17
http://informahealthcare.com/mby ISSN: 1040-841X (print), 1549-7828 (electronic) Crit Rev Microbiol, Early Online: 1–17 ! 2014 Informa Healthcare USA, Inc. DOI: 10.3109/1040841X.2014.929564 REVIEW ARTICLE Microbial degradation of herbicides Baljinder Singh and Kashmir Singh Department of Biotechnology, Panjab University, Chandigarh, Punjab, India Abstract Herbicides remain the most effective, efficient and economical way to control weeds; and its market continues to grow even with the plethora of generic products. With the development of herbicide-tolerant crops, use of herbicides is increasing around the world that has resulted in severe contamination of the environment. The strategies are now being developed to clean these substances in an economical and eco-friendly manner. In this review, an attempt has been made to pool all the available literature on the biodegradation of key herbicides, clodinafop propargyl, 2,4-dichlorophenoxyacetic acid, atrazine, metolachlor, diuron, glypho- sate, imazapyr, pendimethalin and paraquat under the following objectives: (1) to highlight the general characteristic and mode of action, (2) to enlist toxicity in animals, (3) to pool microorganisms capable of degrading herbicides, (4) to discuss the assessment of herbicides degradation by efficient microbes, (5) to highlight biodegradation pathways, (6) to discuss the molecular basis of degradation, (7) to enlist the products of herbicides under degradation process, (8) to highlight the factors effecting biodegradation of herbicides and (9) to discuss the future aspects of herbicides degradation. This review may be useful in developing safer and economic microbiological methods for cleanup of soil and water contaminated with such compounds. Keywords Biodegradation, herbicides, microbes, toxicity History Received 3 April 2014 Revised 22 May 2014 Accepted 27 May 2014 Published online 26 August 2014 Introduction Herbicides are class of chemical compounds that are toxic to plants, especially unwanted ones. Modern agriculture relies heavily on herbicides for the control of weeds and ease out to maximize yield in crops. These compounds have economical benefits to sustain an increasing world population. The development of herbicide-resistant plants has also led to an unexpected increase in the resilience of weeds. Genetically modified crops resistant to herbicides have become so prevalent that resistant weeds are beginning to appear, necessitating new forms of genetic modification. Weeds have become more and more resistant to herbicides, prompting farmers to use a wider variety and larger quantity of herbicides to control them. The introduction of herbicide- tolerant plants at first decreased herbicide use, but afterwards increased its usage and scope. While pesticide use dropped from 22 454 lbs to 15 618 lbs from 2003 to 2008, at a rate of 7000 lbs per acre per year, herbicide use increased from 278 514 000 lbs to 330 422 709 lbs (Cherry, 2010). The world- wide herbicide market value grew by 39% between 2002 and 2011 and is projected to enhance a further 11% by 2016 (Gianessi, 2013). Majority of herbicides are reported to constitute between 40 and 60% of pesticides used for agricultural purpose. Due to excessive use of herbicides, there is great concern about their potential environmental hazard. Herbicides con- tamination can lead to soil and water pollution (Juhler et al., 2001), reduced biodiversity and depression in soil hetero- trophic bacteria (including denitrifying bacteria) and fungi (Song et al., 2013). The environmental fate of herbicides is a matter of recent concern provided that only a small fraction of the chemicals reach the target organisms (Pimentel, 1995) leading to great impacts of residual herbicides in soil and water on human, animal and crop health. Major sources of herbicides contamination appear to be an inadequate man- agement practices specifically involving on-farm handling of herbicides. The chemical properties and quantity of herbi- cides determine their toxicity and persistence in the environ- ment. Their interaction with targeted and nontargeted organisms has extensively damaged the ecosystem through entry into the food chains (Singh et al., 2013). According to the Weed Science Society of America, herbicides have been classified into 29 different classes based on mechanism of action (wssa.net/wp-content/uploads/WSSA-Mechanism-of- Action.pdf). Application method of herbicides generally include foliar applied method (apply to leaf), soil applied (with soil contact), broadcast (contact with entire area) and spot (contact with specified area). Commonly used herbicides include clodinafop propargyl (CF), 2,4-dichlorophenoxyacetic acid (2,4-D), atrazine, meto- lachlor, diuron, glyphosate (GP), imazapyr, pendimethalin, and paraquat (PQ). These herbicides are harmful to organisms even at micro (mg) levels. Their uses have resulted in severe Address for correspondence: Dr. Baljinder Singh, Department of Biotechnology, Punjab University, Chandigarh, Punjab, India. Tel: +91 172 2534076. Fax: +91 172 2541409. E-mail: [email protected], [email protected] Critical Reviews in Microbiology Downloaded from informahealthcare.com by University of Bath on 11/02/14 For personal use only.

Microbial degradation of herbicides

  • Upload
    kashmir

  • View
    218

  • Download
    3

Embed Size (px)

Citation preview

Page 1: Microbial degradation of herbicides

http://informahealthcare.com/mbyISSN: 1040-841X (print), 1549-7828 (electronic)

Crit Rev Microbiol, Early Online: 1–17! 2014 Informa Healthcare USA, Inc. DOI: 10.3109/1040841X.2014.929564

REVIEW ARTICLE

Microbial degradation of herbicides

Baljinder Singh and Kashmir Singh

Department of Biotechnology, Panjab University, Chandigarh, Punjab, India

Abstract

Herbicides remain the most effective, efficient and economical way to control weeds; and itsmarket continues to grow even with the plethora of generic products. With the development ofherbicide-tolerant crops, use of herbicides is increasing around the world that has resulted insevere contamination of the environment. The strategies are now being developed to cleanthese substances in an economical and eco-friendly manner. In this review, an attempt hasbeen made to pool all the available literature on the biodegradation of key herbicides,clodinafop propargyl, 2,4-dichlorophenoxyacetic acid, atrazine, metolachlor, diuron, glypho-sate, imazapyr, pendimethalin and paraquat under the following objectives: (1) to highlight thegeneral characteristic and mode of action, (2) to enlist toxicity in animals, (3) to poolmicroorganisms capable of degrading herbicides, (4) to discuss the assessment of herbicidesdegradation by efficient microbes, (5) to highlight biodegradation pathways, (6) to discuss themolecular basis of degradation, (7) to enlist the products of herbicides under degradationprocess, (8) to highlight the factors effecting biodegradation of herbicides and (9) to discuss thefuture aspects of herbicides degradation. This review may be useful in developing safer andeconomic microbiological methods for cleanup of soil and water contaminated with suchcompounds.

Keywords

Biodegradation, herbicides, microbes, toxicity

History

Received 3 April 2014Revised 22 May 2014Accepted 27 May 2014Published online 26 August 2014

Introduction

Herbicides are class of chemical compounds that are toxic to

plants, especially unwanted ones. Modern agriculture relies

heavily on herbicides for the control of weeds and ease out to

maximize yield in crops. These compounds have economical

benefits to sustain an increasing world population. The

development of herbicide-resistant plants has also led to an

unexpected increase in the resilience of weeds. Genetically

modified crops resistant to herbicides have become so

prevalent that resistant weeds are beginning to appear,

necessitating new forms of genetic modification. Weeds

have become more and more resistant to herbicides,

prompting farmers to use a wider variety and larger quantity

of herbicides to control them. The introduction of herbicide-

tolerant plants at first decreased herbicide use, but afterwards

increased its usage and scope. While pesticide use dropped

from 22 454 lbs to 15 618 lbs from 2003 to 2008, at a rate of

7000 lbs per acre per year, herbicide use increased from

278 514 000 lbs to 330 422 709 lbs (Cherry, 2010). The world-

wide herbicide market value grew by 39% between 2002 and

2011 and is projected to enhance a further 11% by 2016

(Gianessi, 2013). Majority of herbicides are reported to

constitute between 40 and 60% of pesticides used for

agricultural purpose.

Due to excessive use of herbicides, there is great concern

about their potential environmental hazard. Herbicides con-

tamination can lead to soil and water pollution (Juhler et al.,

2001), reduced biodiversity and depression in soil hetero-

trophic bacteria (including denitrifying bacteria) and fungi

(Song et al., 2013). The environmental fate of herbicides is a

matter of recent concern provided that only a small fraction of

the chemicals reach the target organisms (Pimentel, 1995)

leading to great impacts of residual herbicides in soil and

water on human, animal and crop health. Major sources of

herbicides contamination appear to be an inadequate man-

agement practices specifically involving on-farm handling of

herbicides. The chemical properties and quantity of herbi-

cides determine their toxicity and persistence in the environ-

ment. Their interaction with targeted and nontargeted

organisms has extensively damaged the ecosystem through

entry into the food chains (Singh et al., 2013). According to

the Weed Science Society of America, herbicides have been

classified into 29 different classes based on mechanism of

action (wssa.net/wp-content/uploads/WSSA-Mechanism-of-

Action.pdf). Application method of herbicides generally

include foliar applied method (apply to leaf), soil applied

(with soil contact), broadcast (contact with entire area) and

spot (contact with specified area).

Commonly used herbicides include clodinafop propargyl

(CF), 2,4-dichlorophenoxyacetic acid (2,4-D), atrazine, meto-

lachlor, diuron, glyphosate (GP), imazapyr, pendimethalin,

and paraquat (PQ). These herbicides are harmful to organisms

even at micro (mg) levels. Their uses have resulted in severe

Address for correspondence: Dr. Baljinder Singh, Department ofBiotechnology, Punjab University, Chandigarh, Punjab, India. Tel: +91172 2534076. Fax: +91 172 2541409. E-mail: [email protected],[email protected]

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y U

nive

rsity

of

Bat

h on

11/

02/1

4Fo

r pe

rson

al u

se o

nly.

Page 2: Microbial degradation of herbicides

contamination of the environment, and strategies are now

being developed to clean these substances in an economical

and eco-friendly manner. The present review summarizes

information on the toxicity and microbial degradation path-

ways of these nine herbicides.

Clodinafop propargyl

CF (prop-2-ynyl (R)-2-[(5-chloro-3-fluoro-2-pyridyloxy)

phenoxy]propanoate) is a recently introduced aryloxyphenox-

ypropionate herbicide, used for post emergence control of

annual grasses in cereals (Singh, 2013). The widespread use

of CF has resulted in the discharge of large amounts of the

compound into the environment, which eventually reach the

biosphere. CF is absorbed by the leaves and rapidly

translocated to the growing points of leaves and stems. It

interferes with the production of fatty acids needed for plant

growth in susceptible grassy weeds (Singh, 2013). CF acts by

targeting the enzyme acetyl coenzyme-A-carboxylase, essen-

tial for lipid biosynthesis.

Toxicity of CF

The research focusing on the ecotoxicity of this herbicide is

limited. This might be because of its low persistence and half-

life in soil, which is reported to be 5 d, dependent on the soil

type and pH (Singh, 2013). However, few studies have

demonstrated that CF and its derivatives are toxic and

carcinogenic to humans and other living organisms (Gui

et al., 2011; Kashanian et al., 2008). Kashanian et al. (2008)

reported that the CF in low concentration interacts with calf

thymus DNA by an intercalative mode of binding. CF disrupts

the posterior and ventral development of zebrafish embryos

(Gui et al., 2011; Jaquet et al., 2014). Embryos were exposed

to a range of concentrations from 0.2 mM to 5 mM starting at

late cleavage stage (2 h post fertilization (hpf)) or late

gastrulation stage (10 hpf). The results showed that two

exposure strategies had the same minimum teratogenic con-

centration of 0.6 mM but caused different groups of morpho-

genetic malformations. Thus, due to high toxicity of CF, its

degradation from the environment is of great concern.

Biodegradation of CF

Degradation of CF by aerobic bacteria involves esterase

activity that results in the formation of phenols as metabolites

(Figure 1) (Hou et al., 2011; Singh et al., 2013). Hou et al.

(2011) describe for the first time a microbial strain

Rhodococcus sp. T1, which is able to use CF as a source of

energy. They had reported 97.9% CF degradation without

identifying its metabolites. Singh et al. (2013) isolated

Pseudomonas sp. strain B2 from crop field soil and found

that 87.14% CF was degraded out of initial provided 80 mg/L

CF. The two metabolites clodinafop acid and 4-(4-Chloro-

2-fluoro-phenoxy)-phenol are reported within 9 h of incuba-

tion. Recently, Jaquet et al. (2014) studied the metabolic fate

of 14C-phenyl-labeled CF for 28 d in laboratory assays using

a soil from Germany (Ap horizon, silt loam and cambisol) and

reported that metabolic fate of CF was predominated by the

formation of non-extractable residues (about 60% after 28 d of

incubation). Besides these two studies, no other reports are

available on biodegradation of CF. This might be because of

its low persistence and half life in soil.

2,4-Dichlorophenoxyacetic acid

2,4-D is one of the most widely used chlorinated acidic

phenoxy herbicide in the world and is an analogue of a growth

hormone, auxin. 2,4-D was first synthesized in 1941,

marketed commercially and registered for use in the United

States in 1944 and 1948, respectively. It has provided

economical, selective, effective control of broadleaf weeds

in agriculture crops, pastures and forests for the past decades.

This is used both agriculturally and domestically for post-

emergent control of broad-leaf weeds. 2,4-D is formulated as

amine salts (mainly dimethyl-amine salt), which are more

soluble in water than acid, and ester derivatives (2-ethyhexyl

ester), which are readily dissolved in an organic solvent. The

2,4-D amine salts and esters rapidly convert to the 2,4-D acid

under most environmental conditions (US Environmental

Protection Agency (EPA), 2004).

Figure 1. CF degradation by microbes (1) CF(2) acid metabolite, clodinafop acid; (3) 4-(4-Chloro-2-fluoro-phenoxy)-phenol (4) phenol.The scheme is based on articles cited in thetext.

N

Cl

F

O

O

CH3

CH3

CO2 H2O

O

O

CH

N

Cl

F

O

O

O

OH

N

Cl

F

O

OH

OHN

Cl

F

OH

+ +

1

2

3

4

2 B. Singh Crit Rev Microbiol, Early Online: 1–17

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y U

nive

rsity

of

Bat

h on

11/

02/1

4Fo

r pe

rson

al u

se o

nly.

Page 3: Microbial degradation of herbicides

In plants, 2,4-D functions by maintaining high levels of the

plant hormone auxin, resulting in overstimulation of plant

growth and ultimately death. 2,4-D also induced ethylene

production and therefore, it acts as defoliating agent. After

application, 2,4-D is readily absorbed by the leaves and roots,

and is then translocated to all parts of the plant. However, the

exact mechanism by which this herbicide affects cells is not

completely understood. It was reported that 2,4-D is a

peroxisome proliferators and especially in plant cells, the

herbicide induced mitotic and meiotic abnormalities both in

vivo and in vitro (Gonzalez et al., 2005; Marron et al., 2006).

2,4-D is a moderately persistent chemical with a half- life

between 20 and 200 d. Weintraub et al. (1954) studied the

metabolism and persistence of 2,4-D in dormant plant tissue

and reported relatively long persistence of the compound

in bean plants by performing experiments with C14-labeled

2,4-D. The persistence of 2,4-D in soil depends on the type of

soil. It is highly water-soluble acids (pKa¼ 3.11) and have a

low tendency to accumulate in organic matter (El-Bestawy &

Hans-Jorgen, 2007), therefore can enter as contaminants into

streams, rivers or lakes directly from drainage of agricultural

lands (Lagana et al., 2002; Mikov et al., 2010; Muller &

Babel, 2004). In the aqueous environment, 2,4-D is most

commonly found as the free anion (Burns et al., 2012).

Toxicity of 2,4-D

2,4-D is classified by WHO as a hormonal herbicide of level

II toxicity. Exposure to human beings may occur through

inhalation, skin contact or ingestion. In most cases, the

predominant route of occupational exposure has been the

absorption of spills or aerosol droplets through the skin

(Chaudhry & Huang, 1988). 2,4-D is easily adsorbed into the

human organ from the alimentary tract and skin and is

subsequently excreted in the urine in nearly unchanged form.

It was detected in stomach, blood, brain and kidney of four-

day-old rat neonates fed by 2,4-D-exposed mothers (Cenkci

et al., 2010; Sturtz et al., 2000). 2,4-D has been shown to

exert toxic effects on both animals and humans such as

carcinogenic, teratogenic, neurotoxic, immunosuppressive,

cytotoxic and hepatoxic effects. The toxicity of 2,4-D have

been studied and reviewed by many workers (Atamaniuk,

2013; Bortolozzi et al., 2001; Bukowska, 2006; Burns et al.,

2012; Coady et al., 2013; Garabrant & Philbert, 2002; Mikov

et al., 2010; Munro et al., 1992). In mammals, 2,4-D disrupts

energy production (adenosine triphosphate (ATP)) (Palmeira

et al., 1994). 2,4-D is associated with Hodgkin’s disease,

non-Hodgkin’s lymphoma (NHL) and soft tissue sarcoma

(Garabrant & Philbert, 2002; Holland et al., 2002).

In mammalian cells in vitro, 2,4-D inhibits cell growth and

protein synthesis by affecting the DNA (Gonzalez et al.,

2005). Ganguli et al. (2014) have investigated the molecular

mechanism of 2,4-D-induced lung toxicity in A549 and WI38

cell lines and reported IC50 values 126 ± 2.25 mM and

115 ± 4.39 mM, in A549 and WI38 cells, respectively, for 72 h.

In plant, it causes an increase in chromosomal aberration,

multipolar cells and a decrease in mitotic index (Gui et al.,

2011). The genotoxic potential of 2,4-D in plant has been

investigated by using different test systems including micro-

nucleus, comet assay and Random Amplified Polymorphic

DNA (Aksaka et al., 2013; Cenkci et al., 2010).

2,4-D is among the most used herbicides in cultivable

lands and its high toxicity in both animals and plants reinforce

researched to found out the degradation pathways along with

genes involved in such pathways.

Biodegradation of 2,4-D

Several bacterial strains have been described that are able to

use 2,4-D as the sole carbon and energy source (Baelum et al.,

2010; Marron et al., 2006; Sandoval-Carrasco et al., 2013).

The most commonly cited 2, 4-D degrading genera are

Pseudomonas, Alcaligenes, Ralstonia, Delftia, Arthrobacter

and Burkholderia. Degradation of 2,4-D via oxidative cleavage

of ether bond with subsequent chlorophenol hydroxylation

followed by the modified ortho-cleavage pathway of chlor-

ocatechols has been demonstrated for most of these isolates

(Figure 2). The electron effects, the spatial orientation and the

hydrophobic effect of their substituent groups have obvious

influence on their degradation pathway. Evans et al. (1971)

isolated two strains of Pseudomonas sp. capable of degrading

2,4-D and proposed biodegradation pathway of 2,4-D. Plasmid

Cl

Cl

OO

O-

Cl

OO

O-

Cl

HO

Cl

HO

OH

TCA

Cl

Cl

HO

Cl

HO

Cl

OHClO

HO

ClO

O TCA

4Chlorophenoxyacetate

4-chlorophenol

4-chlorocatechol

3,5-DCC 2,4-DC cis muconate2,4-DCP

2,4-D

intradiol

ring cleavage

intradiol

ring cleavage

2,4-dehalogenase

monooxygenase hydroxyalase

tfdA

ketoglutarate deoxygenase

tfdB

hydroxyalasetfdC

1,2-dioxygenase

tfdDEFKR

oxoglutarate

succinate

NADPH

NADH+

Figure 2. Degradation of 2,4 D by microbes. 2,4-DCP, (2,4-dichlorophenol), 3,5-DCC (3.5-Dichlorocatechol) TCA, tricarboxylic acid cycle. Thescheme is based on articles cited in the text.

DOI: 10.3109/1040841X.2014.929564 Microbial degradation of herbicides 3

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y U

nive

rsity

of

Bat

h on

11/

02/1

4Fo

r pe

rson

al u

se o

nly.

Page 4: Microbial degradation of herbicides

involvement in the degradation of 2,4-D was first reported by

Fisher et al. (1978). They reported that ability to degrade 2,4-

D is encoded by a 58-megadalton conjugal plasmid, pJPl.

Genes encoding 2,4-D degradation are often located on

conjugative plasmids, pEMT1 and pJP4 (Bhat et al., 1994;

Chong & Chang, 2009; Don & Pemberton, 1981; Top et al.,

1995), but have been found to be also chromosomally located

in Burkholderia spp (Matheson et al., 1996). Plasmid pEMT1

contains the same degradative genes in an organization similar

to that of plasmid pJP4 but does not belong to the same

incompatibility group as pJP4. Don & Pemberton (1985)

constructed a genetic and biophysical map of pJP4 by using

transposon mutagenesis, deletion analysis, gene cloning and

restriction analyses. Plasmid pJP4 contains the genes for the

degradation of 2, 4-D to chloromaleyl acetic acid, whereas

chromosomal genes of the host are necessary for complete

mineralization of the compound. The modified ortho cleavage

pathway in degradation of 2,4-D is encoded on conjugative

plasmids (Bhat et al., 1994; Top et al., 1995; Poh et al., 2002;

Garabrant & Philbert, 2002), which give information on

movement of genes by horizontal gene transfer. The genes

responsible for 2,4-D degradation pathway (tfdA, -B, -C,-D, E

and F) are localized to the transmissible plasmid, pJP4 is well

understood and the enzymes participating in the pathway have

been purified and characterized (Amy et al., 1985; Annelie

et al., 2001; Atamaniuk et al., 2013; Evans et al., 1971). The

tfd genes encoding these enzymes have been localized, cloned

and sequenced (Don et al., 1985; Hoffmann et al., 2003;

Tiedje et al., 1969; Tsutsui et al., 2013). The regulatory

mechanisms of tfd gene expression have been elucidated

(Inoue et al., 2012). Kitagawa et al. (2002) cloned and

characterized new family of 2,4-D degradation genes,

cadRABKC, from Bradyrhizobium sp. strain HW13. The

cadR gene was inferred to encode an AraC/XylS type of

transcriptional regulator from its deduced amino acid

sequence. The cadABC genes were predicted to encode 2,4-

D oxygenase subunits from their deduced amino acid

sequences. The cadK gene was presumed to encode a 2,4-D

transport protein from its deduced amino acid sequence that

showed 60% identity with the 2,4-D transporter, TfdK, of strain

Ralstonia eutropha JMP134. Gene bioaugmentation with

conjugative plasmids-harboring bacteria capable of degrading

2,4-D on the indigenous microbial community has been

evaluated by several researchers previously (Dejonghe et al.,

2000; Don & Pemberton, 1981; Inoue et al., 2012; Newby

et al., 2000; Tsutsui et al., 2013). Hoffmann et al. (2003)

reported that the putative genes of the complete 2,4-D

degradation pathway are organized in a single genomic unit

in alkalitolerant strain Delftia acidovorans P4a.

Atrazine

Atrazine, [2-chloro-4-(ethylamino)-6-(isopropylamino)-s-tria-

zine], is a selective herbicide belonging to the family of the s-

triazines. Because of its high mobility in soil and its massive

application, atrazine has often been detected in surface and

ground waters at concentrations well above the permissible

limits (Biradar & Rayburn, 1995; Hayes et al., 2002; Kolpin &

Kalkhoff, 1993; Tappe et al., 2002). The moderate persistent,

chemical water solubility (33 mg l�1 at 20 �C) and low soil

sorption partition coefficient (Kd¼ 3.7 L kg�1) are key factors

influencing its potential contamination to aquifers, ground-

water and rain water as it normally percolates to groundwater

or river via infiltration (Siripattanakul et al., 2009). Atrazine

inhibits photosynthesis and its associated noncyclic photopho-

sphorylation is in higher plants (Shelton et al., 1996). All

higher plants probably metabolize atrazine by N-dealkylation

with some species such as corn and wheat, which contain

benzoxazinone, and utilize hydroxylation as well.

Toxicity of atrazine

The main target of atrazine in humans and animals is the

endocrine system. This herbicide is associated with relatively

high chronic toxicity. One study showed the presence of14C-labeled atrazine in soil 22 years after the last application

(Jablonowski et al., 2008), revealing the risk of chronic

atrazine exposure. Atrazine is a carcinogen that affects the

central nervous system, reproductive system, immune system

and cardiovascular function. Exposure to atrazine in animals

is associated with some types of NHL in adult humans. In

2007, the USEPA began reviewing several epidemiological

cancer studies concerning atrazine and its possible association

with carcinogenic effect in humans. Atrazine is a potential

disruptor of normal sexual development in male frogs and

also alter some aspects of the immune response (Yaw-Jian

et al., 1999). Studies in rodents suggested that atrazine causes

alterations in locomotor activity (Bardullas et al., 2011),

decreases striatal catecholamine content and decreases the

number of tyrosine hydroxylase-positive neurons (TH+) in the

substantia nigra pars compacta and ventral tegmental area

(Hayes et al., 2002). Due to ubiquitous and unpreventable

water contamination, the European Union announced a ban of

atrazine in October 2003, but same month, the USEPA

approved continued use of atrazine.

Due to high toxicity at ppb level, bioremediation in

atrazine contaminated soil and water is necessary and

received many attentions around the world.

Biodegradation of atrazine

Microbial metabolism has been regarded as the most important

mechanism of atrazine degradation in soil. A number of

microorganisms with different atrazine degradation efficien-

cies and growth characteristics have been reported (Table 1).

One of the well-known atrazine-degrading bacteria is

Pseudomonas sp. strain ADP, which can decompose atrazine

to NH3 and CO2 by some enzymes encoded by six genes,

atzABCDEF (Siripattanakul et al., 2009). Another best-

characterized strain whose atrazine metabolic track is different

from Pseudomonas sp. strain ADP is Arthrobacter aurescens

TC-1. It can metabolize atrazine to cyanuric acid through a

degrading pathway catalyzed by TrzN, AtzB and AtzC

enzymes (Figure 3). In addition, there are also some

researchers reported the hybrid pathways involving the trzDN

or atzABCDEF genes combinations found in atrazine-degrad-

ing strains or consortiums (Hess & Warren, 2002). To date, the

known genes that encode the enzymes to hydrolyze atrazine are

atzA, atzB, atzC, atzD, atzE, atzF, trzD and trzN. Although

these genes have their own substrate specificity and are

harbored in different atrazine-degrading bacteria isolated from

geographically distinct locations, nearly all of the genes are

4 B. Singh Crit Rev Microbiol, Early Online: 1–17

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y U

nive

rsity

of

Bat

h on

11/

02/1

4Fo

r pe

rson

al u

se o

nly.

Page 5: Microbial degradation of herbicides

Tab

le1

.R

epo

rted

mic

roo

rgan

ism

sca

pab

leo

fd

egra

din

gco

mm

on

lyu

sed

her

bic

ides

.

Atr

azin

e

Mic

roo

rgan

ism

Iso

lati

on

fro

mo

rso

urc

eD

egra

dat

ion

pat

hw

ayG

ene

invo

lved

ind

egra

dat

ion

pat

hw

ayR

efer

ence

sA

gro

ba

cter

ium

rad

iob

act

erJ1

4a,

So

ilM

etab

oli

tes

pro

du

ced

are

hy

dro

xyat

ra-

zin

e,d

eeth

yla

traz

ine

and

dee

thylh

yd

rox

yat

razi

ne

–S

tru

ther

set

al.,

19

98

Kle

bsi

ella

sp.

A1

and

Co

ma

mo

na

ssp

.S

ewag

eo

fa

pes

tici

de

mil

lH

yd

rox

yat

razi

ne,

N-i

sop

rop

yla

mm

elid

ean

dcy

anu

ric

acid

atz

A,

B,

C,

D,

E,

Fgen

ew

ere

clo

ned

Yan

get

al.,

20

10

Art

hro

ba

cter

sp.

stra

inD

NS

10

,B

lack

soil

–a

tzA

,a

tzD

,a

tzE

,a

tzF

and

trzD

trzN

,a

tzB

and

atz

Cgen

ew

ere

clo

ned

Zh

ang

etal

.,2

01

1

Rh

od

oco

ccu

ssp

.st

rain

MB

-P1

So

ilD

e-et

hyla

traz

ine

and

de-

iso

pro

pyla

traz

ine

Pla

smid

-en

cod

edF

azlu

rrah

man

etal

.,2

00

9A

rth

rob

act

ersp

.st

rain

HB

-5In

du

stri

alw

aste

wat

erH

yd

rox

yat

razi

ne

and

cyan

uri

cac

id–

Wan

get

al.,

20

11

Pse

ud

om

on

as

stra

inY

AY

A6

Gar

den

soil

Deg

rad

atio

np

roce

eded

via

dec

hlo

rin

atio

nan

dN

-dea

lkyla

tio

n–

Yan

ze-K

on

tch

ou

&G

schw

ind

,1

99

4A

rth

rob

act

ersp

.st

rain

DA

T1

So

ilD

egra

dat

ion

gen

estr

zN,

atz

Ban

da

tzC

on

pla

smid

DN

AW

ang

etal

.,2

01

3

Art

hro

ba

cter

sp.

TE

S6,

Agri

cult

ura

lso

iltr

zN,

atzB

,an

dat

zC.

Seb

aıet

al.,

20

11

Clo

din

afo

p-p

rop

arg

yl

(CF

)P

seu

do

mo

na

ssp

.st

rain

B2

Cro

pfi

eld

soil

.E

ster

ase

acti

vit

y:

stra

inB

2d

egra

ded

CF

tocl

od

inaf

op

acid

and

4-(

4-c

hlo

ro-2

-fl

uo

ro-p

hen

ox

y)-

ph

eno

lw

ith

in9

h.

–S

ing

het

al.,

20

13

Rh

od

oco

ccu

sw

rati

sla

vien

sis

So

ilC

lod

inaf

op

acid

–Ja

qu

etet

al.,

20

14

2,4

-DB

rad

yrh

izo

biu

msp

.st

rain

HW

13

Bu

ried

soil

-2

,4-D

deg

rad

atio

ngen

es,

cad

RA

BK

CK

itag

awa

etal

.,2

00

2S

erra

tia

ma

rces

cen

san

dP

enic

illi

um

spB

razi

lian

soil

con

tam

inat

edw

ith

2,4

-Dh

erb

icid

e2

,4-d

ich

loro

ph

eno

l(2

,4-D

CF

)–

Sil

va

etal

.,2

00

7

Alc

ali

gen

eseu

tro

ph

us

So

il2

,4-d

ich

loro

ph

eno

l2

,4-D

mo

noxy

gen

ase

Per

kin

s&

Lu

rqu

inet

al.,

19

88

Ra

lsto

nia

eutr

op

ha

JMP

13

4H

awai

ian

soil

Ph

eno

ld

eriv

ativ

esca

dR

AB

KC

Kit

agaw

aet

al.,

20

02

Met

ola

chlo

rF

usa

riu

msp

.,M

.ra

cem

osu

sM

eto

lach

lor-

trea

ted

farm

soil

,2

-ch

loro

-N-(

2-e

thy

l-6

-hy

dro

xy

met

hylp

he-

nyl)

-N

-(2

-met

hox

y-1

-met

hyle

thy

l)-

acet

amid

e

–S

axen

aet

al.,

19

87

Azo

tob

act

ersp

.st

rain

SS

B8

1.

–4

-ch

loro

ph

eno

lan

d4

-ch

loro

cate

cho

l,O

xid

ativ

ep

ath

way

Gau

riet

al.,

20

12

Gly

ph

osa

teO

chro

ba

ctru

ma

nth

rop

iG

PK

3,

Ach

rom

ob

act

ersp

.S

od

-po

dzo

lso

ilH

igh

deg

rad

atio

nca

pac

ity

wit

ho

ut

the

accu

mu

lati

on

of

any

met

abo

lite

,in

par

-ti

cula

r,am

ino

met

hylp

ho

sph

on

icac

id

–E

rmak

ova

etal

.,2

01

0

Tri

cho

der

ma

viri

de

stra

inF

RP

3F

ore

stso

il–

–A

rfar

ita

etal

.,2

01

3F

usa

riu

mst

rain

s9

11

48

(Fu

sari

um

oxy

spo

rum

)Is

ola

ted

fro

msu

gar

can

eA

min

om

eth

ilp

ho

sph

on

icac

id(A

MP

A).

–C

astr

oJr

etal

.,2

00

7

Pse

ud

om

on

assp

.st

rain

PG

29

82

.G

lyp

ho

sate

pro

cess

was

test

ream

Pri

mar

ym

etab

oli

team

ino

met

hylp

ho

-sp

ho

nat

ew

aso

bse

rved

,5

%o

fth

eg

ly-

ph

osa

tew

asd

egra

ded

by

ase

par

ate

pat

hw

ayin

vo

lvin

gb

reak

dow

no

fg

lyp

ho

-sa

teto

gly

cin

e.

–Ja

cob

etal

.,1

99

8

Fla

vo

bac

teri

um

spec

ies.

Ind

ust

rial

acti

vat

edsl

ud

ge

AM

PA

–D

iuro

nA

rth

rob

act

erst

rain

sD

47

Bri

tish

and

Fre

nch

agri

cult

ura

lso

ils

Am

mo

nia

and

CO

2.

–T

urn

bu

llet

al.,

20

01

3,4

-dic

hlo

roan

ilin

e,C

O2

.–

ren

sen

etal

.,2

00

8

(co

nti

nu

ed)

DOI: 10.3109/1040841X.2014.929564 Microbial degradation of herbicides 5

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y U

nive

rsity

of

Bat

h on

11/

02/1

4Fo

r pe

rson

al u

se o

nly.

Page 6: Microbial degradation of herbicides

generally highly conserved and plasmid borne. Furthermore,

associating with transposon, some of the genes could widely

spread between bacteria regardless of their genera. This may

enable more microbes or consortiums to mobilize atrazine and

mobilizes some new catabolic pathways.

The degradation of atrazine occurs predominantly by

biological processes, including N-dealkylation, dechlorination

and ring cleavage. Atrazine biodegradation can be initiated by

N-dealkylation of the ethyl or isopropyl side chains to

produce deethylatrazine (DEA) or deisopropylatrazine (DIA).

Dechlorination has been reported as an early step in atrazine

metabolism, and two different s-triazine hydrolase enzymes

have been characterized. In some microorganisms, complete

biodegradation of atrazine to ammonia and CO2 has been

obtained. Pseudomonas sp. strain ADP might be the best-

characterized atrazine mineralizing one (Mandelbaum et al.,

1993). The genes encoding the three enzymes that are

responsible for the conversion of atrazine to cyanuric acid

were atzA, B and C. The research on atrazine-degrading

microorganisms has been directed to the isolation and

characterization of natural occurrence lineages in environ-

ments contaminated with this pesticide. For the purpose of

potential bioremediation practice, a large variety of atrazine-

degrading bacteria from diverse genera have been isolated

(Rousseaux et al., 2001). Among bacteria, there are reports

on atrazine degradation by individual strains such as

Pseudomonas sp. (Mandelbaum et al., 1995), Rhodococcus

rhodochrous, Acinetobacter spp., Agrobacterium sp.,

Microbacterium sp., Bacillus sp., Micrococcus sp.,

Deinococcus sp. and D. acidovorans (Vargha et al., 2005),

as well as by species consortia including Agrobacterium

tumefaciens, Caulobacter and Pseudomonas sp. ADP is now

the most known and the best-characterized atrazine-degrading

bacterium (Wackett et al., 2002). However, microorganisms

from genus Arthrobacter are well known for their strong

capacity to degrade atrazine and have been isolated from

agricultural and heavily contaminated soils at spill sites and

industrial wastewaters from atrazine production plants (Zhou

et al., 2012). The atrazine-degrading bacteria generally initi-

ate the degradation through a hydrolytic dechlorination,

catalyzed by the enzyme atrazine chlorohydrolase (AtzA),

encoded by the atzA gene, followed by two hydrolytic

deamination reactions catalyzed by hydroxyatrazine ethylami-

nohydrolase (AtzB) and N-isopropylammelide isopropyl-

amino-hydrolase (AtzC), encoded by the genes atzB (trzB) e

atzC (trzC), respectively (De Souza et al., 1998), which

convert atrazine sequentially to cyanuric acid that is then

completely mineralized to CO2 and NH3 by other three

hydrolases. In some bacterial strains, the biodegradation of

atrazine initiate through N dealkylation of the lateral ethyl and

isopropyl chains to DEA and DIA (Kaufman & Blake, 1970).

Pseudomonas sp. ADP is the best-characterized bacterial strain

capable to degrading the herbicide atrazine. The atrazine

catabolic pathway in this bacterium contains six enzymatic

steps encoded by atzABC and the atzDEF genes. The atzABC

genes have been shown to be widespread and plasmid borne in

a number of bacteria isolates (De Souza et al., 1998:

Rousseaux et al., 2001; Topp et al., 2000; Wackett et al.,

2002). In Pseudomonas sp. ADP, the atzABCDE genes are

harbored on the catabolic plasmid pADP-1 (Martinez et al.,

Tab

le1

.C

on

tin

ued

Atr

azin

e

A.

glo

bif

orm

isst

rain

D4

7an

dV

ari

ovo

rax

sp.

SR

S1

6C

on

tam

inat

edso

ilan

dw

ater

reso

urc

esB

.b

ass

ian

a,

C.

eleg

an

s,A

gri

cult

ura

lan

dn

on

-ag

ricu

ltu

ral

soil

s3

,4-d

ich

loro

anil

ine,

Am

mo

nia

and

CO

2.

pu

hA

gen

eC

asti

llo

etal

.,2

00

6

Mic

roco

ccu

ssp

.st

rain

PS

-1D

iuro

nst

ora

ge

site

Deg

rad

atio

np

roce

eded

via

de-

alk

yla

tio

nan

dfi

nal

lysy

nth

esiz

ing

amin

od

eriv

a-ti

zed

met

abo

lite

–S

har

ma

&S

uri

,2

01

1

Par

aquat

Asp

erg

illu

sn

iger

span

dP

seu

do

mo

na

ssp

UK

san

dy

loam

soil

sC

arb

on

dio

xid

eas

the

ult

imat

em

iner

al-

izat

ion

pro

du

ct–

Ric

ket

tes

etal

.,1

99

9

Lip

om

yces

sta

rkey

iC

lay

soil

Ace

tic

acid

and

carb

on

dio

xid

e–

Car

ret

al.,

19

86

Pen

dim

ath

alin

Fu

sari

um

oxy

spo

rum

,A

sper

gil

lus

ory

zae,

Len

tin

ula

edo

des

,P

enic

illi

um

bre

vico

mp

act

um

and

Lec

an

icil

liu

msa

ksen

ae

So

ilD

eset

hy

l-te

rbu

thyla

zin

e–

Pin

toet

al.,

20

12

6 B. Singh Crit Rev Microbiol, Early Online: 1–17

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y U

nive

rsity

of

Bat

h on

11/

02/1

4Fo

r pe

rson

al u

se o

nly.

Page 7: Microbial degradation of herbicides

2001). In Pseudomonas sp. ADP, the atzDEF operon encodes

cyanuric acid amidohydrolase (AtzD), biuret amidohydrolase

(AtzE) and allophanate hydrolase (AtzF), involved in cleavage

of the cyanuric acid to carbon dioxide and ammonia, which is

assimilated as a nitrogen source (De Souza et al., 1998).

Garcia-Gonzalez et al. (2003) have demonstrated that nitrogen

control of atrazine metabolism is functional under soil

conditions and may therefore limit the potential of

Pseudomonas sp. strain ADP for atrazine bioremediation in

nitrogen-fertilized agricultural soils. The atzABC genes are

constitutively expressed and are not regulated either by

induction of atrazine or by repression of other N sources in

this strain (Martinez et al., 2001; Devers et al., 2007). The

atzDEF genes are divergently transcribed from AtzR, predicted

N

N

N

Cl

HN NHCH3

NH2

NH2

NH2

H3C

H2N

H2N

H2N

H3C

H2N

CH3

H3C CH3

CH3

CH3

CH3

CH3

H3C CH3

N

N

N

OH

NH NH

Atrazine

Hydroxyatrazine

N

N

N

OH

NH OH

N-isopropylammelide

N

N

N

OH

O HOH

Cyanuric acid

N

N

N

Cl

HN

deethylatrazine

N

N

N

Cl

NH

deisopropylatrazine

N

N

N

Cl

deisopropyl deethylatrazine

N

N

N

Cl

OH

N

N

N

OH

OH

ammelide

N

N

N

Cl

O HNH

N-ethylammelide

NH

OH

O O

+NHO O

BiuretAllophanate

AC

HAEA

IAIA

AM DEAM

TH

HAEA

IAIA

TC

EAA

CAH

BH AH

2-chloro 4-hydroxy 6-amino 1,3,5-triazine

atzA

atzB

atzC

atzD

atz E atzF

NH3

NH2NH2

CO2

Figure 3. Atrazine biodegradation along with gene and enzymes involved in degradation pathways: AC, atrazine chlorohydrolase; HAEA,hidroxyatrazine ethylaminohydrolase; IAIA, N-isopropylammelide isopropylamidohydrolase; TC, s-triazine chlorohydrolase; AM, atrazinemonooxygenase; DEAM, deethylatrazine monooxygenase; DIHA, deisopropyhidroxylatrazine amidohydrolase; EAA, N-ethylammelide amidohy-drolase, TH, s-triazine hydrolase; CAH, cyanuric acid hydrolase; BH, biuret hydrolase; and AH, allophanate hydrolase. The scheme is based onarticles cited in the text.

DOI: 10.3109/1040841X.2014.929564 Microbial degradation of herbicides 7

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y U

nive

rsity

of

Bat

h on

11/

02/1

4Fo

r pe

rson

al u

se o

nly.

Page 8: Microbial degradation of herbicides

to encode a transcriptional LysR type regulator (LTTR). A

putative LTTR-binding site can in fact be found upstream of

atzD gene, thereby suggesting that transcription of the atzDEF

operon may be regulated and the protein encoded by the orf99

(AtzR) play a role in this regulation. The atzDEF operon

resides in a contiguous cluster adjacent to the orf99, a potential

transcriptional LTTR (Martinez et al., 2001). Atrazine bio-

degradation can be positively or natively affected by the

external addition of organic carbon sources (Abdelhafid et al.,

2000) and is dependent on types of added carbon sources found

that the addition of citrate to soils resulted in enhanced atrazine

degradation by Arthrobacter sp. strain KU001. A good

candidate for bioaugmentation is the Arthrobacter strain

DAT1 that has shown high atrazine-degrading efficiency

both in liquid cultures and soils (Wang & Xie, 2012). This

strain can utilize atrazine as a sole nitrogen source for growth

and harbors the atrazine-metabolic genes trzN, atzB and atzC.

The cited literatures on biodegradation of atrazine help us

to know the metabolic ability and gene characteristics of

strains involved in biodegradation of atrazine. However, there

is need to study strains evolution by comparing the metabolic

ability and characterizing genes involved in atrazine

biodegradation.

Metolachlor

Metolachlor (2-chloro-N-(2-ethyl-6-methylphenyl)-N-(meth-

oxyprop-2-yl)acetamide) is a selective chloroacetamide herbi-

cide used to control broadleaf and annual grass weeds in corn

(Zea mays L.), soybean (Glycine max L. Merr.), peanut

(Arachis hypogaea L.) and potato (Solanum tuberosum L.)

(Gaynor et al., 1993; Thurman et al., 1992). Large quantities

of metolachlor (22 million kg of active ingredient) are applied

to agricultural fields in the United States, particularly in the

Midwest, where most of the corn and soybeans are grown.

Metolachlor is a member of the chloroacetanilide herbicide

chemical family. Other members include acetochlor, alachlor,

butachlor, butenachlor, delachlor, diethatyl, dimethachlor,

metazachlor, propachlor, propisochlor, prynachlor, terbuchlor,

thenylchlor. When it is absorbed through the roots and shoots, it

acts as a growth inhibitor by suppressing synthesis of chloro-

phyll, proteins, fatty acids and lipids, isoprenoids (including

gibberellins) and flavonoids (including anthocyanins).

Toxicity of metolachlor

Studies have shown that metolachlor affects cell growth, and its

low levels cause cytotoxic effects in lymphocytes along with a

significant decrease in mitotic index (Rollof et al., 1992).

Studies with HepG2 cells, an immortalized human liver cell

line, exposed to 100 ppb metolachlor for 24 h had 21% fewer

numbers of cells compared to nontreated control cells (Hartnett

et al., 2013). Recently, Lowry et al. (2013) studied decrease

of HepG2 cell growth after metolachlor exposure and reported

that the levels of the retinoblastoma protein including two of its

hyperphosphorylated forms are decreased in metolachlor

exposed cells possibly leading to cell cycle arrest.

Biodegradation of metolachlor

A major breakdown pathway of metolachlor in the soil

is by both aerobic and anaerobic microorganisms. The

transformation by soil microorganisms of metolachlor to its

primary degradates: metolachlor ethane sulfonic acid (ESA)

and metolachlor oxanilic acid (OA) (Figure 4) has been

suggested to occur as a result of shifting of chlorine atom of the

parent compound by glutathione, followed by the formation of

the ESA and OA degradates by different enzymatic pathways

(Barbash et al., 1999). Degradation of metolachlor in soil

occurs mainly by microbial decomposition (Xu et al., 2008)

and photo-degradation. Microbial degradation rates are

affected by soil depth, organic carbon and dissolved oxygen

concentrations, temperature and size of microbial populations.

In case of sandy soils, a half-life of 81 d for anaerobic microbial

populations and 67 d for aerobic microbial populations were

reported in the laboratory. Photo-degradation occurs only when

metolachlor is present on the soil surface. Fifty percent of

surface-applied metolachlor can be degraded in eight days on

soil, while only 6% degrades over one month in soils where

metolachlor was incorporated into the surface layer.

Diuron

Diuron (3,4-dichlorophenyl)-1,1-dimethylurea) is a systemic

substituted wide-spectrum phenylurea herbicide used for

weed control in agricultural crops and non crops areas

(Castillo, 2006; Stasinakis et al., 2009). Non crops areas

include along fence lines, pipelines, powerlines, railway lines,

roads, footpaths; in timber yards and storage areas; and

around commercial, industrial and farm buildings, electrical

substations and petroleum storage tanks. Diuron is used as an

algaecide in ornamental ponds, fountains and aquaria and

mildewcide in paints. Diuron is often used either alone or in

combination of other herbicides such as bromacil, hexazi-

none, PQ, thiadiazuron, imazapyr, monosodium, sodium

chlorate, sodium metaborate and copper sulfate (USEPA,

2004). Diuron is available in wettable powder, granular,

flowable, pelleted/tableted, liquid suspension and soluble

concentrate formulations. Technical grade (the grade that is

usually used for agricultural purposes) diuron is a white,

crystalline and odorless solid.

Diuron is a systemic substituted phenylurea herbicide. It is

easily taken up from soil solution through the root system of

plants and rapidly translocated into stems and leaves by the

transpiration system, moving mainly via the xylem. Diuron

primarily functions by blocking the Hill reaction in

Metolachlor [2-chloro-N-ethylphenyl)-N-(2-methoxy-1-methylethyl) acetamide]

(2-ethyl-6-methylphenyl)(2-methoxy-1-methylethyl)amino oxo-acetic acid

Metolachlorethane sulfonic acid

Metolachlor oxanilic acid

morpholinone + Carbinol

CO2 + H2O

Figure 4. Biodegadation of metolachlor. The scheme is based on articlescited in the text.

8 B. Singh Crit Rev Microbiol, Early Online: 1–17

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y U

nive

rsity

of

Bat

h on

11/

02/1

4Fo

r pe

rson

al u

se o

nly.

Page 9: Microbial degradation of herbicides

photosynthesis process, limiting the production of high-

energy compounds such as ATP, which is used for several

metabolic processes. Diuron binds to the QB-binding site on

D1 protein of the photosystem-II complex in chloroplast

(thylakoid membranes), thus blocking electron transport from

QA to QB. This process prevents CO2 fixation and the

production of ATP and other high energy compounds, which

are needed for plant growth. The inability to reoxidize QA

promotes the formation of triplet state chlorophyll, which

interacts with ground state oxygen to form singlet oxygen.

Both triplet chlorophyll and singlet oxygen can extract

hydrogen from unsaturated lipids, producing a lipid radical

and initiating a chain reaction of lipid per-oxidation. Lipids

and proteins are attacked and oxidized, resulting in loss of

chlorophyll and carotenoids, and in leaky membranes, which

cause cells and cell organelles to dry and disintegrate rapidly

(Hess & Warren, 2002).

Toxicity of diuron

Diuron is highly toxic to aquatic organisms, LC50 (48 h)

values for diuron range from 4.3 to 42 mg/L in fish, and from

1 to 2.5 mg/L for aquatic invertebrates/organisms. The LC50

(96 h) is 3.5 mg/L for rainbow trout (EXTOXNET et al.,

1996). Residues of diuron was detected in milk, fat body,

muscles, liver and kidney of cows, cows fed with very low

doses of diuron in their diets (EXTOXNET et al., 1996).

Domingues et al. (2011) studied diuron exposure in male

Wister rats and they found that diuron exerts systemic and

target-organ toxicity, mainly at higher concentration. This

research records the potential mutagenic, teratogenic, repro-

ductive and carcinogenic effects of diuron (Federico et al.,

2011) investigated the genotoxic effects of chlorotoluron,

diuron and difenoxuron by analyzing chromosomal and sister

chromatid exchange in exposed mammalian cells (Chinese

hamster). They found that phenylurea herbicides induced

direct genotoxic activity, but the cytogenetic effects were

greatly enhanced after metabolic conversion.

Biodegradation of diuron

Diuron is susceptible to degradation by soil microorganisms,

and enriched cultures of aquatic microorganisms from pond

water could also degrade diuron to 3,4-dichloroaniline as a

major metabolite (Figure 5) (Fratila-Apachitei et al., 1999).

Bogaerts et al. (2000) studied microbial degradation of

diuron and ecotoxicology to investigate its breakdown after

application to soils. Quantitative biodegradation assays were

executed with fungal strains, showing that diuron was

degraded but not entirely. A series of tests were taken out

to select the most efficient fungal strain for diuron degrad-

ation. Among the fungal strains, only four strains were able to

transform diuron to an extent (up to 50%) after seven days of

incubation: Beauveria bassiana, Caenorhabditis elegans,

Phanerochaete chrysosporium and Mordellistena isabellina.

Diuron degradation by the fungal strains led to the formation

of two metabolites obtained in different proportions according

to the microorganism. For the fungal strains, diuron degrad-

ation led to the formation of the demethylated products. The

identified metabolites were synthesized in sufficient amounts

to confirm their structures and determine their non-target

toxicity using four biotests. According to the Microtox test,

the metabolites N-(3,4-dichlorophenyl)-N-methylurea and

N-3, 4-dichlorophenylurea presented a three times higher

toxicity than that of diuron. Dalton et al. (1966) reported that

the removal of N-methyl groups eliminates herbicidal activity

of diuron. Decomposition is followed by the removal of the

urea group, which results in the formation of 3,4-dichloroani-

line, ammonia and CO2.

Glyphosate

GP [N-phosphonomethyl glycine] is an active ingredient of

herbicides applied to annual and perennial weeds. GP inhibits

5-enolpyruvilshikimate-3-phosphate synthase, the enzyme of

shikimic acid synthesis that participates in the synthesis of

aromatic amino acids. GP is a broad-spectrum, non-selective

systemic herbicide that can be used to control most weeds,

both annual and perennial plants, under many varied

situations such as agriculture, forestry, orchards, vineyards,

industry and no-till cropping systems and has been domes-

ticity classified as an easily degradable herbicide in the past

(Celis et al., 2008). Although GP is believed to be a relatively

safe compound, in our days, reports can be found that GP has

negative effects on human health. In addition, nowadays, GP

and one of its principal metabolites (aminomethylphosphonic

acid) have frequently been detected in ground water (Laitinen

et al., 2007; Meza-Joya et al., 2013). The application of GP

results in the yellowing and decay of leaves within 5–10 d

Figure 5. Biodegradation of diuron: DCPU[1-(3,4-dichlorophenyl) urea], DCPMU[3-(3,4-dichlorophenyl)-1-methylurea].

Cl

Cl NH

N

O

Diuron

Cl

Cl NH

NH

O

DCPMU

Cl

Cl NH O

DCPU

Cl

Cl

dichloroaniline

Cl

NH O

monuron

NH2

NH2

NH2

H3C CH3CH3

DOI: 10.3109/1040841X.2014.929564 Microbial degradation of herbicides 9

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y U

nive

rsity

of

Bat

h on

11/

02/1

4Fo

r pe

rson

al u

se o

nly.

Page 10: Microbial degradation of herbicides

(sometimes 30 d) caused by the breakdown of aromatic amino

acids synthesis. Penaloza-Vazquez et al. (1995) reported that

GP remains unchanged in the soil for varying lengths of time,

because of its adsorption on clay particles and organic matter

present in the soil. This condition makes this herbicide very

persistent in soils and sediments.

Toxicity of GP

Application of GP in the agriculture for a long period of time

increases the susceptibility of crops to diseases and in arable

areas causes dieback in hedgerow trees. In soil, GP molecules

may be adsorbed on humus and may also form complexes

with metal cations (Fe2+ Cu2+, Mn2+ and Ni2+) (Veiga et al.,

2001). Soil amendment with P fertilizers – sources of

phosphate ions, which compete with GP for binding sites –

leads to accumulation of free GP, available for plant root

uptake and for microbial metabolism (Simonsen et al., 2008).

GP absorbed into the plant not only mainly through its foliage

but also through soft stalk tissue. It is then translocated to

growing points of the plant where it acts on various enzyme

systems inhibiting aromatic amino acids: tyrosine, tryptophan

and phenylalanine that are essential for protein formation and

secondary products in susceptible plants. This pathway also

works in higher plants and microorganisms but not in animals.

However, GP-containing products are acutely toxic to animals

(Cox, 1995). Plants treated with GP slowly die over a period

of days or weeks, and because the chemical is transported

throughout the plant, no part survives. GP is chosen for early-

season weed control before planting and after harvesting

(Duke & Powles, 2008; Laitinen, 2007). In reduced tillage or

no-till cultivations, GP is used to prepare fields before

planting, during crop development and post harvest.

Biodegradation of GP

GP is an amphoteric and non-volatile compound, no

photodegradation happens and it is stable in air. It is

practically insoluble in most of organic solvents, for instance,

ethanol, acetone and benzene, because of its high polarity, but

it is completely soluble in water (Laitinen et al., 2007). The

degradation of GP is slower in soils with a higher adsorption

capacity. Degradation rate was also affected by the particular

microbial community of each soil (Carlisle & Trevors, 1986).

Microorganisms known to degrade GP include Pseudomonas

sp., Arthrobacter atrocyaneus and Flavobacterium sp.

(Table 1). The primarily metabolite of GP is aminomethyl-

phosphonic acid (Figure 6), which is non-toxic and degraded

by microbes at a somewhat slower rate than the parent

compound (Rueppel et al., 1977; Carlisle & Trevors, 1986).

Imazapyr

Imazapyr [2-(4,5-dihydro-4-methyl-4-(1-methylethyl)-5-oxo-

1H-imidazol-2-yl)-3-pyridinecarboxylic acid] also known as

Chopper, Arsenal and Assault is a non-selective broad-

spectrum systemic herbicide used for control of annual and

perennial grasses, broadleaf branch, sedge weeds and a

variety of shrubs and deciduous trees. It was first registered in

the United States in 1984. Imazapyr is absorbed by the foliage

and roots, with rapid transfer to the xylem and phloem to the

meristematic regions, where it accumulates and causes

disruption of protein synthesis and interferes with cell

growth and DNA synthesis. The result of exposure is death

of new leaves. Imazapyr comes in two forms: weak acid and

isopropylamine salt (49% water solution), although commer-

cial products are almost solely the isopropylamine salt

form. Imazapyr is persistent in the soil with the half-life of

17 months (USEPA, 1984).

Toxicity of Imazapyr

Imazapyr is considered as non-toxic to animals; however, it is

irritating to the eyes and can cause rashes, redness and

swelling at the site of exposure. It stops the biosynthesis of

essential chemical (aliphatic three branched-chain amino

acids) in plants. Animals also require these chemicals but they

do not synthesize it and take directly from plants. Therefore,

specific toxic affect can only happen in a plant, and toxic

affects in animals occur at a much higher dose.

Biodegradation of imazapyr

The herbicide is commonly found in two forms: weak acid and

isopropylamine salt (Figure 7), although commercial products

are almost solely the isopropylamine salt. Biodegradation of

imazapyr by Streptomyces sp. strain PSI/5 was first observed

by Shelton et al. (1996). Xuedong et al. (2005) isolated two

Figure 6. Biodegradation of glyphosate. Thescheme is based on articles cited in the text.

Arthrobacter atrocyaneusEnterobacter aerogenes

P

HO

O

NH

O

OH

NH

O

OH

GlyphosateSarcosine

P

HO OH

OH

OOMAD

FormaldehydeMethlyamineAMPA

C-P lyase

C-P lyase

Intermediary Metabolism

C1 metabolic cycle

GD

Geobacillus caldoxylosilyticusFlavobacterium sp.

NH2NH2

H2CH3C

H3C

10 B. Singh Crit Rev Microbiol, Early Online: 1–17

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y U

nive

rsity

of

Bat

h on

11/

02/1

4Fo

r pe

rson

al u

se o

nly.

Page 11: Microbial degradation of herbicides

bacterial species, Pseudomonas and Bacillus, from soil that

were capable of degrading imazapyr. They found degradation

rate of more than 70% at 50 or 100 ppm concentration of

initially added imazapyr to the culture after 48 h of incubation.

Further studies are needed to study the mechanism of

biodegradation of imazapyr by microbes.

Pendimethalin

Pendimethalin (N-(1-ethylpropyl)-3, 4-dimethyl-2, 6-dinitro-

benzenenamine, a dinitroaniline herbicide) is both a pre-

emergence and early post-emergence herbicide, used for

control of most annual grasses and many annual broad-leaved

weeds in crop fields. Pendimethalin was first registered as a

pesticide in the United States in 1972. Pendimethalin has

relatively long persistence in soil due to immobilization with

low leaching potential or due to hydrophobic nature of

pendimethalin adsorb strongly to organic matter and clay

minerals. It degrades more rapidly in anaerobic soil than in

aerobic soil conditions (Megadi et al., 2010). Its mode of

action is inhibition of mitotic cell division in developing root

systems.

Toxicity of pendimethalin

The USEPA has classified pendimethalin as persistent

bioaccumulatice toxic; it is of low acute toxicity, but causes

thyroid follicular cell adenoma. It is highly toxic to fish and

aquatic invertebrates. It is moderately persistent in aerobic

soil environments. It is therefore important to investigate the

degradation mechanism of pendimethalin.

Biodegradation of pendimethalin

Biodegradation of pendimethalin in soil was reported by

several investigators under both aerobic and anaerobic

conditions with bacteria and fungi in different types of soils

(Zheng & Cooper, 1996). Four metabolites were formed and

identified as N-(1-ethylpropyl)-3,4-dicarboxy 2,6-dinitroben-

zenamine-N-oxide, N-(1-ethylpropyl)-3,4-dimethoxy-2,6-

dinitrobenzenamine and benezimadazole-7-carboxyaldehyde

(Figure 8). The reactions involved were monohydrolysis of 2-

methyl groups followed by dihydrolysis. Further oxidation of

amine groups and hydroxylation of propyl groups produced

the above-mentioned metabolites.

Paraquat

PQ (1,1-dimethyl-4,4-bipyridinium) is a quaternary nitrogen

herbicide widely used for control of broadleaf weed. It is

nonselective compound that destroys green plant tissue on

contact by disrupting photosynthesis and rupturing cell

membranes, which allows water to escape leading to rapid

desiccation of foliage. PQ has strong affinity to bound to clay

minerals and organic matter in the soil.

Toxicity of PQ

PQ is toxic to living organisms including human beings as it is

easily absorbed practically by all routes including the gastro-

intestinal tract, skin, mucous membranes and lungs (Huang

et al., 2012). PQ poisoning is a major medical problem in Asia

where it is used for suicides in rural communities (Wua et al.,

2013). The acute toxic effects of PQ in the lung, liver and

kidney have been well recognized in animals (Huang et al.,

2012; Wua et al., 2013). In toxicity studies using laboratory

animals, PQ has been shown to be highly toxic by the inhalation

route and has been placed in Toxicity Category I (the highest of

four levels) for acute inhalation effects.

Baldwin et al. (1966) studied toxicity of PQ and reported

that PQ caused deterioration in the lungs. A dermal toxicity

study using rabbits resulted in scabbing and inflammation

when tested at the two highest doses (2.6 mg cation/kg group

and 6.0 mg cation/kg group). In an inhalation toxicity study,

rats were exposed to respirable aerosols (particle size less

than 2 um in diameter) of PQ dichloride, which resulted

in lung changes and extensive sores and swelling in the

larynx.

Biodegradation of PQ

PQ is considered a toxic to soil fungi and bacteria causing a

reduction in their population (Sahid et al., 1992). Several

bacterial and fungal isolates obtained from soil and

Figure 7. Biodegradation of imazapyr.

N

O

OH

N

NH

O

N

O

O−

N

NH

O

NH+

Imazapyr acid Imazapyr isopropylamine salt

Pyridine hydroxy-dicarboxylic acid Nicotinic acid

Pyridine dicarboxylic acid (quinolinic acid)

degraded into three possible product

CH3CH3

CH3

CH3CH3

CH3

CH3H3C

DOI: 10.3109/1040841X.2014.929564 Microbial degradation of herbicides 11

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y U

nive

rsity

of

Bat

h on

11/

02/1

4Fo

r pe

rson

al u

se o

nly.

Page 12: Microbial degradation of herbicides

wastewater can degrade PQ. The contact herbicide PQ is used

to control weeds in a wide range of crops. When PQ enters the

soil environment, it is rapidly and strongly bound to clay

minerals and organic matter and deactivated. Various studies

have used pure cultures of soil micro-organisms to elucidate

the degradative pathways of ring-labeled PQ. These studies

established the range of bacteria and fungi able to degrade

PQ (e.g. Corynebacterium fascians Dows, Lipomyces starkeyi

Loo and Rij, Aspergillus niger van Teigh, Penicillium

frequentans West, Fusarium sp and Pseudomonas sp) and

that conditions in soil solution are conducive to the degrad-

ation of PQ. In majority of the cases, however, PQ degrad-

ation was shown to be extremely variable and evidence of

only one degradation product other than CO2 has been

reported (Figure 9), the N-methyl betaine of isonicotinic acid

(Funderburk & Bozarth, 1967). The latter product has,

however, been shown to be a major intermediate in the

photolytic degradation of PQ (Slade, 1965).

NHO

O

O

O

NHO

+

Pendimethalin

6 -Aminopendimethalin

3,4-dimethyl -2,6-dinitroaniline Pentane

dealkylation

nitroreduction

CH3

NHO

O

OH

NHO

O

OOH

NHO

O

NH

N

1 2

3

4

O−

O−

O−

O−O−

O−

O−

O−

O−

O−

O−

N+N+

N+N+

N+

N+ N+

N+ N+

N+ N+

H3C

CH3

CH3

CH3

CH3

CH3

CH3 CH3

CH3

CH3CH3

CH3

CH3

CH3

CH3

CH3

CH3

H2N

CH3

CH3

CH3CH3

Figure 8. Biodegradation of pendimathalin. Four metabolites were formed and identified as N-(1-ethylpropyl)-3,4-dicarboxy 2,6-dinitrobenzenamine-N-oxide, N-(1-ethylpropyl)-3,4-dimethoxy-2,6-dinitrobenzenamine and benezimadazole-7-carboxyaldehyde. The reactions involved weremonohydrolysis of 2-methyl groups followed by dihydrolysis. Further oxidation of amine groups and hydroxylation of propyl groups produced theabove-mentioned metabolites.

12 B. Singh Crit Rev Microbiol, Early Online: 1–17

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y U

nive

rsity

of

Bat

h on

11/

02/1

4Fo

r pe

rson

al u

se o

nly.

Page 13: Microbial degradation of herbicides

Factors effecting microbial degradation of herbicides

Based on cited literature, it has been found that the inoculum

size, amounts of additional co-substrates carbon and nitrogen

compounds, organic matter of soil and pH are the major

factors that affected the extent and rate of herbicides

degradation.

Agricultural soils are rich in nitrogen due to routine

fertilization, and most atrazine-degrading bacteria use atra-

zine as a nitrogen source. However, Yang et al. (2010)

reported members of Klebsiella sp. A1 and Comamonas sp.

A2 capable of degrading atrazine and are insensitive to

exogenous nitrogen sources. Wang et al. (2013) reported

addition of both carbon and nitrogen sources promotes

degradation rate of atrazine. Struthers et al. (1998) reported

that addition of Agrobacterium radiobacter J14a cells in soil

resulted in two to five times higher mineralization of atrazine

than in the noninoculated soil. However, sucrose addition did

not result in significantly faster mineralization rates or shorten

degradation lag times.

The mineralization of GP in soils is individually regulated

and correlated by exchangeable H+, soil pH, oxalate extract-

able Al3+ and bacterial cell numbers. Recently, Al-Rajab &

Hakami (2014) studied GP degradation in three agriculture

soil and reported rapid degradation with a half-life of 14.5 d in

the silt clay loam soil incubated at 20 �C.

Johnson & Sims (2011) studied solvent toxicity in soil for

bioavailability of 2,4-D toward microorganisms and suggested

that solvent toxicity should be balanced with uniformity of

substrate distribution when using organic carriers in soils.

Conclusion and future directions in biodegradationof herbicides

Due to excessive use of herbicides, ecosystems are under

threat of its pollution. Microbes are the main vehicle for

remediation of herbicides, and new discoveries, such as novel

biodegradation pathways, multispecies interactions and com-

munity-level responses to herbicides addition, are helping us

to understand, predict and monitor the fate of herbicides. With

the recent release of new metagenomic information from

herbicides-associated environments determining the micro-

bial players offers the promise of new discovery. Genomic,

transcriptomic and mutant studies conducted with defined co-

cultures are yielding new information about how microbes

interact and benefit energetically via different mechanisms of

interspecies electron transfer. Despite this, there are many

challenges, not least because of the heterogeneity of these

ecosystems and the structure of herbicides. For example, there

is growing awareness about the toxicity of the herbicides,

which are difficult to degrade. Microbial metabolism is a

process of energy conversion, and it is governed by enzymatic

mechanisms, where reaction intermediates play a vital role.

Screening of organisms that degrade herbicides or produce

enzymes or enzyme systems that degrade herbicides may

prove as environmentally profitable in the present time.

Figure 9. Biodegradation of paraquat. M(microorganisms) – L. starkeyi,Corynebacterium fascians, Lipomyces star-keyi, Aspergillus niger, Penicillium frequen-tans, Fusarium sp and Pseudomonas sp.1¼ 1,10-dimethyl-4,40-bipyridinium, 2¼ 1-methyl-4,40-bipyridinium, 3¼ 1,10-dimethyl-2-oxo2,3-dihydro-4,4-bipyridinium, 4¼N-methyl-isonicotinic acid or 4-carboxyl-1-methyl-pyridylium, 5¼ succinic acid, 6¼oxalic acid, 7¼ formic acid and8¼methylamine.

O

N

O

OH

O

OH

O

OHO

OHO

OH O

OH

H +++

++

+

1

23

4

5 6 7 8

hv, Mhv

hv, Mhv

M

+hv

H3C H3C

H3CH3C

H3C

NH2 CO2

CH3

CH3

N+

N+N+N+

N+N+

H3C NH2 CO2

H2O CO2NH3

DOI: 10.3109/1040841X.2014.929564 Microbial degradation of herbicides 13

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y U

nive

rsity

of

Bat

h on

11/

02/1

4Fo

r pe

rson

al u

se o

nly.

Page 14: Microbial degradation of herbicides

A screening program for such organisms and enzymes is

required but will require more universally uniform standards

for assessment of their degradative ability. Current research

provides an understanding of how the evolution of promis-

cuous enzymes and the recruitment of enzymes available from

the metagenome allows for the assembly of biodegradation

pathways. Nevertheless, physicochemical constraints includ-

ing bioavailability, bioaccessibility and the structural vari-

ations of similar chemicals limit the evolution of

biodegradation pathways.

Further study is required in genetic modification of

microorganisms. The use of gene probes for studying the

distribution of set of genes in herbicides contaminated soils

will be useful in identifying niches in which these kinds of

genes prevail and the conditions under which the population

of microbes bearing these genes increases (Singh et al.,

2013). Phytoremediation in conjunction with rhizospheric

microbes may provide a cheap, fast, eco-friendly and efficient

rhizoremediation processes for the removal of explosive waste

from the upper layers of the soil (Singh et al., 2012). The

implementation of advanced technologies such as proteomics

and bioinformatics should be investigated to provide more

knowledge regarding the enzymatic mechanisms and inter-

mediates involved in metabolic activities during biodegrad-

ation. In this regard, research may be also focused on

examining the possibility of using the enzymes rather than the

microorganisms in biological treatment. To sum it up, the

synergistic performance of various microorganisms and

various technologies is to be considered a research topic of

high priority.

References

Abdelhafid R, Houot S, Barriuso E. (2000). Dependence of atrazinedegradation on C and N availability in adapted and non-adapted soils.Soil Biol Biochem 32:389–401.

Aksaka O, Erturk FA, Sunar S, et al. (2013). Assessment of genotoxiceffects of 2,4-dichlorophenoxyacetic acid on maize by using RAPDanalysis. Indus Crops Prod 42:552–7.

Al-Rajab AJ, Hakami OM. (2014). Behavior of the non-selectiveherbicide glyphosate in agricultural soil. Am J Environ Sci 10:94–101.

Amy PS, Schulke JW, Frazier LM, Seidler RJ. (1985). Characterizationof aquatic bacteria and cloning of genes specifying partial degradationof 2,4 dichlorophenoxyacetic. Appl Environ Microbiol 49:1237–45.

Annelie M, Westerberg K, Jernberg C, Janet K. (2001). Use of greenfluorescent protein and luciferase biomarkers to monitor survival andactivity of Arthrobacter chlorophenolicus A6 cells during degradationof 4-chlorophenol in soil. Environ Microbiol 3:32–42.

Arfarita N, Imai T, Kanno A, et al. (2013). The potential use oftrichoderma viride strain frp3 in biodegradation of the herbicideglyphosate. Biotechnol Biotechnol Eq 27:3518–21.

Atamaniuk T, Kubrak I, Storey K, Lushchak V. (2013). Oxidative stressas a mechanism for toxicity of 2,4 dichlorophenoxyacetic acid(2,4-D): studies with goldfish gills. Ecotoxicology 22:1498–508.

Baelum J, Jacobsen CS, Holben WE. (2010). Comparison of 16S rRNAgene phylogeny and functional tfdA gene distribution in thirty-onedifferent 2,4- dichlorophenoxyacetic acid and 4-chloro-2-methylphe-noxyacetic acid degraders. Syst Appl Microbiol 33:67–70.

Baldwin B, Bray M, Geoghegan M. (1966). The microbial decompos-ition of paraquat. Biochem J 101:15.

Barbash, Jack E, Gail P, et al. (1999). Distribution of major herbicides inground water of the United States. U.S. Geological Survey, Water-Resources Investigations 98–4245.

Bardullas U, Giordano M, Rodriguez VM. (2011). Chronic atrazineexposure causes disruption of the spontaneous locomotor activity andalters the striatal dopaminergic system of the male Sprague–Dawleyrat. Neurotoxicol Teratol 33:263–72.

Bhat MA, Tsuda M, Horiike K, et al. (1994). Identification andcharacterization of a new plasmid carrying genes for degradation of2,4-dichlorophenoxyacetate from Pseudomonas cepacia CSV90. ApplEnviron Microbiol 60:307–12.

Biradar DP, Rayburn AL. (1995). Flow cytometric analysis of whole cellclastogenicity of herbicides found in groundwater. Arch EnvironContamin Toxicol 28:13–17.

Bogaerts P, Bohatier J, Bonnemoy F, et al. (2000). Fungal biodegrad-ation of a phenylurea herbicide, diuron: structure and toxicity ofmetabolites. Pest Management Sci 56:455–62.

Bortolozzi A, Evangelista de Duffard F, Daja R, Silveira R. (2001).Intracerebral administration of 2,4-dichlorophenoxyacetic acidinduces behavioral and neurochemical alterations in the rat brain.Neurotoxicology 22:221–32.

Bukowska B. (2006). Toxicity of 2,4-dichlorophenoxyacetic acid –molecular mechanisms. Pol J Environ Stud 15:365–74.

Burns CJ, Swaen GM. (2012). Review of 2,4-dichlorophenoxyacetic acid(2,4-D) biomonitoring and epidemiology. Crit Rev Toxicol 42:768–86.

Carlisle S, Trevors JT. (1986). Effect of the herbicide glyphosate onrespiration and hydrogen consumption in soil. Water Air Soil Pollut27:391–401.

Carr RJG, Bilton RF, Atkinson T. (1986). Toxicity of paraquat tomicroorganisms. Appl Environ Microbiol 52:1112–16.

Castillo MA, Felis N, Aragon P, et al. (2006). Biodegradation of theherbicide diuron by streptomycetes isolated from soil. Int BiodeterBiodegrad 58:196–202.

Castro Joao V, Peralba Maria CR, Ayub Marco AZ. (2007).Biodegradation of the herbicide glyphosate by filamentous fungiin platform shaker and batch bioreactor. J Environ Sci Health 42:883–6.

Celis E, Elefsiniotis P, Singhal N. (2008). Biodegradation of agriculturalherbicides in sequencing batch reactors under aerobic or anaerobicconditions. Water Res 42:3218–24.

Cenkci S, Yldz M, Cigerci IH, et al. (2010). Evaluation of 2,4-D anddicamba genotoxicity in bean seedlings using comet and RAPDassays. Ecotoxicol Environ Saf 73:1558–63.

Chaudhry GR, Huang GH. (1988). Isolation and characterization of anew plasmid from a Flavobacterium sp. which carries the genes fordegradation of 2,4-dichlorophenoxyacetate. J Bacteriol 170:3897–902.

Cherry, B. (2010). GM crops increase herbicide use in the UnitedStates. Institute of Science and Technology. Available from: http://www.i-sis.org.uk/GMcropsIncreasedHerbicide.php [last accessed 10Nov 2010].

Chong NM, Chang HW. (2009). Plasmid as a measure of microbialdegradation capacity for 2,4-dichlorophenoxyacetic acid. BioresourTechnol 100:1174–9.

Coady K, Marino T, Thomas J, et al. (2013). An evaluation of2,4-dichlorophenoxyacetic acid in the amphibian metamorphosisassay and the fish short-term reproduction assay. Ecotoxicol EnvironSafety 90:143–50.

Cox C. (1995). Glyphosate, part 2: human exposure and ecologicaleffects. Herbicide factsheet. J Pestic Reform 15:14–20.

Dalton RL, Evans AW, Rhodes RC. (1966). Disappearance of diuronfrom cotton fields soils. Weeds 14:14–31.

De Souza ML, Newcombe D, Alvey S, et al. (1998). Molecular basis of abacterial consortium: interspecies catabolism of atrazine. ApplEnviron Microbiol 64:178–84.

Dejonghe W, Goris J, El Fantroussi S, et al. (2000). Effect ofdissemination of 2,4-dichlorophenoxyacetic acid (2,4-D) degradationplasmids on 2,4-D degradation and on bacterial community structurein two different soil horizons. Appl Environ Microbiol 66:3297–304.

Devers M, Rouard N, Martin-Laurent F. (2007). Genetic rearrangementof the atzAB atrazinedegrading gene cassette from pADP1: Tn5 to thechromosome of Variovorax sp. MD1 and MD2. Gene 392:1–6.

Domingues I, Soares AM, Loureiro S. (2011). Growth rate ofPseudokirchneriella subcapitata exposed to herbicides found insurface waters in the Alqueva reservoir (Portugal): a bottom-upapproach using binary mixtures. Ecotoxicology 20:1167–75.

Don RH, Pemberton JM. (1981). Properties of six pesticide degradationplasmids isolated from Alcaligenes paradoxus and Alcaligeneseutrophus. J Bacteriol 145:681–6.

Don RH, Pemberton JM. (1985). Genetic and physical map of the2,4 dichlorophenoxyacetic acid degradative plasmid pJP4. J Bacteriol161:466–8.

14 B. Singh Crit Rev Microbiol, Early Online: 1–17

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y U

nive

rsity

of

Bat

h on

11/

02/1

4Fo

r pe

rson

al u

se o

nly.

Page 15: Microbial degradation of herbicides

Duke SO, Powles SB. (2008). Glyphosate: a once in a century herbicide.Pest Manag Sci 64:319–25.

El-Bestawy E, Hans-Jorgen A. (2007). Effect of nutrient amendmentsand sterilization on mineralization and/or biodegradation of14C-labeled MCPP by soil bacteria under aerobic conditions. IntBiodeterior Biodegrad 59:193–201.

Ermakova IT, Kiseleva NI, Shushkova T, et al. (2010). Bioremediation ofglyphosatecontaminated soils. Appl Microbiol Biotechnol 88:585–94.

Evans W, Smith W, Fernely H, Davis J. (1971). Bacterial metabolism of2,4 dichlorophenoxyacetic acid. J Biochem 122:543–51.

EXTOXNET. (1996). Pesticide information profiles diuron, ExtensionToxicology Network. Available: http://extoxnet.orst.edu/pips/diur-on.htm. Revised June 1996.

Fazlurrahman B, Pandey M, Suri J, Jain RK. (2009). Isolation andcharacterization of an atrazine-degrading Rhodococcus sp. strainMB-P1 from contaminated soil. Lett Appl Microbiol 49:721–9.

Federico C, Motta S, Palmieri C, et al. (2011). Phenylurea herbicidesinduce cytogenetic effects in Chinese hamster cell lines. Mutat Res721:89–94.

Fisher PR, Appleton J, Pemberton JM. (1978). Isolation and character-ization of the pesticide-degrading plasmid pJP1 from Alcaligenesparadoxus. J Bacteriol 135:798–804.

Fratila-Apachitei LE, Hirst JA, Siebel MA, Gijzen HJ. (1999). Diurondegradation by Phanerochaete chrysosporium BKM-F-1767 in syn-thetic and natural media. Biotechnol. Lett 21:147.

Funderburk H, Bozarth GA. (1967). Review of the metabolism anddecomposition of diquat and paraquat. J Agric Food Chem 15:563–8.

Ganguli A, Choudhury D, Chakrabarti G. (2014). 2,4-Dichlorophenoxyacetic acid induced toxicity in lung cells by disruptionof the tubulin-microtubule network. Toxicol Res, 3:118–30.

Garabrant DH, Philbert MA. (2002). Review of 2,4-dichlorophenox-yacetic acid (2,4-D) epidemiology and toxicology. Crit Rev Toxicol32:233–57.

Garcia-Gonzalez V, Govantes F, Shaw LJ, et al. (2003).Nitrogencontrolof atrazineutilizationin Pseudomonas sp. strain ADP.Appl Environ Microb 69:6987–93.

Gauri SS, Mandal SM, Dey S, Pati BR. (2012). Biotransformation of p-coumaric acid and 2,4-dichlorophenoxy acetic acid by Azotobacter sp.strain SSB81. Bioresour Technol 126:350–3.

Gaynor JD, Hamill AS, MacTavish DC. (1993). Efficacy, fruit residues,and soil dissipation of the herbicide metolachlor in processing tomato.J Am Soc Hortic Sci 118:68–72.

Gianessi LP. (2013). The increasing importance of herbicides inworldwide crop production. Pest Manag Sci 69:1099–105.

Gonzalez M, Soloneski S, Reigosa MA, Larramendy ML. (2005).Genotoxicity of the herbicide 2,4-dichlorophenoxyacetic acid and acommercial formulation, 2,4-dichlorophenoxyacetic acid dimethyla-mine salt I. Evaluation of DNA damage and cytogenic endpoints inChinese Hamster ovary (CHO) cells. Toxicol In Vitro 19:289–97.

Gui W, Dong Q, Zhou S, et al. (2011). Waterborne exposure toclodinafop-propargyl disrupts the posterior and ventral developmentof zebrafish embryos. Environ Toxicol Chem 30:1576–81.

Hartnett S, Musah S, Dhanwada KR. (2013). Cellular effects ofmetolachlor exposure on human liver (HepG2) cells. Chemosphere90:1258–66.

Hayes TB, Collins A, Lee M, et al. (2002). Hermaphroditic,demasculinized frogs after exposure to the herbicide atrazine at lowecologically relevant doses. Proc Natl Acad Sci USA 99:5476–80.

Hess D, Warren F. (2002). The herbicide handbook of the Weed ScienceSociety of America. 8th ed., Weed Science Society of America.Herbicide Handbook Committee. 159–61.

Hoffmann D, Kleinsteuber S, Muller RH, Babel W. (2003). A transposonencoding the complete 2,4-dichlorophenoxyacetic acid degradationpathway in the alkalitolerant strain Delftia acidovorans P4a.Microbiology 149:2545–56.

Holland NT, Duramand P, Rothman N, et al. (2002). Micronucleusfrequency and proliferation in human lymphocytes after exposure toherbicide 2,4-dichlorophenoxyacetic acid in vitro and in vivo. MutatRes 521:165–78.

Hou Y, Tao J, Shen W, et al. (2011). Isolation of the fenoxaprop-ethyl(FE)-degrading bacterium Rhodococcus sp. T1, and cloning of FEhydrolase gene feh. FEMS Microbiol Lett 323:196–203.

Huang CL, Lee YC, Yang YC, et al. (2012). Minocycline preventsparaquat-induced cell death through attenuating endoplasmic reticu-lum stress and mitochondrial dysfunction. Toxicol Lett 209:203–10.

Inoue D, Yamazaki Y, Tsutsui H, et al. (2012). Impacts of genebioaugmentation with pJP4-harboring bacteria of 2,4-D-contaminatedsoil slurry on the indigenous microbial community. Biodegradation23:263–76.

Jablonowski ND, Koeppchen S, Hofmann D, et al. (2008).Spatial distribution and characterization of long-term aged14C-labeled atrazine residues in soil. J Agricul Food Chem 56:9548–54.

Jacob GS, Garbow JR, Hallas LE, et al. (1988). Metabolism ofglyphosate in Pseudomonas sp. strain LBr. Appl Environ Microbiol54:2953–8.

Jaquet J, Weitzel P, Junge T, Schmidt B. (2014). Metabolic fate of 14 Clabeled herbicide clodinafop-propargyl in soil. J Environ Sci Health49:245–54.

Johnson TA, Sims GK. (2011). Introduction of 2,4-dichlorophenoxya-cetic acid into soil with solvents and resulting implications forbioavailability to microorganisms. World J Microbiol Biotechnol 27:1137–43.

Juhler R, Sorensen S, Larsen L. (2001). Analysing transformationproducts of herbicide residues in environmental samples. Water Res35:1371–8.

Kashanian S, Askari S, Ahmadi F, et al. (2008). In vitro study of DNAinteraction with clodinafop-propargyl herbicide. DNA Cell Biol 27:581–6.

Kaufman DD, Blake J. (1970). Degradation of atrazine by soil fungi. SoilBiol Biochem 2:73–80.

Kitagawa W, Takami S, Miyauchi K, et al. (2002). Novel 2,4-dichlorophenoxyacetic acid degradation genes from oligotrophicBradyrhizobium sp strain HW13 isolated from a pristine environment.J Bacteriol 184:509–18.

Kolpin DW, Kalkhoff SJ. (1993). Atrazine degradation in a small streamin Iowa. Environ Sci Technol 27:134–9.

Lagana A, Bacaloni A, De Leva I, et al. (2002). Occurrence anddetermination of herbicides and their major transformation products inenvironmental waters. Analytica Chimica Acta 462:187–98.

Laitinen P, Ramo S, Siimes K, et al. (2007). Glyphosate translocationfrom plants to soil-does this constitute a significant proportion ofresidues in soil? Plant Soil 300:51–60.

Lowry DM, Greiner D, Fretheim M, et al. (2013). Mecha-nism ofmetolachlor action due to alterations in cell cycle progression. CellBiol Toxicol 29:283–91.

Mandelbaum RT, Allan DL, Wackett LP. (1995). Isolation and charac-terization of a Pseudomonas sp. that mineralizes the s-triazineherbicide atrazine. Appl Environ Microbiol 61:1451–7.

Mandelbaum RT, Wackett LP, Allan DL. (1993). Mineralization of thes-triazine ring of atrazine by stable bacterial mixed cultures. ApplEnviron Microbiol 59:1695–701.

Marron E, Ruiz N, Rubio C, et al. (2006). 2,4-D Degrading bacterialconsortium. Isolation, kinetic characterization in batch and continuousculture and application for bioaugmentating an activated sludgemicrobial community. Process Biochem 41:1521–8.

Martinez B, Tomkins J, Wackett R, et al. (2001). Complete nucleo-tide sequence and organization of the atrazine catabolicplasmid pADP-1 from Pseudomonas sp. strain ADP. J Bacteriol183:5684–97.

Matheson VG, Forney LJ, Suwa Y, et al. (1996). Evidence for acquisitionin nature of a chromosomal 2,4-dichlorophenoxyacetic acid/a-keto-glutarate dioxygenase gene by different Burkholderia spp. ApplEnviron Microbiol 62:2457–63.

Megadi VB, Tallur PN, Hoskeri RS, et al. (2010). Biodegradation ofpendimethalin by Bacillus circulans. Indian J Biotechnol 9:173–7.

Meza-Joya FL, Ramirez-Pinilla MP, Fuentes-Lorenzo JL. (2013). Toxic,cytotoxic, and genotoxic effects of a glyphosate formulation(Roundup�SL-Cosmoflux�411F) in the direct-developing frogEleutherodactylus johnstonei. Environ Mol Mutagen 54:362–73.

Mikov I, Vasovic V, Mikov A, et al. (2010). Hypoglycemic effect ofherbicide 2,4 dichlorophenoxyacetic acid (2,4-D). Pestic Phytomed(Belgrade) 25:349–52.

Muller RH, Babel W. (2004). Delftia acidovorans MC1 resists highherbicide concentrations e a study of nutristat growth on (RS)-2-(2,4-dichlorophenoxy) propionate and 2,4 dichlorophenoxyacetate. BiosciBiotechnol and Biochem 68:622–30.

Munro IC, Carlo GL, Orr JC, et al. (1992). A comprehensive, integratedreview and evaluation of the scientific evidence relating to the safetyof the herbicide 2,4-D. J Am Coll Toxicol 11:559–664.

DOI: 10.3109/1040841X.2014.929564 Microbial degradation of herbicides 15

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y U

nive

rsity

of

Bat

h on

11/

02/1

4Fo

r pe

rson

al u

se o

nly.

Page 16: Microbial degradation of herbicides

Newby DT, Gentry TJ, Pepper IL. (2000). Comparison of 2,4-dichlorophenoxyacetic acid degradation and plasmid transfer in soilresulting from bioaugmentation with two different pJP4 donors. ApplEnviron Microbiol 66:3399–407.

Palmeira CM, Moreno AJ, Madeira VMC. (1994). Interactions ofherbicides 2,4-D and dinoseb with liver mitochondrial bioenergetics.Toxicol Appl Pharmacol 127:50–7.

Penaloza-Vazquez A, Mena GL, Herrera-Estrella L, Bailey AM.(1995). Cloning and sequencing of the genes involved in glyphosateutilization by Pseudomonas pseudomallei. Appl Environ Microbiol61:538–43.

Perkins EJ, Lurquin PF. (1988). Duplication of a 2,4-dichlorophenox-yacetic acid monooxygenase gene in Alcaligenes eutrophusJMP134(pJP4. J Bacteriol 170:5669–72.

Pimentel D. (1995). Amounts of pesticides reaching target pests:environmental impacts and ethics. J Agr Environ Ethic 8:17–29.

Pinto AP, Serrano C, Pires T, et al. (2012). Degradation ofterbuthylazine, difenoconazole and pendimethalin pesticides byselected fungi cultures. Sci Total Environ 435–6, 402–10.

Poh RPC, Smith ARW, Bruce IJ. (2002). Complete characterizationof Tn5530 from Burkholderia cepacia strain 2a (pIJB1) and studiesof 2,4-dichlorophenoxyacetate uptake by the organism. Plasmid 48:1–12.

Rickettes DC. (1999). The microbial biodegradation of Paraquat in soil.Pestic Sci 55:596–614.

Rollof B, Belluck D, Meiser L. (1992). Cytogenic effects of cyanazineand metolachlor on human lymphocytes exposed in vitro. Mut ResLett 281:295–8.

Rousseaux S, Hartmann A, Soulas G. (2001). Isolation and character-isation of new Gram-negative and Gram-positive atrazine degradingbacteria from different French soils. FEMS Microbiol Ecol 36:211–22.

Rueppel ML, Brightwell BB, Schaefer J, Marvel JT. (1977). Metabolismand degradation of glyphosate in soil and water. J Agric Food Chem25:517–28.

Sahid I, Hamzah A, Aris PM. (1992). Effects of paraquat and alachlor onsoil microorganisms in peat soil. Pertanika 15:121–5.

Sandoval-Carrasco CA, Ahuatzi-Chacon D, Galındez-Mayer J, et al.(2013). Biodegradation of a mixture of the herbicides ametryn, and2,4-dichlorophenoxyacetic acid (2,4-D) in a compartmentalizedbiofilm reactor. Bioresour Technol 145:33–6.

Saxena A, Zhang RW, Bollag JM. (1987). Microorganisms capable ofmetabolizing the herbicide metolachlor. Appl Environ Microbiol 53:390–6.

Sebaı TE, Devers-Lamrani M, Changey F, et al. (2011). Evidence ofatrazine mineralization in a soil from the Nile Delta: isolation ofArthrobacter sp. TES6, an atrazine-degrading strain. Int BiodeteriorBiodegrad 65:1249–55.

Sharma P, Suri CR. (2011). Biotransformation and biomonitoring ofphenylurea herbicide diuron. Bioresource Technol 102:3119–25.

Shelton DR, Khader S, Karns JS, Pogell BM. (1996). Metabolism oftwelve herbicides by Streptomyces. Biodegradation 7:129–36.

Silva TM, Stets MI, Mazzetto AM, et al. (2007). Degradation of 2,4-Dherbicide by microorganisms isolated from Brazilian contaminatedsoil. Braz J Microbiol 38:522–5.

Simonsen L, Fomsgaard IS, Svensmark B, Spliid NH. (2008). Fate andavailability of glyphosate and AMPA in agricultural soil. J Environ SciHealth 43:365–75.

Singh B. (2013). Degradation of clodinafop propargyl by Pseudomonassp. strain B2. Bull Environ Contam Toxicol 6:730–3.

Singh B, Kaur J, Singh K. (2012). Microbial remediation of explosivewaste. Crit Rev Microbiol 38:152–67.

Singh B, Kaur J, Singh K. (2013). Microbial degradation of anorganophosphate pesticide, malathion. Crit Rev Microbiol 40:146–54.

Siripattanakul S, Wirojanagud W, McEvoy J, et al. (2009). Atrazinedegradation by stable mixed cultures enriched from agricultural soiland their characterization. J Appl Microbiol 106:986–92.

Slade P. (1965). The photochemical degradation of paraquat. Nature(London) 207:515–16.

Song J, Gu J, Zhai Y, et al. (2013). Biodegradation of nicosulfuron by aTalaromyces flavus LZM1. Bioresour Technol 140:243–8.

Sorensen S, Albers C, Aamand J. (2008). Rapid mineralization of thephenylurea herbicide diuron by Variovorax sp. SRS16 in pure cultureand within a two-member consortium. Appl Environ Microbiol 74:2332–40.

Stasinakis AS, Kotsifa S, Gatidou G, Mamais D. (2009). Diuronbiodegradation in activated sludge batch reactors under aerobic andanoxic conditions. Water Res 43:1471–9.

Struthers JK, Jayachandran K, Moorman TB. (1998). Biodegradation ofatrazine by Agrobacterium radiobacter J14a and use of this strain inbioremediation of contaminated soil. Appl Environ Microbiol 64:3368–75.

Sturtz N, Evangelista De Duffard AM, Duffard R. (2000). Detection of2,4-dichlorophenoxyacetic acid (2,4- D) residues in neonates breast-fed by 2,4-D exposed dams. Neurotoxicol 2:147.

Tappe W, Groeneweg J, Jantsch B. (2002). Diffuse atrazine pollution inGerman aquifers. Biodegradation 13:3–10.

Thurman EM, Goolsby DA, Meyer MT, et al. (1992). A reconnaissancestudy of herbicides and their metabolites is surface water of theMidwestern United States using immunoassay and gas chromatog-raphy/mass spectrometry. Environ Sci Technol 26:2440–7.

Tiedje JM, Duxbury JM, Alexander M, Dawson JE. (1969). 2,4-Dmetabolism: pathway of degradation of chlorocatechols byAthrobacter sp. J Agr Food Chem 17:1021–6.

Top EM, Holben WE, Forney LJ. (1995). Characterization of diverse2,4-dichlorophenoxyacetic acid-degradative plasmids isolated fromsoil by complementation. Appl Environ Microbiol 61:1691–8.

Topp E, Zhu H, Nour SM, et al. (2000). Characterization of an atrazinedegrading Pseudaminobacter sp. isolated from Canadian and Frenchagricultural soils. Appl Environ Microbiol 66:2773–82.

Tsutsui H, Anami Y, Matsuda M, et al. (2013). Plasmid-mediatedbioaugmentation of sequencing batch reactors for enhancement of2,4-dichlorophenoxyacetic acid removal in wastewater using plasmidpJP4. Biodegradation 24:343–52.

Turnbull GA, Cllington JE, Walker A, Morgan J. (2001). Identificationand characterisation of a diuron-degrading bacterium. Biol Fertil Soils33:472–6.

United States Environmental Protection Agency (USEPA). (2004).Environmental date and effects division’s risk assessment for thereregistration eligibility document for 2,4-dichlorophenoxyaceticacid (2,4-D). Available from: http://www.regulations.gov/fdmspub-lic/component/main (Docket IDEPA-HQ-OPP-2004-0167) [lastaccessed 21 Nov 2006].

USEPA-OPP. (1984). Memo from S. Creeger, Hazard EvaluationDivision, to R. Taylor, Registration Division.

Vargha M, Takats Z, Marialigeti K. (2005). Degradation of atrazine in alaboratory scale model system with Danube river sediment. Water Res39:1560–8.

Veiga F, Zapata JM, Fernandez Marcos ML, Alvarez E. (2001).Dynamics of glyphosate and aminomethylphosphonic acid in aforest soil in Galicia, north-west Spain. Sci Total Environ 271:135–44.

Wackett LP, Sadowsky MJ, Martinez B, Shapir N. (2002).Biodegradation of atrazine and related s-triazine compounds: fromenzymes to field studies. Appl Microbiol Biotechnol 58:39–45.

Wang JH, Zhu LS, Liu AJ, et al. (2011). Isolation and characterization ofan Arthrobacter sp. strain HB-5 that transforms atrazine. EnvironGeochem Health 33:259–66.

Wang Q, Xie S. (2012). Isolation and characterization of a high-efficiency soil atrazine-degrading Arthrobacter sp. strain. Int BiodeterBiodegr 71:61–6.

Wang QF, Xie SG, Hu R. (2013). Bioaugmentation with Arthrobacter spstrain DAT1 for remediation of heavily atrazine-contaminated soil. IntBiodeterior Biodegrad 77:63–7.

Weintraub RL, Reinhart JH, Scherff RA, Schisler LC. (1954).Metabolism of 2,4-dichlorophenoxyacetic acid. III. Metabolism andpersistence in dormant plant tissue. Plant Physiol 29:303–4.

Wua B, Song B, Yang H, et al. (2013). Central nervous system damagedue to acute paraquat poisoning: an experimental study with ratmodel. Neuro Toxicology 35:62–70.

Xu J, Yang M, Dai J, et al. (2008). Degradation of acetochlor by fourmicrobial communities. Bioresource Technol 99:7797–802.

Xuedong W, Huili W, Defang F. (2005). Biodegradation of imazapyr byfree cells of Pseudomonas fluorescene biotype II and Bacillus cereusisolated from soil. Bull Environ Contam Toxicol 74:350–5.

Yang CY, Li Y, Zhang K, et al. (2010). Atrazine degradation by a simpleconsortium of Klebsiella sp. A1 and Comamonas sp. A2 in nitrogenenriched medium. Biodegradation 21:97–105.

Yanze-Kontchou C, Gschwind N. (1994). Mineralization of the herbicideatrazine as a carbon source by a Pseudomonas strain. Appl EnvironMicrobiol 60:4297–302.

16 B. Singh Crit Rev Microbiol, Early Online: 1–17

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y U

nive

rsity

of

Bat

h on

11/

02/1

4Fo

r pe

rson

al u

se o

nly.

Page 17: Microbial degradation of herbicides

Yaw-Jian L, Karuppiah M, Shaw A, Gupta G. (1999). Effect of simulatedsunlight on atrazine and metolachlor toxicity of surface waters.Ecotoxicol Environ Saf 43:35–7.

Zhang Y, Jiang Z, Cao B, et al. (2011). Metabolic ability and genecharacteristics of Arthrobacter sp strain DNS10, the sole atrazinedegrading strain in a consortium isolated from black soil. IntBiodeterior Biodegrad 65:1140–4.

Zheng SQ, Cooper JE. (1996). Adsorption, desorption anddegradation of three pesticides in different soils. Arch EnvironToxicol 30:15–20.

Zhou ZS, Song JB, Liu ZP, Yang H. (2012). Molecular dissection ofatrazine-responsive transcriptome and gene networks in rice by high-throughput sequencing. J Hazard Mater 15:57–68.

DOI: 10.3109/1040841X.2014.929564 Microbial degradation of herbicides 17

Cri

tical

Rev

iew

s in

Mic

robi

olog

y D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y U

nive

rsity

of

Bat

h on

11/

02/1

4Fo

r pe

rson

al u

se o

nly.