81
MICROBIAL TECHNOLOGY FOR BIOREMEDIATION OF ORGANOPHOSPHORUS PESTICIDES By KAROLIN K.P 2012-11-165 SEMINAR REPORT Submitted in partial fulfillment of the requirement of the course MICRO 591 1

MICROBIAL TECHNOLOGY FOR BIOREMEDIATION

Embed Size (px)

Citation preview

MICROBIAL TECHNOLOGY FOR BIOREMEDIATION OF ORGANOPHOSPHORUS PESTICIDES

ByKAROLIN K.P2012-11-165

SEMINAR REPORTSubmitted in partial fulfillment of the requirement of the

course MICRO 591

1

Department of Agricultural MicrobiologyCollege of Agriculture

VellayaniThiruvanathapuram

2014

DECLARATION

I hereby declare that this seminar report entitled

“Microbial Technology for Bioremediation of

Organophosphorus Pesticides” is a bonafide record of work

done by me during the course of MSc Ag. Programme.

Place: Date:

Karolin K.P

2012-11-165

2

CERTIFICATE

I hereby certify that this seminar report entitled

“Microbial Technology for Bioremediation of

Organophosphorus Pesticides” has been prepared by Ms.

KAROLIN K.P. (2012-11-165) for the partial fulfillment of

the M.Sc Ag. Programme

Place: Dr. K.S.MeenakumariDate: (Chairman, Advisory committee) Professor and HOD Department of Agricultural Microbiology. College of Agriculture, vellayani

3

4

SI NO. CONTENTS PAGE NUMBER

1 INTRODUCTION 5

2 ORGNOPHOSHORUS PESTICIDES 5

3 BIOREMEDIATION 10

4 CHLORPYRIFOS 17

5 DEGRADATION OF OP’s BY

MICROBIAL CONSORTIA

25

6 ENZYMATIC DEGRADATION OF

ORGANOPHOSPHORUS PESTICIDES

27

7 GENETIC BASIS OF OP DEGRADATION 29

8 SIGNIFICANCE OF MICROBIAL

TECHNOLOGY FOR BIOREMEDIATION

OF OP PESTICIDES

31

9 DISCUSSION 33

10 REFERENCES 34

11 ABSTRACT 46

5

INTRODUCTION

Pesticides constitute the key control strategy for crop

pest and disease management. Continuous application of

pesticides to the soil and aquatic system resulted in

health hazards and environmental pollution which has

triggered much public concern. The wide spread use of these

pesticides over the years has resulted in problems caused

by their interaction with the biological systems in the

environment. Pesticides will continue to be an

indispensable tool for the management of pests in the years

to come, as there is no suitable alternative to replace

them totally.

1. Organophosphorus pesticides

Organophoshorus compounds are the most widely used

insecticides, accounting for 34% of world-wide insecticide

sales. These compounds possess high mammalian toxicity and

it is therefore essential to remove them from the

environment. Considering the toxic effect of these

6

pesticides it is essential to remove them from the

environment employing suitable remedial measures. The first

commercialized OP insecticide Bladan, containing TEPP

(tetraethyl pyrophosphate) was synthesized by German

chemist Gerhard Schrader, who later synthesized insecticide

parathion in 1944 (Gallo and Lawryk 1991). Thus, their

development as insecticides took place in the late 1930s

and early1940s. Bioremediation exploiting microbial

technology is one of the recent techniques for

environmental clean-up. In this process heterotrophic

microorganisms breakdown hazardous compounds to obtain

carbon and energy. The first micro-organism that could

degrade organophoshorus compounds was isolated in 1973

(Dragun et al., 1984) and identified as Flavobacterium sp.

Since then several bacterial and a few fungal species have

been isolated which can degrade a wide range of

organophoshorus compounds in liquid cultures and soil

systems. The wide varieties of microorganisms have been

found to possess the enzyme catalyzing hydrolysis of7

organophoshates. This enhances the feasibility of using

bioremediation to treat organophosphate compounds. Moreover

application of molecular biology to genetically engineer

microorganism contain appropriate genes can potentially

contribute to making bioremediation more efficient.

1.1 Structure of organophosphorus pesticides

The general structure of OP pesticide is represented as:

R1 and R2 are the aryl or alkyl groups which either

directly bonded to phosphorous (P) as in case of

phosphinates or through oxygen (phosphates) or sulphur

(phosphothioates) atoms.

In phosphonates R1 is directly bonded to P but R2 is

bonded to oxygen. In phosphonothioates R1 is directly

bonded to P but R2 is bonded to sulphur atom.

In phosphoramidates both or atleast one R group is

attached to –NH2.

X is the ‘‘leaving group’’ (as it gets displaced upon

hydrolysis of OP).

8

Oxygen and sulphur (O/S) are directly attached to P by

a double bond

1.2 Use of Organophosphorus pesticides

Organophosphate pesticides are widely used in variety

of cereal crops, oil seed crops, vegetable crops and fruit

crops for effective control of insect such as termites,

jassids, aphids, white fly, leaf hoppers etc. They are also

used in some veterinary and human medicines to remove

parasitic insects, in various public hygiene products

generally for the control of cockroaches and termites

(Racke et al. 1994). Among them chlorpyrifos (O, O-diethyl-

O-3,5,6 trichloro-2-pyridylphosphorothioate) is being

extensively used in developing countries like India where

it was the fourth highest consumed pesticide after

monocrotophos, acephate and endosulfan, in the year 2000

(Ansaruddin, 2003).

1.3 Bioaccumulation of organophosphorus pesticides

The bioaccumulation is the primary cause of toxicity

of pesticides as their higher concentrations in biological9

systems leads health problems. The persistence of OP in

soil has also been related to the organic matter, clay

content and iron and/or aluminiumoxy(hydro)oxide content of

the soil (Weber, 1972). These have higher affinity to

absorb/adsorb the pesticides and act as a sink for such

hydrophobic compounds and discharge into other phases

included living organisms and plants (Boonsaner et al.

2002). Parathion, an organophosphate pesticide, has been

found to persist in soil for more than 16 year (Stewart et

al. 1971). The bioaccumulation of OPs is primarily related

to their molecular weight and aqueous phase solubility.

There are reports regarding their bioaccumulation in

living systems ranging from blue green algae to higher

systems. Chlorpyrifos due to their very low aqueous phase

solubility (2ppm) had the highest bioaccumulation factor in

the blue-green algae, aquatic plants, gold fish, and

mosquito fish and Mytilus galloprovincialis, a

Mediterranean mussel (Spacie and Hamelink, 1985; Lal et al.

1987). Thus, being at the top of the food chain humans10

indirectly gets much adversely affected by bioaccumulation

of chlorpyrifos and other pollutants in aquatic fauna

(Serrano et al.1997).

Chlorpyrifos has least water solubility and presence

of three chlorine atoms attached to its ring makes it more

resistant to microbial degradation. Due to these structural

features the limiting rate of degradation results in its

accumulation in environment (Volkering et al. 1998;

Angelova and Schamander 1999)

1.4 Fate of Organophosphate pesticides

Pesticides undergo various changes in environment

including their adsorption transmission and degradation

depending on the physicochemical nature of the pesticide

and the soil (Redondo et al. 1997). The predominant

processes involved in transformation of such molecules is

mediated by microbes (Vink and Van der Zee, 1997) followed

by photolysis or photodegradation and chemical

transformations (Roberts, 1998; Stangroom et al. 2000).

11

Thus, generally fate of pesticide involves both biological

and non-biological agents.

1.5 Transformation of OP pesticides

1.5.1 Abiotic transformation of OP pesticides

The abiotic processes involved either their transport

where parent compound remains unchanged and simply

transferred from one matrix to another depends on the

physicochemical properties of pesticides itself (Stangroom

et al. 2000) e.g. volatilization, leaching (Laabs et al.

2000), runoff (Moore et al. 2002), absorption and

adsorption of pesticides (Yu et al. 2006) or by abiotic

transformations by photodegradation (Walia et al. 1988) and

chemical hydrolysis (Liu et al. 2001). Volatilization can

transform pesticides from liquid or solid into gaseous form

depending on temperature of the surface and air currents in

that area (Yates et al. 2002; Haith et al. 2002). The run

off/soil erosion causes movement of pesticides either on

their dissolution in the water or through attachment with

soil particles or sediments and add a fairly large part12

(e.g., 5–30% for chlorpyrifos) of the overall mass transfer

from ground soil to surface water aquifers (Bailey et al.

1974; Richards and Baker, 1993; Wood and Stark, 2002; Luo

and Zhang 2009). Leaching also plays role in transmission

of higher water soluble pesticides to ground water through

fractures, root holes and earthworm burrows in earth crust

(Stagnitti et al. 1994; Magri and Haith, 2009). Pesticides

can also be absorbed/adsorption by plants and soils depend

on their water solubility, soil characteristics and

properties of pesticide itself (Gan et al. 1996; Trapp,

2000; Davis et al. 2002; Paranychianakis et al. 2006;

Johnson et al. 2007).

1.6 Biotic transformation of OP pesticides

Xenobiotic compounds like organophosphate pesticides

are man made compounds and were not previously present in

nature. Consequently the natural microflora does not have

potential to metabolize these pesticides due to lack of

enzyme and proper transport processes. But over the year

due to excessive use of xenobiotic compounds microbes have13

evolved the new degradation pathways resulting in

accelerated degradation of such compounds (Seffernick and

Wackett, 2001; Johnson and Spain, 2003). For instance

Seffernick et al. in 2001 suggests that enzyme Atrazine

chlorohydrolase (AtzA) responsible for dechlorination of

herbicide atrazine under strong selective pressure was

evolved from enzyme melamine deaminase (TriA) and is 98%

similar to the TriA. The accelerated bioremediation under

natural conditions is also helped by transfer of genes

among different microbial cultures by transformation,

transduction and conjugation (Ghigo 2001 and Fux 2005).

2. Bioremediation

Bioremediation techniques for treatment of hazardous

waste have attracted attention over physical-chemical

processes because of their environmentally sound and cost

effective characteristics. Bioremediation is a clean-up of

pollution from soil, ground water, surface water and air,

using biological, usually microbiological processes (Philip

et al., 2001). Biological processes generally produce14

benign end-products, unlike conventional physical-chemical

processes which often create other potential pollution

hazards, such as emission of toxic byproducts from

incineration and leaching of toxic substances from land

disposal. By properly establishing and maintaining

microbial communities and environmental conditions

necessary for biodegradation, most hazardous organic

compounds can be biologically detoxified or mineralized.

A waste is considered as hazardous when it poses a

substantial present or potential danger to human health or

the environment regardless of the form of the waste. In the

United States, a waste is called hazardous when it has any

of the characteristics of corrosivity, ignitability,

reactivity, or toxicity. Nowadays, sources of hazardous

wastes are not limited to large chemical companies but

include a wide range of industrial, agricultural, and

household activities. Amounts produced by various sources

vary widely.

15

The publication of Silent Spring by Rachel Carson in

1962, which emphasized the ecological consequences of

pesticide contamination that were widely ignored, began a

great public debate on pesticide usage. She reported the

hazards of the pesticide DDT (dichloro-diphenyl-trichloro-

ethane), the most powerful pesticide in world history. DDT,

an organochlorine, was originally synthesized in 1874, but

it was in 1939 that DDT was discovered to have insecticide

properties by a Swiss scientist, Paul Müller, who was later

awarded the Nobel Prize. In 1945, DDT was released for

civilian use to become a miracle compound with its

capability of killing hundreds of different kinds of

insects at once. However, extensive usage of DDT created

problems in the environment and in humans because of its

toxicity and its bioaccumulation in fatty tissue. In 1970,

Norway and Sweden banned DDT, and in 1972 the US followed

suit. However, DDT is still used in some countries.

2.1 Factors affecting degradation of OPs

16

The most important parameters involved in pesticide

biodegradation are pesticide concentration, inoculum size,

pH, temperature and its bioavailability, (Karpouzas and

Walker, 2000; Singh et al. 2006; Getzin, 1981; Chapman and

Chapman, 1986; Singh et al., 2002).

2.2.1 Effect of substrate concentration

The chlorpyrifos concentration found in the upper 10

mm of the soil sediment after its application was observed

to be the highest (Brock et al. 1992). Cink and Coats

(1993) observed that after the use of agricultural

application rate of 10 mg kg-1 of chlorpyrifos 5% of

chlorpyrifos persisted in the soil. However, at termite

infestation sites where chlorpyrifos might be used at

higher level almost 57% of appended chlorpyrifos remained

in soil. The ability of microbes to degrade a pollutant

depends on the available concentration of polluting

chemicals, as high concentrations are usually toxic for

microbial degraders and low concentrations may not be able

to induce the enzymes involved in degradation (Block et al.17

1993; Morra, 1996). Menon et al. (2005) reported delayed

dehydrogenase activity in the soil after chlorpyrifos

application at 0.20 μg g-1 indicated inhibition of

microbial growth at the polluted site. A similar

observation was reported by Pandey and Singh (2004) where a

dose of 4 L/hm2 chlorpyrifos showed a short-term inhibitory

effect on the total microbial population. Shan et al.

(2006) also indicated that the application of chlorpyrifos

lead to decrease in bacterial, fungal and actinomycete

populations with increasing chlorpyrifos concentration (2,

4, and 10 mg kg-1) in the soil. Hua et al. (2009) reported

that soil ammended with chlorpyrifos at the initial level

of 4, 8, and 12 mg kg-1, the chlorpyrifos was degraded

after 35 days with average half-live of 14.3, 16.7, and

18.0 respectively. The initial inhibition of soil microbial

communities was followed by recolonization of soil

microbial communities after two weeks.

2.2.2 Effect of inoculum size

18

Ramadan et al. 1990 observed that at low inoculum

levels, <104 cells g-1 of soil, significantly lower number

of inoculated cells could survive the initial exposure to

effectively carry out pesticide degradation. Comeau et al.

(1993) suggested that for the bioremediation of pesticide

contaminated sites the inoculum level in the range of 106–

108 Cells g-1 of soil was adequate. A higher initial

inoculum level can counterbalance the initial decline in

viable cell number due to various environmental factors and

toxic substrate concentrations, so that the survivors can

multiply and degrade the pesticides (Comeau et al. 1993;

Duquenne et al. 1996). Rousseaux et al. (2003) also

reported that inoculum size of Chelatobacter heintzii Cit 1

is one of the important factors for efficient

biodegradation of applied atrazine. Singh et al. 2006

observed that inoculum size has a noticible effect on

biodegradation of fenamiphos and chlorpyrifos. Fenamiphos

was not degraded when soils was inoculated with less than

105 cells g-1. Similarly, degradation of chlorpyrifos by19

Enterobacter sp. below an inoculum density 103 cells g-1

was not observed

2.2.3 Effect of pH

Many organophosphorus insecticides are succeptable to

enhanced basecatalyzed hydrolysis at pH values above 7.5

(Greenhalgh et al. 1980). The change of rate of hydrolysis

is about a factor of 10 for each pH unit (Schwarzen bach et

al. 1993). Singh et al. 2003 reported that chlorpyrifos

degradation in the acidic soils with pH 4.7 was slow where

the half-life of chlorpyrifos was 256 days. The

chlorpyrifos degradation was rapid in alkaline soils (pH

7.7 and 8.4) with a half-life of 16 days and leads to

formation of TCP up to 10 mg kg-1. Initially, it was

assumed that the high rate of chlorpyrifos degradation in

alkaline soils was due to chemical hydrolysis. In general,

higher the pH higher is the rate of hydrolysis of OP

pesticides) which may be due to higher copy numbers of opd

(organophosphate degrading) gene (Sparks, 1989; Singh et

al. 2003; Singh et al. 2003). Wang et al. (2006) had also20

reported that chlorpyrifos degradation rate by B.

laterosporus DSP was increased with increase in pH from 7.0

to 9.0. Al-Qurainy and Abdel-Megeed 2009 observed the

effect of pH on two OP pesticides malathion and dimethoate.

They reported complete degradation of malathion and

dimethoate by Pseudomonas frederiksbergensis at pH 7.0

after 6 days of incubation with half lives accounted by 3

and 2.3 days respectively. However, when the medium pH was

set at 8.0, biodegradation began on the first day and

complete degradation was observed after 3 days. Wang et al.

(2005) reported that the biodegradation rates of

chlorpyrifos in the pH range 6.5–9.0 by Fusarium LK. ex Fx.

WZ-I were higher.

2.2.4 Effect of solubility/bioavailability

Many researchers have reported that high organic

matter content of the soil leads to an absorption of

pesticide to soil particles resulting in lower

bioavailability of organophosphorus pesticides and hence

decreases their degradation rate (Barriuso et al. 1992;21

Weber and Huang, 1996; Karpouzas and Walker, 2000; Nelson

et al. 2000; Ben-Hur et al. 2003). Knuth and Heinis (1992)

and Brock et al. (1992) also reported high absorption of

chlorpyrifos to sediments in static aquatic systems

sediments due to the hydrophobic nature of aquatic

sediments and lower solubility of chlorpyrifos in aqueous

phase. Similarly, Civilini, 1994 reported that it is easy

to remove more water soluble lighter hydrophilic compounds

than heavier hydro-phobic PAHs. In general pesticides

having low water solubility are less susceptible to

accelerated degradation due to their limited dissolution

rates (Alexander, 1999). The fate and behavior of pesticide

in the environment is determined by its solubility, half-

life and partitioning coefficients (Neitsch et al. 2005).

The hydrophobic compounds become non-available to microbial

degraders as these compounds get entrapped in nanopores of

the solid phase of organic matter (Arbeli and Fuentes

2007). However, many researchers used chemical surfactant

or biosurfactants produced by microorganisms to desorb22

chemical compounds from soil organic matter so as to make

them bioavailable for their consumption (Aronstein et al.

1991; Brown and Jaffe, 2006; Zhou and Zhu, 2008).

2.2 Microbial degradation/transformations of

Organophosphorus pesticides

Use of organochlorine pesticides such as dichloro-

diphenyl-trichloroethane (DDT), lindane, etc., has been

reduced drastically in developed countries due to their

long persistence, tendency towards bioaccumulation and

potential toxicity towards non-target organisms. This group

of compounds has been replaced by the less persistent and

more effective organophosphorus compounds. However, most of

the organophosphorus compounds possess high mammalian

toxicity. Among the organophosphorus compounds, glyphosate,

chlorpyrifos, parathion, methyl parathion, diazinon,

coumaphos, monocrotophos, fenamiphos and phorate have been

used extensively and their efficacy and environmental fate

have been studied in detail. The phosphorus is usually

present either as a phosphate ester or as a phosphonate.23

Being esters they have many sites which are vulnerable to

hydrolysis. The principal reactions involved are

hydrolysis, oxidation, alkylation and dealkylation (Singh et

al., 1999). Microbial degradation through hydrolysis of P-O-

alkyl and P-O-aryl bonds is considered the most significant

step in detoxification. Both co-metabolic and bio-

mineralization of organophosphorus compounds by isolated

bacteria have been reported.

Hydrolysis of organophosphorus compounds leads to a

reduction in their mammalian toxicity by several orders of

magnitude. Since most of the research has been directed

towards detoxification, studies on the further metabolism

of the phosphorus containing products have not been

extensive. Hypothetical phospho-ester hydrolysis steps can

be postulated, yielding mono-ester and finally inorganic

phosphate, but this pathway has not been specifically

studied. Analogous phospho-monoesterase and diesterase,

which degrade methyl and dimethyl phosphate, respectively,

have been reported in Klebsiella aerogenes (Wolfenden & Spence,24

1967) and are produced only in the absence of inorganic

phosphate from the growth medium. The final enzyme in the

postulated degradative pathway is bacterial alkaline

phosphatase, which can hydrolyze simple monoalkyl

phosphates and is also regulated by the level of phosphate

available to the cell (Wolfenden & Spence, 1967). A similar

mechanism of metabolism has been reported for phosphonates

(Kertesz et al., 1994a). The way in which metabolism is

regulated depends very strongly on what role the

organophosphorus compound plays for the particular

organisms studied. Most often these compounds are used to

supply only a single element (carbon, phosphorus or sulfur)

and the relevant gene cannot be expressed as a response to

starvation for another of these elements (Kertesz et al.,

1994a). For example, a strain of Pseudomonas stutzeri isolated

to utilize parathion as a carbon source released the

diethylphosphorothioanate products quantitatively and could

not metabolize them further, even when alternative source

of phosphorus or sulfur were removed (Daughton & Hsieh,25

1977). Similarly, a variety of isolates that could use

phosphorothionate and phosphorodithionate pesticides as a

sole source of phosphorus were unable to utilize these

compounds as a source of carbon (Rosenberg & Alexander,

1979). Shelton (1988) isolated a consortium that could use

diethylthiophosphoric acid as a carbon source but was

unable to utilize it as a source of phosphorus or

sulfur. Kertesz et al. (1994a) explained possible underlying

reasons for this phenomenon. They suggested that the

conditions under which environmental isolates enriched were

crucial in selecting for strains not only with the desired

degradative enzyme systems but also with specific

regulation mechanisms for the degradation pathways.

3. Chlorpyrifos

Currently among the various groups of pesticides,

organophosphates form the major, accounting for more than

36 percent of the total world market (Kanekar et al.,

2004).Among the insecticides, monocrotophos, quinalphos and

chlorpyrifos top the list of organophosphorus insecticides26

in Indian market. In India the estimated consumption of

technical grade chlorpyrifos during 2002-03 was 5000 MT

(Singhal, 2003). In the light of restricted or banned use

of organochlorine compounds, chlorpyrifos is gaining

importance in Agriculture. It is a broad spectrum pesticide

displaying insecticidal activity against a wide range of

insects and pests. The most commonly available formulations

include emulsifiable concentrates (EC), granules (GR) and

wettable powders (WP). Aerial application of chlorpyrifos

is a common method followed against surface feeding insects

of cotton, rice, mustard, bengal gram etc. (Dhawan and

Simwat, 1996; Gupta et al., 2001). Soil applications are used

for control of root damaging insect larvae attacking crops

such as vegetables, cardamom, tobacco, cole crops,

groundnut and onion (Rouchaud et al., 1991; Bhatnagar and

Gupta, 1992).The practice of application of 1-2%

concentration of chlorpyrifos to soil surrounding building

structures against termite invasion has aggravated the

residue problems (Sundararaj et al., 2003).27

In soil, chlorpyrifos may remain biologically active

for periods ranging from days to months. It is moderately

persistent in nature as its residues were detected in soil

even after 3 months and hence causes potential

environmental hazards (Chapman et al., 1984). Extensive use

of chlorpyrifos contaminates air, groundwater, rivers,

lakes, rainwater and fog water. The contamination has been

found up to about 24 kilometers from the site of

application. Considerable residues of chlorpyrifos were

found in tomatoes (Aysal et al., 1999), cotton seed (Blossom

et al., 2004) and oil of oil seed crop like groundnut,

safflower and mustard (Bhatnagar and Gupta, 1998; Gupta et

al., 2001). It is speculated that the bioaccumulation ability

of chlorpyrifos and other organophosphorus pesticides in

living tissues may spell a potential environmental risk to

marine organisms and humans as well (Serrano et al., 1997;

Tilak et al., 2004).

When organophoshates are released in to the

environment, their fate is decided by various environmental28

conditions and microbial degradation is the key factor for

the disappearance of these pesticides, since they possess

the unique ability to completely mineralize many aliphatic,

aromatic and heterocyclic compounds.

3.1 Microbiological transformation of chlorpyrifos and its

metabolites

In general, microorganisms demonstrate considerable

capacity for the metabolism of many pesticides. Although

they are capable of catalyzing similar metabolic reactions

as mammals and plants, they possess the unique ability to

completely mineralize many aliphatic, aromatic, and

heterocyclic compounds. There are two major types of

microbial degradation of organic chemicals. The first,

termed catabolism is a type of degradation in which the

organic chemical or a portion thereof is completely

degraded (e.g. mineralized) and the energy or nutrient

gained contributes to cell growth. The second, incidental

metabolism or cometabolism, involves the partial

degradation of an organic chemical with no net benefit to29

the organism, the compound being merely caught up in some

metabolic pathway during the normal metabolic activities of

the microorganisms (Racke 1993).Studies conducted in soil

have generally reported significantly longer dissipation

half–lives under sterilized versus natural conditions and

led to the conclusion that microbial activities are

important in the degradation of chlorpyrifos in soil (Miles

et al. 1984). Evidence from soil degradation studies

indicates that cleavage and mineralization of the

heterocyclic ring occurs in soil due to the activities of

microorganisms (Racke &Coats 1990). However, the singularly

most important microbial role in the chlorpyrifos

degradation pathway may be the further metabolism and

mineralization of 3, 5, 6- trichloro-2-pyrinidinol (TCP)

and 3, 5, 6-trichloro-2-methoxypyridine (TMP) metabolites

(Racke 1993).

Microbial enzymes have been shown to hydrolyze

chlorpyrifos under controlled conditions. Munnecke and his

co-worker in 1975 first reported the ability of parathion30

hydrolase, an organophosphorus ester-hydrolyzing enzyme

isolated from a mixed microbial culture, to hydrolyze

chlorpyrifos.

Jones and Hastings (1981) reported the metabolism of

50-ppm chlorpyrifos in cultures of several forest fungi

(Trichoderma harzianum, Penicillium vermiculatum, and Mucor sp.).

After 28 days, chlorpyrifos and its metabolite TCP were

present in all cultures at levels of 2-5 % and 1-14% of

applied, respectively. Ivashina (1986) studied chlorpyrifos

degradation by several microbial cultures maintained in

liquid media containing 10 ppm chlorpyrifos. Dissipation

was more rapid in a sucrose-supplemented media containing

Trichoderma sp. and glucose supplemented media containing

Bacillus sp. than in control media containing no

microorganisms. Chlorpyrifos disappeared from the microbial

cultures in a linear fashion over a 2-week period. Lal and

Lal (1987) observed some degree of degradation by the yeast

Saccharomyces cervisiae. Only half the initial chlorpyrifos was

recovered 12 h after the cultures were inoculated with 1-1031

ppm. The possible metabolism by two lactic acid bacteria

(Lactobacillus bulgaricus and Streptococcus thermophilus) was reported

by Shaker et al. (1988). The synthetic culture medium, in

which the bacteria were grown initially contained 7.4 ppm,

but displayed a 72-83% loss in chlorpyrifos after 96 h.

Havens and Rase (1991) circulated a 0.25 % aqueous (EC)

solution of chlorpyrifos through a packed column containing

immobilized parathion hydrolase enzyme obtained from

Pseudomonas diminuta. Approximately 25 % of the initial dose

was degraded after 3 h of constant recirculation through

the column. Strains of Aspergillus flavus and Aspergillus niger

isolated from agricultural soil with previous history of

chlorpyrifos use were, also reported to biomineralise

chlorpyrifos in liquid culture medium (Swati & Singh 2002).

Yu et al. (2006) isolated and characterized a fungal strain

capable of degrading chlorpyrifos.18S rDNA gene analysis

revealed that they showed that the fungal strain had a high

level of homology (99%) to those from other Verticillium

species. They found that the degradation of chlorpyrifos in32

by the fungal strain in mineral salt medium increased

almost linearly with increasing concentrations of

chlorpyrifos (r 2 =0.9999), suggesting that the degradation

is subjected to pseudo-first order kinetics. With the first

order kinetic function, the DT50 of chlorpyrifos at

concentrations of 1, 10, and 100 mg/l, were calculated to

be 2.03, 2.93, and 3.49 days, respectively. In the controls

the hydrolysis percentages of chlorpyrifos were found to be

less than 5%. They further used the cell free extracts of

the strain to detoxify chlorpyrifos in vegetables and

reported that the cell free extracts of the fungus can used

for enhanced degradation in vegetables.

Some evidence also indicates that, the metabolites of

chlorpyrifos are also degraded and mineralized by soil

microorganisms. Several researchers have noted the

extensive mineralization of TCP and TMP to carbon dioxide

in soil. Racke et al. (1988) reported that approximately

65-85 % of TCP applied (5ppm) to several soils was

mineralized within 14 days. The initially accelerating rate33

of mineralization observed in these soils was indicative of

microbial enzyme induction or adaptation. Racke and Robbins

(1991) probed a suite of soils for evidence of the presence

of TCP-catabolizing microorganisms. Of the 25 soils

investigated, only two displayed significant degradation of

TCP within 21 days of inoculation into mineral salts medium

containing 5-ppm TCP as the sole carbon source. Isolation a

pure culture of bacteria capable of using 3, 5, 6-

trichloro-2-pyridinol (TCP) as the sole source of carbon

and energy under aerobic conditions was reported for the

first time by Feng and his co-workers in 1998. The

bacterium was identified as a Pseudomonas sp. and designated

as ATCC 700113. The TCP degradation yielded CO2, chloride

and some unidentified polar metabolites. They further

reported that the degradation of the parent compound, TCP,

by the Pseudomonas sp. appeared to involve a reductive de-

chlorination mechanism.

3.2 Microbial degradation of chlorpyrifos

34

When organophosphates are released in to the

environment, their fate is decided by various environmental

conditions and microbial degradation. Microbial degradation

is the key factor for the disappearance of these

pesticides. Micro organisms possess the unique ability to

completely mineralize many aliphatic, aromatic and

heterocyclic compounds. Several studies conducted in soil

indicated significantly longer dissipation half-lives under

sterilized versus natural conditions, and led to the

conclusion that microbial activities are important in

degradation of chlorpyriphos (Getzin, 1981a; Miles et al.,

1983). Schmimmel et al. (1983), based on laboratory

degradation studies with aqueous solution and sediments,

concluded that microorganisms play an important role.

Cleavage and mineralization of heterocyclic ring occur in

soil due to activities of microorganisms (Somasundaram et

al., 1987; Racke et al., 1988).

In literature there are many reports regarding

screening and isolation of microbes capable of degrading35

pollutants under laboratory conditions. However, their use

at the contaminated sites under field scale application has

not been successful (Pilon-Smits 2005; Dua et al. 2002;

Kuiper et al. 2004). The reasons behind this include the

competition faced from the natural microflora and

microfauna of the soil, suboptimal nutrition or nutritional

deficiency leading to low microbial growth, non

availability or less bioavailability of the pollutant

desired to be degraded and the growth inhibitory

concentration of pollutant itself (Kuiper et al. 2004;

Dillewijn et al.2007). Shan et al. (2006) reported the

suppressed growth of bacterial, fungal, and actinomycete

populations in the presence of chlorpyrifos (10 mg kg-1).

Vischetti et al. (2007) reported reduction in soil

microbial biomass in an Italian soil field by 25% and 50%

after chlorpyrifos treatment at 10 mg kg-1and 50 mg kg-1,

respectively. Inspite of these limitations, microbial

degradation of organophosphate pesticides is an important

process responsible for their biotic degradation in36

environment (Felsot, 1989). There are reports regarding

ability of microbes to degrade pesticides co-metabolically

or as source of carbon, nitrogen and phosphorous.

3.2.1 Bacterial degradation of chlorpyrifos

Bacteria are dominantly involved in accelerated

biodegradation of pesticides (Racke and Coats, 1990). The

bacterial strains from different taxonomic groups with

potential to degrade the organophosphorus insecticides have

been reported (Yasouri, 2006; Li et al., 2008).

Sethunathan and Yoshida (1973) reported Flavobacterium

sp. having the ability to degrade chlorpyrifos in liquid

medium by cometabolism. Similarly, Serdar et al. in 1982

isolated Pseudomonas diminuta degrading chlorpyrifos co-

metabolically rather than as a source of carbon On the

other hand, Ohshiro et al. (1996) reported that Arthrobacter

sp. strain B- 5 can use chlorpyrifos as a substrate rather

than a co-metabolite. Singh et al. (2003) isolated six

chlorpyrifos degrading bacteria capable of degrading

chlorpyrifos in both liquid medium and soil. Dutta et al.37

(2004) observed increase in net microbial biomass carbon

(MBC) in chlorpyrifos treated soils as compared to the

control containing no chlorpyrifos. Singh et al. (2004)

also reported degradation of chlorpyrifos by pure culture

Enterobacter sp. B-14 in liquid as well as in soils. Wang et

al. (2006) reported degradation of chlorpyrifos by pure

culture of Bacillus laterosporus DSP. Li et al. (2008) reported

isolation of chlorpyrifos degrading bacterial strains Dsp-

2, Dsp-4, Dsp-6 and Dsp-7 identified as Sphingomonas sp.,

Stenotrophomonas sp., Bacillus sp. and Brevundimonas sp.

respectively and few other strains distinguished as members

of Pseudomonas sp. From chlorpyrifos-contaminated samples.

There are many chlorpyrifos degraders reported but

very few degraders are known to degrade the compound at

higher rates. Mallick et al. (1999) reported complete

degradation of 10 mg l-1 of chlorpyrifos in the mineral

salts medium by Flavobacterium sp. ATCC 27551 and Arthrobacter

sp. within 24 h and 48 h respectively. As described

earlier, the major limitation in the process of38

chlorpyrifos degradation is the formation an anti-microbial

compound 3,5,6-trichloro-2-pyridinol (TCP) which may also

affect the growth of chlorpyrifos-transforming

microorganisms (Racke et al. 1990). A report by Racke and

Coats (1990) indicate that transformation of 30 mg kg-1 of

chlorpyrifos in the soil resulted in production of TCP

which repress the proliferation of microbes introduced into

the soil. The accelerated degradation of chlorpyrifos was

observed either due to the ability of degraders to tolerate

TCP or their potential to mineralize TCP efficiently at a

rate higher rapidly than the rate of its formation in the

medium.

There are reports regarding degradation of both

chlorpyrifos and TCP in aqueous phase (Feng et al. 1998;

Mallick et al. 1999; Horne et al. 2002; Bondarenko et al.

2004). A Stenotrophomonas sp. isolated by Yang et al. (2006)

was found to be a degrader of both chlorpyrifos and TCP. On

the other hand, Singh et al. (2004) isolated Enterobacter sp.

capable of degrading chlorpyrifos was not able to degrade39

TCP but utilize diethylthiophosphate as carbon and

phosphorus source. The Enterobacter species in this case

showed the tolerance against TCP even at higher

concentrations (150 mg l-1), which might be the reason of

effective chlorpyrifos degradation.

3.2.2 Fungal degradation of chlorpyrifos

There are not many reports of OP degradation by fungal

species as compared to those by their bacterial

counterparts. Furthermore the organophosphates degradation

rate by fungal isolates was found to be slower. Jones and

Hastings (1981) reported 95% to 98% degradation of 50 ppm

chlorpyrifos by a group of forest fungi namely Trichoderma

harzianum, Penicillium vermiculatum, and Mucor sp. after 28 days

of incubation along with accumulation of its metabolite

TCP. Bumpus et al. (1993) reported a fungal strain

Phanerochaete chrysosporium able to mineralize only 26.6% of

added chlorpyrifos after 18 days of incubation. Bending et

al. (2002) reported Hypholoma fasciculare and Coriolus versicolor

degraded chlorpyrifos in soil bio-bed after 42 days.40

Studies had also reported the chlorpyrifos degradation in

soil by Aspergillus sp., Trichoderma sp. (Liu et al. 2003) and

Fusarium sp. (Wang et al. 2005).

4. Degradation of OPs by microbial consortia

Most of the lab scale microcosm based in-situ

bioremediation studies involving addition of pure cultures

to polluted soil are prone to problem because of poor

survival or low activity of these cultures in the natural

environmental conditions. Munnecke and Hsieh (1974)

reported a mixed bacterial consortium consisting of

Pseudomonas sp., Xanthomonas sp., Azotomonas sp. and a

Brevibacterium sp. capable of hydrolysing 50 mg l-1 of

parathion. The biodegradation of pesticides, such as 4-

nitrophenol (Laha & Petrova 1998), endosulfan (Awasthi et

al. 2000), 1,3- dichloropropene (Ou et al. 2001) and

diazonin degradation (Cycon et al. 2009) by a microbial

consortium has been reported.

The entire pesticide degradation pathways generally may

not be present in individual species. However, different41

components of microbial consortia can work in concerted

manner to acheive effeciant degradation of these compounds

(Macek et al. 2000; Kuiper et al. 2004; Chaudhry et al.

2005). The rhizosphere soil contains 10–100 times more

microbes than un-vegetated soil due to presence of plant

exudates such as sugars, organic acids, and larger organic

compounds in the soil (Lynch 1990; Kumar et al. 2006).

However, there are certain factors those can interfere with

the enrichment process for development of degradative

consortium including (a) the complex molecular and

structural features of the degrading compound that may

limit its degradability e.g. polyhalogenated compounds

(Wackett et al. 1994), (b) natural dominance of a non-

productive metabolic pathway (Oh and Bartha 1997), (c) low

frequency of an essential degradative gene (Shapir et al.

1998), (d) poor bioavailability e.g. polycyclic aromatic

hydrocarbons (Bastiaens et al. 2000) and (c) production of

recalcitrant intermediates (Van Hylckama Vlieg and Janssen

2001). The problem may be overcome by developing42

genetically engineered microbial strains or by developing

an efficient consortium from natural degraders. The

metabolic synergism between different microbial species

encourages the biodegradation of recalcitrant molecules via

aerobic and anaerobic reactions. Gilbert et al. 2003

developed a consortium comprised of two engineered strains,

Escherichia coli SD2 with plasmids encoding a gene for parathion

hydrolase and Pseudomonas putida KT2440 having pSB337 plasmid

contained a p-nitrophenol-inducible operon encoding the

genes for p-nitrophenol mineralization, to hydrolyze 500 μM

of organophosphate insecticide parathion without the

accumulation of p-nitrophenol in suspended culture.

Vidya Lakshmi et al. 2008 developed a microbial

consortium consisting of Pseudomonas fluorescence, Brucella

melitensis, Bacillus subtilis, Bacillus cereus, Klebsiella sp., Serratia marcescens

and Pseudomonas aeruginosa supported 75–87% degradation of

chlorpyrifos after 20 days of incubation.

Immobilization of the pure cultures and consortium may

help to improve bioremediation potential as immobilized43

cells have prolonged microbial cell viability ranging from

weeks to months and improved capacity to tolerate higher

concentrations of pollutants (Richins et al. 2000; Chen and

Georgiou 2002). Karamanev et al. 1998 immobilized bacterial

consortium in alginate beads and on tezontle (a porous

igneous rock) by biofilm for the removal of a pesticide

mixture form liquid medium composed of methyl-parathion and

tetrachlorvinphos.

5. Enzymatic degradation of organophosphorus pesticides

Microorganisms degrading xenobiotic chemicals are

equipped with elaborate enzyme systems. Biodegradation of

organophosphates involves activities of phosphatase,

esterase, hydrolase and oxygenase enzymes. Munnecke (1976)

observed that crude cell free extract from a mixed

bacterial culture growing on parathion, hydrolysed the same

at a rate of 416 n mol/min/mg of protein. The enzyme

phosphotriesterase (parathion hydrolase) was also reported

from Streptomyces lividans (Rowland et al., 1991) partially

purified enzyme can use as a field inoculum to44

decontamination of soil. Parathion hydrolase is also

isolated from Pseudomonas diminuta, Monomeric, spherical

protein having molecular weight of 39,000 Hydrolyze

paraxon, cyanophos, fensulfothion and coumaphos. (Dumas et

al., 1989).The enzyme cloned from a Flavobacterium sp. Into

Streptomyces lividans was secreted at high levels and consisted

of a single polypeptide with an apparent molecular weight

of 35,000. Substrate specificity showed kms of 68 micro M

for parathion and 599 micro M for methyl parathion.

Biodegradation of organophosphorus pesticides by surface

expressed organophosphorus hydrolase was studied by Richins

et al. (1997). Organophosphorus hydrolase (OPH) was displayed

and anchored on to the surface of Escherichia coli using an Lpp-

OmpA fusion system. OPHs are members of the amidohydrolase

superfamily and share the same (α–β) 8 barrel structural

folds and an active site with two transition metal ions,

such as zinc, iron, manganese or cobalt. The OPHs purified

from B. diminuta and Flavobacterium sp. have identical, or very

45

similar, amino-acid sequences (Serdar et al., 1982; Mulbry,

1989; Siddavattam et al., 2003).

A variant of OPH called OPDA (OP-degrading enzyme) has

been purified from A. radiobacter21. Production of OPH on cell

surface is host specific it is highly dependent on growth

conditions of microorganisms.OPH activity in Nocardia simplex

NRRL B-24074 located in the cytoplasm ( Mulbry ,2000). More

than 80 per cent of the activity was found to be located on

the cell surface. The precise conditions for surface

targeting or pesticide degradation were further studied by

Kaneva et al. (1998). Optimum OPH activity was observed when

cells were grown in Luria-Bertani buffered medium at 37°C.

The resulting culture grown under optimized conditions had

an eight fold increase in parathion degradation.

Sureshkumar et al. (1998) reported that the cell free extract

of Flavobacterium balustinum hydrolysed a variety of

organophosphorus pesticides like fenitrothion, quinalphos

and monocrotophos. The enzyme responsible for hydrolysis

was identified as a phosphotriesterase. Cho et al. (2002)46

observed that the effectiveness of degradation by OPH

varied dramatically ranging from highly efficient with

paraoxon to relatively slow with methyl parathion. A solid

phase top agar method based on detection of the yellow

product P-nitrophenol was developed for the rapid pre-

screening of potential variants with improved hydrolysis of

methyl parathion. One variant 22AII hydrolysed methyl

parathion 25 folds faster than did the wild strain.

Organophoshorus acid anhydrolase (OPAA) was isolated

from Alteromonas undina and Alteromonas haloplankti.(Chen et al.,

1993; Chen et al., 1998). Enzyme OPAA belongs to the

dipeptidase family and does not share enzyme or gene-

sequence similarity with OPH.

6. Genetic basis of OP degradation

Molecular biological studies have thrown light on

organophosphate degrading (opd) genes and involvement of

plasmid in degradation of organophosphates. Involvement of

plasmid in parathion degradation was reported by Serdar et

al. (1982). Pseudomonas diminuta cells grown for 48 hours47

contained 3400 U of parathion hydrolase activity per litre

of broth and the activity was found to be associated with

the plasmid pSC1 of molecular mass 44 x 106 daltons. In

Flavobacterium sp. strain ATCC 27551, a 43 kb plasmid

possessed significant homology to the opd sequence (Mulbry

et al., 1986). Chaudhary et al. (1988) described two mixed

bacterial cultures utilizing methyl parathion and

parathion. The hybridisation data showed that the DNA from

Pseudomonas sp. and from mixed culture had homology with opd

gene from a previously reported parathion hydrolysing

bacteria Flavobacterium sp.

Kumar et al. (1996) have reviewed microbial degradation

of pesticides with special emphasis on the role of

catabolic genes and the application of recombinant DNA

technology in the development of an organism which could

simultaneously degrade several xenobiotics. Possible

involvement of plasmids in degradation of malathion and

chlorpyriphos was reported by Guha et al. (1997). Two

plasmids harbouring strains of Micrococcus sp. (M-36 and AG-48

43) degraded malathion and parathion. Chen and Mulchandani

(1998) studied use of live biocatalysts for pesticides

detoxification. They discussed the use of a genetically

engineered E. coli with surface expressed organophosphorus

hydrolase and suggested the ultimate creation of super

biocatalyst capable of degrading several pesticides rapidly

and cost effectively. Moraxella sp. growing on P-nitrophenol

was genetically engineered for the simultaneous degradation

of OP pesticides and P-nitrophenol. The truncated ice

nucleation protein anchor was used to target the

organophosphorus hydrolase on to the surface and the

resulting Moraxella spp. degraded organophosphates rapidly,

all within an hour (Shimazu et al., 2001). Walker and

Keasling (2002) reported that transformation of Pseudomonas

putida with the plasmids harbouring opd and the p-

nitrophenol operons allowed the organism to utilize 0.8 mM

parathion as a source of carbon and energy.

Somara et al. (2002) amplified the opd gene of

Flavobacterium balustinum using PCR and the resulting product49

was cloned in to PUC 18. The protein sequence predicted

from the nucleotide sequence was almost identical to

parathion hydrolase. A transposon like organization of the

plasmid borne opd gene cluster in Flavobacterium sp. was

reported by Siddavattam et al. (2003).

Cho et al. (2004) selected five improved variants based on

the formation of clear halos on Luria-Bertani plates

overlaid with chlorpyriphos. Of these, B3561 exhibited a

725 fold increase in the K (cat)/K (m) value for

chlorpyriphos hydrolysis as well as enhanced hydrolysis

rates for several other OP compounds. Lei et al. (2005)

reported improved degradation of OP nerve agents by

Pseudomonas putida JS444 with surface expressed

organophosphorus hydrolase.

7. Significance of microbial technology for bioremediation

of OP pesticides

Among various pesticide used by farmers,

organophoshorus pesticides (OP) form the major group.

Although these are thought to be easily biodegradable,50

literature survey indicates that complete degradation or

mineralization of the pesticides is rarely achieved. Many

researchers have studied biodegradation of OP particularly

parathion, malathion in laboratory nutrient media wherein

it is easy to detect metabolites of degradation. However it

is difficult to detect metabolites in soils contaminated

with the pesticides. There is a mixed population in soil

and the pesticides are attacked by a number and variety of

microorganisms. The metabolites formed by one type of

microorganisms may be utilized by other group of organisms.

Some of the metabolites like acetic acid are assimilated by

the microbial cells. Such transient metabolites may not be

easily detected in soil. Depending upon solubility and

mobility, the pesticides and their metabolites may reach

deeper layers of soil. There is always uneven distribution

of the pesticides in soil. Many bind to soil particles.

Hence soil sampling for estimation of biodegradation of

pesticides becomes critical.

51

Since organophosphorus pesticides and some of their

metabolites are toxic, their degradation in contaminated

soil is necessary. Although indigenous soil microflora do

attack on the pesticide residue in soil, the process in

slow. To enhance the process in soil, it is desirable to

inoculate soils with cultures efficient in degradation of

such pesticides. This bio- augmentation of soil with

selected cultures will enhance bioremediation of

contaminated soil. Since industrial waste waters are one of

the source of pollution of water bodies. The studies so far

carried out suggest that micro-organisms endowed with this

property of degradation of toxic pollutants are a boon to

mankind.

Conclusion

The pollution of the environment by pesticides is a

consequence of the continuous agricultural expansion,

combined with the population increase. Pesticides are used

in sizeable areas and applied to soil surfaces and

accumulate beneath the ground surface, reaching rivers and52

seas. The natural microbiota is continuously exposed to

pesticides therefore, it is no surprise that these

microorganisms, that inhabit in polluted environments, are

armed with resistance by catabolic processes to remove the

toxic compounds. Biological degradation by organisms

(fungi, bacteria, viruses, protozoa) can efficiently remove

pesticides from the environment, especially

organochlorines, organophosphates and carbamates used in

agriculture. The enzymatic degradation of synthetic

pesticides with microorganisms represents the most

important strategy for the pollutant removal, in comparison

with non-enzymatic processes. The degradation of persistent

chemical substances by microorganisms in the natural

environment has revealed a larger number of enzymatic

reactions with high bioremediation potential. These

biocatalysts can be obtained in quantities by recombinant

DNA technology, expression of enzymes, or indigenous

organisms, which are employed in the field for removing

pesticides from polluted areas. The microorganisms53

contribute significantly for the removal of toxic

pesticides used in agriculture and in the absence of

enzymatic reactions many cultivable areas would be

impracticable for agriculture.

Isolation and characterization of pesticide degrading

microorganisms is crucial for enhancing our understanding

of the array of mechanisms and biodegradative pathways

relating to their enhanced degradation in the environment.

Bioremediation technologies are aimed in the process of

degrading toxic compounds and related nerve agents using

organophosphorus hydrolase enzymes. Future studies on the

genes responsible for enhanced biodegradation will enable

us to elucidate the exact degradative pathway involved in

its microbial biodegradation.

DISCUSSION

Q) Is any microorganisms produce toxic products in to soil?

54

Ans) No. microorganisms use the soil as its food source

they use the enzymes and other soil related materials

for its growth and metabolism.

Q) What is the relevance of bioremediation in kerala?

Ans) we can use this technology for the waste water

treatment.

Q) Can we use this technology for purify ganga river?

Ans) of course we can use this technology for purify ganga

also.

Q) Is bioremediation is mainly for pesticide degradation?

Ans) No. bioremediation is the technology mainly used for

all type of xenobiotic compounds; it’s not only for

pesticides.

Q) What is meant by abiotic factors? How they involving in

bioremediation?

Ans) abiotic factors mainly include the photodegradation

and chemical hydrolysis of xenobiotics. They degrade

the xenobiotics and produce them to simpler molecular

55

forms that can easily degraded by microorganisms as

carbondioxide and hydrogen, methane etc

References

Ansaruddin, P.A. and Vijayalakshmi, K. 2003. The womb is not

safe anymore. Indegenous Agriculture News 2: 3.

Available: http://www.ciks.org/march-03.pdf.[3 March

2003].

Aysal, P., Gozek, K., Artik, N. and Tuncbilek, A.S. 1999.

14C-chlorpyrifos residues in tomatoes and tomato

products. Bulletin Environ Cont Toxicol., 62: 122-125.

Bailey, G. W., Swank, Jr. R. R. and Nicholson, H. P. 1974.

Predicting pesticide runoff from agricultural land: a

conceptual model. J. Environ. Quality 3: 95-102.

Bhatnagar, A. and Gupta, A. 1992. Persistence of

chlorpyriphos residues in soil and groundnut seed.

International Arachis Newsletter., 11: 16-17.

56

Bhatnagar, A. and Gupta, A. 1998. Lindane, chlorpyrifos and

quinalphos residues in safflower seeds and oil.

Pesticide Residue J., 10: 127-128.

Block, R., Stroo, H. and Swett, G. H. 1993. Bioremediation-

why doesn't it work sometimes? Chem. Engng. Prog. 89:

44-50.

Blossom, Singh, B. and Gaganjyot, 2004. Monitoring of

insecticide residues in cotton seed in Punjab, India.

Bulletin Environ. Cont. Toxicol., 73: 707-712.

Boonsaner, M., Kaewumput, T. and Boonsaner, A. 2002.

Bioaccumulation of organophosphate pesticides in

vetiver grass (Vetiveria zizanioides Nash). Thammasat Int. J. of

Sci. and Technol. 7: 30-39.

Brock, T. C. M., Crum, S. J. H., Van Wijngaarden, R.,

Budde, B. J., Tijink, J., Zuppelli, A. and Leeuwangh,

P. 1992. Fate and effects of the insecticide Dursban

4E in Indoor Elodea-dominated and macrophyte free

freshwater model ecosystems: I Fate and primary

57

effects of the active ingredient Chlorpyrifos. Arch.

Environ. Contamination and Toxicology 23: 69–84.

Chapman, R. A. and Chapman, P. C. 1986. Persistence of

granular and EC formulation of chlorpyrifos in a

mineral and an organic soil incubated in open and

closed containers. J. Environ. Sci. Hlth. Part B 21: 447–456.

Chapman, R. A., Harris, C. R., Svec, H. J. and Robinson,

J.R. 1984. Resistance and mobility of granular

insecticides in an organic soil following furrow

application for onion maggot. J. Environ. Sci. Health., 19:

259-270.

Chen, W. and Mulchandani, A .1998. The use of live

biocatelysts for pesticide detoxification; Trends

Biotechnol. 16:71-76

Cink, J. H. and Coats, J. R. 1993. Effect of concentration,

temperature, and soil moisture on the degradation of

chlorpyrifos in an Urban Iowa soil, In: Racke, K.D and

Leslie, A.R, (eds.) Pesticides in Urban Environments: Fate and

58

Significance. ACS Symposium Series 522, American Chemical

Society Washington, DC pp 62-70.

Comeau, Y., Greer, C. W. and Samson, R. 1993. Role of

inoculum preparation and density on the bioremediation

of 2,4-D contaminated soil by bioagumentation. Appl.

Microbial Technol. 38: 681–687.

Daughton CG and Hsieh DP 1977 Parathion utilization by

bacterial symbionts in a chemostat. Appl. Environ. Microbiol.

34: 175-184.

Davis, L. C., Castro-Diaz, S., Zhang, Q. and Erickson, L.

E. 2002. Benefits of vegetation for soils with organic

contaminants. Crit. Rev. Pl. Sci. 21: 457–491.

Dhawan, A. K. and Simwat, G. S. 1996.Pattern of

insecticides use on cotton by farmers in Punjab,

Pestology., 20: 30-34.

Dragun, J., Kuffner, A. C., and Schneiter, R.W. 1984.

Ground water contamination. Transport and

transformation of organic chemicals. Chem. Eng. 91:65-70

59

Dumas, D. P., Caldwelol, S.S.R., Wild, J. R., and Raushel,

F. M. 1989. Purification and properties of the

phosphotriesterase from Pseudomonas diminuta;

J.Biol.Chem.264: 19659-19665

Duquenne, P., Parekh, N. R., Gatroux, G. and Fournier, J.

C. 1996. Effect of inoculant density, formulation,

dispersion and soil nutrient amendment on the removal

of carbofuran residues from contaminated soils. Soil Biol.

Biochem. 28: 1805–1811.

Fernando, T. and Aust, S. D. 1994. Biodegradation of toxic

chemicals by white rot fungi. In: Chaudhry, G.R.

(Ed.), Biological Degradation and Bioremediation of

Toxic Chemicals. Chapman &amp; Hall, London, pp. 386–

402

Fux, C. A. 2005. Survival strategies of infectious

biofilms. Trends in Microbiol. 13: 34–40.

Gan, J., Becker, R. L., Koskinen, W. C. and Buhler, D. D.

1996. Degradation of atrazine in two soils as a

60

function of concentration. J. Environ. Quality 25: 1064–

1072.

Getzin, L.W. 1981. Degradation of chlorpyrifos in soil:

influence of autoclaving, soil moisture, and

temperature. J. Econ. Ent. 74: 158– 162.

Ghigo, J. M. 2001. Natural conjugative plasmids induce

bacterial biofilm development. Nature 412: 442–445

Greenhalgh, R., Dhavan, K. L. and Weinberger, P. 1980.

Hydrolysis of fenitrothion in model and natural

aquatic systems. J. Agric. Food Chem. 28: 102–105.

Gupta, A., Parihar, N. S. and Bhatnagar, A. 2001. Lindane,

chlorpyriphos and quinalphos residues in mustard seed

and oil. Bulletin Environ. Cont. Toxicol., 67: 122-125.

Havens, P.L. and Rase, H.F. 1991. Detoxification of

organophosphorus pesticide solutions. In: Tedder,

D.W., Pohland, F.G. (eds.) Emerging technologies in Hazardous

waste management II. ACS, Washington, DC, pp 261-281.

Hua, F., Yunlong, Y., Xiaoqiang, C., Xiuguo, W., Xiaoe, Y.

and Jingquan, Y. 2009. Degradation of chlorpyrifos in61

laboratory soil and its impact on soil microbial

functional diversity. J. Environ. Sci. 21: 380–386.

Ivashina, S.A. 1986. Interaction of Dursban with soil

micro-organisms, Agrokhimia 8:75- 76.

Johnson, G. R. and Spain, J. C. 2003. Evolution of

catabolic pathways for synthetic compounds: bacterial

pathways for degradation of 2,4-dinitrotoluene and

nitrobenzene. Appl. Microbiol. Biotech. 62: 110–123.

Johnson, R. L., Anschutz, A. J., Smolen, J. M., Simcik, M.

F. and Penn, R.L. 2007. The adsorption of

perfluorooctane sulfonate onto sand, clay, and iron

oxide surfaces. J. Chem. Engng. Data 52: 1165-1170.

Jones, A.S. and Hastings, F.L. 1981. Soil microbe studies.

In: Hastings F.L. and Coster, J.E. (eds.) Field and

laboratory evaluations of insecticides for southern pine beetle control.

USDA. Southern Forest Experiment Station, Forest

Service, SE 21, pp 13-14.

Kanekar, P.P., Bhadbhade, B., Deshpande, N.M. and Sarnaik,

S.S. 2004. Biodegradation of organophosphorus62

pesticides. Proceedings of Indian National Science Acadamy. 70:

57-70.

Karpouzas, D. G. and Walker, A. 2000. Factors influencing

the ability of Pseudomonas putida epI to degrade

ethoprophos in soil. Soil Biol. Biochem. 32: 1753–1762.

Kertesz MA,Cook AM and Leisinger T 1994 Microbial

metabolism of sulfur and phosphorus – containing

xenobiotics. FEMS Microbiol. Rev. 15: 195-215.

Kuffner, A. C and Schneiter, R. W.1984. Ground water

contamination.1.Transport and transformation of

organic chemicals. Chem. Engng. 91:65-70.

Lal, S. and Lal, R. 1987. Bioaccumulation, metabolism, and

effects of DDT, fenitrothion, and chlorpyrifos on

Saccharomyces cerevisae. Arch. Environ. Contam. Toxicol. 16:753-

757.

Li, X., Jiang, J. D., Gu, L., Ali, S. W., He, J., and Li,

S.P. 2008. Diversity of chlorpyrifos-degrading

bacteria isolated from chlorpyrifos-contaminated

63

samples. International Biodeterioration and Biodegradation 62:

331–335.

Luo, Y. and Zhang, M. 2009. Multimedia transport and risk

assessment of organophosphate pesticides and a case

study in the northern San Joaquin Valley of

California. Chemosphere 75: 969–978.

Magri, A. and D.A. Haith. 2009. Pesticide decay in turf: A

review of processes and experimental data. J. Environ.

Quality 38: 4-12.

Miles, J.R.W., Harris, C.R., and Tu. C.M. 1984 Influence of

moisture on the persistence of chlorpyrifos and

chlorfenvinphos in sterile and natural mineral and

organic soils. J.Environ. Sci. Hlth B19, 237-243.

Moore, M. T., Schulz, R., Cooper, C. M., Smith, S. and

Rodgers, J. H. 2002. Mitigation of chlorpyrifos runoff

using constructed wetlands. Chemosphere 46: 827–835.

Mulbry. 2000. Charecterization of a novel organophosphorus

hydrolase from Nocardiodes simplex NRRL B-24074 Microbiol.

Res. 154: 285-308 64

Munnecke, D.M. and Hsieh, D.P.H. 1975. Development of

microbial systems for the disposal of concentrated

pesticide suspensions. Med. Gent.40:1237-1247.

Munnecke, D .M. 1976. Enzymatic hydrolysis of

organophosphate insecticides, a possible pesticide

disposal method; Appl. Environ.Microbiol.32-13

Ohshiro, K., Kakuta, T., Sakai, T., Hirota, H., Hoshino, T.

and Uchiyama, T. (1996). Biodegradation of

organophosphorus insecticides by bacteria isolated

from turf green soil. J. Fermentation Bioengng. 82: 299–

305.

Pandey, S. and Singh, D. K. 2004. Total bacterial and

fungal population after chlorpyrifos and quinalphos

treatments in groundnut (Arachis hypogaea L.) soil.

Chemosphere 55: 197–205.

Paranychianakis, N. V., Angelakis, A. N., Leverenz, H. and

Tchobanoglous, G. 2006. Treatment of wastewater with

slow rate systems: a review of treatment processes and

plant functions. Crit. Rev.Environ. Sci. Technol. 36: 187–259. 65

Philp, T.M., Liu, D., Seech, A .G., Lee, H., and Trevors,

J.T. 2000. Monitoring Bioremediation in Creosote-

Contaminated Soils Using Chemical Analysis and

Toxicity Tests. J. Ind. Microbiol. Biotech. 24:132-135.

Racke, K.D. and Coats, J.R. 1990. Enhanced biodegradation

of insecticides in Midwestern corn soils. In: Racke,

K.D. and Coats J R (eds.) Enhanced biodegradation of

pesticides in the environment. ACS Symp Series 426,

Washington, DC, pp 68-81.

Racke, K.D. 1993 Environmental fate of chlorpyrifos. Rev.

Environ. Contam. Toxicol. 131, 1-151.

Racke, K. D., Fontaine, D. D., Yoder, R. N., and Miller, J.

R. 1994. Chlorpyrifos degradations in soil at

termiticidal application rates. Pesticide Sci. 42: 43– 51.

Ramadan, M. A., EL-Tayeb, O. M. and Alexander, M. (1990).

Inoculum size as a factor limiting success of

inoculation for biodegradation. Appl. Environ. Microbiol. 56:

1392–1396.

66

Richards, R. P. and Baker, D. B. 1993. Pesticide

concentration patterns in agricultural drainage

networks in the lake Erie basin. Environ. Toxicology and

Chem. 12: 13–26.

Rosenberg A and Alexander M 1979 Microbial cleavage of

various organophosphorus insecticides. Appl. Environ.

Microbiol. 37: 886-891.

Rouchaud, J., Gustin, F. and Vanparys, L. 1991. Influence

of slurry and ammonium nitrate fertilizations on soil

and plant metabolism of chlorpyriphos in field

cauliflower. Bulletin Environ. Cont. Toxicol., 46: 705-712.

Rousseaux, S., Hartmann, A., Lagacherie, B., Piutti, S.,

Andreux, F. and Soulas, G. 2003. Inoculation of an

atrazine-degrading strain, Chelatobacter heintzii Cit

1, in four different soils: effects of different

inoculum densities. Chemosphere 51: 569–576.

Rowland,S. S., Speedie, M. K., and Pogell, B. M. 1991.

Purification and characterization of a secreted

67

recombinant phosphotriesterase (parathion hydrolase)

from Streptomyces lividens. Environ. Mirobiol.57: 440-444

Seffernick, J. L. and Wackett, L. P. 2001. Rapid evolution

of bacterial catabolic enzymes: a case study with

atrazine chlorohydrolase. Biochem. 40: 12747– 12753.

Serdar, C. M., Gibson, D. T.,Munnecke, D. M., and

Lancaster, J. H. 1982. Plasmid involvement in

parathion hydrolysis by Pseudomonas diminuta; Appl. Environ.

Microbiol. 44:246-249

Serrano, R., Lopez, F. J., Hernandez, F. and Pena, J.B.

1997. Bioconcentration of chlorpyrifos,

chlorfenvinphos and methidathion in Mytilus galloprovincialis.

Bulletin Environ. Cont. Toxicol., 59: 968-975.

Sethunathan, N. N. and Yoshida, T. 1973. A Flavobacterium

that degrades diazinon and parathion. Canadian J.

Microbiol. 19: 873–875.

Shaker, N., Abo-Donia, S., El-Shaheed, Y.A., and Ismail.

1988. Effect of lactic acid bacteria and heat

68

treatments on pesticides contaminated milk. Egypt J. Diary

Sci. 16:309-317.

Shan, M., Fang, H., Wang, X., Feng, B., Chu, X. Q. and Yu,

Y. L. 2006. Effect of chlorpyrifos on soil microbial

populations and enzyme activities. J. Environ. Sci. 18: 4–5.

Shelton DR 1988 Mineralization of diethylthiophosphoric

acids by an enriched consortium from cattle dip. Appl.

Environ. Microbiol. 54:2572-2573.

Singh BK and Kuhad RC 1999 Biodegradation of lindane by the

white rot fungus Trametes hirsutus. Lett. Appl. Microbiol. 28:

238-241.

Singh, B. K., Walker, A., and Wright, D. J. 2002.

Persistence of chlorpyrifos, fenamiphos,

chlorothalonil, and pendimethalin in soil and their

effects on soil microbial characteristics. Bull. Environ.

Contamination and Toxicol. 69: 181–188.

Singh, B. K., Walker, A., Morgan, A. J. W. and Wright, D.

J. (2003). Effects of soil pH on the biodegradation of

69

chlorpyrifos and isolation of a chlorpyrifosdegrading

bacterium. Appl. Environ. Microbiol. 69: 5198– 5206.

Singh, B. K., Walker, A. and Wright, D. J. 2006.

Bioremedial potential of fenamiphos and chlorpyrifos

degrading isolates: Influence of different

environmental conditions. Soil Biol.Biochem. 38: 2682–2693.

Singhal, V. 2003. Indian Agriculture published by Indian

Economic Data Research Centre, New Delhi, pp.85-93.

Spacie, A. and Hamelink, J. L. 1985. Bioaccumulation, In:

Rand, P.M., Petrocelli, S.R. (eds.), Fundamentals of

Aquatic Toxicology: Methods and Applications. Hemisphere

Publishers, New York. pp. 497–534.

Stagnitti, F., Parlange, J.-Y., Steenhuis, T. S., Nijssen,

B. and Lockington, D. 1994. Modeling the migration of

water-soluble contaminants through preferred paths in

the soils. In: Kovar, K., Soveri, J. (eds.),

Groundwater Quality Management IAHS, Wallingford, U.K.220:

367–379.

70

Stangroom, S. J. C., Collins, C. D. and Lester, J. N. 2000.

Abiotic behaviour of organic micropollutants in soils

and the aquatic environment. A review: II.

Transformations. Environ. Technol. 21: 865–882.

Sundararaj, R., Remadevi, O. K. and Muthukrishnan, R. 2003.

Comparative efficacy of some insecticides in ground

contact against subterranean termites. Pestology., 27:

17-20. Stewart, D. K., Chrisholm, D. and Ragab, M. T.

1971. Long term persistence of parathion in soil.

Nature 229: 47-48.

Tilak, K. S., Veeraiah, K. and Rao, D. K. 2004. Toxicity

and bioaccumulation of chlorpyrifos in Indian carp

Catla catla (Hamilton), Labeo rohita (Hamilton), and Cirrhinus

mrigala (Hamilton). Bulletin Environ. Cont. Toxicol., 73: 933-

941.

Trapp, S. 2000. Modelling uptake into roots and subsequent

translocation of neutral and ionisable organic

compounds. Pest Manage. Sci. 56: 767–778.

71

Vink, J. P. M. and Van der Zee, S. E. A. T. M. 1997.

Pesticide biotransformation in surface waters:

multivariate analyses of environmental factors at

field sites. Water Res. 31: 2858–2868.

Volkering, F., Breura, A. M. and Rulkens, W. H. 1998.

Microbiological aspects of surfactant use for

biological soil remediation. Biodegradation 8: 401-417.

Walia, S., Dureja, P. and Mukerjee, S. K. 1988. New

photodegradation products of chlorpyrifos and their

detection on glass, soil and leaf surface. Arch. of

Environ.Contamination and Toxicology 17: 183–188.

Wang, X., Chu, X. Q., Yu, Y. L., Fang, H., Chen, J. and

Song, F. M. (2006). Characteristics and function of

Bacillus latersprorus DSP in degrading chlorpyrifos.

Acta Pedologica Sinica 43: 648–654.

Weber, J. B. 1972. Interaction of organic pesticides with

particulate matter in aquatic and soil systems. In:

Fate of Organic Pesticides in the Aquatic Environment. Am. Chem.

Soc. Adv. in Chem. Ser. 111: pp.55-120. 72

Wolfenden R and Spence G 1967 Depression of

phosphomonoesterase and phophodiestease activities in

Aerobacter aerogenes. Biochem. Biophys. Acta. 146: 296-298

Wood, B. and Stark, J. D. 2002. Acute toxicity of drainage

ditch water from a Washington State cranberry-growing

region to Daphnia pulex in laboratory bioassays.

Ecotoxicology and Environ. Safety 53: 273–280.

Yadav, J. S., Soellner, M. B., Loper, J. C., and Mishra, P.

K. 2003. Tandem cytochrome P450 monooxygenase genes

and splice variants in the white rot fungus

Phanerochaete chrysosporium: cloning, sequence analysis

and regulation of differential expression. Fungal Genet.

Biol. 38: 10–21.

Yang, C., Liu, N., Guo, X. M. and Qiao, C. 2006. Cloning of

mpd gene from a chlorpyrifos-degrading bacterium and

use of this strain in bioremediation of contaminated

soil. FEMS Microbiol. Lett. 265: 118–125.

Yasouri, F. N. 2006. Plasmid mediated degradation of

diazinon by three bacterial strains Pseudomonas sp.,73

Flavobacterium sp. and Agrobacterium sp. Asian J. Chem. 18:

2437– 2444

Yates, S. R., Wang, D., Papiernik, S. K. and Gan, J. 2002.

Predicting pesticide volatilization from soils.

Environmetrics 13: 569-578.

Yu, Y. L., Wu, X. M., Li, S. N., Fang, H., Zhan, H. Y. and

Yu, J. Q. 2006. An exploration of the relationship

between adsorption and bioavailability of pesticides

in soil to earthworm. Environ. Pollut. 141: 428–433.

KERALA AGRICULTURAL UNIVERSITYCOLLEGE OF AGRICULTURE, VELLAYANI

Agricultural MicrobiologyMicro 591

ABSTRACT

KAROLIN K.P.16 -11- 13

2012 – 11-16510 – 11 AM

74

MICROBIAL TECHNOLOGY FOR BIOREMEDIATION OF ORGANOPHOSPHORUSPESTICIDES

Pesticides constitute the key control strategy forcrop pest and disease management. Continuous application ofpesticides to the soil and aquatic system resulted inhealth hazards and environmental pollution which hastriggered much public concern. The wide spread use of thesepesticides over the years has resulted in problems causedby their interaction with the biological systems in theenvironment. Pesticides will continue to be anindispensable tool for the management of pests in the yearsto come, as there is no suitable alternative to replacethem totally.Organophosphorus pesticides

Organophosphorus compounds are the most widely usedinsecticides, accounting for 34% of world-wide insecticidesales. These compounds possess high mammalian toxicity andit is therefore essential to remove them from theenvironment. Considering the toxic effect of thesepesticides it is essential to remove them from theenvironment employing suitable remedial measures.Bioremediation exploiting microbial technology is one ofthe recent techniques for environmental clean-up. In thisprocess heterotrophic microorganisms breakdown hazardouscompounds to obtain carbon and energy. The first micro-organism that could degrade organophosphorus compounds wasisolated in 1973 (Dragun et al., 1984) and identifiedas Flavobacterium sp. Since then several bacterial and a fewfungal species have been isolated which can degrade a widerange of organophosphorus compounds in liquid cultures andsoil systems.

75

Pesticide transformationsAbiotic

The abiotic processes involved either their transportwhere parent compound remains unchanged and simplytransferred from one matrix to another depending on thephysicochemical properties of pesticides itself. Apart fromvolatilization, leaching, runoff, absorption and adsorptionof pesticides photodegradation, chemical hydrolysis etc.help in transformations.

Biotic Xenobiotic compounds like organophosphate pesticides

are man made compounds and were not previously present innature. Consequently the natural microflora does not havepotential to metabolize these pesticides due to lack ofenzyme and proper transport processes. Over the years, dueto excessive use of xenobiotic compounds microbes haveevolved new degradation pathways resulting in accelerateddegradation of such compounds.Bioremediation

Technology used for clean-up of pollutants from soil,groundwater, surface water and air, using biological,usually microbiological process (Philp et al., 2000).Microbiological transformation of chlorpyrifos and itsmetabolites

Chlorpyrifos is one of the world's most widely usedorganophosphorus pesticides in agriculture. Exposure tochlorpyrifos and its metabolites have been related to avariety of nerve disorders in humans. Microbial degradationis considered to be an efficient and cost effective methodfor decontamination of toxic organophosphorus pesticidesfrom the environment. Chlorpyrifos previously shown to be

76

immune to enhanced biodegradation has now been proved toundergo enhanced microbe mediated decay into less harmfuland non-toxic metabolites, under a set of favourableabiotic conditions. Recently, research activities in thisarea has shown that a diverse range of microorganisms areresponsible for chlorpyrifos degradation.

Bacteria are dominantly involved in acceleratedbiodegradation of pesticides(Racke and Coats, 1990).Thebacterial strains from different taxonomic groups withpotential to degrade the organophosphorus insecticides havebeen reported (Yasouri,2006; Li et al., 2008).

Co-metabolic process of OP degradation by lignolyticor cytochrome 450 associated enzymes of fungi is alsoprevalent (Fernado and Aust,1994;Yadav et.al.,2003)

Degradation of OPs by microbial consortiaThe entire pesticide degradation pathways generally

may not be present in individual species. However,different components of microbial consortia can work inconcerted manner to achieve efficient degradation ofdifferent metabolites.Genetic basis of OP degradation

Molecular biological studies have thrown light onorganophosphate degrading (opd) genes and involvement ofplasmid in degradation of organophosphates.Conclusion

Isolation and characterization of pesticide degradingmicroorganisms is crucial for enhancing our understandingof the array of mechanisms and biodegradative pathwaysrelating to their enhanced degradation in the environment.Bioremediation technologies are aimed in the process ofdegrading toxic compounds and related nerve agents using

77

organophosphorus hydrolase enzymes. Future studies on thegenes responsible for enhanced biodegradation will enableus to elucidate the exact degradative pathway involved inits microbial biodegradation. References

Chapman, R. A. and Chapman, P. C. 1986. Persistence ofgranular and EC formulation of chlorpyrifos in amineral and an organic soil incubated in open andclosed containers. J. Environ. Sci. Hlth. Part B 21:447–456.

Chen, W. and Mulchandani, A .1998. The use of livebiocatelysts for pesticide detoxification; TrendsBiotechnol. 16:71-76

Dragun, J., Kuffner, A. C., and Schneiter, R.W. 1984.Ground water contamination. Transport and transformation of organic chemicals. Chem. Eng.91:65-70

Dumas, D. P., Caldwelol, S.S.R., Wild, J. R., and Raushel,F. M. 1989. Purification and properties of thephosphotriesterase from Pseudomonas diminuta;J.Biol.Chem.264: 19659-19665.

Fernando, T. and Aust, S. D. 1994. Biodegradation of toxicchemicals by white rot fungi. In: Chaudhry, G.R.(Ed.), Biological Degradation and Bioremediation ofToxic Chemicals. Chapman &amp; Hall, London, pp. 386–402.

Getzin, L.W. 1981. Degradation of chlorpyrifos in soil:influence of autoclaving, soil moisture, andtemperature. J. Econ. Ent. 74: 158– 162.

Havens, P.L. and Rase, H.F. 1991. Detoxification oforganophosphorus pesticide solutions. In: Tedder,

78

D.W., Pohland, F.G. (eds.) Emerging technologies in Hazardouswaste management II. ACS, Washington, DC, pp 261-281.

Ivashina, S.A. 1986. Interaction of Dursban with soilmicro-organisms, Agrokhimia 8:75- 76.

Jones, A.S. and Hastings, F.L. 1981. Soil microbe studies.In: Hastings F.L. and Coster, J.E. (eds.) Field andlaboratory evaluations of insecticides for southern pine beetle control.USDA. Southern Forest Experiment Station, ForestService, SE 21, pp 13-14.

Kuffner, A. C and Schneiter, R. W.1984. Ground watercontamination.1.Transport and transformation oforganic chemicals. Chem. Engng. 91:65-70.

Lal, S. and Lal, R. 1987. Bioaccumulation, metabolism, andeffects of DDT, fenitrothion, and chlorpyrifos onSaccharomyces cerevisae. Arch. Environ. Contam. Toxicol. 16:753-757.

Li, X., Jiang, J. D., Gu, L., Ali, S. W., He, J., and Li,S.P. 2008. Diversity of chlorpyrifos-degradingbacteria isolated from chlorpyrifos-contaminatedsamples. International Biodeterioration and Biodegradation 62:331–335.

Miles, J.R.W., Harris, C.R., and Tu. C.M. 1984. Influenceof moisture on the persistence of chlorpyrifos andchlorfenvinphos in sterile and natural mineral andorganic soils. J.Environ. Sci. Hlth. 19: 237-243.

Mulbry. 2000. Charecterization of a novel organophosphorushydrolase from Nocardiodes simplex NRRL B-24074Microbiol.Res. 154: 285-308

Munnecke, D.M. and Hsieh, D.P.H. 1975. Development ofmicrobial systems for the disposal of concentratedpesticide suspensions. Med. Gent.40:1237-1247.

79

Munnecke, D .M. 1976. Enzymatic hydrolysis oforganophosphate insecticides, a possible pesticidedisposal method; Appl. Environ.Microbiol.32-13

Philp, T.M., Liu, D., Seech, A .G., Lee, H., and Trevors,J.T. 2000. Monitoring Bioremediation in Creosote-Contaminated Soils Using Chemical Analysis andToxicity Tests. J.Ind.Microbiol.Biotech. 24:132-135.

Racke, K.D. and Coats, J.R. 1990. Enhanced biodegradationof insecticides in Midwestern corn soils. In: Racke,K.D. and Coats J R (eds.) Enhanced biodegradation ofpesticides in the environment. ACS Symp Series 426,Washington, DC, pp 68-81.

Racke, K.D. 1993. Environmental fate of chlorpyrifos. Rev.Environ. Contam. Toxicol. 131:1-151.

Racke, K. D., Fontaine, D. D., Yoder, R. N., and Miller, J.R. 1994. Chlorpyrifos degradations in soil attermiticidal application rates. Pesticide Sci. 42: 43– 51.

Rowland,S. S., Speedie, M. K., and Pogell, B. M. 1991.Purification and characterization of a secretedrecombinant phosphotriesterase (parathion hydrolase)from Streptomyces lividens. Environ.Mirobiol.57: 440-444

Serdar, C. M., Gibson, D. T.,Munnecke, D. M., andLancaster, J. H. 1982. Plasmid involvement inparathion hydrolysis by Pseudomonas diminuta;Appl.Environ.Microbiol. 44:246-249

Shaker, N., Abo-Donia, S., El-Shaheed, Y.A., and Ismail.1988. Effect of lactic acid bacteria and heattreatments on pesticides contaminated milk. Egypt J. DiarySci. 16:309-317.

Singh, B. K., Walker, A., and Wright, D. J. 2002.Persistence of chlorpyrifos, fenamiphos,

80

chlorothalonil, and pendimethalin in soil and theireffects on soil microbial characteristics. Bull. Environ.Contamination and Toxicol. 69: 181–188.

Yadav, J. S., Soellner, M. B., Loper, J. C., and Mishra, P.K. 2003. Tandem cytochrome P450 monooxygenase genesand splice variants in the white rot fungusPhanerochaete chrysosporium: cloning, sequence analysisand regulation of differential expression. Fungal Genet.Biol. 38: 10–21.

Yasouri, F. N. 2006. Plasmid mediated degradation ofdiazinon by three bacterial strains Pseudomonas sp.,Flavobacterium sp. and Agrobacterium sp. Asian J. Chem. 18:2437– 2444

81