 Effect of temperature on AQS mediated Cr(VI) reduction

The non-mediated Cr(VI) reduction rate (k, h-1

) of Bacillus sp. BT1 increased from 0.017 to 0.034

h-1

with increase in temperature from 25 to 35 ºC. The mediated Cr(VI) reduction rate increased

from 0.038 to 0.164 h-1

, with increase in temperature from 25 to 35 ºC (Table 3.10).

The AQS mediated Cr(VI) reduction rate of Bacillus sp. BT1 were found to be 2.2-, 3.7- and 4.8-

fold higher than non- mediated Cr(VI) reduction at 25, 30 and 35 ºC respectively (Table 3.10).

Table 3.10 First-order rate constant k (h-1

) of non-mediated and AQS-mediated Cr(VI) reduction.

Temperature (ºC) First order rate constant, k (h-1

)

Non Mediated AQS-Mediated

25 0.017 0.038

30 0.023 0.087

35 0.034 0.164

REDUCTION OF Cr(VI)

64

Figure 3.18 Arrhenius plot of AQS mediated and non-mediated Cr(VI) reduction monitored at

incubation temperatures from 25 to 35 ºC.

Activation energy, Ea was calculated using Arrhenius plot of ln k (h-1

) against 1/T (kelvin). The

calculated Ea value of AQS-mediated and non-mediated Cr(VI) reduction were found to be 52.7 and

111.4 KJ/mol, respectively (Figure 3.18). The 2.1-fold lower activation energy of AQS mediated in

comparison to non-mediated Cr(VI) reduction reflects the difference in Cr(VI) reduction

mechanism of Bacillus sp. BT1 in presence and absence of AQS. This also suggests that mechanism

of Cr(VI) reduction in Bacillus sp. BT1 is significantly different from that in O. intermedium

BCR400.

3.3.18 Influence of nitrate (NO3-) on Cr(VI) reduction by Bacillus sp. BT1

Class II family chromate reductases (NfsA in E.coli) have been reported to exhibit nitroreductase

activity as well (Ackerley et al., 2004a; Whiteway et al., 1998; Park et al., 2002). It was

hypothesized that if chromate and quinone reductase activity of Bacillus sp. BT1 also possessed

NO3- reductase activity, then presence of NO3

- in medium should influence its chromate reduction

activity. In order to test the hypothesis, we investigated chromate reduction by growing cells of

Bacillus sp. BT1 in presence of varying concentrations of NO3-. The chromate reduction rate was

not affected by presence of NO3- upto 0.5% (w/v) in the medium (Figure 3.19).

REDUCTION OF Cr(VI)

65

Figure 3.19 Effect of nitrate (NO3-) concentration on Cr(VI) reduction by Bacillus sp. BT1. The

cells were cultured in Luria Bertani broth amended with Cr(VI) (100 mg Cr(VI)/L) and NO3- (0-0.5

%) supplemented with/without AQS.

This means that putative chromate reductase of Bacillus sp. BT1 either has very low affinity for

NO3- or does not possess nitrate reductase activity at all. However, chromate reduction rate of

Bacillus sp. BT1 in cultures augmented with 1.0 mM AQS with varying concentration of NO3-, was

found to decrease with increase in concentration of NO3- (Figure 3.19). This decrease in chromate

reduction rate in the presence of AQS and NO3- may be attributed to competition between NO3

- and

Cr(VI) for accepting electrons from reduced AQS. Furthermore, NO3- is having oxidation reduction

potential (+0.78 mv) intermediate to oxidation reduction potential of AQS and Cr(VI); it will

scavenge electrons with higher affinity from AQS in comparison to Cr(VI).

Similar trend was observed for growth rate of Bacillus sp. BT1 in medium containing AQS (1.0

mM) and varying concentration of NO3- (Figure 3.20). This may be explained as follows: in

presence of AQS, electrons from NADH will be transferred to AQS in reaction catalysed by

quinone reductase and reduced AQS would preferably reduce NO3- as a result, Cr(VI) would

continue exerting toxic effect, thereby reducing growth rate. In absence of AQS, growth rate of

Bacillus sp. BT1 was not affected significantly with increasing NO3- concentration upto 0.5% (w/v).

This suggests that in absence of AQS, NO3- does not compete with Cr(VI) for electron donors in

Bacillus sp. BT1.

REDUCTION OF Cr(VI)

66

Figure 3.20 Effect of nitrate (NO3-) concentration on growth rate of Bacillus sp BT1. The cells

were cultured in Luria Bertani broth amended with Cr(VI) (100 mg Cr(VI)/L) and NO3- (0-0.5 %)

supplemented with/without AQS.

Thus from above experiment, it becomes clear that AQS prevents toxicity of Cr(VI), causing its

rapid reduction. Furthermore, this process involves reduction of AQS to AQS(H2) (enzymatic,

probably involving quinone reductase?) and subsequent reduction of Cr(VI) by reduced AQS in a

purely redox reaction (non enzymatic).

3.3.19 Chromate reduction in soil microcosm

In soil microcosms containing 100 mg Cr(VI)/kg soil inoculated with Bacillus sp. BT1 (104 CFU/g

soil), up-to ~ 80% Cr(VI) reduction was observed after 15 days of incubation (estimated residual

Cr(VI) was 21.4 ± 1.5 mg Cr(VI)/ kg soil). In un-inoculated soil microcosm containing 100 mg

Cr(VI) / kg soil, negligible Cr(VI) reduction (0.075 %) was observed which can be attributed to

abiotic reduction (Figure 3.21).

The extent of Cr(VI) reduction dropped to 59.3% in microcosm spiked with 300 mg Cr(VI) / kg soil

after 15 days of incubation with Bacillus sp. BT1 (104

cfu/g) (Table 3.11). The viable count of

Bacillus sp. BT1 in microcosms with initial 100 mg and 300 mg Cr(VI) / kg soil increased from 1.7

x 104 cfu/g soil to 7.2 x 10

4 and 4.5 x 10

4 cfu/g soil, respectively during 15 days of incubation

(Table 3.12). The lower rate of bacterial growth in microcosms with initial 300 mg Cr(VI) / kg soil

explains the lower efficiency of Cr(VI) in comparison to soil microcosms with initial 100 mg

Cr(VI)/kg soil (Table 3.12). These results correlate well with our observations on liquid cultures in

LB broth.

REDUCTION OF Cr(VI)

67

Figure 3.21 Profile of residual Cr(VI) upon reduction by Bacillus sp. BT1 in microcosm amended

with 100 mg Cr(VI)/kg soil.

Table 3.11 Profile of Cr(VI) reduction by Bacillus sp. BT1 in soil microcosm amended with two

different concentrations of Cr(VI) (100 and 300 mg Cr(VI)/kg soil).

Time (d) Residual Cr(VI)

(mg/kg soil)

Uninoculated Inoculated Uninoculated Inoculated

0 108.3±3 106.7±1.5 325.33±8.6 329.3±11.01

3 108.9±6.5 92.6±2.8 357.66±13 271.8±26.19

6 113±12 73.16±4.06 332.36±13.9 243.3±7.6

9 104±2.7 48.2±6.4 329.4±3.2 218.1±4.4

12 99.8±1.9 34.16±2.1 319.13±3.27 196.8±5.7

15 99.2±2 33.9±3.2 318.2±3.1 182.8±4.8

REDUCTION OF Cr(VI)

68

Table 3.12 Influence of varying Cr(VI) concentration on growth (cfu/g soil) of Bacillus sp. BT1 in

soil microcosms amended with either 100 or 300 mg Cr(VI)/kg soil.

Time (d) Cr(VI) mg/kg soil

100 300

0 1.7 X 104

1.9 X 104

3 1.7 X 104 1.9 X 10

4

9 3.9 X 104 2.1 X 10

4

12 5.4 X 104 3.3 X 10

4

15 7.2 X 104

4.5 X 104

Since, Cr(VI) reduction in Bacillus sp. BT1 was found to be associated with growth we attempted

to stimulate growth by supplementing soil microcosms with malt extract and thereby enhance the

Cr(VI) reduction efficiency. The malt extract was added to mimic nutrients in plant root exudates

and thus simulating microcosm experiment more closer to field conditions (Ellis, 2004). The

augmentation of soil microcosms with malt extract above 5 g/kg soil stimulated growth of Bacillus

sp. BT1 as well as its Cr(VI) reduction efficiency (Table 3.13).

Table 3.13 Influence of varying malt extract concentration on Cr(VI) reduction by Bacillus sp.

BT1 and its growth (cfu/g soil) in soil microcosms amended with 100 mg Cr(VI)/kg soil.

Time

(d)

Malt Extract concentration in microcosms (%, w/w)

0.05 0.1 0.2 0.5 1.0

Cr(VI),

mg/kg

soil

CFU/g

soil

Cr(VI),

mg/kg

soil

CFU/g

soil

Cr(VI),

mg/kg

soil

CFU/g

soil

Cr(VI),

mg/kg

soil

CFU/g

soil

Cr(VI),

mg/kg

soil

CFU/g

soil

0 102.96 1.0x104

102.9 1.4x104 100 1.2x10

4 104.24 1.7x10

4 101.46 1.4x10

4

5 97.7 3.2x104 95.08 8.5x10

4 90 8.6x10

4 93.97 8.9x10

4 86.16 9.1x10

4

10 79.99 8.5x104 67.7 9.4x10

4 52.7 1.0x10

5 52.57 1.2x10

5 33.54 1.2x10

5

15 40.7 1.3x105 27.9 1.5x10

5 17.46 1.6x10

5 9.01 1.7x10

5 8.05 1.7x10

5

REDUCTION OF Cr(VI)

69

3.3.20 Seed germination assay

The evaluation of Cr(VI) toxicity on germination of the test seeds (mustard and mungbean seeds)

was carried out in petri-plates containing filter papers soaked with Cr(VI) solution (0 to 300 mg

Cr(VI)/L). The increasing Cr(VI) concentration was found to be associated with gradual decrease in

the length of seedling for both mustard and mungbean seeds (Figure 3.22). Mungbean seeds

exhibited 100% germination irrespective of the Cr(VI) concentration. However, negative correlation

was observed between Cr(VI) concentration and post-germinative developments in mungbean

seeds. The inhibition in mungbean seedling length was observed even at the lowest concentration

tested (50 mg Cr(VI)/L). In case of mustard seeds, the increasing concentration of Cr(VI) affected

both seedling length and seed germination, suggesting the higher sensitivity of mustard seeds

towards Cr(VI).

Srivastava and Thakur (2006) demonstrated similar reduction in seedling length and germination

(%) of wheat seeds with increasing Cr(VI) concentration.

Figure 3.22 Effect of Cr(VI) (0 to 300 mg/L) concentration on germination (b) and seedling growth

(a) of mungbean and mustard seeds. Inset in Fig. 7(a) is photograph of mungbean (upper) and

mustard (lower) seedling development after 48 h in presence of increasing concentration of Cr(VI)

(0 to 300 mg/L). (The values are significantly different at p≤0.05).

In order to test the efficiency of Bacillus sp. BT1 treatment to Cr(VI) contaminated microcosms,

seed germination assay was done using aqueous soil extracts obtained from uninoculated soil

microcosm without Cr(VI) (control soil), un-inoculated soil spiked with Cr(VI) (abiotic control) and

soil microcosm (containing 100 mg Cr(VI) / kg soil) treated with 104 cfu of Bacillus sp. BT1/g soil

for 15 days (bioremediated soil).

REDUCTION OF Cr(VI)

70

The length of mungbean seedlings in control and abiotic control (amended with 100 mg Cr(VI)/ kg

soil) soil microcosm extracts was found to be 5.18 and 2.41 cm, respectively. The inhibition in

development of mungbean seedlings observed for aqueous extracts of abiotic control soil

microcosms was found to be alleviated upon treatment of soil microcosm (amended with 100 mg

Cr(VI)/ kg soil) with Bacillus sp. BT1 for 15 days (bioremediated microcosms). Similar

observations were made in case of mustard seed germination assay (Figure 3.23).

Salunkhe et al. (1998) demonstrated that Pseudomonas mendocina could immobilize 100 µg/g soil

by converting it to Cr(III). Furthermore, Cr(VI) contaminated soil microcosm inoculated with

Pseudomonas mendocina supported the growth of wheat seedlings without exerting any toxic

effects.

Figure 3.23 Growth of mungbean and mustard seedlings during 48 h of incubation in presence of

aqueous extract of control microcosms (without Cr(VI)); uninoculated abiotic control microcosms

(amended with 100 mg Cr(VI)/kg soil) and microcosms (amended with 100 mg Cr(VI)/kg soil)

bioremediated by Bacillus sp. BT1. (The values are significantly different at p≤0.05).

REDUCTION OF Cr(VI)

71

3.3 Summary

The physiological characterization of Cr(VI) reduction by O. intermedium and Bacillus sp. BT1

indicated that both the bacterial strains could reduce 100 mg Cr(VI)/L completely in Luria Bertani

broth at optimum pH 7.0 and temperature 30 to 37 ºC. The Cr(VI) reduction by both O.

intermedium and Bacillus sp. BT1 remained either un-affected or slightly affected in presence of

co-contaminant metal ions.

The efficiency and rate of Cr(VI) reduction of both the bacterial strains was enhanced in presence of

AQS. The higher efficiency of Cr(VI) reduction in presence of AQS was accompanied with higher

intracellular levels of quinone and chromate reductase.

Hence, a hypothetical model for AQS mediated two-step Cr(VI) reduction by O. intermedium

BCR400 and Bacillus sp. BT1 has been proposed. In the first-step, AQS may be reduced to AQH2S

(hydra-quinone) by quinone reductase. In the second-step, the AQH2S reduces Cr(VI) by 2e-

transfer, thereby reducing it to Cr(IV) in a purely redox chemical reaction. The direct reduction of

Cr(VI) to Cr(IV) by AQH2S would also prevent ROS generation, thereby exert protective effect on

the growth of O. intermedium BCR400. The further reduction of Cr(IV) to Cr(III) may be mediated

by single electron transfer catalysed by specific chromate reductase or non-specific reductases. The

model is based on several lines of evidences:

1. The supplementation of AQS to nutrient growth medium amended with Cr(VI), not only had

protective effect against toxicity due to Cr(VI) but also enhanced the rate of Cr(VI)

reduction by both O. intermedium BCR400 and Bacillus sp. BT1.

2. Increase in both quinone and chromate reductase activity (U/mg of total proteins) was

observed in cell free lysate prepared from O. intermedium BCR400 and Bacillus sp. BT1

grown in presence of either AQS or Cr(VI)+AQS.

3. The cell-free extract (CFE) of both the bacterial strains grown in presence of variable AQS

concentration (0 to 1mM) displayed high, non linear positive correlation between the

concentration of AQS and activities of both quinone reductase and chromate reductase (γ >

+ 90).

Finally, the soil microcosm studies with Bacillus sp. BT1 demonstrated its potential application in

bioremediation of Cr(VI) contaminated soils

REDUCTION OF Cr(VI)

72

Reference

Ackerley Df., Gonzalez CF., Keyhan M., Blake R. and Matin A. 2004a. Mechanism of chromate

reduction by the Escherichia coli, NfsA, and the role of different chromate reductases in

minimizing oxidative stress during chromate reduction. Environmental Microbiology 6:851-

860.

Ackerley DF., Gonzalez CF., Park CH., Blake R., Keyhan M. and Matin A. 2004b. Chromate

reducing properties of soluble flavoproteins from Pseudomonas putida and Escherichia coli.

Applied and Environmental Microbiology. 70:873-882.

Agrawal A., Kumar V. and Pandey BD. 2008. Remediation options for the treatment of

electroplating and leather tanning effluent containing chromium-A Review. Mineral

Processing Extractive Metallurgy Reviews. 272:99-130.

Bond DR. and Lovley DR. 2002. Reduction of Fe(III) oxide by methanogens in the presence and

absence of extracellular quinones. Environmental Microbiology. 4:115-124.

Branco R., Alpoim MC. and Morais PV. 2004. Ochrobactrum tritici strain 5bvI1-characterization of

a Cr(VI) resistant and Cr(VI) reducing strain. Canadian Journal of Microbiology. 50:697-

703.

Burgos W., Fang Y., Royer RA., Yeh GT., Stone JJ., Jeonn BH. and Dempsey BA. 2003. Reaction

based modelling of quinone mediated bacterial iron(III) reduction. Geochemica et

Cosmochimica Acta. 67:2735-2748.

Camargo FAO., Okeke BC., Bento FM. and Frankenberger WT. 2004. Hexavalent chromium

reduction by immobilized cells and cell free-extracts of Bacillus sp. ES 29. Bioremediation

Journal. 8:23-30.

Camargo FAO., Okeke BC., Bento FM and Frankenberger WT. 2003. In-vitro reduction of

hexavalent chromium by a cell-free extract of Bacillus sp. ES 29 stimulated by Cu2+

.

Applied Microbiology and Biotechnology. 62:569-573.

Campos J., Martinez PM. and Cerventes C. 1995. Hexavalent chromium reduction by chromate

resistant Bacillus sp. strain. Antonie Van Leeuwenhoek. 68:203-208.

Cheung KH. and Gu JD. 2005. Chromate reduction by Bacillus megaterium TKW3 isolated from

marine sediments. World Journal of Microbiology and Biotechnology. 16:855-862.

Clark DP. 1994. Chromate reductase activity of Enterobacter aerogenes is induced by nitrite.

FEMS Microbiology Letters. 122:233-238.

REDUCTION OF Cr(VI)

73

Ellis RJ. 2004. Artificial soil microcosms: a tool for studying microbial autecology under controlled

conditions. Journal of Microbiological Methods. 56:287-290.

Elongovan R., Abhipsa S., Rohit B., Ligy P. and Chandraraj K. 2006. Reduction of Cr(VI) by

Bacillus sp. Biotechnology Letters. 28:247-252.

Elongovan R., Philip L. and Chandraraj K. 2008. Hexavalent chromium reduction by free and

immobilized cell-free extract of Arthrobacter rhombi-RE. Applied Biochemistry and

Biotechnology. 160:81-97.

Fredrickson JK., Kostandarithes HM., Li SW., Plymale AE. and Daly MJ. 2006. Reduction of

Fe(III), Cr(VI), U(VI) and Tc(VII) by Deinococcus radiodurans R1. Applied and

Environmental Microbiology. 66:2006-2011.

Ganguli A. and Tripathi AK. 2002. Bioremediation of toxic chromium from electroplating effluent

by chromate reducing Pseudomonas aeroginosa A2Chr in two bioreactors. Applied

Microbiology and Biotechnology. 58:416-420.

Garbisu C., Alkorta I., Llama MJ. and Serra JL. 1998. Aerobic chromate reduction by Bacillus

subtilis. Biodegradation. 9:133-141.

Gonzalez CF., Ackerley DF., Lynch SV. and Matin A. 2005. ChrR, a soluble quinone reductase of

Pseudomonas putida that defends against H2O2. The Journal of Biological Chemistry.

280:22590-22595.

Gonzalez CF., Ackerley DF., Park CH. and Matin A. 2003. A soluble flavoprotein contributes to

chromate reduction and tolerance by Pseudomonas putida. Acta Biotechnologica. 23:233-

239.

He Z., Gao F., Sha T., Hu Y. and He C. 2009. Isolation and characterization of a Cr(VI)-reduction

Ochrobactrum sp. strain CSCr-3 from chromium landfill. Journal of Hazardous Materials.

163:869-873.

Hussein H., Farag S., Kandil K. and Moawad H. 2005. Tolerance and uptake of heavy metals by

Pseudomonads. Process Biochemistry. 40:955-961.

Ishibashi Y., Cerventes C. and Silver S. 1990. Chromium reduction in Pseudomonas putida.

Applied and Environmental Microbiology. 6:2268-2270.

Kwak YH., Lee DS. and Kim HB. 2003. Vibrio harveyi nitroreductase is also a chromate reductase.

Applied and Environmental Microbiology. 69:4390-4395.

REDUCTION OF Cr(VI)

74

Ling J., Hong LU., Jiti Z. and Jing W. 2009. Quinone mediated decolourization of sulfonated azo

dyes by cells and cell extracts from Sphingomonas xenophaga. Journal of Environmental

Sciences. 21:503-508.

Liu YG., Xu WH., Zeng GM., Li X. and Gao H. 2006. Cr(VI) reduction by Bacillus sp. isolated

from chromium landfill. Process Biochemistry. 41: 1981-1986.

Liu G., Yang H., Wang J., Jin R., Zhou J. and Lv H. 2010. Enhanced chromate reduction by resting

Escherichia coli cells in the presence of quinone redox mediators. Bioresource Technology.

101:8127-8131.

Liu G., Zhou J., Wang J., Zhou M., Lu H. and Jin R. 2009. Acceleration of azo dye decolorization

by using quinone reductase activity of azoreductase and quinone redox mediator.

Bioresource Technology. 100:2781-2795.

Lovley DR., Fraga JL., Blunt-Harris EL., Hayes LA., Phillips EJP. and Coates JD. 1998. Humic

substance as a mediator for microbially catalyzed metal reduction. Acta Hydrochimica Et

Hydrobiologica. 26:152-157.

Lovley DR., Fraga JL., Coates JD. and Blunt-Harris EL. 1999. Humics as an electron donor for

anaerobic respiration. Environmental Microbiology. 1:89-98.

Lowe KL., Straube W., Little B. and Meehan J. 2003. Aerobic and anaerobic reduction of Cr(VI) by

Shewanella oneidensis: Effects of cationic metals, sorbing agents and mixed microbial

cultures. Acta Biotechnologica. 23:161-178.

Lowry OH., Rosebroough NJ., Farr AL. and Randall RJ. 1951. Protein measurements with folin

phenol reagents. The Journal of Biological Chemistry. 193:265-275.

Maithreepala RA. and Doong R. 2009. Transformation of carbon tetrachloride by biogenic iron

species in the presence of Geobacter sulfurreducens and electron shuttles. Journal of

Hazardous Materials. 164:337-344.

Mazoch J., Tesar R., Sedlacek V., Kucera I. and Turanek J. 2004. Isolation and biochemical

characterization of two soluble iron (III) reductases from Paracoccus denitrificans.

European Journal of Biochemistry. 271:553-562.

McLean J. and Beveridge TJ. 2001. Chromate reduction by a Pseudomonad isolated from a site

contaminated with chromate copper arsenate. Applied and Environmental Microbiology.

67:1076-1084.

Megharaj M., Avudainayagam S. and Naidu R. 2003. Toxicity of hexavalent chromium and its

reduction by bacteria isolated from soil contaminated with tannery waste. Current

Microbiology. 27:51-54.

REDUCTION OF Cr(VI)

75

Michel C., Brugna M., Aubert C., Bernadac A. and Brushi M. 2001. Enzymatic reduction of

chromate: comparative studies using sulphate reducing bacteria. Key role of polyheme

cytochrome c and hydrogenases. Applied Microbiology and Biotechnology. 55:95-100.

Ozdemir G., Ozturk T., Ceyhan N., Isler R. and Cosar T. 2003. Heavy metal biosorption by biomass

of Ochrobactrum anthropi producing exopolysaccharide in activated sludge. Bioresource

Technology. 90:71-74.

Park CH., Gonzalez CF., Ackerley DF., Keyhan M. and Matin A. 2002. Molecular engineering of

soluble bacterial proteins with chromate reductase activity. In Proceedings of the 1st

International Conference on Remediation of Contaminated Sediments, Venice, Italy, Vol.

III. Hinchee, R.E.,Porta, A., and Pellei M. (eds). Columbus, OH: Batelle Press, pp. 103–111.

Puzon GJ., Petersen JN., Roberts AG., Kramer DM. and Xun L. 2002. A bacterial flavin reductase

system reduces chromate to a soluble chromium (III)-NAD+ complex. Biochemical and

Biophysical Research Communications. 294:76-81.

Rau J., Knackmuss HJ. and Stolz A. 2002. Effects of different quinoid redox mediators on the

anaerobic reduction of azo dyes by bacteria. Environmental Science and Technology.

36:1497-1504.

Rau J. and Stolz A. 2003. Oxygen- insensitive nitroreductases NfsA and NfsB of Escherichia coli

function under anaerobic conditions as lawsone dependent azo reductases. Applied and

Environmental Microbiology. 69:3448-3455.

Rehman A., Zahoor A., Muneer B. and Hasnain S. 2008. Chromium tolerance and reduction

potential of a Bacillus sp. ev3 isolated from metal contaminated wastewater. Bulletin of

Environmental Contamination and Toxicology. 81:25-29.

Romanenko VI. and Korenkov VN. 1977. A pure culture of bacterial cells assimilating chromates

and bichromates as hydrogen acceptors when grown under anaerobic conditions.

Mikrobiologiya. 46:414-417.

Salunkhe PB., Dhakephalkar PK. and Paknikar KM. 1998. Bioremediation of hexavalent chromium

in soil microcosms. Biotechnology Letters. 20:749-751.

Santos ABD., Bisschops IAE., Cerventes FJ. and Lier JBV. 2004a. Effect of different redox

mediators during thermophillic azo dye reduction by anaerobic granular sludge and

comparative study between mesophillic (30 ºC) and thermophillic (55 ºC) treatments for

decolourization of textile wastewaters. Chemosphere. 55:1149-1157.

Santos ABD., Cervantes FJ. and Lier JB. 2004b. Azo dye reduction by thermophilic anaerobic

granular sludge and the impact of the redox mediator anthraquinone - 2, 6 - disulfonate

REDUCTION OF Cr(VI)

76

(AQDS) on the reductive biochemical transformation. Applied Microbiology and

Biotechnology. 64:62-69.

Sarangi A. and Krishnan C. 2008. Comparison of in vitro Cr(VI) reduction by CFEs of chromate

resistant bacteria isolated from chromate contaminated soil. Bioresource Technology.

99:4130-4137.

Sedlacek V. and Kucera I. 2010. Chromate reductase activity of the Paracoccus denitrificans ferric

reductase B (FerB) protein and its physiological relevance. Archives of Microbiology.

192:919-926.

Sendelbach LE. 1989. A review of the toxicology and carcinogenicity of anthraquinone derivatives.

Toxicology. 57:227-240.

Shen H. and Wang YT. 1993. Characterization of enzymatic reduction of hexavalent chromium by

Escherichia coli ATCC 33456. Applied and Environmental Microbiology. 59:3771-3777.

Srivastava S. and Thakur IS. 2006. Evaluation of bioremediation and detoxification potentiality of

Aspergillus niger for removal of hexavalent chromium in soil microcosm. Soil Biology

Biochemistry. 38: 1904-1911.

Sultan S. and Hasnain S. 2006. Characterization of an Ochrobactrum intermedium strain STCr-5

manifesting high level Cr(VI) resistance and reduction potential. Enzyme and Microbial

Technology. 39:833-888.

Sultan S. and Hasnain S. 2007. Reduction of toxic hexavalent chromium by Ochrobactrum

intermedium strain SDCr - 5 stimulated by heavy metals. Bioresource Technology. 98:340-

344.

Suzuki T., Miyata N., Horitsu H., Kawai K., Takamizawa K., Tai Y. and Okazaki M. 1992.

NAD(P)H-dependent chromium(VI) reductase of Pseudomonas ambigua G-1: a Cr(V)

intermediate is formed during the reduction of Cr(VI) to Cr(III). Journal of Bacteriology.

174:5340-5345.

Thacker U. and Madamwar D. 2005. Reduction of toxic chromium and partial localization of

chromium reductase activity in bacterial isolate DM1. World Journal of Microbiology and

Biotechnology. 21:891-899.

Timmis KN. and Pieper DH. 1999. Bacteria designed for bioremediation. Trends in Biotechnology.

17:200-204.

Wang PC., Mori T., Komoril K., Sasatsu M., Toda K. and Ohtake H. 1989. Isolation and

characterization of an Enterobacter cloacae strain that reduces hexavalent chromium under

anaerobic conditions. Applied and Environment Microbiology. 55:1665-1669.

REDUCTION OF Cr(VI)

77

Wang PC., Mori T., Toda K. and Ohtake H. 1990. Membrane-associated chromate reductase

activity from Enterobacter cloacae. Journal of Bacteriology. 172:1670-1672.

Wang Y., Wu C., Wang X. and Zhou S. 2009. The role of humic substances in the anaerobic

reductive dechlorination of 2,4-dichlorophenoxyacetic acid by Comamonas korensis strain

CY01. Journal of Hazardous Materials. 164:941-947.

Whiteway J., Koziarz P., Veall J., Sandhu N., Kumar P. and Hoecher B. 1998. Oxygen-insensitive

nitroreductases: analysis of the roles of nfsA and nfsB in development of resistance to 5-

nitrofuran derivatives in Escherichia coli. Journal of Bacteriology. 180:5529-5539.

Zee FPVD. and Cervantes FJ. 2009. Impact and application of electron shuttles on the redox (bio)

transformation of contaminants: A Review. Biotechnology Advances. 27:256-277.