Bacterial Pathways for Degradation of tics

download Bacterial Pathways for Degradation of tics

of 6

Transcript of Bacterial Pathways for Degradation of tics

  • 8/9/2019 Bacterial Pathways for Degradation of tics

    1/6

    HIGHLIGHT www.rsc.org/npr | Natural Product Reports

    Bacterial pathways for degradation of nitroaromatics

    Zoe C. Symons and Neil C. Bruce*

    Received (in Cambridge, UK) 27th June 2005

    First published as an Advance Article on the web 6th October 2006

    DOI: 10.1039/b502796a

    Covering: up to 2006

    The last one hundred years have seen a massive expansion in the chemicals industry; however, with this

    progress came the concomitant pollution of the environment with a significant range of xenobiotics.

    Nitroaromatic compounds form one such category of novel environmental contaminants and are

    produced through a large number of industrial processes, most notably the pesticides, dyes and

    explosives industries. Whilst singly nitrated aromatic compounds are usually mineralised in the

    environment, multiply nitrated aromatics, such as the explosive 2,4,6-trinitrotoluene (TNT), are

    recalcitrant and highly toxic. The predominant route of biological transformation of aromatic

    compounds is oxidation; however, the presence of three electron-withdrawing nitro-groups around the

    ring prevents oxidation, rendering such compounds resistant to biodegradation. The subsequent

    accumulation of these contaminants has stimulated much research leading to the isolation of bacteriathat possess, to varying extents, the ability to remediate explosives and other nitroaromatic pollutants.

    The extreme environments created by these toxic substances accelerate the evolutionary process and

    examples of bacteria that have conscripted metabolic enzymes for novel remediatory pathways are

    included. This Highlight ends with a discussion of the future of nitroaromatic bioremediation including

    engineering plants to express bacterial enzymes for use in bioremediation programs.

    Introduction

    The chemical versatility of the nitrogroup means nitroaromatic

    compounds are widely used in industry in the synthesis of

    dyestuffs, pesticides, explosives and polyurethane foams. Ni-

    troaromatics also form the basis of explosives such as 2,4,6-

    trinitrotoluene (TNT) and 2,4,6-trinitrophenol (picric acid)and have historically been produced on a massive scale. 2,4-

    CNAP, The Department of Biology, AREA 8, The University of York, POBox 373, York, England YO10 5YW. E-mail: [email protected]; Fax: 44-(0)1904 328801; Tel: 44-(0)1904 328777

    Zoe Symons obtained her Master of Biochemistry from the University of Bath, studied for her PhD at the Institute of Biotechnology,

    University of Cambridge, and is currently a post-doctoral research fellow in CNAP, University of York.

    Zoe C. Symons Neil C. Bruce

    Neil Bruce is currently Professor of Biotechnology at CNAP in the

    Department of Biology at the University of York. Following a PhDin Microbial Biochemistry at the University of Kent in 1987, he

    joined the Institute of Biotechnology at the University of Cambridge

    as a postdoctoral research associate with Prof. Chris Lowe and was

    subsequently awarded a Research Fellowship at Wolfson College. He

    was appointed to a University Lectureship in Biotechnology in the same

    departmentin 1990 and subsequently appointed to a Fellowship at Trinity

    Hall. He was promoted to a Personal Readership in Biotechnology in

    2001. In 2002 he was appointed to the Chair of Biotechnology at the

    University of York. The major research themes of his laboratory are

    microbial metabolism, biocatalysis and environmental biotechnology.

    Dinitrotoluene (DNT), an intermediate in TNT manufacture

    and starting material for the synthesis of toluene diisocyanate,

    is also produced on a large scale. The extensive use of many

    nitroaromatic compounds in such chemical processes has led to

    their widespread contamination of the environment. Five hundred

    thousand U. S.gallons of water, contaminated with trinitrotoluene

    and other nitroaromatics, may be released into the environment

    by one TNT-manufacturing plant in one day.1 At some munitions

    manufacturing and processing sites the contamination can be

    as high as 200g TNT per 1 kg of soil.2 Many of the multiply-

    nitrated compounds are cytotoxic and have been classified as

    probable carcinogens. TNT, for example, is documented as causing

    This journal is The Royal Society of Chemistry 2006 Nat. Prod. Rep., 2006, 23, 845850 | 845

  • 8/9/2019 Bacterial Pathways for Degradation of tics

    2/6

    erythrocyte abnormalities, dysfunction of the liver and cancer

    in mammals. Unfortunately, once released into the environment

    many of these compounds persist for long periods of time.

    This is likely to be because similar compounds rarely occur

    in the Natural world, their structural novelty rendering them

    resistant to breakdown as indigenous organisms have no means of

    metabolising them.3

    The inherent toxicity of these anthropogenic compounds means

    they should not be permitted to remain in the environment;

    however, the widespread contamination of the soil and ground

    water worldwide means that inexpensive and sustainable tech-

    nologies to remove these toxic compounds need to be developed

    and employed. Current methods of remediating contaminated

    sites, such as incineration and composting, can be expensive

    and harmful. Incineration of soil to rid it of explosives can

    result in the exposure of workers to high levels of toxins.3,4,5 One

    solution to the problem of nitroaromatic contamination, which

    is considered both economically feasible and environmentally

    sound, is bioremediation. This is the use of organisms, such as

    microbes or plants, to degrade or detoxify hazardous materials. To

    these ends research groups have employed selective enrichments,

    with the nitroaromatic supplied as the sole carbon and/or nitrogen

    source for growth, to isolate specialist microbes with the ability

    to transform nitroaromatics. In this way, bacteria and fungi have

    been isolated with the ability to transform nitroaromatic com-

    pounds. Enzymes responsible for conferring the nitroaromatic-

    transforming activity on a number of microbes have been cloned

    and characterised, offering an enhanced understanding of the bac-

    terial metabolism of nitroaromaticsthe subject of this Highlight.

    Enzymes associated with nitroaromatic transformation

    Many enzymes have been characterised as being capable of trans-

    forming nitroaromatic compounds via oxidative and reductiveattack. These include monooxygenases, dioxygenases, hydride

    transferases and nitroreductases. Most recently, an iron-only

    hydrogenasefrom Clostridium acetobutylicum hasalso been shown

    capable of reducing TNT to its dihydroxylamino derivative in a

    hydrogen-dependent manner and has been demonstrated to be

    responsible for the great extent of the nitroreductase activity ofC.

    acetobutylicum.6

    The evolution of nitroaromatic degradation pathways

    Recent work on the biodegradation of dinitrotoluene by

    Burkholderia cepacia R34 offers one of the most elegant examples

    of how bacteria have evolved to not only tolerate but to metabolise

    nitroaromatic compounds as a nutrient source for growth.7 B.

    cepacia R34 was isolated from surface water from an ammunition

    plant and was therefore subject to significant environmental selec-

    tive pressures. Analysisof theenzymes it employs to transform2,4-

    DNT has revealed that the bacterium has co-opted enzymes from

    diverse pre-existing pathways to produce one specific to 2,4-DNT.7

    The genes encoding the enzymes required for degradation are

    found on a contiguous 27 kb region of DNA, which also includes

    regions devoted to the regulation of the pathway. Considering that

    2,4-DNT has only been produced for the last one hundred years

    the ability of bacteria to piece together a new metabolic pathway

    to utilise this compound as a carbon source for growth is quite

    remarkable.

    The enzymes which form the pathway have been characterised

    individually and their reactions are illustrated in Fig. 1.

    Fig. 1 The metabolism of 2,4-dinitrotoluene catalysed by Burkholderia

    cepacia R34, adapted from Johnson et al.7 Compounds in brackets remain

    hypothetical.

    The first denitration is catalysed by DNT dioxygenase (encoded

    by dntAaAbAcAd), a Rieske non-haem iron oxygenase recruited

    from a naphthalene degradation pathway. This oxygenase has

    three parts, an oxidoreductase (Aa), an iron-sulfur ferredoxin pro-

    tein (Ab) and a terminal oxygenase centre (Ac and Ad form a and

    b subunits). The second denitration is catalysed by the 4-methyl-

    5-nitrocatechol monooxygenase (dntB) believed to be seconded

    from a degradation pathway originally for chloroaromatic com-

    pounds. The 2,4,5-trihydroxytoluene then undergoes ring fission

    catalysed by 2,4,5-trihydroxytoluene oxygenase (dntD). The 2,4-

    dihydroxy-5-methyl-6-oxo-2,4-hexadienoic acid, the ring fission

    product, is metabolised by the enzyme dntG to yield pyruvate and

    methylmalonic acid semialdehyde in turn transformed by dntE,

    although the final product propionyl-CoA remains hypothetical.

    All the last three enzymes have been hijacked from amino acid

    846 | Nat. Prod. Rep., 2006, 23, 845850 This journal is The Royal Society of Chemistry 2006

  • 8/9/2019 Bacterial Pathways for Degradation of tics

    3/6

    degradation pathways. The enzymes which form this novel 2,4-

    DNT degradation pathway are all encoded on a single contiguous

    length of DNA, producing a pathway specific to an anthropogenic

    compound.

    The correct functioning of a pathway that has adapted or been

    created to transform nitroaromatic compounds relies not only

    on possessing enzymes with the appropriate catalytic activity but

    also on the expression of these enzymes in the presence of the

    nitroaromatic compound. The first example of a nitroaromatic-

    triggered lysR transcriptional activator was reported recently.8

    Two nitroarene dioxygenases were studied, nitrobenzene dioxyge-

    nase (NBDO) in Comomonas sp. strain JS765 and 2-nitrotoluene

    dioxygenase (2NTDO) from Acidovorax sp. strain JS42. They are

    under the control of identical LysR related proteins, NbzR and

    NtdR, homologous to the previously reported NahR.

    The work revealed the expression of 2NTDO and NBDO can

    be induced by a plethora of nitroaromatic compounds including

    nitrobenzene, 2-, 3- and 4-nitrotoluene, 2,4- and 2,6-dinitrotoluene

    and aminodinitrotoluenes in addition to the endogenous salicylate

    and anthranilate. The metabolism of these compounds was not

    necessary for induction suggesting the nitroaromatic compounds

    themselves are the actual effector molecules with the NtdR

    essential for induction. This represents a pathway in evolution.

    Enzymes capable of transforming nitroaromatic compounds are

    being organised to be expressed in response to the presence

    of these novel substrates, although at present the regulatory

    proteins still respond to their endogenous effectors salicylate and

    anthranilate.

    Nitrogen assimilationby bacteria has also been studied with ref-

    erence to TNT degradation. Recent work employing Pseudomonas

    putida JLR11 has shown that the nitrogen containing products

    released on TNT degradation are assimilated via the glutamine

    synthetaseglutamate synthase (GSGOGAT) pathway.9

    The metabolism of naturally occurring nitroaromatic

    compounds

    Very few naturally occurring nitroaromatic compounds are

    known. Recently, for the first time, the degradation pathway

    of one of these compounds, 3-nitrotyrosine, a product of the

    interaction between reactive nitrogen species and amino acids,

    was elucidated.10 Using 3-nitrotyrosine as sole growth substrate in

    soil enrichment cultures, multiple bacteria were isolated capable

    of catalysing the degradation of 3-nitrotyrosine with the release

    of nitrite and ammonia. Two species, Burkholderia sp. strain

    JS165 and Variovorax paradoxus JS171 were chosen for further

    study.

    The pathway for the breakdown of 3-nitrotyrosine was found

    to be both separate from the tyrosine degradation pathway and

    inducible in its entirety by cell growth on 3-nitrotyrosine or

    its breakdown product 4-hydroxy-3-nitro-phenylacetate (HNPA)

    (Fig. 2).

    The pathway starts with an a-ketoglutarate dependent deam-

    ination to produce the conjectural intermediate 4-hydroxy-

    3-nitrophenyl pyruvate. This is decarboxylated to yield HNPA,

    in turn denitrated to homoprotocatechuate by a monooxygenase.

    The homoprotocatechuate is cleaved by homoprotocatechuate-

    2,3-dioxygenase to yield 5-carboxymethyl-2-hydroxy-cis,cis-

    Fig. 2 The proposed pathway for the breakdown of the naturally occur-

    ring nitroaromatic compound 3-nitrotyrosine, adapted from Nishino and

    Spain.10 Compounds in brackets are proposed intermediate compounds.

    muconic semialdehyde, the metabolism of which has been

    thoroughly characterised. This pathway utilised for 3-nitrotyrosine

    metabolism has been demonstrated to be analogous to the tyro-

    sine degradation pathway, which also proceeds via homoproto-

    catechuate, and may have derived from the tyrosine pathway.

    Nitroreductases

    The enzymatic transformation of triply nitrated aromatic com-

    pounds, such as the explosives TNT and picric acid, is gener-

    ally limited to reductive transformations. Enzymes capable of

    catalysing such reactions are exemplified by the Enterobacter

    cloacae PB2 enzyme, PETN reductase (PETNr), a monomeric

    40 kDa flavoenzyme,11,12 which is capable of catalysing both the

    nitroreduction of TNT and the ring reduction, as illustrated in

    Fig. 3.

    PETNr was isolated through its ability to confer resistance

    to PETN, a nitrate ester explosive, on the bacterium.11 PETNr

    was found to be a member of the Old Yellow Enzyme family

    of flavoenzymes and was the first of a series of enzymes from

    This journal is The Royal Society of Chemistry 2006 Nat. Prod. Rep., 2006, 23, 845850 | 847

  • 8/9/2019 Bacterial Pathways for Degradation of tics

    4/6

    Fig. 3 The biotransformation of 2,4,6-trinitrotoluene as catalysed by

    PETN reductase from Enterobacter cloacae PB2.

    this family that have since proven capable of transforming

    explosives.1216 PETNr is capable of reductively liberating nitrite

    not only from PETN and nitroglycerine, but from nitroaromatic

    explosives such as TNT and picric acid. The nitroreduction of

    triply nitrated aromatics (Fig. 3, Pathway A) is more common.

    Many bacteria/enzymes have proven capable of opportunistic

    nitroreduction of nitroaromatic compounds. A recent example

    is that of PnrA,17 an NADPH dependent nitroreductase, ho-

    mologous to nfsA, the Escherichia coli nitroreductase, capable

    of transforming TNT to yield 4-hydroxylamino-2,6-DNT via a

    ping-pong bibi mechanism. It is capable of reducing multiple

    nitroaromatic substrates including 2,4-DNT, 3-nitrotoluene, 3-

    and 4-nitrobenzoate, 3,5-dinitrobenzamide and 3,5-dinitroaniline,

    although it was most efficient with TNT.

    The nitroreduction of TNT results in the production of

    metabolites that are still toxic. Hydroxylamino-dinitrotoluenes

    (HADNTs) damage DNA causing mutations18 and result in

    the redox cycling of the TNT derivatives with the concomitant

    production of harmful reactive oxygen species.19 This has been

    put to good use in the field of cancer therapy as discussed later.

    Interestingly, the nitroso intermediate inherent in this pathway

    has not been isolated. Recent work testing the nfsB encoded

    E. coli nitroreductase against nitroso- and nitro-toluene and

    benzene (Fig. 4) demonstrated that the nitrosoderivatives are

    far better substrates for the enzyme, having specificity constants

    2030 000 times that of the nitro-substrates.20 This offers an

    explanation as to why the nitrosoderivatives of these and other

    nitrocompounds are seldom seen in biotransformations.

    Fig. 4 The structures of nitro- and nitroso-toluene and benzene.

    Hydride transferases

    A second reductive biotransformationof TNT (Fig. 3, Pathway B)

    has been observed,which offers hope for the complete degradation

    of TNT. It is catalysed bysome members of theOld YellowEnzyme

    family16,21,22 including PETNr. This second pathway features the

    reduction of the aromatic ring of TNT. Aromatic transformations

    are intuitively associated with oxidative attack due to the high

    electron density of the benzene ring; however, the three NO2groups present in TNT are sufficiently electron-withdrawing to

    render the ring electron-poor and subject to reductive attack. The

    result is the production of hydride- and dihydride-Meisenheimer

    TNT complexes (H-TNT and 2H-TNT) and hence the loss

    of aromaticity. The ultimate products of the pathway are not

    known; however, the reduction of the dihydride-Meisenheimer

    TNT complexes has been confirmed to be enzyme-catalysed,23 and

    in the course of the transformation, nitrite is released. The final

    transformation products remain elusive, although recent work has

    characterised them as lacking visible absorbance and being highly

    polar. They contain less nitrogen than TNT, are non-aromatic and

    water soluble, all of which suggests that they willbe less recalcitrant

    to degradation and less toxic than both TNT and the products of

    the nitroreduction pathway.5,16

    A second example of such a hydride transfer reaction with a

    nitroaromatic compound comes in the form of picric acid degra-

    dation by Rhodococcus (opacus) erythropolis and Nocardiodes

    simplex FJ21A, reviewed in Heiss and Knackmuss24 (Fig. 5).

    Picric acid is more easily degraded than TNT. Substantial

    work characterising the entire pathway of picric acid degradation

    has been performed by the Knackmuss group in their work on

    Rhodococcus (opacus) erythropolis HL PM-1 and Nocardioides

    simplex FJ21A.2427 The initial reaction is hydride addition to

    the aromatic ring, to create a hydride Meisenheimer complex,

    as discussed with TNT; however, with picric acid no nitrore-

    duction is observed, only the hydride addition as catalysed

    b y a F420

    -dependent hydride transferase (NpdI) and its F420

    reductase (NpdG).25,27 A second hydride addition is catalysed by

    NpdC/NpdG in R. (opacus) erythropolis; whereas, in N. simplex,

    both hydride transfers are catalysed by the same enzyme.

    Following the production of the dihydride Meisenheimer com-

    plex it is hypothesised that the multiple products observed are due

    to the tautomerisation of the protonated dihydride Meisenheimer

    complex, catalysed by NpdH.27 The hydride Meisenheimer com-

    plex of 2,4-dinitrophenol has been identified as an intermediate

    of picric acid degradation in Nocardioides sp. strain CB 222,

    which would suggest that nitrite is released from the dihydride

    Meisenheimer complex or one of its protonated tautomers.24 Inter-

    estingly, R. erythropolis can also catalyse the ring hydrogenation

    848 | Nat. Prod. Rep., 2006, 23, 845850 This journal is The Royal Society of Chemistry 2006

  • 8/9/2019 Bacterial Pathways for Degradation of tics

    5/6

    Fig. 5 The transformation of picric acid by Rhodococcus (opacus)

    erythropolis HLPM1, adapted from Heiss and Knackmuss.24

    of TNT, similar to the reaction catalysed by PETNr; however,

    unlike PETNr, in spite of the production of 2H-TNT, no nitrite

    release is observed.28

    Bacterial enzymes applied

    It seems that the most widespread source of nitroaromatic pollu-

    tion is from the explosives industry, either through the active use

    of explosives or through the manufacture and decommissioning of

    munitions. Having characterised bacteria capable of transforming

    and degrading explosives it seems a logical next step to use these

    bacteria to remove the toxic explosives from soil. There are,

    however, disadvantages in using bacteria for this purpose. Firstly,

    when in competition with the local bacterial ecosystem, in non-

    optimal growth conditions, there may be a significant drop in

    the biomass of the applied bacteria. Equally, co-metabolism can

    necessitate the addition of extra substrates to induce the enzyme

    required for pollutant transformation.29 Additionally, although

    many strains of explosives-detoxifying microorganisms have been

    isolated from contaminated soil, pollution remains a problem

    at these sites. This can be overcome using genetic engineering;

    however, overwhelming anti-GMO public opinion makes such

    a course of action difficult with genetically modified microbes.

    Considerable research has been devoted to the phytoremediation

    of explosives, that is usingplants to biotransform explosives. Plants

    are uniquely suited to the removal of explosives contamination

    due to their penetrating root systems, comparably high biomass

    and compatibility with restorative ecology. Original work sought

    to employ genetically unmodified plant species in this process;

    however, major disadvantages were the inherently low levels of

    detoxification coupled with the susceptibility of these plants to

    the toxicity of the explosives. For phytoremediation to be a com-

    petitive alternative to current methods of explosives degradation,

    plants need to efficiently take up the contaminants and break

    them down or transform them to less toxic intermediates. This has

    been achieved by the constitutive expression of bacterial enzymes

    in plants.30 The expression of both nitroreductase and PETN

    reductase from Enterobacter cloacae PB2 has been achieved in

    arabidopsis and tobacco plants.The resultant PETNr plants could

    survive in concentrations of GTN and TNT (1 mM and 0.05 mM

    respectively) that were toxic to genetically unmodified tobacco

    plants.30 Similarly, aromatic nitroreductases, when expressed in

    plants, provided even greater resistance to the toxic effects of

    TNT.31

    Bacterial enzymes are similarly being used outside their natural

    environment in novel cancer therapies. Most particularly E. coli

    oxygeninsensitive nitroreductase (nfsB)is beingused to activate ni-

    troaromatic compounds, such as the dinitrobenzamide prodrugs,

    to their cytotoxic hydroxylamino derivatives to destroy tumours.

    Mammalian nitroreductases do not catalyse the activation of these

    compounds, so the bacterial nitroreductases, when targeted to

    the tumour using ADEPT (antibody directed enzyme prodrug

    therapy) or GDEPT (gene-directed enzyme prodrug therapy),

    exclusively activate the compounds in the vicinity of cancerous

    growth. Recent determination of the nitroreductase structure in

    complex with a series of these dinitrobenzamide drugs now paves

    the way not only for the rational design of novel prodrugs but

    also for the rational design of the enzyme specifically for cancer

    therapies.32

    Conclusions

    An understanding of the bacterial transformation of nitroaro-

    matic compounds has clear benefits for bioremediation research,

    medicine and industry. In addition to a more detailed understand-

    ing of how the initial nitroaromatic oxidases or reductases operate,

    an enhanced appreciation of the role of enzymes that function

    downstream of these initial attack enzymes will facilitate the

    optimisation of remediation pathways through the upregulation ofkey downstream metabolic enzymes. In addition to a contribution

    to the field of bioremediation, the elucidation of dissimilation

    pathways for nitroaromatic compounds is likely to continue to

    furnish a rich source of enzymes as biocatalysts for the synthesis

    of high value chemicals and drugs.

    References

    1 J. Yinon, Toxicity & Metabolism of Explosives, CRC Press Inc., BocaRaton, Florida, USA, 1990.

    2 B. S. Hooker and R. S. Skeen, Nat. Biotechnol., 1999, 17, 428.3 T. Gorontzy, O. Drzyzga, M. W. Kahl, D. Bruns-Nagel, J. Breitlung, E.

    von Loew and K.-H. Blotevogel, Crit. Rev. Microbiol., 1994, 20, 265.

    This journal is The Royal Society of Chemistry 2006 Nat. Prod. Rep., 2006, 23, 845850 | 849

  • 8/9/2019 Bacterial Pathways for Degradation of tics

    6/6

    4 A. Esteve-Nunez, A. Caballero and J. L. Ramos, Microbiol. Mol. Biol.Rev., 2001, 65, 335.

    5 A. Basran, C. E. French, R. E. Williams, S. Nicklin and N. C. Bruce,Biochem. Soc. Trans., 1998, 26, 680.

    6 M. M. Watrous, S. Clark, R. Kutty, S. Huang, F. B. Rudolph, J. B.Hughes and G. N. Bennett, Appl. Environ. Microbiol., 2003, 69, 1542.

    7 G. R. Johnson, R. K. Jainand J. C.Spain, J. Bacteriol., 2002, 184, 4219.8 D. J. Lessner, R. E. Parales, S. Narayan and D. T. Gibson, J. Bacteriol.,

    2003, 185, 3895.9 A. Caballero, A. Esteve-Nunez, G. J. Zylstra and J. L. Ramos,

    J. Bacteriol., 2005, 187, 396.10 S. F. Nishino and J. C. Spain, Appl. Environ. Microbiol., 2006, 72, 1040.11 P. R. Binks, C. E. French, S. Nicklin and N. C. Bruce, Appl. Environ.

    Microbiol., 1996, 62, 1214.12 C. E. French, S. Nicklin and N. C. Bruce, J. Bacteriol., 1996, 178, 6623.13 J. R. Snape, N. A. Walkley, A. P. Morby, S. Nicklin and G. F. White,

    J. Bacteriol., 1997, 179, 7796.14 D. S. Blehert, B. G. Fox and G. H. Chambliss, J. Bacteriol., 1999, 181,

    6254.15 T. B. Fitzpatrick, N. Amrhein and P. Macheroux, J. Biol. Chem., 2003,

    278, 19891.16 R. E. Williams, D. A. Rathbone, N. S. Scrutton and N. C. Bruce, Appl.

    Environ. Microbiol., 2004, 70, 3566.17 A. Caballero, J. J. Lazaro, J. L. Ramos and A. Esteve-Nunez, Environ.

    Microbiol., 2005, 7, 1211.18 S. Homma-Takeda, Y. Hiraku, Y. Ohkuma, S. Oikawa, M. Murata, K.

    Ogawa, T. Iwamuro, S. Li, G. F. Sun, K. Kumagai, N. Shimojo and S.Kawanishi, Free Radical Res., 2002, 35, 555.

    19 J. Sarlauskas, A. Nemeikaite-Ceniene, Z. Anusevicius, L. Miseviciene,M. M. Julvez, M. Medina, C. Gomez-Moreno and N. Cenas, Arch.Biochem. Biophys., 2004, 425, 184.

    20 P. R. Race, A. L. Lovering, R. M. Green, A. Ossor, S. A. White, P. F.Searle, C. J. Wrighton and E. I. Hyde, J. Biol. Chem., 2005, 280, 13256.

    21 C. E. French, S. Nicklin and N. C. Bruce, Appl. Environ. Microbiol.,1998, 64, 2864.

    22 J. W. Pak, K. L. Knoke, D. L. Noguera,B. G. Foxand G. H. Chambliss,Appl. Environ. Microbiol., 2000, 66, 4742.

    23 H. Khan, T. Barna, R. J. Harris, N.C. Bruce, I. Barsukov, A. W. Munro,P. C. E. Moody and N. S. Scrutton, J. Biol. Chem., 2004, 279, 30563.

    24 G. Heiss and H.-J. Knackmuss, Curr. Opin. Microbiol., 2002, 5,282.

    25 S. Ebert, P.-G. Rieger and H.-J. Knackmuss, J. Bacteriol., 1999, 181,2669.

    26 P.-G. Rieger, V. Sinnwell, A. Preu, W. Francke and H.-J. Knackmuss,J. Bacteriol., 1999, 181, 1189.

    27 G. Heiss, K. W. Hofmann, N. Trachtmann, D. M. Walters, P. Rouvieand H.-J. Knackmuss, Microbiology, 2002, 148, 799.

    28 C. Vorbeck, H. Lenke, P. Fischer, J. C. Spain and H.-J. Knackmuss,Appl. Environ. Microbiol., 1998, 64, 246.

    29 M. Vidali, Pure Appl. Chem., 2001, 73, 1163.30 C. E. French, S. J. Rosser, G. J. Davies, S. Nicklin and N. C.Bruce, Nat.

    Biotechnol., 1999, 17, 491.31 N. Hannink, S. J. Rosser, C. E. French, A. Basran, J. A. Murray, S.

    Nicklin and N. C. Bruce, Nat. Biotechnol., 2001, 19, 1168.

    32 E. Johansson, G. N. Parkinson, W. A. Denny and S. Neidle, J. Med.Chem., 2003, 46, 4009.

    850 | Nat. Prod. Rep., 2006, 23, 845850 This journal is The Royal Society of Chemistry 2006