Exploiting plant metabolism for the phytoremediation of persistent herbicides

11
Persistent Herbicides Phytoremediation Phytoremediation: Persistent Herbicides Exploiting Plant Metabolism for the Phytoremediation of Persistent Herbicides Julian O. D. Coleman 1, Carla Frova 2, Peter Schr6der 3" and Michel Tissut 4 1School of Biological and Molecular Sciences, Oxford Brookes University, Gipsy Lane, Oxford OX3 0BP, UK 2Department of Genetics and Microbiology, University of Milano, Via Celoria 26, 1-20133 Milano, Italy 3GSF-Institute for Soil Ecology, GSF Research Centre for Environment and Health, D-85764 Neuherberg, Germany 4Laboratoire Ecosyst~mes et Changements Environnementaux, Universit6 J. Fourier, BP 53, F-38041 Grenoble cedex 9, France * Corresponding author ([email protected] DOh http://dx.doi.org/10.1065/esDr2001.09.084.5 Abstract. Weed control by herbicides has helped us to create the green revolution and to provide food for at least the major- ity of human beings living today. However, some herbicides re- main in the environment and pose an ecological problem. The present review describes the properties and fate of four repre- sentative herbicides known to be presistent in ecosystems. Meta- bolic networks are depiced and it is concluded that removal of these comopounds by the ecologically friendly technique of phytoremediation is possible. The largest problem is seen in the uptake of the compounds into suitable plants and the time needed for such an approach. Keywords: Atrazine; chloroacetanilides, isoproturon; metolachlor; persistent herbicides; phytoremediation; plant metabolism Introduction Over the last 50 years weed control by herbicides has be- come an integral part of commercial agriculture and because of its cost-effectiveness, herbicide use is likely to remain an important agronomic strategy for the foreseeable future. In order to ensure that agrochemical compounds produce no 'unacceptable risk' to human health or the environment, all herbicides are subject to an approval procedure under EU and national legislation that imposes detailed controls on the use of each product. However, one of the inevitable re- percussions of the widespread use of herbicides is contami- nation of soil, surface water and ground water by parent compounds or their degradation products. Heavy environ- mental contamination by herbicides may arise from indus- trial point sources such as accidental spillage during pro- duction, storage and transport, or as leachates from former dumping sites. In contrast, sources of pollution arising from agricultural uses of herbicides (e.g. crop spraying) are con- sidered to be diffuse as the compounds are distributed over large areas. During herbicide spray application less than five percent of the compound will reach the intended target due to spray drift, volatilisation or washout. Herbicides that are sprayed on the foliage of crops and weeds (e.g. post-emer- gence herbicides) reach surface waters via the air and soil. Herbicides that are applied to the soil (pre-emergence herbi- cides) through leaching or run-off may reach deeper soil lay- ers or adjacent surface water and groundwater aquifers. The concentration and persistence of herbicides in the envi- ronment depend on several factors including crop and soil type, application rates, degradation rates and the physico- chemical properties of the compound (Fig. 1). Good exam- ples of the latter are the properties of soil applied pre-emer- Fig. 1 : Differential entry paths of per- and post-emergence herbicides into agricultural ecosystems 18 ESPR - Environ Sci & Poliut Res 9 (1) 18 - 28 (2002) 9 ecomed publishers, D-86899 Landsberg, Germany and Ft. Worth/TX, USA 9 Tokyo, Japan 9 Mumbai, India 9 Seoul, Korea

Transcript of Exploiting plant metabolism for the phytoremediation of persistent herbicides

Persistent Herbicides Phytoremediation

Phytoremediation: Persistent Herbicides

Exploiting Plant Metabolism for the Phytoremediation of Persistent Herbicides Julian O. D. Co leman 1, Car la Frova 2, Peter Schr6der 3" and Michel Tissut 4

1 School of Biological and Molecular Sciences, Oxford Brookes University, Gipsy Lane, Oxford OX3 0BP, UK 2Department of Genetics and Microbiology, University of Milano, Via Celoria 26, 1-20133 Milano, Italy 3 GSF-Institute for Soil Ecology, GSF Research Centre for Environment and Health, D-85764 Neuherberg, Germany 4Laboratoire Ecosyst~mes et Changements Environnementaux, Universit6 J. Fourier, BP 53, F-38041 Grenoble cedex 9, France

* Corresponding author ([email protected]

D O h h t t p : / / d x . d o i . o r g / 1 0 . 1 0 6 5 / e s D r 2 0 0 1 . 0 9 . 0 8 4 . 5

A b s t r a c t . Weed control by herbicides has helped us to create the green revolution and to provide food for at least the major- ity of human beings living today. However, some herbicides re- main in the environment and pose an ecological problem. The present review describes the properties and fate of four repre- sentative herbicides known to be presistent in ecosystems. Meta- bolic networks are depiced and it is concluded that removal of these comopounds by the ecologically friendly technique of phytoremediation is possible. The largest problem is seen in the uptake of the compounds into suitable plants and the time needed for such an approach.

Keywords: Atrazine; chloroacetanilides, isoproturon; metolachlor; persistent herbicides; phytoremediation; plant metabolism

I n t r o d u c t i o n

Over the last 50 years weed control by herbicides has be- come an integral part of commercial agriculture and because of its cost-effectiveness, herbicide use is likely to remain an important agronomic strategy for the foreseeable future. In order to ensure that agrochemical compounds produce no 'unacceptable risk' to human health or the environment, all

herbicides are subject to an approval procedure under EU and national legislation that imposes detailed controls on the use of each product. However, one of the inevitable re- percussions of the widespread use of herbicides is contami- nation of soil, surface water and ground water by parent compounds or their degradation products. Heavy environ- mental contamination by herbicides may arise from indus- trial point sources such as accidental spillage during pro- duction, storage and transport, or as leachates from former dumping sites. In contrast, sources of pollution arising from agricultural uses of herbicides (e.g. crop spraying) are con- sidered to be diffuse as the compounds are distributed over large areas. During herbicide spray application less than five percent of the compound will reach the intended target due to spray drift, volatilisation or washout. Herbicides that are sprayed on the foliage of crops and weeds (e.g. post-emer- gence herbicides) reach surface waters via the air and soil. Herbicides that are applied to the soil (pre-emergence herbi- cides) through leaching or run-off may reach deeper soil lay- ers or adjacent surface water and groundwater aquifers.

The concentration and persistence of herbicides in the envi- ronment depend on several factors including crop and soil type, application rates, degradation rates and the physico- chemical properties of the compound (Fig. 1). Good exam- ples of the latter are the properties of soil applied pre-emer-

Fig. 1 : Differential entry paths of per- and post-emergence herbicides into agricultural ecosystems

18 ESPR - Environ Sci & Poliut Res 9 (1) 18 - 28 (2002) �9 ecomed publishers, D-86899 Landsberg, Germany and Ft. Worth/TX, USA �9 Tokyo, Japan �9 Mumbai, India �9 Seoul, Korea

Phytoremediation Persistent Herbicides

gence herbicides. Pre-emergence herbicides are designed to kill emerging weed seedlings soon after germination. As weed germination can occur over extended periods pre-emergence herbicides are formulated to bind to upper soil components and to degrade slowly. These design properties inevitably contribute to the persistence of these compounds.

Herbicides differ in many significant ways from other or- ganic pollutants and these differences have important impli- cations for their removal or detoxification during phyto- remediation. Herbicides by their very nature are designed to kill unwanted plants and do so by blocking vital biochemical and physiological processes. Because the processes affected by herbicides may be common to a wide range of plant species and to other non-target organisms, herbicides can be severe pollutants even at low concentrations. Typically herbicides are designed to be selective. This means that some tolerant crop plants can metabolise particular herbicides or a family of herbicides. In addition some weed species have acquired herbicide resistance by point mutations at the target site or by the selection of genes that lead to en- hanced metabolism of the herbicide. Consequently for phytoremediation, plants capable of removing and trans- forming herbicides are already known. However, a formu- lated herbicide consists of a mixture of biologically active ingredients and additive chemicals such as solvents and surfactants. These additives are devised to enhance plant uptake and may improve the chances for phytoremediation. On the other hand, some active compounds may be present as a racemic mixture of isomers and the enantomeric selec- tivity of plants may affect the process of phytoremediation by removal of only one isomer.

In many countries all herbicides are subject to approval and regulatory procedures and standards for use are set interna- tionally (e.g. EU legislation and W H O recommendations) and nationally. The European Commission's Directive on Drinking Water 80-778-EEC (1980) established a maximum allowable concentration of 0.1 ~tg/L for individual pesticides and 0.5 ~tg/L for total pesticides. Many European countries have adopted these statutory limits.

In Europe (EU countries), 33 herbicides being approved for use have been detected and monitored in ground water. Of these, approximately 12 compounds from 4 chemical fami- lies of herbicides have been detected at some sampling sites with concentrations >0.1 ~tg/L. and were classified as im- portant pollutants. Families of herbicides and examples of member compounds are shown in Table 1.

In this review we describe the general properties of four of these compounds (atrazine, metolachlor, isoproturon and

Table 1 : Families of herbicides and examples of member compounds

2,4-D) and their uptake and metabolism in plants. On the basis of these case histories we assess the prospects of using phytoremediation technology for the in situ treatment of herbicide-contaminated water, sediments and soil.

1 Atrazine

The triazines are a widely used family of herbicides that were developed during the 1950s and 1960s. The characteristic struc- tural feature of triazine herbicides is the symmetrical 1,3,5- triazine ring (Fig. 2). These compounds are photosynthesis inhibitors, acting on the electron transfer between photosystem II and the plastoquinone pool, in chloroplast thylakoids (Good 1963). Since they were first introduced, several triazine de- rivatives have been developed with selectivity for use in many important crops such as maize, sorghum, sugar cane, vines, oil palms, coffee and in forestry. The wide spread use of triazines is exemplified by the case of atrazine whose application reached 100,000 tons per year by the end of the eighties. However, due to several resistance problemsand connected to its high persistence, atrazine was banned in some european countries at the beginning of the nineties.

Atrazine (Fig. 2, Table 2) is a selective systemic herbicide used for the pre-and post-emergence control of annual grasses and broad-leaf weeds in a variety of cultivated crops. However the major use of atrazine is the pre-emergence control of weeds in maize and sorghum. Atrazine absorbed by foliage and roots, is translocated acropetally in the xylem and accumulates in the apical meristems where it develops its herbicidal action.

~N I~CI (CH3~CHNH~T~NyC' N~N NHCH2CH3

A. B. Fig. 2: Chemical structures of triazines. A. 1.3.5-triazine nucleus. B. Atrazine

Table 2: Physical properties of Atrazine

Molecular formula CsH14CINs

Molecular weight 215.7

Water solubility 33 (22 C; pH 7)

Vapour pressure 0.039 mPa (25 C)

Log Kow 2.5 (25 C) (non-ionised form

PK, 1.7 (base)

Family Common Name

Triazines Atrazine

Chloroacetanilides I Metolachlor

Ureas I Isoproturon

Aryloxyalkanoic acids I 2,4-D

Compound I Chemical name (IUPAC)

I 6-chioro-N2-ethyl-N4-isopropyi- 1,3,5-t riazine-2,4-diamine

I -chloro-N-(2-ethyl-6-methylphenyl)-N-(2-methoxy-1-methylethyl)acetamide

I -(4-isopropylphenyl)- 1 !1 -dimethylurea

(2,4-dichlorophenoxy)aceticacid

ESPR - Environ Sci & Pollut Res 9 (1) 2002 19

Pers is ten t H e r b i c i d e s Phy to remed ia t ion

1.1 Atrazine metabolism in plants

In corn, atrazine is metabolised by three major pathways:

a) dechlorination of the C 2 followed by hydroxylation takes place in the presence of a secondary metabolite (Lamoureux et al. 1998) that is accumulated to high concentrations in maize. This chemical mechanism was first described by Castelfranco et al. (1961) and its role demonstrated by Raveton et al. (1997). This process is a very active meta- bolic reaction in the young seedling. The hydroxylated atrazine is practically devoid of herbicidal activity.

b) In the aerial parts of the maize plant, a large pool of glutathione-s-transferases catalyses the formation of glu- tathione conjugates with several xenobiotic substrates (Marrs 1996). One isoenzyme of this pool catalyses glu- tathione conjugation to the C a of atrazine to give a highly hydrophilic derivative (Shimabukuro et al. 1971).

c) In many species N-dealkylation occurs catalysed by an N- dealkylase (Lamoureaux et al. 1997), leading to the for- mation of the hydrophilic derivatives N-deethylatrazine and N-deisopropylatrazine. These derivatives are not phy- totoxic. In maize leaves the N-dealkylase seems to be much less active than the glutathione-S-transferase.

1991) sometimes at levels in excess of the regulatory limit (Bintein and Devillers 1996). Moreover, atrazine traces can be detected in the atmosphere and in rain (Glotfelty 1978, Glotfelty et al. 1990, Goolsby et al. 1995, Bester et al. 1995). Although atrazine is neither a stable compound in the bio- sphere nor an accumulated one, the conditions of its world- wide agronomic use has led to its appearance in several en- vironmental compartments. Therefore for all practical pur- poses it may be considered as persistent pollutant.

Among the most interesting possible uses of phytoremedia- tion of atrazine contamination would be: 1) To prevent leaching to groundwater from unplanted

fields, for instance from the end of August to winter in maize cultivation. This would also help to attenuate ni- trate leaching;

2) to prevent surface run-off to the rivers either by main- taining buffer zones planted with species capable of me- tabolising atrazine or by cultivating detoxifying aquatic plants in the ditches, storm basins or wetlands.

Of course, the choice of plants for this purpose must be some- what restricted to monocotyledoneous species, as most broadleaf plants are killed by the compound.

1.2 Atrazine degradation in soils

Several studies carried out on seedlings under controlled conditions or in the field show that the metabolic potential of maize for atrazine is less than 100 g ha -1 year -1 for a field with 90,000 maize seeds ha -1 and treated with 1500 g atrazine ha -1 (Tasli et al. 1996). Under agronomic condi- tions, most of the metabolism of atrazine occurs in the up- per layer of the soil and is carried out by micro-organisms (Mandelbaum et al. 1995, Mirgain 1993, Shao and Behki 1995, Anderson and Coats 1995, Ostrofsky et al. 1997, Masaphy et al. 1996). Generally, the first step in soil micro- bial metabolism is N-dealkylation leading to the formation and accumulation of the metabolites N-deethylatrazine, N- deisopropylatrazine and N-didealkylatrazine in the soil. In maize fields treated repeatedly with atrazine, the capacity of the soil to metabolise this herbicide increases as the result of selective pressure on the microflora. It could be anticipated that all these transformation proc- esses might ensure the rapid disappearance of atrazine from the treated areas. However, more than thirty years of repeti- tive usage of this herbicide has led to contaminating traces of this product and its metabolites in water (Buffaut 1992) (Table 1, p. 17), food (Gunderson 1995, Wittmann and Hock

2 Isoproturon

Most members of the urea group of herbicides are com- pounds with dimethyl or methyl/methoxy substituent on the terminal N atom (Fig. 3, Table 3).

The phenylurea herbicide, isoproturon (IPU), [N-(4-isopro- pylphenyl)N',N'-dimethyl urea], is mainly used as post emer- gence herbicide in the control of grasses (e.g. blackgrass) and dicot weeds in cereals at application rates of up to 2 kg/ha. The mode of action is the plastoquinone antagonism at photosystem II. Its selectivity is based upon enhanced rhe- tabolism in crops compared to the susceptible weeds.

o ,R 3 R ~ N H - C - - N

�9 ~2 Y R4 . ,~--~ o ,CH~

P 0 ~ N H - - C - N , c H 3

A. a .

Fig. 3: Chemical structures ureas. A. Urea nucleus. B. Isoproturon

Table 3- Some chemical properties of the herbicide, isoproturon (from Perkow, 1988, Behrend and Br0ggemann 1993, modified)

Formula C12HlaN20

Molar mass 206.29 g/tool

Melting point/boiling point 151-153~ distillable

Stability Stable against light and acids, hydrolysis in the presence of strong bases

Vapour pressure 0.31 10 -'~ mbar at 77~ 0.17 10 -2 mbar at 90~

Solubility 170 mg/L in water at 25~ good in all solvents

Partition coefficients: aWwater; log kow

Sorption to soil (equilibrium coeff. K d (cm3/g)

KAw = 3.9 104, Iog Kow = 2.3

sand: 0.31; loam: 0.92

20 ESPR - Environ Sci & Pollut Res 9 (1) 2002

Phytoremediation Persistent Herbicides

In animals, IPU exhibits mutagen and potential hepatocarcino- gen properties. Connected to its ability to reach groundwater and surface waters easily (Schiilein et al. 1997), it is a com- pound with possible negative impact on human health.

Typically, IPU is applied to the soil as pre-emergence herbi- cide. Phenylureas are taken up into the plant via the leaves or the roots, but root uptake is the predominant uptake path. After transport in the xylem, the chemicals accumulate in the edges of the leaves where first necroses are observed. In the soil, IPU seems to persist for longer periods of time. In several investigations, autumn applications of IPU were still found after 10 weeks (30%; Kulshrestha 1982), 30 weeks (2-8%; Moss 1979), and a similar period (40-80%; Anon 1979). However, heavy rainfall may render the compound and its metabolites mobile as has been shown in a report examining the migration of IPU on a bavarian research farm. It was found that IPU was translocated to depths down to 170 cm at concentrations exceeding 4 lag/L and its metabo- lites at concentrations of 1 lag/L.

In any case, the breakdown of IPU in soil follows first order kinetics wi th respect to the herbicide concent ra t ion (Kulshrestha 1982). It is significant from these reports, that the presence of plants (wheat) on the sites does not contrib- ute to an increased breakdown of the compound. This in- creased persistence is apparently due to an increase in the strength of adsorption to soil matrices with time (Blair et al. 1990). The adsorption seems to be a function of organic matter but not governed by the clay content of the soils in- vestigated (Yadak and Malik 1988). Similarly, isoproturon uptake into wheat plants was found to be more rapid from soils with low sorption coefficients ( K a ) than from those with high K a (Blair et at. 1991).

From these reports it seems as if the desorption from the soil and the subsequent uptake of the parent compound into the plant represented a major barrier to remediation. Ex- periments on the addition of Triton X-100 to the applied herbicide indicate that increased solubility leads also to higher penetration into tissues (Ponte-Freitas et al. 1991).

Once inside the plant, IPU tolerance is conferred by metabo- lism rates. Generally N-demethylation or C-hydroxylation are typical plant reactions on phenylurea herbicides. This has been shown for monuron and diuron (Frear et al. 1972), and chlortoluron (Gross et al. 1979). Accordingly, also plant me- tabolism of IPU is mediated by P450 monooxygenases. As first steps of metabolism hydroxylation and demethylation have been shown (Haas 1997). Although both reactions are catalyzed by monooxygenases, evidence has been presented that two distinct P450s are responsible for these steps. It is significant to note that the product of the demethylation reac- tion, Monodesmethyl-isoproturon, exhibits residual phytotox- icity, whereas the metabolite, 2-OH-IPU, is non toxic. In ki- netic studies with cell cultures it could be shown that wheat and soybean differ exactly in this respect in their detoxifica- tion capacity (Gliif~gen et al. 1999). Whereas soybean cell cul- tures accumulate the toxic desmethyl IPU, predominantly hydroxylated metabolites are found in wheat. In a later step,

wheat builds up 2-OH-monodesmethyl-IPU as intermediary metabolite. In both cultures, 45 to 54% of the applied IPU were transformed to more polar metabolites within 48 hrs. After 7 days, wheat metabolized up to 87% of the parent compound. Contrary to the wheat cell culture studies, mi- crosome preparations from field grown wheat plants reveiled the desmethyl-IPU as major metabolite. This contradiction is not well understood and ha's not been unravelled by Gl~it~gen and coworkers (1999).

Whereas the closely related phenylurea chlortoluron is fur- ther me tabo l i zed to glucosides , no evidence for a glycosidation of IPU or its metabolites has been detected in wheat and soybean (Glfit~gen et al. 1999). It may be dis- cussed that glucosyl transfer is sterically inhibited in the case of isoproturon. Consequently, the amount of bound residues formed in the plant is also very small (below 5%).

As for phytoremediation, it has to be concluded that the re- moval of IPU from contaminated land seems to be governed mainly by its bioavailability in various soil types and by the uptake into the plants. Once metabolised by soil bacteria, the uptake of break down products of IPU has to be regarded as to be poor. However, special attention has to be focused on the fact that phenylureas are metabolized to chloroanilins, which might themselves be a threat to biosphere. In water, there might be a net uptake of the isoproturon observed de- pending on the conditions set. Fuertet-Mazel and co-workers (2000) found that IPU was bioconcentrated in the freshwater macrophytes Elodea densa and Ludwigia natans up to 60fold. However, this process might be exploited for bioremediation only for a short period of time as the plants showed massive growth reduction.

In conclusion, phenylureas tend to become persistent in soils with high organic carbon contents. Even if they are biode- gradable, a good proportion may remain present in the soil in the form of chloroanilins which have their own toxicity and persistence problems (see review of Harvey et al. this issue). For phytoremediation, most cereals would be good targets as many of them are tolerant due to their metabolic features. However, uptake of the isoproturon into the plant root has to be facilitated. Hence experimental and practical approaches for phytoremediation could include the use of biotensides such as rapeseed oils and esters. Another possibility that has not been experimentally proven so far is to set the area under con- sideration aside and leave it as fallow, but cut and remove biomass. Under these conditions, organic matter will dimin- ish with time and phenylurea herbicide residues could become accessible to soil organisms or plant roots.

3 Metolachlor

The chloroacetanilides are a large group of pre- or early post- emergence herbicides, active against a wide spectrum of an- nual grasses and broad-leaved weeds. Metolachlor is one of the most extensively used chloroacetanilides for selective weed control in maize, sorghum, soybeans, peanuts, sugarcane, cot- ton, potatoes, vegetables and woody ornamentals.

ESPR - Environ Sci & Pollut Res 9 (1) 2002 21

Persistent Herbicides Phytoremediation

N-~,3-CH2--CI

R2

Et O ~-CHz--Cl

N" / - - ( ~ ' - "CH--CH;t-OMe

Me Me

A. g. Fig. 4: Chemical structure of chloroacetanilides. A. Chloroacetanilide nucleus. B. Metolachlor

Fig. 4 shows the chemical structure of cloroacetanilides in general (Fig 4a) and of metolachlor in particular (Fig 4b). The presence of the cx-chlorine atom is necessary for herbi- cide activity (Le Baron 1988).

Chloroacetanilides are active as general plant growth inhibi- tors (Fedke 1982). Inhibition of protein and lipid synthesis has been proposed as a mechanism of phytotoxicity for all chloroacetanilides, but their primary biochemical mode of action is still not defined. Moreover, the physiological ef- fects on plants depend on the herbicide used, plant species, tissue type, herbicide concentration and treatment duration (reviewed in Le Baron 1988). Metolachtor has been shown to affect seed germination, root and shoot development, cell division, protein and lipid synthesis, membrane permeabil- ity (Fuerst 1987, Le Baron 1988, Matthes et al. 1998, Schmal- fuss et al. 2000).

The differences in tolerance to chloroacetanilides observed between plant species may be due to differences in either uptake/translocation of the herbicide or in its metabolism. In general, however, the major mechanism responsible of differential tolerance to chloroacetanilides is thought to be differential metabolism or rate of metabolism. The major route for chloroacetanilides metabolism starts with a conju- gation reaction with glutathione (see paragraph on metabo- lism). Accordingly, tolerance has been correlated with a higher availability of reduced glutathione and/or with the expression of specific glutathione S-transferase isozymes with a greater activity towards defined herbicidal molecules (Fuerst 1987, O'Connell et al. 1988, Farago et al. 1993, Cottingham and Hatzios 1992).

Several major crops, including corn and soybean, are gener- ally tolerant to metolachlor doses effective in weed control. However, intraspecific variability for tolerance is common: in corn, for instance, significant differences in chloroaceta- nilides tolerance both between inbreds and hybrid varieties have been detected (Sari Gorla et al. 1993, Cottingham et al. 1993, Cottingham & Hatzios 1992), meaning that also in generally tolerant species these herbicides are not always completely selective. To overcome this problem safeners can be used in conjunction with the herbicide itself, in order to enhance the tolerance of the crop. Safeners like cytometrinil, flurazole, oxabetrinil, fluxotenim, benoxacor have been suc- cessfully used with metolachlor in sorghum and other non completely tolerant crops. The biochemical and molecular mechanisms of action of herbicide safeners are not fully understood yet. However, at least for chloroacetanilide

Table 4: Physical properties of Metolachlor

Molecular formula C~sH=CINO=

Molecular weight 283.8

Water solubility 488 mg L -~ (25 C)

Vapour pressure 4.2 x 10 -~ Pa (25 C)

Log Kow 2.9 (25 C)

safeners, it seems clear that they mainly act by enhancing metabolic detoxification (Hatzios 1997, Jepson et al. 1994, Irzyk and Fuerst 1997, Wu et al. 1996, Han and Hatzios 1991) (Table 4).

3.1 Metolachlor uptake and translocation

Plants absorb chloroacetanilides through both roots and shoots. Corn absorbs metolachlor mostly through shoots, and variability in uptake between genotypes has been shown (Kreuz et al. 1989). Translocation of metolachlor and alachlor in plants is mostly by acropetal movement, with some limited basipetal movement (Obrigawitch et al. 1980). [14C]metolaclor applied to roots of corn seedlings grown in nutrient solution was taken up rapidly. After one week 30% of the radioactivity was found in aerial parts and 70% still in roots (Gross, reported in Le Baron 1988). Corn grown in a field plot treated preemergence with [14C]metolachlor showed, at maturity, some radioactivity in the stalk and al- most none in seeds (Sumner & Cassidy, reported in Le Baron 1988). When metolachlor was applied to seedling coleoptiles, most of it (65%) remained in the tissue, some (24%) was translocated to mesocotyls, and only marginal proportions could be detected in leaves, roots and seeds (Kreuz et al. 1989, Cottingham and Hatzios 1992).

3.2 Metolachlor metabolism

Detoxification of chloroacetanilides in planta has been shown to occur by different metabolic pathways. By far the major is the glutathione S-transferase (GST) mediated conjugation of the herbicide to reduced glutathione, GSH, (step 1), followed by vacuolar compartmentation of the conjugate through the action of a GS- ATP-dependent pump (step 2). Evidence exists for further metabolization of the glutathione conjugate in the vacuole (step 3) and possibly final excretion of the metaboIites with inclusion in the cell wall (step 4). The first two steps are so far the best characterized and recognized as of major im- portance. GSH, GSTs and the GS-pump are therefore critical for chloroacetanilides detoxification.

GSTs are a very abundant and ubiquitous class of enzymes encoded by a large multigene family. They are active as homo- or hetero-dimers and catalyse the conjugation of electrophilic and frequently hydrophobic toxic compounds with glutath- lone to form less or non toxic peptide derivatives. As for chloroacetanilides, they operate the glutathione (or homo- glutathione in some legumes) displacement of the electro- philic chlorine atom, thus drastically reducing their herbi- cidal activity.

22 ESPR - Environ Sci & Pollut Res 9 (1) 2002

Phytoremediation Persistent Herbicides

Plant GSTs classification on a functional basis has proved difficult and confusing, mainly because several members of the family are multifunctional enzymes that show broad substrate specificity. A recent classification based on aminoacid sequence similarity and intron-exon organization of the encoding genes (Dixon et al. 1998, Edwards et al. 2000) has recognized four classes of plant GSTs: Phi (F), Zeta (Z), Tau (U) and Theta (T). The Phi class is plant spe- cific and includes most of the herbicide-detoxifying GSTs; their genes contain 3 exons. The Tau class, 2 exons, is also plant specific and includes many auxin-induced GSTs. These two classes account for the majority of plant GSTs. The other two classes, Zeta (10 exons) and Theta, are close to mam- malian Zeta and Theta classes respectively, but are much less abundant in plants. GSTs with similar function and substrate specificity often belong to the same class, but sub- stantial overlapping has been observed. For instance, in maize, alachlor and metolachlor are mainly detoxified by four phi GSTs, ZmGSTFI-1, F1-2, F2-2 and F3-3 (Jepson et al. 1994, O'Connel et al. 1988, Timmermann 1989, Irzyk and Fuerst 1997, Rossini et al. 1996), but also by the tau ZmGSTUI-1 (Irzik and Fuerst 1997). GSTs involved in the metabolism of dimethenamid and metolachlor in wheat and soybean, respectively, are tau class (Riechers et al. 1997, Skipsey et al. 1997).

On the whole a very large number of plant GST enzymes/ genes have been identified in the last years, many of them as a result of sequencing projects. It now appears likely that most plant species possess up to 25 or more different Gst genes. Table 5 shows a few examples, with the relative abun- dance of the different classes, when known.

Although GS-conjugates of herbicides are usually much less toxic than the original compound, their accumulation in the cytosol could still be harmful, either because, as potential inhibitors of GSTs, they may prevent conjugation of addi- tional xenobiotics, or because they might be further metabo- lized to toxic compounds by cytosolic enzymes. Thus re- moval of conjugates from the cytosol and their compart- mentation and immobilization in the cell is usually required for efficient detoxification.

The vacuole has long been supposed to be a primary deposi- tion site of GS-conjugates. The transport mechanism has recently been elucidated with the discovery, in isolated bar- ley vacuoles, of an active GS-X vacuolar pump, capable to specifically direct the transfer of a variety of GS-conjugates across the tonoplast (Martinoia et al. 1993). This pump is functionally equivalent to the GS-conjugate Mg 2§ ATPases of mammalian cells, which are coded by members of the MRP subfamily of the ABC transporter genes (Rea 1999, Sfinchez-Fernfindez et al. 2001).

The fate of GS-conjugates after vacuolar import is less de- fined. One possibility is that they are simply trapped as such in an organelle with limited metabolic activity to eliminate them from the cytosol, where their accumulation is poten- tially harmful. Such a process has been termed 'storage ex- cretion' (Martinoia et al. 1993). However, there is evidence that GS-conjugates within the vacuole can go through fur- ther processing and degradation, and their metabolites can be recicled to the cytosol and/or definitively excreted to the apoplast. Wolf and coworkers (1996) showed that, in bar- ley, a vacuolar carboxypeptidase degrades GS-conjugates, including alachlor-GS, by cleaving the glycine residue of the GSH moiety. Also the glutamic acid residue can be succes- sively removed (Lamoureux and Rusness 1989a), possibly by a vacuolar dipeptidase (Schr6der 1997). The identifica- tion of several sulphonated chloroacetanilide metabolites in diverse crops (Lamoureoux and Rusness 1989 b, Lamoureux et al. 1991), further suggest that the cysteine conjugates are reexported from the vacuole, metabolized by cysteine S-lyases and perhaps excreted to the apoplast where they may form residues bound to lignin and cellulose.

Sulphonated chloroacetanilide metabolites have been detected in both surface and groundwaters in areas characterized by heavy use of these herbicides like the corn belt in the USA. Runoff studies from alachlor treated experimental corn fields showed that the herbicide is rapidly metabolized to its ethansulphonic acid metabolite (ESA) that, because of its wa- ter solubility, can be easily transported by surface waters (Thurman et al. 1996). How these compounds are formed in soil, whether they are microbial or reactions from plant exu- dates is not well defined yet. On the other hand data from parallel studies on hydroponically and soil grown soybeans treated with propachlor suggest that the likely major source of ESA metabolites is microbial degradation in the soil, through a metabolic pathway involving GSH conjugation as a first step, and that the metabolites detected in plants may, at least in part, be first formed in the soil and subsequently taken up by crops (Lamoureux and Rusness 1989). Since ESA metabo- lites of chloroacetanilides could be toxic (toxicity data are still scant), the possibility that they could be removed from the soil by plant uptake is intriguing.

A secondary pathway for chloroacetanilide metabolism initi- ates with their hydroxylation/oxidation and involves P450 enzymes. O-demethylation of metolachlor by microsomal P450 monooxygenases has been observed in corn, mung bean and sorghum (Moreland et al. 1990,1993 1995). The O-demethyl- ated metabolite may be further conjugated with glucose (Moreland et al. 1990) and, in analogy with other glucose- conjugated herbicides (Sandermann et al. 1991, Schmitt and Sandermann 1992, Gaillard et al. 1994) translocated into el-

Table 5: Putative number and class subdivision of expressed or potentially expressed Gst genes in three representative plant species. Arabidopsis data were obtained from EST and genomic sequences present in Gene Bank and EMBL databases

~ S ~ i ~ ' : : : ; : c l a s s ' ReferenCes

putative number phi tau zeta theta Arabidopsis 40-50 12-15 25-30 2 3-4 MiT_zi L, pers. comm.

Soybean 25 4 20 1 0 McGonigle et al. 2000

Maize 42 12 28 2 0 McGonigle et al. 2000

ESPR - Environ Sci & Pollut Res 9 (1) 2002 23

Persistent Herbicides Phytoremediation

ther the vacuole or the apoplast. The involvement of this sec- ondary pathway for chloroacetanilides detoxification is sup- ported by recent studies with transgenic potato plants co-ex- pressing three human P450s. These transgenic plants show a marked tolerance increase towards several herbicides, includ- ing metolachlor and acetochlor. Moreover the microsomal frac- tion of recombinant yeast strains expressing the same human P450s were able to metabolise the herbicides and one of the identified metolachlor metabolites was O-demethylated metolachlor (Inui et al. 2000).

4 Enhancement of Chloroacetanilide Metabolism

Since plant tolerance to chloroacetanilides is fundamentally related to their detoxification capacity via specific biochemi- cal reactions, it is reasonable to expect that higher tolerance may be achieved by two ways. i) An increase in metabolism rate and/or in the activity of the enzymes/genes involved; and ii) de novo acquisition of metabolic capabilities through introduction and expression of novel genes. As for the first point only few cases have been reported of induction by the herbicide itself: in maize, ZmGst2 transcription was specifi- cally enhanced by alachlor (Rossini et al. 1998) and in wheat GST 25 and 26 were induced by alachlor and metolachlor (Mauch and Dudler 1993). On the contrary, the inductive effects of several chloroacetanilide safeners on different metabolic pathways is well documented, and seems to be the major mechanism involved in the protective action of these compounds. In particular, safeners may affect one or more of the following: glutathione metabolism, herbicide- glutathione conjugation, herbicide hydroxylation/oxidation and the ensuing conjugation with glucose, GS- or glucoside- conjugate vacuolar transport (see Hatzios 1997 for a review). For instance, fenchlorim, which is routinely used to protect rice from pretilachlor injury, has been shown to increase GSH levels, through the induction of glutathione reductase (Han and Hatzios 1991), specific GST activity (Wu et al. 1996) and Gst mRNA levels (Wu et al. 1999). In several cereal crops a strong correlation between the protective ac- tion of chloroacetanilide safeners and their ability to increase GST activity has been demonstrated (Gronwald 1989, Cole et al. 1997, Marrs 1996). In maize, the induction of expres- sion of specific Gst genes was the most common response to safeners observed (Wiegand et al. 1986, Jepson et al. 1994, Irzik and Fuerst 1997).

The second way can be the result of different approaches, i.e. interspecific crosses, selection of spontaneous or induced mutants, genetic manipulation via recombinant DNA tech- nology. Of these the most promising to da te is the third. Depending on the mode of action of the different classes of herbicides and the mechanisms of tolerance, three main routes have been suggested or applied to engineer tolerance in the plants of interest:

1) Overexpression of the target enzyme, such as, for in- stance, EPSPS in the case of glyphosate (Shah et al. 1986);

2) modification of the target enzyme, such as ALS in the case of sulphonylureas (Mazur and Falco 1989);

3) introduction of herbicide detoxifying enzymes from di- verse origin. Nice examples of this is the obtainment of

metolachlor tolerant transgenic tobacco plants express- ing the maize Gstl and Gst2 genes (Jepson et al. 1997) and tolerant transgenic potato plants expressing human P450 genes.

5 (2,4-Dichlorophenoxy)acetic acid;2,4-D

The phenoxyalkanoic (or aryloxyalkanoic) acids are a family of selective herbicides that were developed in the 1940s. Some members of this group notably the phenoxyacetic acids like 2,4-dichlorophenoxyacetic acid (2,4-D) were among the first commercially successful synthetic herbicides. 2,4-D, is still one of the most widely applied herbicide throughout the world for both crop and non-crop uses (Fig. 5, Table 6).

2,4-D is a systemic hormone-type herbicide that acts as an antagonist to natural plant growth regulators causing ab- normal growth and development in susceptible dicotyle- donous plants.

O--CH2-COOH

CI

Fig. 5: 2,4-dichlorophenoxyacetic acid (2,4-D)

Table 6: Physical properties of 2,4-D

Molecular formula C8HECI203

Molecular weight 221.03

Water solubility 45,000mg/L (Na salt 20~ 890 (free acid at 25oC)

Vapour pressure 1 x 10 -s Pa (20~

Partition coefficients Log Kow = 2.58-2.53 (pH 1)

The wide use of 2,4-D and fears that it might be a danger to human health has prompted extensive research and evalua- tion of its environmental toxicology and biodegradation. The acid form (2,4-dichlorophenoxy)acetic acid is the herbicidal moiety. However most commercial formulations contain 2,4- D as an inorganic or amine salt or as an ester. Results from a number of studies have indicated that the rates of break- down of the amine and ester forms in soil are similar (Wilson et al. 1997). In soils the 2,4-D ester underwent hydrolysis to the acid at a rate dependent on pH. For example the half- life at pH 8.0 for the butoxy ester was 9 hours, whereas at pH 5.0 it was more than one year (Smith 1988).

5.1 2,4-D metabolism in soils

In the soil 2,4-D is relatively immobile (lowest average depth detected 60 cm.) and has short half-life. In studies in the

24 ESPR - Environ Sci & Pollut Res 9 (1) 2002

Phytoremediation Persistent Herbicides

USA, its average half-life in soils ranged from 6.4 days in southern US soils to 8.3 days in soils with high organic mat- ter (Roberts 1998). The half-life in natural water was two to four weeks, although in areas such as a treated rice paddy, the half-life was as short as one day. The acid form of 2,4- D, as well as the amine and ester chemical groups, is me- tabolised by soil microorganisms. Microbial degradation is brought about by a variety of microorganisms and soils pre- viously treated with 2,4-D may become enriched in these microorganisms and degradation may take place without a significant lag period. Typically microbial metabolism leads either to the formation of Krebs cycle acids or by complete degradation to carbon dioxide. As a result some of the car- bon will become associated with the microbial biomass.

Details of the metabolism of 2,4-D by soil microbes are be- yond the scope of this review. In general the main steps in the pathways of microbial degradation are the enzymatic cleav- age of the side chain forming glyoxylate and the parent phe- nol, followed by oxidation and cleavage of the aromatic ring.

5.2 2,4-D metabolism in plants

2,4-D is taken up, systemically translocated and actively metabolised in several dicotyledonous and monocotyle- donous plant species, for example bean, soybean and maize and wheat respectively. In contrast to the degradative me- tabolism of soil microbes, the major metabolic pathways in plants involve hydroxylation and conjugation with sugars or amino acids. The principal pathways are shown in Fig. 6.

The pathways used and the mixture of products formed de- pend on the plant species. In addition some pathways are complicated further by steps that include migration of the chlorine substituent to another site along with hydroxylation of the aromatic ring. For example in bean and soybean 2,3 dichloro-4-phenoxyacetic acid is a major product of metabo- lism. In several species the hydroxylated products are con-

0 II O-CH2-C--NH--asp ~ CI ql

CI

O-CH2-COOH @c, OH

CI

l, O--CH2-CH2-CH2-COOH @c, CI

O--CH2-COOH

CI o--glucose

O-CH2-COOH

2 I, ~ C I

CI OH

Fig. 6: Metabolism of 2,4-D. Reactions shown are 1. Hydroxylation, 2. Hydroxylation and chlorine migration, 3. Glucose conjugation, 4. Side chain extension and 5. Amino acid conjugation

jugated to form O-~-D-glucosides. Conjugates of aspartate and glutamate are the most frequently observed amino acid conjugates. Formation of these conjugates may be followed by subcellular compartmentation in the vacuole. The ]]-D- glucoside conjugate was the first reported example of vacu- olar sequestration of a herbicidal conjugate (Schmitt and Sandermann 1982). Recent studies suggest that the glucosyl conjugation takes place in the cytosol and members of ATP- binding cassette (ABe) transporter family catalyses the trans- port of the glucoside into the vacuole. In addition to the hydroxylation and conjugation reactions, some grasses can metabolise 2,4-D by lengthening of the side chain. For ex- ample in alfalfa (Medicago sativa L.) side chain extension forming 2,4-dichlorobutyric acid is serves as resistant mecha- nism to 2,4-D (Roberts 1998).

At some sites 2,4-D has been identified as an important pol- lutant. However under normal conditions contamination from the agricultural use of the herbicide does not lead to persistence in soil, water and vegetation as both plants and microorganisms can metabolise 2,4-D. Whereas microorgan- isms can effect complete biodegradation, metabolism of 2,4- D in plants leads to chemical transformation of the com- pound followed by sequestration of the non-toxic products. Remediation of sites affected for example by point source contamination of the herbicide will require the concerted activities of soil microorganisms (bioremediation) and plants (phytoremediation). Harvesting of the plants, followed by a post-harvest polishing step such as composting could bring about complete degradation of the parent herbicide. In ad- dition the genes encoding the enzymes that catalyse degra- dation in microbes are potential resources for enhancement of the biodegradation of the herbicide.

6 Conclusions

Persistent herbicides from four compound classes have metab- olism in plants via distinct pathways. All of them follow patterns that have been well described during recent years. The metabolic net includes the well known three phases cas- cade (Coupland 1991) and compounds are detoxified via conjugation reactions. Fig. 7 shows that upon conjugation a number ofenzymatic reactions convert the compounds into metabolites the final fate of which is either the apoplastic space and cell wall or the vacuole. The latter compartment is thought as a locus for storage excretion, and thus metab- olites will become available to the food chain when the plant is eaten or dies. On the basis of current literature, it is un- likely to find total mineralization of herbicides to CO 2 in higher plants. Bound residues may, however, be unavailable to the members of the food web.

Soil applied herbicides tend to be adsorbed in soil particles and pose more problems for bioremediation. No valid con- cept exists for the solubilization of soil bound herbicides and their subsequent uptake into plants for remediation. Quite frequently, microbial metabolism is found to alter the active compounds and to partially mineralize them (Scheu- nert and Schr6der 1998). However, microbes do so only in the course of co-metabolism and leave residues behind once they get access to other energy sources.

ESPR - Environ Sci & Pollut Res 9 (1) 2002 2 5

Persistent Herbicides Phytoremediation

Phase I

Phase II

Fate o f herbicides in plants

Chloroacet- Ureas Aryloxyalkanoic anilides acids

Triazines

�9 demethylatiin ' 1 hydroxylation

demethylation

I I Phase HI vacuole vacuole vacuole vacuole

soluble metabolites . . . . -1, secondary conjugates

Fig. 7" Fate of herbicides in ~lants. The thickness and length of the arrows indicates the importance of the respective metabolic step for the corresponding herbicide. Generally herbicides are activated, conjugated and further metabolized according to Coupland's three phase scheme (Coupland 1991) and finally stored as soluble non toxic compounds or bound residues

For the removal of diffuse pollution of persistent herbicides f rom agricultural or set aside land, agronomic practice has to be applied. Farming techniques will provide the correct seed density and nutrient supply, and the growing canopy will pre- vent the pollutant f rom leaching into ground water or surface runoff. Treatment with herbicide antidotes may strengthen the plant performance, but ways to enhance the uptake f rom soil have to be found. As some compounds are adsorbed to or- ganic matter in the soil, setting aside the land before agro- nomic phytoremediat ioon programs are started seems help- ful, as this might solubilize the herbicidal pollutants for phytoextract ion. In any case, care should be taken to prevent the plants used for phytoremediat ion being used as fodder.

Wi th view to the growing demands of the wor ld popula t ion for heal thy food, clean water, space for housing and recrea- tion, phytoremedia t ion might present an ecologically and economical ly sound wa y for the removal of persistent her- bicides f rom ecosystems.

R e f e r e n c e s

Anderson TA, Coats JR (1995): Screening rhizosphere soil samples for the ability to mineralize elevated concentrations of atrazine and metolachlor. Journal of Science and Health B30, 473-484

Anon A (1979): Herbicide residue levels and redrilling of failed cere- als. MAFF ADAS South East Region Agricultural Science Service. Annual Report, HMSO, London

Behrendt H, Brtiggemann R (1993): Modelling the fate of organic chemicals in the soil plant environment: Model study of root up- take of pesticides. Chemosphere 12, 2325-2332

Bester K, Hiihnerfuss H, Brockmann U, Rick HJ (1995): Biological effects of triazine herbicide contamination on marine phytoplank- ton, Arch Environ Contain Toxicol 23, 277-283

Bintein S, Devillers J (1996): Evaluating the environmental fate of lindane in France. Chemosphere 32, 2427-2440

Blair AM, Martin TD, Brain P, Cotterill EG (1991): The interaction between planting depth of four winter wheat cultivars, Alopecurus

myosuroides Huds. and Bromus sterilis L. and their susceptibility to post-emergent applications of isoproturon and chlortoluron. Weed Res 31,285-293

Blair AM, Martin TD, Walker A, Welch SJ (1990): Measurement and prediction of isoproturon movement and persistence in three soils. Crop Protection 9, 289-294

Buffaut P (1992): Produits phytosanitaires: teneures actuelles des eaux en France. Colloque Phyt'eau - Produits Phytosanitaires - Usages Agricoles et Connexe 21, 5-19

Castelfranco P, Foy CL, Deutsch DB (1961): Nonenzymatic detoxi- cation of 2- chloro-4,6-bis (etymalmino)-s-triazine (simazine) by extracts of Zea mays. Weed 9, 580

Cole DJ, Cummings I, Hatton PJ, Dixon D, Edwards R (1997): Glu- tathione transferases in crops and major weeds. In: Hatzios KK (Ed): Regulation of enzymatic systems detoxifying xenobiotics in plants, NATO ASI series Vol 37, Kluwer Academic Publishers, Dordrecht, pp 139-154

Cottingham CK, Hatzios KK (1992): Basis of differential tolerance of two corn hybrids (Zea mays) to metolachlor. Weed Sci 40, 359-363

Cottingham CK, Hatzios KK, Meredith SA (1993): Comparative re- sponses of selected corn (Zea mays) hybrids to EPTC and metola- chlor. Weed Res 33, 161-170

Coupland D (1991): Detoxification of herbicides in plants. In: Caseley JC, Cussan GW, Atkin RK (Eds): Herbicide resistance in weeds and crops, Wiley, New York, pp 263-278

Dixon DP, Cole DJ, Edwards R (1998): Purification, regulation and cloning of a glutathione transferase (GST) from Maize resem- bling the auxin-inducible type-HI GSTs. Plant Molecular Biology 36, 75-87

Edwards R, Dixon DP, Walbot V (2000): Plant glutathione S-trans- ferases: Enzymes with multiple functions in sickness and in health. Trends in Plant Science 5, 193-198

European Commission's Directive on Drinking Water 80-778-EEC (1980): European Commission, Brussels

Farago S, Kreuz K, Brunold C (1993): Decreased glutathione levels enhance the susceptibility of maize seedlings to metolachlor. Pestic Biochem Physiol 47, 199-205

Fedke C (1982): Biochemistry and physiology of herbicide action, Springer Verlag, Berlin Heidelberg New York, 148-158

26 ESPR - Environ Sci & Pollut Res 9 (1) 2002

Phytoremediation Persistent Herbicides

Frear DS, Swanson HR, Tanaka FS (1972): Herbicide metabolism in plants. In: Runeckles VC, Tso TC (Eds): Structural and functional aspects of phytochemistry. Recent advances in phytochemistry, Acad Press, New York, 225-247

Fuerst EP (1987): Understanding the mode of action of the chloro- acetanilide and thiocarbammate herbicides. Weed Technol 1, 270-277

Fuertet-Mazel A, Grollier T, Grouselle M, Ribeyre F, and Boudou A (2000): Experimental study of bioaccumulation and effects of the herbicide isoproturon on freshwater rooted macrophytes (Elodea densa and Ludwigia natans). Chemosphere 32, 1499-1512

Galliard C, Dufaud A, Tommasini R, Kreuz K, Amrhein N, Martinoia E (1994): A herbicide antidote (safener) induces the activity of both the herbicide detoxifying enzyme and of a vacuolar transporter for the detoxified herbicide. FEBS Letters 352 [FEBS 14574], 219-221

Gl~iflgen WE, Komot~a D, Bohnenk~imper O, Haas M, Hertkorn N, May RG, Szymczak W, Sandermann H (1999): Metabolism of the herbicide Isoproturon in wheat and soybean cell suspension cultures. Pestic Biochem Physiol 63, 97-113

Good NE (1963) Carbon dioxide and the Hill reaction. Plant Physiol 38, 298-304

Goolsby DA, Thurman EM, Pomes ML, Meyer MT, Battaglin WA (1997): Herbicides and their metabolites in rainfall-Origin, trans- port, and deposition patterns across the Midwestern and North- eastern United States, 1990-1991: Environmental Science & Tech- nology 31, 1325-1333

Glotfelty DE (1978) The atmosphere as a sink for applied pesticides. J Air Pollut Control Assoc 28,917-921

Glotfelty DE, Williams GH, Freeman HP, Leech MM (1990) Re- gional atmospheric transport and deposition of pesticides in Mary- land. In: Kurtz D (Ed): Long Range Transport of Pesticides, Lewis Publishing Co, Chelsea, pp 199-222

Gronwald JW (1989): Influence of herbicide safeners on herbicide metabolism. In: Hatzios KK, Hoagland RE (Eds): Crop safeners for herbicides: development, uses and mechanisms of action, Aca- demic Press, San Diego, pp 163-175

Gross D, Laanio T, Dupuis G, Esser HO (1979): The metabolic behav- iour of chlortoluron in wheat and soil. Pestic Biochem Physio110, 49

Gunderson EL (1995): FDA Total Diet Study, July 1986-April 1991, Dietary Intakes of Pesticides, Selected Elements, and Other Chemi- cals, Journal of AOAC Intl 78, 6

Haas M (1997): Metabolisierung yon Xenobiotika durch pflanzliche ZeUkulturen und Enzyme. Dissertation TU-Miinchen. Shaker Verlag, Aachen, Germany

Han S, Hatzios KK (1991): Effects of the herbicide pretilachlor and the safener fenclorium on glutathione content and glutathione- dependent enzyme activity of rice. Zeitschrift ffir Naturforschung 46c, 861-865

Hatzios KK (1997): Regulation of xenobiotics degrading enzymes with herbicide safeners. In: Hatzios KK (Ed): Regulation of enzymatic systems detoxifying xenobiotics in plants, NATO ASI series Vol 37, Kluwer Academic Publishers, Dordrecht, pp 275-288

Inui H, Kodama T, Ohkawa Y, Ohkawa H (2000): Herbicide me- tabolism and cross-tolerance in transgenic potato plants co-ex- pressing human CYPIA1, CYP2B6, and CYP2C19. Pestic Biochem Physiol 66, 116-129

Izryk GP, Fuerst EP (1997): Characterization and induction of maize glutathione S-transferases involved in herbicide detoxification. In: Hatzios KK (Ed): Regulation of enzymatic systems detoxifying xenobiotics in plants, NATO ASI Series Vo137, Kluwer Academic Publishers, Dordrecht, pp 155-170

Jepson I, Lay VJ, Holt DC, Bright SWJ, Greenland AJ (1994): Clon- ing and characterization of maize herbicide safener-induced cDNAs encoding subunits of glutathione S-transferase isoforms I, II and IV. Plant Molecular Biology 26, 1855-1866

Jepson I, Holt DC, Roussel V, Wright SY, Greenland AJ (1997): Transgenic plant analysis as a tool for the study of glutathione S- transferases. In: Hatzios K.K. (Ed): Regulation of enzymatic sys- tems detoxifying xenobiotics in plants, NATO ASI Series Vol 37, Kluwer Academic Publishers, Dordrecht, pp 313-323

Kreuz K, Gaudin J, Ebert E (1989): effects of the safeners CGA 154281, oxabetrinil and fenclorim on uptake and degradation of metolachlor in corn (Zea mays L.) seedlings. Weed Research 29, 399-405

Kulshresta G (1982): Hydrolysis of isoproturon in aqueous medium and its persistence in soil and plants. Ind J Weed Sci 14, 96-102

Lamoureux GL and Rusness DG (1989a): The role of glutathione and glutathione S-transferases in pesticide metabolism, selectiv- ity and mode of action in plants and insects. In: Dolphin D, Poulson R, Avramovic O (Eds): Glutathione: Chemical biochemical and medical aspects, Vol IIIB, Set: Enzyme and Cofactors, J. Wiley & Sons, New York, pp 153-196

Lamoureux GL and Rusness DG (1989b): Propachlor metabolism in soybean plants, excised soybean tissues, and soil. Pestic Biochem Physiol 34, 187-204

Lamoureux GL, Rusness DG, Schrtder P, Rennenberg H (1991): Di- phenyl ether herbicide metabolism in a spruce cell suspension cul- ture: The identification of two novel metabolites derived from a glutathione conjugate. Pestic Biochem Physiol 39,291-301

Lamoureux GL, Simoneaux B, Larson J (1998): The metabolism of atrazine and related 2-chloro-4,6-bis(alkylamino)-s-triazines in plants. In: Ballantine LG, McFarland JE, Hacket DS (Eds): Tri- azine Herbicides: Risk Assessment, Washington, D.C., American Chemical Society

Le Baron MH, McFarland JE, Simoneaux BJ (1988): Metolachlor. In: Kearney PC, Kaufmann DD (Eds): Herbicides: Chemistry, Deg- radation and Mode of Action, Vol 3, Marcel Dekker, New York, pp 336-383

Mandelbaum RT, Allan DL, Wackett LP (1995): Isolation and char- acterization of a Pseudomonas sp. that mineralizes the s-ttiazine herbicide atrazine. Appl Environ Microbiol 61,1451-1457

Marrs KA (1996): The functions and regulation of glutathione S- transferases in plants. Annu Rev Plant Physiol 47, 127-158

Martinoia E, Grill E, Tommasini R, Kreuz K, Amrhein N (1993): ATP-dependent glutathione S-conjugate 'export' pump in the vacu- olar membrane of plants. Nature 364, 247-249

Masaphy S, Henis Y, Levanon D (1996): Manganese-enhanced bi- otransformation of atrazine by the white rot fungus Pleurotus pulmonarius and its correlation with oxidation activity. Appl Environ Microbiol 62, 3587-3593

Marthes B, Schmalfuss J, Boger P (1998): Chloroacetamide mode of action; II: Inhibition of very long chain fatty acid synthesis in higher plants. Z Naturforsch 53c, 1004-1011

Mauch F, Dudler R (1993): Differential induction of distinct glutath- ione-s-transferases of wheat by xenobiotics and by pathogen at- tack. Plant Physiol 102, 1193-1201

Mazur B J, Falco SC (1989): The development of herbicide resistant crops. Annu Rev Plant Physiol Plant Mol Biol 40, 441-470

McGonigle B, Keeler S, Lau S-MC, Koeppe M, O'Keefe DP (2000): A genomics approach to the comprehensive analysis of the glu- tathione S-transferase gene family in soybean and maize. Plant Physiology 124, 1105-1120

Mirgain I, Schenck C, Monteil H (1993): Atrazine Contamination of Groundwaters in Eastern France in Relation to the Hydrogeolog- ical Properties of the Agricultural Land. Environmental Technol- ogy 14, 741-751

Moreland DE, Corbin FT, Novitzky WP, Parker CE, Tomer KB (1990): Metabolism of metolachlor by a microsomal fraction isolated from grain sorghum (Sorghum bicolor) shoots. Z Naturforsch 45c, 558-564

Moreland DE, Corbin FT, McFarland JE (1993): Effects of safeners on the oxidation of multiple substrates by grain sorghum micro- somes. Pestic Biochem Physiol 45, 43-53

ESPR - Environ Sci & Pollut Res 9 (1) 2002 27

Persistent Herbicides Phytoremediation

Moreland DE, Corbin FT, Fleischmann TJ, McFarland JE (1995): Partial characterization of microsomes isolated from mung bean cotyledons. Pestic Biochem Physiol 52, 98-108

Moss SR (1979): The influence of tillage and method of straw dis- posal on the survival and growth of black grass Alopecurus myosuroides and its control by chlortoluron and isoproturon. Ann Appl Biol. 91, 91-100

O'Connell K, Breaux E, FraIey R (1988): Different rates of metabo- lism of two chloroacetamide herbicides in pioneer 3320 corn. Plant Physiol 86, 359-363

Obrigawitch T, Abernathy JR, Gipson JR (1980): Response of yel- low ( Cyperus esculentus) and purple ( Cyperus rotundus) nutsage to metolachlor. Weed Sci 28,708-715

Ostrofsky A, Jellison J, Smith K, Shortle W (1997): Cation concen- trations in red spruce wood undergoing fungal biodegradation. Can J For 27, 567-571

Perkow W (1988): Wirksubstanzen der Pflanzenschutz- und Sch~id- lingsbek~impfungsmittel. Paul Parey, Berlin-Hamburg

Ponte-Freitas A, Haddad G, Tissut M, Ravanel P (1991): Distribu- tion of isoproturon, a photosystem II inhibitor, inside wheat leaf fragments. Plant Physiol Biochem 29, 67-74

Raveton M, Ravanel P, Serre AM, Nurit F, Tissut M (1997): Kinetics of uptake and metabolism of atrazine in model plant systems. Pestic Sci 49, 157-163

Rea PA (1999): MRP subfamily ABC transporters from plants and yeast. J Exp Bot 50, 895-913

Riechers DE, Irzyk GP, Jones SS, Fuerst EP (1997): Partial charac- terization of glutathione S-transferases from wheat (Triticum spp) and purification of a safener-induced glutathione S-transferase from Triticum tauschii. Plant Physiol 114, 1461-1470

Roberts TE (1998): Metabolic pathways of agrochemicals. Part 1: Herbicides and and plant growth regulators. Royal Society of Chemistry, Cambridge, UK

Rossini L, Jepson J, Greenland AE, Sari Gorla M (1996): Characteri- zation of glutathione S-transferase isoforms in three maize inbred lines exhibiting differential sensitivity to alachlor. Plant Physiol 112, 1595- 1600

Rossini L, Frova C, P~ ME, Mizzi L, Sari Gorla M (1998): Alachlor regulation of maize glutathione S-transferase genes. Pestic Biochem Physiol 60, 205-211

Sfinchez-Fernfindez R, Davies TGE, Coleman, JOD, Rea PA (2001): The Arabidopsis thaliana ABC protein superfamily: a complete inventory. J Biol Chem 276, 30231-30244

Sandermann H, Schmitt H, Eckey R, Bauknecht T (1991): Plant bio- chemistry of xenobiotics: Isolation and properties of soybean O- and N- glucosyl- and O- and N-malonyltransferases for chlorin- ated phenols and anilines. Arch Biochem Biophys 287, 341-350

Sari Gorla M, Ferrario S, Rossini L, Frova C, Villa M (1993): De- velopmental expression of glutathione S-transferase in maize and its possible connection with herbicide tolerance. Euphytica 67, 221-230

Schmalfuss J, Matthes B, Knuth K, Boger P (2000): Inhibition of Acyl-CoA elongation by chloroacetamide herbicides in microsomes from leek seedlings. Pestic Biochem Physiol 67, 25-35

Schmitt R, Sandermann H (1982): Specific localisation of ~-gluco- side conjugates of 2,4-dichlorophenoxyacetic acid in soybean vacuoles, Z. Naturforsch 37c, 772-777

Schr6der P (1997): Fate of glutathione S-conjugates in plants: Cleav- age of the glutathione moiety. In: Hatzios KK (Ed): Regulation of enzymatic systems detoxifying xenobiotics in plants, NATO ASI Series Vo137, Kluwer Academic Publishers, Dordrecht, The Neth- erlands, 233-244

Scheunert I, Schr6der P (1998): Formation, Characterization and Re- lease of Non-Ectractable Residues of [C-14]-Labeled Organic Xenobiotics in Soils. ESPR -Environ Sci & Pollut Res 5,238-245

Schiilein J, Gl~iflgen WE, Hertkorn N, Schr6der P, Sandermann H, Kettrup A (1996): Detection and identification of the herbicide isoproturon and its metabolites in soil solution, runoff and sur- face water after a heavy rainfall event. Int J Env Analyt Chem 65, 193-202

Shah DM, Horsch RB, Klee HJ, Kishore GM, Winter JA, Tumer NE, Hironaka CM, Sanders PR, Gasser CS, Aykent S, Siegel NR, Rogers SG, Fraley RT (1986): Engineering herbicide tolerance in transgenic plants. Science 233,478-481

Shao ZQ, Behki R (1995): Cloning of the genes for degradation of the herbicides EPTC (S-Ethyl Dipropytthiocarbamate) and Atra- zine from Rhodococcus sp. Strain TEl. Applied and Environmental Microbiology 61, 2061-2065

Shimabukuro RH, Frear DS, Swanson HR, Walsh WC (1971): Glu- tathione conjugation. An enzymatic basis for atrazine resistance in corn. Plant Physiol 47, 10-14

Skipsey M, Andrews CJ, Towson JK, Jepson I, Edwards R (1997): Substrate and thiol specificity of a stress-inducible glutathione transferase from soybean. FEBS Lett 409, 370-374

Smith AE (1988): Transformations in soil, environmental chemistry of herbicides, R. Grover, Boca Raton, FLA, CRC Press, pp 171-200

Tasli S, Patty L, Boetti H, Ravanel P, Vachaud G, Scharff C, Favre- Bonvin J, Kaouadji M, Tissut M (1996): Persistence and leaching of atrazine in corn culture in the experimental site of La Cote Saint Andre (Isere, France). Arch Environ Contam Toxicol 30, 203-212

Thurman E. M., Goolsby D.A., Aga D.S., Pomes M.L., Meyer M.T. (1996): Occurrence of alachlor and its sulphonated metabolite in rivers and reservoirs of the midwestern United States: the impor- tance of sulphonation in the transport of chloroacetanilide herbi- cides. Environ. Science & Technology 30, 569-574

Timmerman KP (1989): Molecular charaterization of corn glutath- lone S-transferase isozymes involved in herbicide detoxication. Physiologia Plantarum 77, 465--471

Wiegand RC, Shah DM, Mozer TJ, Harding El, Diaz-CoUier J, Saunders C, Jaworsky EG, Tiemeier DC (1986): Messenger RNA encoding a glutathione S-transferase responsible for herbicide tol- erance in maize is induced in response to safener treatment. Plant Mol Biol 7, 235-243

Wilson RD, Geronimo J, Armbruster JA (1997)" 2,4-D Dissipation in field soils after applications of 2,4-D dimethylamine salt and 2,4-D 2-ethylhexyl ester. Environmental Toxicology and Chemis- try 16, 1239-1246

Wittmann C, Hock B (1991 ): Development of an ELISA for the analy- sis of atrazine metabolites deethylatrazine and deisopropylatrazine. J Agric Food Chem 39, 1194-1200

Wolf AE, Dietz K J, Schroeder P (1996): Degradation of glutathione S-conjugates by a carboxypeptidase in the plant vacuole. FEBS Letters 384, 31-34

Wu J, Omokawa H, Hatzios KK (1996): Glutathione S-transferase activity in unsafened and fenchlorim-safened rice (Oryza sativa). Pestic Biochem Physiol 54, 220-229

Wu J, Cramer CL, Hatzios KK (1999): Characterization of two cDNAs encoding glutathione S-transferases in rice and induction of their transcription by the herbicide safener fenchlorim. Physiol Plantarum 105, 102-108

Yadav A, Malik RK (1988): Persistence of isoproturon in different soils. Haryana Agric Univ J Res 18, 198-206

Received: December 17th, 2001 Accepted: December 20th, 2001

OnlineFirst: December 21st, 2001

28 ESPR - Environ Sci & Pollut Res 9 (1) 2002